ArticlePDF AvailableLiterature Review

Abstract and Figures

Chemerin is a small chemotactic protein and a key player in initiating the early immune response. As an adipokine, chemerin is also involved in energy homeostasis and the regulation of reproductive functions. Secreted as inactive prochemerin, it relies on proteolytic activation by serine proteases to exert biological activity. Chemerin binds to three distinct G protein-coupled receptors (GPCR), namely chemokine-like receptor 1 (CMKLR1, recently named chemerin 1 ), G protein-coupled receptor 1 (GPR1, recently named chemerin 2 ), and CC-motif chemokine receptor-like 2 (CCRL2). Only CMKLR1 displays conventional G protein signaling, while GPR1 only recruits arrestin in response to ligand stimulation, and no CCRL2-mediated signaling events have been described to date. However, GPR1 undergoes constitutive endocytosis, making this receptor perfectly adapted as decoy receptor. Here, we discuss expression pattern, activation, and receptor binding of chemerin. Moreover, we review the current literature regarding the involvement of chemerin in cancer and several obesity-related diseases, as well as recent developments in therapeutic targeting of the chemerin system.
Content may be subject to copyright.
Review
Tobias F. Fischer and Annette G. Beck-Sickinger*
Chemerin exploring a versatile adipokine
https://doi.org/10.1515/hsz-2021-0409
Received November 6, 2021; accepted December 23, 2021;
published online January 18, 2022
Abstract: Chemerin is a small chemotactic protein and a key
player in initiating the early immune response. As an adi-
pokine,chemerin is also involved in energy homeostasis and
the regulation of reproductive functions. Secretedas inactive
prochemerin, it relies on proteolytic activation by serine
proteases to exert biological activity. Chemerin binds to
three distinct G protein-coupled receptors (GPCR), namely
chemokine-like receptor 1 (CMKLR1, recently named chem-
erin
1
), G protein-coupled receptor 1 (GPR1, recently named
chemerin
2
), and CC-motif chemokine receptor-like 2 (CCRL2).
Only CMKLR1 displays conventional G protein signaling,
while GPR1 only recruits arrestin in response to ligand
stimulation, and no CCRL2-mediated signaling events have
been described to date. However, GPR1 undergoes consti-
tutive endocytosis, making thisreceptor perfectly adaptedas
decoy receptor. Here, we discuss expression pattern, acti-
vation, and receptor binding of chemerin. Moreover, we re-
view the current literature regarding the involvement of
chemerin in cancer and several obesity-related diseases, as
well as recent developments in therapeutic targeting of the
chemerin system.
Keywords: adipose tissue; cytokine; immunity; inflam-
mation; obesity.
Expression and tissue distribution
of chemerin
In 1997, Chandraratna and colleagues reported the identi-
fication of a novel gene in psoriatic skin lesions. Expression
of this gene was upregulated in response to treatment
with the anti-psoriatic retinoic acid tazarotene, and it was
therefore named tazarotene-induced gene 2 (TIG2) or reti-
noid acid responder protein 2 (RARRES2) (Nagpal et al.
1997). After identifying the TIG2 gene product as the nat-
ural ligand of CMKLR1, Wittamer et al. (2003) named the
protein chemerin.
Chemerin is expressed as a biologically inactive pro-
form named preprochemerin, consisting of 163 amino
acids. Upon truncation of the 20 amino acid N-terminal
signal peptide, it is secreted as prochemerin, which is the
main form of chemerin found in the circulation (Chang
et al. 2016; Wittamer et al. 2003).
The highest expression of chemerin mRNA in mice has
been detected in the liver, which is most likely the main
source of circulating chemerin in lean individuals (Banas
et al. 2015; Ernst et al. 2010; Weigert et al. 2010). Chemerin
is also found in the skin, where it is expressed in the
epidermis (Banas et al. 2015; Schultz et al. 2013), and in the
lung (Luangsay et al. 2009). Articular chondrocytes, a
specialized cell type found in joints, have been demon-
strated to express chemerin (Berg et al. 2010), and in-
creased chemerin levels are found in the synovial uid of
inamed joints (Wittamer et al. 2003). Moreover, chemerin
mRNA is found in the uterus, ovary, testes, colon, and
spleen of mice (Luangsay et al. 2009).
On the protein level, chemerin displays the highest
expression in white adipose tissue (Banas et al. 2015).
Expression of chemerin in murine adipose tissue has been
conrmed by Roh et al. (2007), who additionally demon-
strated that the level of chemerin expressed by the murine
pre-adipocyte cell line 3T3-L1 increases during their dif-
ferentiation into mature adipocytes. Using murine mesen-
chymal stromal cell-derived adipocytes, Dranse et al. (2016)
found the same results, showing increasing expression of
chemerin over the course of differentiation. Consequently,
chemerin is also found at high concentrations in human
adipose tissue, as demonstrated by Chang et al. (2016).
Structure and proteolytic
processing of chemerin
Chemerin is a protein without any known close sequence
homolog, and no crystal structure has been published to
date. Structurally, the closest homolog as determined by
*Corresponding author: Annette G. Beck-Sickinger, Institute of
Biochemistry, University of Leipzig, Brüderstraße 34, D-04103 Leipzig,
Germany, E-mail: abeck-sickinger@uni-leipzig.de.
https://orcid.org/0000-0003-4560-8020
Tobias F. Fischer, Institute of Biochemistry, University of Leipzig,
Brüderstraße 34, D-04103 Leipzig, Germany
Biol. Chem. 2022; aop
fold recognition is the cathelin-like domain of the human
cathelicidin LL-37, a member of the innate immune system
(Kelley et al. 2015; Pazgier et al. 2013). LL-37 acts as cationic
antimicrobial protein, and is chemotactic for monocytes,
mast cells, neutrophils, and T-cells (Zanetti 2004). Both
chemerin and LL-37 share structural homology with
cystatins, a large family of cysteine protease inhibitors
(Rawlings and Barrett 1990).
The first structural characterization of chemerin was
an NMR assignment by Allen et al. (2007). They predicted a
central β-sheet anked by two α-helices, with an unstruc-
tured, exible C-terminus, and postulated the presence of
three disulde bridges. Employing a combination of ho-
mology modeling and de novo protein structure prediction,
Schultz et al. (2013) constructed a model of prochemerin
(Figure 1), which closely matches the results from these
NMR studies and is in close agreement with the recently
published structural model by AlphaFold (Jumper et al.
2021).
Prochemerin relies on C-terminal protease cleavage to
exert biological functions through the chemokine-like re-
ceptor 1 (CMKLR1). Incorporating mutagenesis data with
computational modeling, Schultz et al. (2013) showed that
this protease-mediated activation cleaves off a C-terminal
helical segment, resulting in a exible C-terminus. The
most active chemerin species terminate at S157 or F156 and
accordingly are named ChemS157 or ChemF156, respec-
tively (Schultz et al. 2013). Short peptides derived from the
C-termini of these two species are sufcient to trigger
activation at the CMKLR1 and the second chemerin receptor
GPR1 (de Henau et al. 2016; Wittamer et al. 2004). A slightly
longer chemerin form, ChemK158, is signicantly less
active at the CMKLR1, while chemerin species lacking F156,
i.e., ChemA155 and shorter variants, are completely inac-
tive (Wittamer et al. 2004; Yamaguchi et al. 2011). Inter-
estingly, ChemA155 even shows weak antagonism at
CMKLR1, highlighting the complex interplay between the
different chemerin species (Yamaguchi et al. 2011).
Several proteases are known that can process proche-
merin, many of which are present in blood plasma: Two
enzymes of the coagulation cascades, factors Xa and VIIa
both generate active chemerin from prochemerin, offering a
link between tissue damage and chemerin activation (Zabel
et al. 2005). Plasmin, an enzyme of the brinolytic cascade,
generates the biologically inactive ChemK158, which can
then be converted to active ChemS157 bycarboxypeptidase
N, a constitutively active serine protease (Du et al. 2009).
These results suggest that active chemerin is generated
during the process of blood clotting and de-clotting after
blood vessel damage. Apart from these proteases, immune
cell-derived enzymes can activate and de-activate chem-
erin: Two proteases released from neutrophils, elastase and
cathepsin G, can generate active ChemS157 from proche-
merin, while the neutrophil-derived enzyme proteinase 3
converts prochemerin to inactive ChemA155 (Guillabert
et al. 2008; Wittamer et al. 2005; Zabel et al. 2005). Addi-
tionally, chymase, an enzyme secreted by mast cells, in-
activates ChemS157 by cleavage of the two ultimate amino
acids, generating ChemF154 (Guillabert et al. 2008). Next to
these plasma and immune cell-derived proteases, kalli-
krein (KLK) 7 is the only tissue-specic enzyme reported
to process chemerin. Predominantly found in the skin,
it converts prochemerin to active ChemF156 (Schultz
et al. 2013).
Figure 1: Structure and sequence of human chemerin.
(A) The homology model of human chemerinS157 displays a central β-sheet flanked by two α-helices (Schultz et al. 2013). The three putative
disulde bridges are highlighted as yellow spheres. (B) Primary sequence of human preprochemerin, the N-terminal signal peptide is
highlighted with gray background, disulde bridges are indicated by square brackets. The C-terminal peptide chemerin-9 is highlighted.
2T.F. Fischer and A.G. Beck-Sickinger: Chemerin exploring a versatile adipokine
While prochemerin is by far the most abundant
chemerin form in blood plasma, the active ChemS157 is
dominant in adipose tissue (Chang et al. 2016). This may be
due to the co-expression of prochemerin and the chemerin-
activating enzymes elastase and tryptase by adipocytes, as
demonstrated by Parlee et al. (2012). An overview over post-
secretary processing of chemerin is given in Figure 2 and by
Mattern et al. (2014).
Taken together, these results demonstrate that inac-
tive prochemerin is permanently circulating in the blood-
stream, waiting to be activated upon tissue damage or
inflammation. Skin and adipose tissue may play a special
role, as they can be expected to dispose active chemerin
species independent of triggering events.
The chemerin receptors
Chemerin binds to three distinct G protein-coupled re-
ceptors (GPCR), and each of these chemerin receptors has
a distinct pharmacological prole. CMKLR1, recently also
named chemerin receptor 1 (chemerin
1
) was the rst re-
ceptor reported to bind to chemerin in 2003 (Kennedy and
Davenport 2018; Wittamer et al. 2003). G protein-coupled
receptor 1 (GPR1), the closest sequence homolog of CMKLR1,
and recently named chemerin receptor 2 (chemerin
2
) was
identied as a receptor for chemerin shortly after (Barnea
et al. 2008; Kennedy and Davenport 2018). CMKLR1 and
GPR1 share the highest sequence identities with the
formyl peptide receptors (Figure 3A), pattern-recognizing
G
i
-coupled receptors that are activated by a wide range of
formylated bacterial peptides and are involved in host de-
fense (Chen et al.; Weiß and Kretschmer 2018; Zhuang
et al.). However, in an in-depth phylogenetic analysis, Fre-
drikson et al.grouped GPR1 and CMKLR1 with the receptors
for the complement 5a (C5a) anaphylatoxin (Fredriksson
et al. 2003). C5a receptor 1 (C5aR1) and C5a receptor 2 (C5aR2,
formerly C5L2) play an important role in complement-
mediated innate immunity (Pandey et al. 2020) and show
functional similarities to GPR1 and CMKLR1, which will be
discussed in the following sections.
In contrast to CMKLR1 and GPR1, CC-motif chemokine
receptor-like 2 (CCRL2), the third chemerin receptor, is
grouped with the chemokine receptors (Figure 3A) (Fre-
driksson et al. 2003). However, no chemokine ligands for
CCRL2 have been described to date, making chemerin the
only known ligand for CCRL2 so far (Zabel et al. 2008).
CMKLR1
CMKLR1 or chemerin
1
is a receptor expressed by several
types of immune cells: Macrophages and dendritic
cells express CMKLR1 and display CMKLR1-dependent
migration towards increasing concentrations of chemerin
(Wittamer et al. 2003). Natural killer (NK) cells also express
CMKLR1 and decreased expression of this receptor has
been associated with reduced migration towards chemerin
(Parolini et al. 2007). Apart from these immune cells,
CMKLR1 expression has also been detected in human and
murine adipocytes, and RNA-targeted knockdown of this
receptor resulted in an impaired differentiation of 3T3-L1
Figure 2: Proteolytic activation and deactivation of chemerin by immune cell-derived (mast cell and neutrophil-secreted), plasma, or tissue-
specific proteases.
Differently processed chemerin species vary in the positions of their respective C-termini.
T.F. Fischer and A.G. Beck-Sickinger: Chemerin exploring a versatile adipokine 3
cells into mature adipocytes (Goralski et al. 2007). More-
over, CMKLR1 is expressed by primary human hepatocytes
and vascular endothelial cells (Döcke et al. 2013; Kennedy
et al. 2016).
Studies with CMKLR1 knock-out mice came to con-
flicting results: Rouger et al. (2013) describe a higher fat
mass and body weight, but unaltered glucose tolerance
for CMKLR1/mice compared to the control group. In
contrast, Ernst et al. (2012) report a lower food intake and
body weight, combined with lower glucose tolerance and
insulin secretion for CMKLR1/mice compared to the
control group, independent of diet. Moreover, Ernst et al.
(2012) describe decreased levels of the pro-inammatory
cytokines tumor necrosis factor (TNF) αand interleukin (IL)
6 in white adipose tissue (WAT) of CMKLR1/mice fed a
high-fat diet compared to the control group, which is not
conrmed by the results by Rouger et al. (2013).
Chemerin was independently discovered as a natural
ligand for CMKLR1 by two different groups (Meder et al.
2003; Wittamer et al. 2003): Wittamer et al. (2003) were the
rst to engage in detailed structure-activity studies of this
interaction, revealing that the agonistic properties of
chemerin are preserved in its C-terminus. Employing
radioligand-binding and Ca2+mobilization assays, they
demonstrated that the C-terminal aromatic residues, and
especially F156, are crucial for binding to and activation
of CMKLR1 (Wittamer et al. 2004). They identied the
C-terminal nonapeptide as the minimal activation motif
and named it chemerin-9. This peptide has since been used
since then as a tool compound for numerous studies
(Bandholtz et al. 2012; de Henau et al. 2016; Peyrassol et al.
2016; Shimamura et al. 2009; Tu et al. 2020). Recently, we
further characterized the interaction between chemerin-9
and CMKLR1 and demonstrated that chemerin-9 interacts
with several conserved aromatic residues in the extracel-
lular loops of CMKLR1, folding into a hairpin conformation
to induce receptor activation (Fischer et al. 2021c). Thus,
the terminal phenylalanine in chemerin-9 (F8, corre-
sponding to F156 in the full-length protein) binds to a hy-
drophobic patch in the second extracellular loop (ECL2) of
CMKLR1, while F6 (corresponding to F154 in the full-length
protein) interacts with Y2.63 and Y2.68 in the extracellular
domain of transmembrane helix 2 and ECL1, respectively
(nomenclature according to Ballesteros and Weinstein
[1995]). This results in a binding mode where both termini
of the ligand face the extracellular solvent, while the cen-
tral stretch of amino acids reaches deep down into the
transmembrane bundle (Figure 4).
Based on experiments using nanobodies directed
against CMKLR1, Peyrassol et al. (2016) hypothesized that
chemerin binds to CMKLR1 similar to the two-step binding
model postulated for chemokines (Thiele and Rosenkilde
2014). According to this hypothesis, the cystatin motif of
chemerin interacts with the extracellular parts of the re-
ceptor, positioning the chemerin C-terminus in the trans-
membrane bundle to trigger activation of the receptor. This
Figure 3: Chemerin binds to the three GPCR CMKLR1, GPR1, and CCRL2.
(A) Phylogenetic tree using a sequence alignment of selected human protein, peptide, and fatty acid receptors from the γbranch of rhodopsin-
like GPCR generated by GPCRdb. The dendrogram was calculated using the Clustal Omega simple phylogenytool and visualized in R using
the ggtree library (Yu 2020). (B) Activation of CMKLR1 and GPR1 is mediated through the chemerin C-terminus, which does not bind to CCRL2.
Instead, chemerin might bind to CCRL2 exclusively through the cystatin domain.
4T.F. Fischer and A.G. Beck-Sickinger: Chemerin exploring a versatile adipokine
binding mode could be similar to the interaction of C5aR1
with its ligand C5a (Siciliano et al. 1994).
The anti-inflammatory lipid mediator resolvin E1
(RvE1) has been described as an additional ligand for
CMKLR1 (Arita et al. 2005; Ohira et al. 2010). However,
while an independent group has reported RvE1-induced
arrestin 3 recruitment (Krishnamoorthy et al. 2010), others
could not conrm binding of RvE1 to CMKLR1 (Kennedy
and Davenport 2018; Luangsay et al. 2009). Unable to
reproduce RvE1 binding to CMKLR1, Bondue et al. (2011)
suggest that instead of CMKLR1 the related leukotriene B4
receptor 1 (BLTR1) is the actual target of RvE1.
CMKLR1 couples to Gα
i
proteins, which inhibit the
synthesis of cyclic adenosine monophosphate (cAMP) by
adenylyl cyclases (Baltoumas et al. 2013; de Henau et al.
2016). Moreover, stimulation with chemerin leads to the
recruitment of arrestin 3 and subsequent internalization
of the receptor (de Henau et al. 2016; Zhou et al. 2014). In
inammatory macrophages, this internalization is regu-
lated by G protein-coupled receptor kinase (GRK) 6 and
arrestin 3 (Seran et al. 2019). Consequently, GRK6 and
arrestin 3-decient macrophages display prolonged
CMKLR1-signaling and increased migration towards
increasing concentrations of chemerin (Seranetal.
2019). Further downstream, activation of CMKLR1 by
chemerin leads to phosphorylation of the p44/p42
mitogen-activated protein kinases (MAPK) ERK1/2, p38
MAPK, and Akt in chondrocytes (Berg et al. 2010), he-
patocytes (Feder et al. 2020), human umbilical vein
endothelial cells (HUVEC) (Nakamura et al. 2018), human
granulosa cells (Reverchon et al. 2012), and in broblast-
like synovial cells (Kaneko et al. 2011). De Henau et al.
(2016) demonstrated that CMKLR1-mediated ERK1/2
phosphorylation is Gα
i/o
and arrestin 3-dependent.
GPR1
GPR1 or chemerin
2
was identied as a receptor for chemerin
by Barnea et al. (2008). Initially detected in the human
hippocampus, GPR1 mRNA has since been found in murine
ovaries, muscle, and white adipose tissue (Marchese et al.
1994; Rourke et al. 2014; Yang et al. 2016). In the latter,
GPR1 mRNA was detected only in the stromal vascular
Figure 4: According to our previously published model, chemerin-9 (red) binds to CMKLR1 (gray) through distinct aromatic and hydrophobic
sites, adopting a loop conformation.
(A) Overall conformation of chemerin-9 in the CMKLR1 ligand-binding pocket.
(B) F8 binds to a hydrophobic patch in ECL2 consisting of A4.67,L
4.69,F
4.76, and F4.79. (C) F6 interacts with Y2.63 and Y2.68 in TM2 and ECL1,
respectively.
T.F. Fischer and A.G. Beck-Sickinger: Chemerin exploring a versatile adipokine 5
fraction (SVF), but not in adipocytes (Rourke et al. 2014).
Interestingly, while GPR1 expression generally has not
been demonstrated in immune cells, tissue-specic alve-
olar macrophages express GPR1, suggesting that GPR1
expression may be induced when macrophages have
migrated into the lung (Farzan et al. 1997). This could also
be the case for immune cells present in adipose tissue,
explaining the detection of GPR1 in SVF (Rourke et al.
2014). Additionally, expression of GPR1 was also demon-
strated in murine and human skin (Banas et al. 2015).
In a recent study by Rourke et al. (2014), knock-out of
GPR1 in mice had no inuence on body weight, fat mass, or
energy expenditure, but led to decreased insulin sensitivity
and exacerbated glucose intolerance in animals fed a high-
fat diet. Unfortunately, no parameters of inammation
were reported, the role of GPR1 in inammation remains
unclear.
Chemerin displays a higher affinity to GPR1 than for
CMKLR1, which also translates into higher potency of
chemerin at GPR1 (Barnea et al. 2008; de Henau et al. 2016).
As described for CMKLR1, chemerin-9 is also sufcient to
induce receptor activation at GPR1, and the overall binding
mode of chemerin-9 is shared between these two receptors
(Barnea et al. 2008; Fischer et al. 2021a; de Henau et al.
2016). Several peptides have been proposed as additional
ligands for GPR1 recently: Osteocrin, cholecystokinin,
gastrin-releasing peptide, and the neuropeptide FAM19A1
(Foster et al. 2019; Zheng et al. 2018). Moreover, previous
results from our group showed that GPR1 undergoes
constitutive internalization, enabling receptor-mediated
endocytosis of peptides that are unable to activate GPR1.
These results suggest that GPR1 may not only act as a re-
ceptor for chemerin, but also as a broad-range decoy re-
ceptor for other peptides (Fischer et al. 2021a). However, the
afnity of GPR1 is highest to chemerin (Foster et al. 2019).
GPR1 is an unusual GPCR and it is unclear whether it
may induce G protein signaling. While Barnea et al. (2008)
report weak Ca2+ux in cells transfected with GPR1 and the
promiscuous G protein Gα
15
upon stimulation with 1 µM
chemerin-9, others did not detect any G protein recruitment
to GPR1 (de Henau et al. 2016; Rourke 2015). However,
activated GPR1 rapidly recruits arrestin, followed by sub-
sequent internalization (de Henau et al. 2016). It was sus-
pected that the altered DR3.50Y motif (DHY in GPR1) may be
responsible for the lack of G protein signaling, however,
restoring the canonical DRY motif did not restore G protein
signaling (Rourke 2015). Interestingly, the related C5aR2
shows similar characteristics: This receptor rapidly recruits
arrestin and internalizes in response to stimulation with
C5a, but does not signal through G proteins (Li et al. 2019b).
It is generally assumed that C5aR2 acts as a decoy receptor,
regulating inammation by scavenging C5a (Scola et al.
2009). A similar role may be conceivable for GPR1, a hy-
pothesis that is supported by recent ndings demonstrating
activation-independent, constitutive internalization as an
additional mechanism for GPR1 (Fischer et al. 2021a).
Downstream of arrestin, activation of GPR1 leads to
signaling through the RhoA/ROCK pathway, while no
activation of the MAPK/ERK pathway has been reported
(Rourke et al. 2015). The absence of ERK signaling would
agree with current ndings that this pathway cannot be
activated in the absence of active G proteins (Grundmann
et al. 2018). Thus, it is clear that GPR1 does not display
conventional GPCR signaling characteristics, but it re-
mains debated to which extent signaling events triggered
by this receptor play a role in a physiological context. In the
light of recent results, a role for GPR1 as a decoy receptor for
chemerin, and potentially other peptides, seems likely.
CCRL2
CCRL2, the third receptor for chemerin, is expressed on
a wide range of immune cells, i.e., neutrophils, B and T
lymphocytes, macrophages, NK cells, mast cells, and
dendritic cells (Auer et al. 2007; Catusse et al. 2010; Galli-
gan et al. 2004; Hartmann et al. 2008; Migeotte et al. 2002;
Otero et al. 2010; Patel et al. 2001; Zabel et al. 2008). In
most of these cases, CCRL2 expression is upregulated in
response to inammatory stimuli such as lipopolysaccha-
rides (LPS), interferon (IFN) γ, or TNFα(Schioppa et al.
2020). Expression of CCRL2 has also been demonstrated in
skin (Banas et al. 2015) and a range of endothelial cells,
including inamed bronchial epithelium (Oostendorp et al.
2004), murine lung endothelial cells, and human primary
endothelial cells (Monnier et al. 2012). Moreover, CCRL2 is
expressed in adipocytes (Muruganandan et al. 2010), he-
patic stellate cells (Zimny et al. 2017), and different tumors
including breast and prostate cancer (Reyes et al. 2017;
Sarmadi et al. 2015).
Consequently, CCRL2 knock-outmice display disturbed
inflammatory responses: However, while Regan-Komito
et al. (2017) report an exaggerated immune response for
CCRL2-decient mice in a zymosan-induced model of in-
ammation, others demonstrate a protective effect of CCRL2
loss in models of inammatory arthritis (Del Prete et al.
2017).
Strikingly, no signaling events initiated by CCRL2 have
been detected so far, although it has been confirmed by
several groups that chemerin binds to CCRL2 (de Henau
et al. 2016; Mazzotti et al. 2017; Zabel et al. 2008). Neither
Ca2+mobilization in HEK293 or murine mast cells nor G
6T.F. Fischer and A.G. Beck-Sickinger: Chemerin exploring a versatile adipokine
protein or arrestin recruitment in CHO cells are triggered by
stimulation with chemerin (de Henau et al. 2016). CCRL2
displays slow, constitutive internalization, which is, how-
ever, not altered by the addition of chemerin (Mazzotti et al.
2017). In contrast to the two other chemerin receptors
CMKLR1 and GPR1, CCRL2 does not interact with the
C-terminus of chemerin (de Henau et al. 2016; Zabel et al.
2008). Instead, it is suspected that CCRL2 binds to the
cystatin motif of chemerin through its N-terminus (Zabel
et al. 2008).
An important additional mechanism by which CCRL2
contributes to the orchestration of the immune response is
by heterodimerization with the chemokine receptor CXCR2.
These CCRL2/CXCR2 heterodimers have been shown to
fine-tune neutrophil migration, a concept that is emerging
for an increasing number of chemokine receptors (Del Prete
et al. 2017; Martínez-Muñoz et al. 2018).
Based on these findings, the current understanding is
that CCRL2 contributes to chemerin-mediated chemotaxis
by elevating local concentrations of chemerin, thereby
forming stable membrane-bound chemerin gradients. The
unique interaction of CCRL2 with the cystatin domain of
chemerin suggests that the chemerin C-terminus is free to
activate CMKLR1 when bound to CCRL2, making CCRL2 a
presenting receptor (Schioppa et al. 2020).
The physiological role of chemerin
in healthy individuals
Since the initial discovery of chemerin in 1997, the impor-
tance of this protein to regulate the immune response
has been conrmed by a multitude of different studies
(Ernst and Sinal 2010; Nagpal et al. 1997; Zabel et al. 2006).
As the chemerin receptors are closely related to other
inammation-regulating GPCR and the closest structural
homolog of chemerin itself is involved in innate immunity,
this may be the most conserved evolutionary role of the
chemerin system.
In the skin, the first barrier against infection, chem-
erin, its receptors, and activating proteases are highly
expressed (Banas et al. 2015; Schultz et al. 2013). Upon
bacterial infection, chemerin can be processed into peptide
fragments with direct antimicrobial activity, offering a rst
defense mechanism against infections (Banas et al. 2013).
The same has been demonstrated for the oral cavity, a
major entry point for pathogens (Godlewska et al. 2017). In
the blood stream, inactive prochemerin is constantly
circulating, and is activated upon tissue damage by en-
zymes of the coagulation cascade, or upon inammation by
neutrophil and mast cell-secreted proteases (Ge et al. 2018;
Zabel et al. 2005). With the help of CCRL2-expressing
endothelial cells and neutrophils, a stable chemotactic
chemerin gradient is formed, attracting CMKLR1-expressing
dendritic cells, NK cells, and macrophages (Parolini et al.
2007; Schioppa et al. 2020; Wittamer et al. 2003). These
leukocytes either directly engage in phagocytosis of the
pathogen (macrophages), or induce apoptosis of virus-
infected or tumorous cells (NK cells). Moreover, as antigen-
presenting cells, dendritic cells and macrophages act by
aiding in the initiation of the adaptive immune response
(Kashem et al. 2017). These results demonstrate that
chemerin plays a critical role in early immunity.
Reproductive functions have emerged as a further field
regulated by chemerin. The two active chemerin receptors
CMKLR1 and GPR1 were both found in the ovary of mice (Li
et al. 2014a). Plasma levels of chemerin undergo signicant
changes during normal pregnancy in humans and rats,
where chemerin levels decrease towards the end of preg-
nancy, but are signicantly increased in pregnant women
with preeclampsia (Duan et al. 2011; Garces et al. 2012,
2013). Chemerin, CMKLR1, and GPR1 are also expressed in
the testes of male rats, where they may play a role in
regulating steroidogenesis (Li et al. 2014b).
The exact role of chemerin in energy homeostasis is less
clear, with competing evidence pointing in several di-
rections (Helfer and Wu 2018). While it has been conrmed
that chemerin promotes adipogenesis through CMKLR1, the
inuence of chemerin signaling on the metabolism seems
context dependent (Ernst et al. 2012; Rouger et al. 2013).
Moreover, chemerin inuences insulin signaling, but the
mechanism is still unexplored (Kralisch et al. 2009; Taka-
hashi et al. 2008).
Thus, chemerin is an important regulator of inflamma-
tion, energy homeostasis, and reproductive functions
complex processes that inuence each other. An imbalance
in chemerin activity can, therefore, lead to a wide range of
complicationsasexemplied in Figure 5, which will be
discussed in the following sections.
Chemerin in obesity and related
diseases
During evolution, an excess of nutrients was barely an
issue. These days, however, the modern western lifestyle
has paved the way for a novel epidemic obesity. The US
center for disease control estimates that around 40% of
adults in the US were obese in 2018, with a clear upward
trend (Hales et al. 2020). In recent years, it has become
T.F. Fischer and A.G. Beck-Sickinger: Chemerin exploring a versatile adipokine 7
increasingly evident that excess body weight and fat mass
have a major inuence on a plethora of diseases ranging
from inammatory disorders over cardiovascular diseases
to cancer, reproductive dysfunctions, and impeded meta-
bolic functions (Guh et al. 2009). While adipose tissue was
initially seen as mere energy storage, it is now recognized
as an active endocrine organ, expressing a multitude of
signaling molecules; the so called adipocytokines or adi-
pokines (Scheja and Heeren 2019).
Several studies show that chemerin levels of non-
diabetic individuals rise with increasing body mass index
(BMI), a robust correlation that has been independently
confirmed by different groups (Bozaoglu et al. 2007, 2009;
Chang et al. 2016; Sledzinski et al. 2013). Physical exercise
such as high-intensity interval training decreased chem-
erin levels in obese women (Taheri Chadorneshin et al.
2019), and a combined strength and endurance training led
to lower chemerin levels independent of BMI (Stefanov
et al. 2014). Another study showed that bariatric surgery
and a low caloric diet both independently decreased
circulating chemerin levels (Chakaroun et al. 2012),
strengthening the notion that an increased serum con-
centration of chemerin may be a consequence, rather than
a cause, of high-fat mass. More importantly, studies sug-
gest that obesity not only leads to higher total levels of
chemerin but also a higher proportion of bioactive chem-
erin (Chang et al. 2016; Haberl et al. 2018). Excess levels of
chemerin may thus contribute to the chronic low-grade
inammation associated with obesity and, ultimately, to
the pathogenesis of the many co-morbidities of obesity
(Reilly and Saltiel 2017). Many of these conditions can
occur independently of obesity or each other, but display
complex interactions. The role of chemerin in these dis-
eases will be discussed in the following sections.
Hypertension and atherosclerosis
Obesity, hypertension, and atherosclerosis are tightly con-
nected risk factors for cardiovascular diseases (DeMarco
et al. 2014; Rocha and Libby 2009). Only recently, the role of
chemerin in the development of these complications has
gained the attention of different groups (Ferland and Watts
2015). Current evidence suggests that chemerin contributes
to hypertension by promoting vasoconstriction through
activation of CMKLR1 (Kennedy et al. 2016). This effect has
also been observed in isolated rat aorta, where treatment
with chemerin-9 induced the CMKLR1-dependent contrac-
tion of the isolated rat thoracic aorta, superior mesenteric
artery, and mesenteric resistance artery (Watts et al. 2013).
In line with these results, Kunimoto et al. (2015) demon-
strated that chronic treatment with chemerin increased
blood pressure in mice. Increased chemerin levels are also
an independent risk factor for atherosclerosis in hyperten-
sive patients, and atherosclerotic mice treated with chem-
erin displayed increased lipid accumulation in atheroscle-
rotic plaques (Gu et al. 2015; Jia et al. 2020). Moreover,
van der Vorst et al. (2019) report that apolipoprotein
E-decient CMKLR1-knockout mice displayed an increasing
proportion of alternatively activated, anti-inammatory M2
macrophages in atherosclerotic lesions, combined with
attenuated dendritic cell recruitment.
Figure 5: The physiological roles of
chemerin.
In a healthy state, chemerin contributes to
the regulation of a variety of processes
(center). An excess (right) or lack of
chemerin (left) can lead to or promote
several disease states.
8T.F. Fischer and A.G. Beck-Sickinger: Chemerin exploring a versatile adipokine
Diabetes
Perhaps the most prominent co-morbidity of obesity, type 2
diabetes is extremely prevalent among adults, impacting
almost half a billion people in 2019 as estimated in a meta-
study by the international diabetes federation (Saeedi
et al. 2019). Insulin resistance and glucose intolerance
are associated with type 2 diabetes, leading to impaired
glucose uptake and hyperglycemia (Kahn et al. 2006). Pa-
tients with type 2 diabetes display signicantly elevated
chemerin expression in adipose tissue compared to healthy
control subjects (Chakaroun et al. 2012). Models of obese
mice display decreased glucose tolerance after chemerin
treatment, suggesting that chemerin may be a driving force
in the pathogenesis of type 2 diabetes (Ernst et al. 2010). On
the other hand, chemerin knockout in mice also leads to
decreased glucose tolerance (Fang et al. 2021). Studies on
the impact of chemerin on glucose uptake by 3T3-L1 adi-
pocytes also come to conicting results: While Takahashi
et al. (2008) demonstrated increased insulin-dependent
glucose uptake upon stimulation with chemerin, Kralisch
et al. (2009) showed decreased glucose uptake in response
to chemerin treatment. Together, these studies suggest
a complex, context-dependent role for chemerin in the
regulation of glucose uptake.
Polycystic ovary syndrome
Polycystic ovary syndrome (PCOS) is the most common
cause of infertility among women, and obesity is a major
risk factor for this reproductive disease (Chen et al. 2013;
Gambineri et al. 2002). PCOS refers to a heterogeneous
group of gynecologic disorders that are characterized by
hyperandrogenism (typically excess testosterone), chronic
inammation, and anovulation (Goudas and Dumesic
1997). Independent of BMI, women with PCOS display
elevated serum chemerin levels compared to healthy in-
dividuals (Abruzzese et al. 2020; Huang et al. 2015; Wang
et al. 2014). Knock-out of CMKLR1 in mice attenuated
dihydrotestosterone-induced clinical signs of PCOS, sug-
gesting a role of chemerin in the pathogenesis of PCOS
(Tang et al. 2016). Lima et al. (2018) suggest that hyper-
androgenism increases chemerin expression in the ovaries,
leading to excess inltration of inammatory M1 macro-
phages and, ultimately, granulosa cell apoptosis. This is
supported by earlier ndings that chemerin expression is
upregulated in female rats treated with dihydrotestoster-
one, and that chemerin treatment alone was sufcient to
induce follicular growth arrest (Kim et al. 2013). Moreover,
chemerin induced insulin resistance in human granulosa-
lutein cells isolated from PCOS patients, thus potentially
aggravating the abnormalities in ovulation (Li et al. 2019a).
Psoriasis
Psoriasis is an auto-inflammatory skin disease, character-
ized by the formation of thickened plaques on the skin.
Current evidence suggests that obesity predisposes to the
development of this condition and aggravates existing
psoriasis (Jensen and Skov 2016). The pathophysiology of
psoriasis involves early inltration with macrophages,
dendritic cells, and NK T-cells, leading to the activation of
a T-cell-mediated inammatory cascade (Armstrong and
Read 2020). Consequently, chemerin expression marks
early psoriatic skin lesions and correlates with dendritic
cell recruitment. In patients with psoriasis, circulating
chemerin levels are elevated, and decrease during treat-
ment with the TNFαantibody iniximab (Lora et al. 2013;
Nakajima et al. 2010). Chemerin and its receptors are
expressed in healthy skin, but pronounced co-localization
of the activating protease KLK7 with chemerin is limited to
psoriatic lesions, suggesting an increased local bioactivity
in psoriatic skin (Banas et al. 2013, 2015; Schultz et al. 2013).
Rheumatoid arthritis
Rheumatoid arthritis (RA) is a chronic inflammatory joint
disease, which can cause bone damage, pain, and ulti-
mately, disability (Smolen et al. 2016). Obesity is a risk
factor for developing RA, and arthritis is so frequently
associated with psoriasis that psoriatic arthritis is a distinct
medical diagnosis (Crowson et al. 2013; Veale and Fearon
2018). Chemerin was rst identied as the ligand for
CMKLR1 from arthritic synovial uid, where it is present at
high concentrations (Wittamer et al. 2003). Chemerin and
CMKLR1 are both expressed by chondrocytes, and treat-
ment of these cells with chemerin leads to increased
secretion of inammatory cytokines (Berg et al. 2010).
Activation of chemerin in synovial uids was demon-
strated by Zhao et al. (2018), who found elevated levels of
active ChemF156 in synovial uids of arthritis patients.
Chemerin treatment induced the secretion of IL-6 and CCL2
in broblast-like synoviocytes, which may additionally
promote joint inammation (Kaneko et al. 2011). Interest-
ingly, polymorphonuclear neutrophils, the predominant
cell type recruited to synovial uid, demonstrate signi-
cantly upregulated expression of CCRL2 in arthritis, which
T.F. Fischer and A.G. Beck-Sickinger: Chemerin exploring a versatile adipokine 9
could further contribute to chemerin-mediated recruitment
of CMKLR1-expressing leukocytes to the inamed joint
(Auer et al. 2007).
Nonalcoholic steatohepatitis
Excessive accumulation of fat in the liver, accompanied
by chronic liver inammation and brosis are hallmarks
of nonalcoholic steatohepatitis (NASH), and the risk of
developing NASH is greatly increased in obese patients
(Neuschwander-Tetri 2017). The pathology of NASH in-
volves inltration of monocytes, neutrophils, and T cells
into the liver and expansion of liver-resident Kupffer and
stellate cells (Nati et al. 2016). Plasma levels of chemerin are
signicantly elevated in obese NASH patients compared to
healthy control subjects (Kukla et al. 2010; Sell et al. 2010).
In rodents, NASH correlated with signicantly elevated
levels of chemerin in the liver (Krautbauer et al. 2013).
Hepatic expression of chemerin and CMKLR1 mRNA
is increased in NASH, possibly reecting a chemerin-
mediated accumulation of CMKLR1-expressing leukocytes
(Döcke et al. 2013).
Chemerin in cancer
Over the past years, chemerin has gained increasing in-
terest regarding its role in cancer. Competing roles have
been proposed for chemerin in the context of tumor pro-
gression, either directly by modulating tumor growth and
metastasis through chemerin receptors expressed on tu-
mor cells, or indirectly by recruiting immune effector cells
to the tumor (Shin et al. 2018).
Downregulating local levels of chemerin seems to
contribute to the immune evasion strategy employed by
several cancer types: Chemerin mRNA levels are signifi-
cantly reduced in breast cancer tissues compared to sur-
rounding healthy tissue, and forced overexpression of
chemerin leads to infiltration of dendritic cells, NK cells,
and macrophages into the tumor microenvironment (TME),
thereby suppressing tumor growth (Pachynski et al. 2019).
In melanoma, chemerin-mediated recruitment of NK cells
to the TME suppressed metastasis (Pachynski et al. 2012).
Knockout of CCRL2 was associated with a reduced immune
response to lung cancer in mice, leading to enhanced tumor
growth (DelPrete et al. 2019). This underlines theimportance
of CCRL2 for generating stable chemotactic gradients of
chemerin, which can efciently recruit CMKLR1-expressing
leukocytes to sites of inammation In hepatocellular carci-
noma, decreased local expression of chemerin was
associated with poor patient prognosis, and administration
or forced overexpression of chemerin in mice reduced he-
patocellular carcinoma metastasis (Haberl et al. 2019; Li
et al. 2018). Low expression levels of chemerin also correlate
with decreased patient survival in prostate cancer and sar-
coma, where chemerin induces the expression of PTEN, a
critical tumor suppressor gene (Rennier et al. 2020). Conse-
quently, forced overexpression of chemerin in the prostate
cancer cell line DU145 inhibited tumor growth in vivo and
increased T cell-mediated cytotoxicity (Rennier et al. 2020).
Thus, downregulating chemerin expression seems a
widespread strategy among cancers to evade innate immu-
nity, especially NK cells the rst responders of the immune
system to emerging tumors (Guillerey et al. 2016). However,
tumor-expressed CMKLR1 seems to be directly involved in
the progression of some cancers: High CMKLR1 expression
by neuroblastoma cells correlates with decreased patient
survival, and blocking CMKLR1 reduced viability of these
cells in vitro and tumor progression in vivo (Tümmler et al.
2017). Colorectal cancer tissue expresses increased levels
of CMKLR1, correlating with tumor size and angiogenic
markers, suggesting that chemerin may promote angiogen-
esis during tumor growth (Kiczmer et al. 2020). Moreover,
chemerin promotes the invasion of oesophageal squamous
cancer cells in vitro through CMKLR1 (Kumar et al. 2016).
The chemerin system as a
therapeutic target state of the art
GPCR constitute the largest family of membrane proteins,
and current estimates suggest that around 35% of ap-
proved drugs target GPCR (Sriram and Insel 2018). Thus, the
chemerin receptors, mediating chemerin bioactivity in a
wide range of diseases, are attractive for therapeutic inter-
vention as well. As the majority of evidence suggest CMKLR1
as the receptor responsible for mediating the inammatory
properties of chemerin, most studies have focused on this
receptor for drug development.
Two compounds targeting CMKLR1 have been tested in
clinical trials to date: CCX832, a CMKLR1 antagonist by
ChemoCentryx and GlaxoSmithKline was tested in phase I
clinical trials in 2011, but was discontinued (Chemocentryx
2011). Although its structure remains undisclosed, it is
frequently used as a tool compound to study CMKLR1
pharmacology (Kennedy and Davenport 2018). In diabetic
ob/ob mice, CCX832 decreased body weight, insulin and
glucose levels, and oxidative stress, suggesting that
blocking CMKLR1 activation could be benecial for treating
type 2 diabetes (Neves et al. 2018). A phase II clinical study
10 T.F. Fischer and A.G. Beck-Sickinger: Chemerin exploring a versatile adipokine
of the resolvin E1 analog RX-10045 in 2019 concluded
insufcient efcacy in the treatment of ocular inamma-
tion (Sablinski 2019). However, although Hauser et al.
(2017) report CMKLR1 as the target of RX-10045, this pairing
was not experimentally conrmed.
Graham et al. (2014) identied 2-(α-naphtoyl) ethyl-
trimetyhlammonium iodide (α-NETA) as an inhibitor of
chemerin-induced recruitment of arrestin 3 to CMKLR1.
This compound was effective in suppressing experimental
autoimmune encephalomyelitis, a murine model of mul-
tiple sclerosis (MS) (Graham et al. 2014). Later, more potent
derivatives of this molecule were identied in structure-
activity relation studies, which proved superior to the
FDA-approved MS drug Tecdera in the same model sys-
tem (Kumar et al. 2019). Moreover, treatment with α-NETA
reduced clonogenicity and cell viability of neuroblastoma
cells in vitro and tumor growth in vitro (Tümmler et al.
2017). However, no binding data are reported for this
compound, and the mechanism of action (orthosteric or
allosteric) at the CMKLR1 remains unclear.
Imaizumi et al. (2019) engaged in extensive SAR
studies of 2-aminobenzoxazol analogs as inhibitors of
CMKLR1-mediated Ca2+ux and migration of the dendritic
cell line CAL-1. These analogs display high oral bioavail-
ability and good pharmacokinetics in monkeys (Imaizumi
et al. 2020). However, the authors state that these com-
pounds inhibit CMKLR1-mediated Ca2+ux by inducing
CMKLR1 internalization (Imaizumi et al. 2019, 2020). This is
in contrast to the mode of action of other known modula-
tors of GPCR signaling, asking for a more detailed phar-
macological investigation of these compounds (Burkert
et al. 2017; Konieczny et al. 2020; Yang et al. 2018). An
overview of small molecule compounds targeting CMKLR1
is given in Figure 6.
Chemerin-derived peptides have also been investi-
gated regarding their therapeutic potential for various
diseases: Cash and coworkers reported that the C-terminal
chemerin fragment C15, terminating at A155, suppressed
inflammation and promoted wound healing through
CMKLR1 (Cash et al. 2008, 2014). However, C15 lacks the
C-terminal F156, which is critical for binding to CMKLR1,
and others have not conrmedanyactivityofC15at
CMKLR1 (Fischer et al. 2021c; Luangsay et al. 2009; Wit-
tamer et al. 2003, 2004; Yamaguchi et al. 2011). Thus, it
remains unclear how C15 mediates its benecial effects.
Using a murine model of pancreatic diabetes mellitus, Tu
et al. (2020) demonstrated that treatment with chemerin-9
signicantly alleviated glucose intolerance and insulin resis-
tance through increased expression of the glucose transporter
GLUT2. However, the plasma half-life of chemerin-9 is less
than 10 min, suggesting that stabilized variants would be
even more benecial (Bandholtz et al. 2012; Shimamura
etal.2009).Suchstabilized,68Ga-labeled derivatives have
demonstrated high potential as PET tracers in murine breast
cancer xenograft models and may develop into future thera-
peutics (Erdmann et al. 2019). The same principle has been
applied by using metabolically stable, cyclic analogs of
the chemerin C-terminus, which can efciently shuttle toxic
payloads into CMKLR1-expressing cells (Fischer et al. 2021b).
Such methotrexate-coupled cyclic chemerin derivatives
demonstrated superior toxicity than free methotrexate on
CMKLR1-expressing cells, while no toxic effect was observed
in cells lacking the receptor. Thus, peptide-drug conjugates
based on cyclic chemerin-9 analogs may be promising for
targeted cancer therapies.
In conclusion, therapeutics targeting the chemerin
system have focused on CMKLR1 so far. Their potential
indications are as diverse as the morbidities that chemerin
Figure 6: Structures of selected small molecule compounds targeting CMKLR1.
The Resolvin E1 analog RX-10045 (left) was tested in phase II clinical trials for the treatment of ocular inflammation. The antagonist α-NETA
(center) is described as a sub-micromolar inhibitor of chemerin-induced arrestin 3 recruitment to CMKLR1 and is widely used in experimental
studies. The recently developed 2-aminobenzoxazol analog 2(right) demonstrated promising potency and pharmacokinetics, but its mode of
action remains to be claried.
T.F. Fischer and A.G. Beck-Sickinger: Chemerin exploring a versatile adipokine 11
is involved in, but no drugs targeting the chemerin system
have been approved to date. This may be, in part, due to the
lack of information on the molecular interactions govern-
ing the bioactivity of chemerin, which are still not suffi-
ciently understood.
Conclusions
Chemerin is an adipokine involved in adipose tissue ho-
meostasis, immunity, and the regulation of reproductive
functions. A wide range of inflammatory and obesity-
related diseases is driven by an excess of active chemerin,
while downregulation of chemerin expression helps several
cancer types to avoid recognition by the innate immune
system. Multiple studies have allowed a glimpse at the
therapeutic potential that lies in addressing the chemerin
system, but so far, no compounds have successfully
exploited this potential to become approved drugs.
The regulation of chemerin activity takes place on the
levels of expression, proteolytic activation and deactiva-
tion, as well as receptor binding. With three chemerin re-
ceptors displaying distinct signaling properties, the latter is
particularly complex. CMKLR1 is the most intensely studied
receptor for chemerin, and seems to mediate most of its
functions. Recruiting NK cells, macrophages, and dendritic
cells to sites of inflammation, the pro-inflammatory role of
chemerin-induced CMKLR1 activation is firmly established
in most contexts. Controversy exists concerning the role of
GPR1, the second chemerin receptor. Displaying uncon-
ventional signaling properties, i.e., reduced or absent G
protein signaling, it is unclear whether this receptor exerts
distinct chemerin-mediated functions, or if it predomi-
nantly acts as a decoy receptor. CCRL2, the third chemerin
receptor, is essential to build up stable chemotactic
chemerin gradients, and no CCRL2-mediated chemerin
signaling has been described to date.
Successful therapeutic interventions targeting the
chemerin system will rely on improved understanding of
the molecular mechanisms governing the bioactivity of this
adipokine.
Author contributions: All the authors have accepted
responsibility for the entire content of this submitted
manuscript and approved submission.
Research funding: This study was nancially supported
by the Deutsche Forschungsgemeinschaft (DFG), German
Research Foundation, project number 209933838
SFB1052 Z05.
Conict of interest statement: The authors declare no
competing nancial interests.
References
Abruzzese, G.A., Gamez, J., Belli, S.H., Levalle, O.A., Mormandi, E.,
Otero, P., Grafgna, M.N., Cerrone, G.E., and Motta, A.B. (2020).
Increased chemerin serum levels in hyperandrogenic and
normoandrogenic women from Argentina with polycystic ovary
syndrome. Gynecol. Endocrinol. 36: 10571061.
Allen, S.J., Zabel, B.A., Kirkpatrick, J., Butcher, E.C., Nietlispach, D.,
and Handel, T.M. (2007). NMR assignment of human chemerin, a
novel chemoattractant. Biomol. NMR Assign. 1: 171173.
Arita, M., Bianchini, F., Aliberti, J., Sher, A., Chiang, N., Hong, S., Yang,
R., Petasis, N.A., and Serhan, C.N. (2005). Stereochemical
assignment, antiinammatory properties, and receptor for the
omega-3 lipid mediator resolvin E1. J. Exp. Med. 201: 713722.
Armstrong, A.W. and Read, C. (2020). Pathophysiology, clinical
presentation, and treatment of psoriasis: a review. J. Am. Med.
Assoc. 323: 19451960.
Auer, J., Bläss, M., Schulze-Koops, H., Russwurm, S., Nagel, T.,
Kalden, J.R., Röllinghoff, M., and Beuscher, H.U. (2007).
Expression and regulation of CCL18 in synovial uid neutrophils
of patients with rheumatoid arthritis. Arthritis Res. Ther. 9: R94.
Ballesteros, J.A. and Weinstein, H. (1995). Integrated methods for the
construction of three-dimensional models and computational
probing of structure-function relations in G protein-coupled
receptors. In: Sealfon, S.C. (Ed.), Methods in neurosciences:
receptor molecular biology. Academic Press, Cambridge,
Massachusetts, USA, pp. 366428.
Baltoumas, F.A., Theodoropoulou, M.C., and Hamodrakas, S.J. (2013).
Interactions of the α-subunits of heterotrimeric G-proteins with
GPCRs, effectors and RGS proteins: a critical review and analysis
of interacting surfaces, conformational shifts, structural
diversity and electrostatic potentials. J. Struct. Biol. 182:
209218.
Banas, M., Zabieglo, K., Kasetty, G., Kapinska-Mrowiecka, M.,
Borowczyk, J., Drukala, J., Murzyn, K., Zabel, B.A., Butcher, E.C.,
Schroeder, J.M., et al. (2013). Chemerin is an antimicrobial agent
in human epidermis. PLoS One 8: e58709.
Banas, M., Zegar, A., Kwitniewski, M., Zabieglo, K., Marczynska, J.,
Kapinska-Mrowiecka, M., LaJevic, M., Zabel, B.A., and Cichy, J.
(2015). The expression and regulation of chemerin in the
epidermis. PLoS One 10: e0117830.
Bandholtz, S., Wichard, J., Kühne, R., and Grötzinger, C. (2012).
Molecular evolution of a peptide GPCR ligand driven by articial
neural networks. PLoS One 7: e36948.
Barnea, G., Strapps, W., Herrada, G., Berman, Y., Ong, J., Kloss, B.,
Axel, R., and Lee, K.J. (2008). The genetic design of signaling
cascades to record receptor activation. Proc. Natl. Acad. Sci.
U.S.A. 105: 6469.
Berg, V., Sveinbjörnsson, B., Bendiksen, S., Brox, J., Meknas, K., and
Figenschau, Y. (2010). Human articular chondrocytes express
ChemR23 and chemerin; ChemR23 promotes inammatory
signalling upon binding the ligand chemerin 21-157. Arthritis
Res. Ther. 12: 112.
12 T.F. Fischer and A.G. Beck-Sickinger: Chemerin exploring a versatile adipokine
Bondue, B., Wittamer, V., and Parmentier, M. (2011). Chemerin and its
receptors in leukocyte trafcking, inammation and metabolism.
Cytokine Growth Factor Rev. 22: 331338.
Bozaoglu, K., Bolton, K., McMillan, J., Zimmet, P., Jowett, J., Collier, G.,
Walder, K., and Segal, D. (2007). Chemerin is a novel adipokine
associated with obesity and metabolic syndrome. Endocrinology
148: 46874694.
Bozaoglu, K., Segal, D., Shields, K.A., Cummings, N., Curran, J.E.,
Comuzzie, A.G., Mahaney, M.C., Rainwater, D.L., VandeBerg, J.L.,
MacCluer, J.W., et al. (2009). Chemerin is associated with
metabolic syndrome phenotypes in a Mexican-American
population. J. Clin. Endocrinol. Metab. 94: 30853088.
Burkert, K., Zellmann, T., Meier, R., Kaiser, A., Stichel, J., Meiler, J.,
Mittapalli, G.K., Roberts, E., and Beck-Sickinger, A.G. (2017). A
deep hydrophobic binding cavity is the main interaction for
different Y2 R antagonists. ChemMedChem 12: 7585.
Cash, J.L., Bass, M.D., Campbell, J., Barnes, M., Kubes, P., and
Martin, P. (2014). Resolution mediator chemerin15 reprograms
the wound microenvironment to promote repair and reduce
scarring. Curr. Biol. 24: 14061414.
Cash, J.L., Hart, R., Russ, A., Dixon, J.P.C., Colledge, W.H., Doran, J.,
Hendrick, A.G., Carlton, M.B.L., and Greaves, D.R. (2008).
Synthetic chemerin-derived peptides suppress inammation
through ChemR23. J. Exp. Med. 205: 767775.
Catusse, J., Leick, M., Groch, M., Clark, D.J., Buchner, M.V., Zirlik, K.,
and Burger, M. (2010). Role of the atypical chemoattractant
receptor CRAM in regulating CCL19 induced CCR7 responses in
B-cell chronic lymphocytic leukemia. Mol. Cancer 9: 297.
Chakaroun, R., Raschpichler, M., Klöting, N., Oberbach, A., Flehmig,
G., Kern, M., Schön, M.R., Shang, E., Lohmann, T., Dreßler, M.,
et al. (2012). Effects of weight loss and exercise on chemerin
serum concentrations and adipose tissue expression in human
obesity. Metab. Clin. Exp. 61: 706714.
Chang, S.-S., Eisenberg, D., Zhao, L., Adams, C., Leib, R., Morser, J.,
and Leung, L. (2016). Chemerin activation in human obesity.
Obesity 24: 15221529.
Chemocentryx (2011). ChemoCentryx initiates phase I clinical trial of
CCX832, the rst small molecule inhibitor of ChemR23, Available
at: <https://ir.chemocentryx.com/news-releases/news-release-
details/chemocentryx-initiates-phase-i-clinical-trial-
ccx832-rst-small>.
Chen, X., Jia, X., Qiao, J., Guan, Y., and Kang, J. (2013). Adipokines in
reproductive function: a link between obesity and polycystic
ovary syndrome. J. Mol. Endocrinol. 50: 37.
Chen, T., Xiong, M., Zong, X., Ge, Y., Zhang, H., Wang, M., Han, G.W.,
Yi, C., Ma, L., Ye, R.D., et al. (2020). Structural basis of ligand
binding modes at the human formyl peptide receptor 2. Nat.
Commun. 11: 19.
Crowson, C.S., Matteson, E.L., Davis, J.M., and Gabriel, S.E. (2013).
Contribution of obesity to the rise in incidence of rheumatoid
arthritis. Arthritis Care Res. 65: 7177.
Del Prete, A., Martínez-Muñoz, L., Mazzon, C., Toffali, L., Sozio, F.,
Za, L., Bosisio, D., Gazzurelli, L., Salvi, V., Tiberio, L., et al. (2017).
The atypical receptor CCRL2 is required for CXCR2-dependent
neutrophil recruitment and tissue damage. Blood 130:
12231234.
Del Prete, A., Sozio, F., Schioppa, T., Ponzetta, A., Vermi, W., Calza, S.,
Bugatti, M., Salvi, V., Bernardini, G., Benvenuti, F., et al. (2019).
The atypical receptor CCRL2 is essential for lung cancer immune
surveillance. Cancer Immunol. Res. 7: 17751788.
DeMarco, V.G., Aroor, A.R., and Sowers, J.R. (2014). The
pathophysiology of hypertension in patients with obesity. Nat.
Rev. Endocrinol. 10: 364376.
Döcke, S., Lock, J.F., Birkenfeld, A.L., Hoppe, S., Lieske, S., Rieger, A.,
Raschzok, N., Sauer, I.M., Florian, S., Osterhoff, M.A., et al.
(2013). Elevated hepatic chemerin mRNA expression in human
non-alcoholic fatty liver disease. Eur. J. Endocrinol. 169:
547557.
Dranse, H.J., Muruganandan, S., Fawcett, J.P., and Sinal, C.J. (2016).
Adipocyte-secreted chemerin is processed to a variety of
isoforms and inuences MMP3 and chemokine secretion through
an NF-κB-dependent mechanism. Mol. Cell. Endocrinol. 436:
114129.
Du, X.-Y., Zabel, B.A., Myles, T., Allen, S.J., Handel, T.M., Lee, P.P.,
Butcher, E.C., and Leung, L.L. (2009). Regulation of chemerin
bioactivity by plasma carboxypeptidase N, carboxypeptidase B
(activated thrombin-activable brinolysis inhibitor), and
platelets. J. Biol. Chem. 284: 751758.
Duan, D.-M., Niu, J.-M., Lei, Q., Lin, X.-H., and Chen, X. (2011). Serum
levels of the adipokine chemerin in preeclampsia. J. Perinat.
Med. 40: 121127.
Erdmann, S., Niederstadt, L., Koziolek, E.J., Gómez, J.D.C., Prasad, S.,
Wagener, A., von Hacht, J.L., Reinicke, S., Exner, S., Bandholtz, S.,
et al. (2019). CMKLR1-targeting peptide tracers for PET/MR
imaging of breast cancer. Theranostics 9: 67196733.
Ernst, M.C., Haidl, I.D., Zúñiga, L.A., Dranse, H.J., Rourke, J.L.,
Zabel, B.A., Butcher, E.C., and Sinal, C.J. (2012). Disruption of the
chemokine-like receptor-1 (CMKLR1) gene is associated with
reduced adiposity and glucose intolerance. Endocrinology 153:
672682.
Ernst, M.C., Issa, M., Goralski, K.B., and Sinal, C.J. (2010). Chemerin
exacerbates glucose intolerance in mouse models of obesity and
diabetes. Endocrinology 151: 19982007.
Ernst, M.C. and Sinal, C.J. (2010). Chemerin: at the crossroads of
inammation and obesity. Trends Endocrinol. Metabol. 21:
660667.
Fang, P., Han, L., Yu, M., Han, S., Wang, M., Huang, Y., Guo, W.,
Wei, Q., Shang, W., and Min, W. (2021). Development of
metabolic dysfunction in mice lacking chemerin. Mol. Cell.
Endocrinol. 535: 111369.
Farzan, M., Choe, H., Martin, K., Marcon, L., Hofmann, W., Karlsson, G.,
Sun, Y., Barrett, P., Marchand, N., Sullivan, N., et al. (1997).
Two orphan seven-transmembrane segment receptors which are
expressedin CD4-positive cells support simian immunodeciency
virus infection. J. Exp. Med. 186: 405411.
Feder, S., Bruckmann, A., McMullen, N., Sinal, C.J., and Buechler, C.
(2020). Chemerin isoform-specic effects on hepatocyte
migration and immune cell inammation. Int. J. Mol. Sci. 21,
https://doi.org/10.3390/ijms21197205.
Ferland, D.J. and Watts, S.W. (2015). Chemerin: a comprehensive
review elucidating the need for cardiovascular research.
Pharmacol. Res. 99: 351361.
Fischer, T.F., Czerniak, A.S., Weiß, T., Schoeder, C.T., Wolf, P., Seitz, O.,
Meiler, J., and Beck-Sickinger, A.G. (2021a). Ligand-binding
and -scavenging of the chemerin receptor GPR1. Cell. Mol. Life Sci.
78: 62656281.
T.F. Fischer and A.G. Beck-Sickinger: Chemerin exploring a versatile adipokine 13
Fischer, T.F., Czerniak, A.S., Weiß, T., Zellmann, T., Zielke, L., Els-
Heindl, S., and Beck-Sickinger, A.G. (2021b). Cyclic derivatives of
the chemerin C-terminus as metabolically stable agonists at the
chemokine-like receptor 1 for cancer treatment. Cancers 13:
3788.
Fischer, T.F., Schoeder, C.T., Zellmann, T., Stichel, J., Meiler, J., and
Beck-Sickinger, A.G. (2021c). Cyclic analogues of the chemerin
C-terminus mimic a loop conformation essential for activating
the chemokine-like receptor 1. J. Med. Chem. 64: 30483058.
Foster, S.R., Hauser, A.S., Vedel, L., Strachan, R.T., Huang, X.-P.,
Gavin, A.C., Shah, S.D., Nayak, A.P., Haugaard-Kedström, L.M.,
Penn, R.B., et al. (2019). Discovery of human signaling systems:
pairing peptides to G protein-coupled receptors. Cell 179:
895908.e21.
Fredriksson, R., Lagerström, M.C., Lundin, L.-G., and Schiöth, H.B.
(2003). The G-protein-coupled receptors in the human genome
form ve main families. Phylogenetic analysis, paralogon
groups, and ngerprints. Mol. Pharmacol. 63: 12561272.
Galligan, C.L., Matsuyama, W., Matsukawa, A., Mizuta, H., Hodge,
D.R., Howard, O.M.Z., and Yoshimura, T. (2004). Up-regulated
expression and activation of the orphan chemokine receptor,
CCRL2, in rheumatoid arthritis. Arthritis Rheum. 50: 18061814.
Gambineri, A., Pelusi, C., Vicennati, V., Pagotto, U., and Pasquali, R.
(2002). Obesity and the polycystic ovary syndrome. Int. J. Obes.
26: 883896.
Garces, M.F., Sanchez, E., Acosta, B.J., Angel, E., Ruíz, A.I., Rubio-
Romero, J.A., Diéguez, C., Nogueiras, R., and Caminos, J.E.
(2012). Expression and regulation of chemerin during rat
pregnancy. Placenta 33: 373378.
Garces, M.F., Sanchez, E., Ruíz-Parra, A.I., Rubio-Romero, J.A.,
Angel-Müller, E., Suarez, M.A., Bohórquez, L.F., Bravo, S.B.,
Nogueiras, R., Diéguez, C., et al. (2013). Serum chemerin levels
during normal human pregnancy. Peptides 42: 138143.
Ge, X., Yamaguchi, Y., Zhao, L., Bury, L., Gresele, P., Berube, C.,
Leung, L.L., and Morser, J. (2018). Prochemerin cleavage by factor
XIa links coagulation and inammation. Blood 131: 353364.
Godlewska, U., Brzoza, P., Sroka, A., Majewski, P., Jentsch, H.,
Eckert, M., Eick, S., Potempa, J., Zabel, B.A., and Cichy, J. (2017).
Antimicrobial and attractant roles for chemerin in the oral cavity
during inammatory gum disease. Front. Immunol. 8: 353.
Goralski, K.B., McCarthy, T.C., Hanniman, E.A., Zabel, B.A.,
Butcher, E.C., Parlee, S.D., Muruganandan, S., and Sinal, C.J.
(2007). Chemerin, a novel adipokine that regulates adipogenesis
and adipocyte metabolism. J. Biol. Chem. 282: 2817528188.
Goudas, V.T. and Dumesic, D.A. (1997). Polycystic ovary syndrome.
Endocrinol Metab. Clin. N. Am. 26: 893912.
Graham, K.L., Zhang, J.V., Lewén, S., Burke, T.M., Dang, T., Zoudilova, M.,
Sobel, R.A., Butcher, E.C., and Zabel, B.A. (2014). A novel CMKLR1
small molecule antagonist suppresses CNS autoimmune
inammatory disease. PLoS One 9: e112925.
Grundmann,M., Merten, N.,Malfacini, D., Inoue,A., Preis, P., Simon,K.,
Rüttiger, N., Ziegler, N., Benkel, T., Schmitt, N.K., et al. (2018).
Lack of beta-arrestin signaling in the absence of active G proteins.
Nat. Commun. 9: 116.
Gu, P., Cheng, M., Hui, X., Lu, B., Jiang, W., and Shi, Z. (2015). Elevating
circulation chemerin level is associated with endothelial
dysfunction and early atherosclerotic changes in essential
hypertensive patients. J. Hypertens. 33: 16241632.
Guh, D.P., Zhang, W., Bansback, N., Amarsi, Z., Birmingham, C.L., and
Anis, A.H. (2009). The incidence of co-morbidities related to
obesity and overweight: a systematic review and meta-analysis.
BMC Publ. Health 9: 88.
Guillabert, A., Wittamer, V., Bondue, B., Godot, V., Imbault, V.,
Parmentier, M., and Communi, D. (2008). Role of neutrophil
proteinase 3 and mast cell chymase in chemerin proteolytic
regulation. J. Leukoc. Biol. 84: 15301538.
Guillerey, C., Huntington, N.D., and Smyth, M.J. (2016). Targeting
natural killer cells in cancer immunotherapy. Nat. Immunol. 17:
10251036.
Haberl, E.M., Pohl, R., Rein-Fischboeck, L., Feder, S., Eisinger, K.,
Krautbauer, S., Sinal, C.J., and Buechler, C. (2018). Ex vivo
analysis of serum chemerin activity in murine models of obesity.
Cytokine 104: 4245.
Haberl, E.M., Pohl, R., Rein-Fischboeck, L., Feder, S., Sinal, C.J.,
Bruckmann, A., Hoering, M., Krautbauer, S., Liebisch, G., and
Buechler, C. (2019). Overexpression of hepatocyte chemerin-156
lowers tumor burden in a murine model of diethylnitrosamine-
induced hepatocellular carcinoma. Int. J. Mol. Sci. 21,
https://doi.org/10.3390/ijms21010252.
Hales, C.M., Carroll, M.D., Fryar, C.D., and Ogden, C.L. (2020).
Prevalence of obesity and severe obesity among adults: United
States, 20172018 NCHS Data Brief. National Center for Health
Statistics, Hyattsville, MD, USA.
Hartmann,T.N.,Leick,M.,Ewers,S.,Diefenbacher,A.,Schraufstatter,I.,
Honczarenko,M.,andBurger,M.(2008).HumanBcellsexpress
the orphan chemokine receptor CRAM-A/B in a maturation-stage-
dependent and CCL5-modulated manner. Immunology 125:
252262.
Hauser, A.S., Attwood, M.M., Rask-Andersen, M., Schiöth, H.B., and
Gloriam, D.E. (2017). Trends in GPCR drug discovery: new agents,
targets and indications. Nat. Rev. Drug Discov. 16: 829842.
Helfer, G. and Wu, Q.-F. (2018). Chemerin: a multifaceted adipokine
involved in metabolic disorders. J. Endocrinol. 238: R79R94.
de Henau, O., Degroot, G.-N., Imbault, V., Robert, V., de Poorter, C.,
Mcheik, S., Galés, C., Parmentier, M., and Springael, J.-Y. (2016).
Signaling properties of chemerin receptors CMKLR1, GPR1 and
CCRL2. PLoS One 11: e0164179.
Huang, R., Yue, J., Sun, Y., Zheng, J., Tao, T., Li, S., and Liu, W. (2015).
Increased serum chemerin concentrations in patients with
polycystic ovary syndrome: relationship between insulin
resistance and ovarian volume. Clin. Chim. Acta 450:
366369.
Imaizumi, T., Kobayashi, A., Otsubo, S., Komai, M., Magara, M., and
Otsubo, N. (2019). The discovery and optimization of a series of
2-aminobenzoxazole derivatives as ChemR23 inhibitors. Bioorg.
Med. Chem. 27: 115091.
Imaizumi, T., Otsubo, S., Komai, M., Takada, H., Maemoto, M.,
Kobayashi, A., and Otsubo, N. (2020). The design, synthesis and
evaluation of 2-aminobenzoxazole analogues as potent and
orally efcacious ChemR23 inhibitors. Bioorg. Med. Chem. 28:
115622.
Jensen, P. and Skov, L. (2016). Psoriasis and obesity. Dermatology
(Basel) 232: 633639.
Jia, J., Yu, F., Xiong, Y., Wei, W., Ma, H., Nisi, F., Song, X., Yang, L.,
Wang, D., Yuan, G., et al. (2020). Chemerin enhances the
adhesion and migration of human endothelial progenitor cells
and increases lipid accumulation in mice with atherosclerosis.
Lipids Health Dis. 19: 207.
Jumper,J.,Evans,R.,Pritzel,A.,Green,T.,Figurnov,M.,Ronneberger,O.,
Tunyasuvunakool, K., Bates, R., Žídek, A., Potapenko, A.,
14 T.F. Fischer and A.G. Beck-Sickinger: Chemerin exploring a versatile adipokine
et al. (2021). Highly accurate protein structure prediction with
AlphaFold. Nature 596: 583589.
Kahn, S.E., Hull, R.L., and Utzschneider, K.M. (2006). Mechanisms
linking obesity to insulin resistance and type 2 diabetes. Nature
444: 840846.
Kaneko,K.,Miyabe,Y.,Takayasu,A.,Fukuda,S.,Miyabe,C.,Ebisawa,M.,
Yokoyama, W., Watanabe, K., Imai, T., Muramoto, K., et al. (2011).
Chemerin activates broblast-like synoviocytes in patients with
rheumatoid arthritis. Arthritis Res. Ther. 13: 114.
Kashem, S.W., Haniffa, M., and Kaplan, D.H. (2017). Antigen-
presenting cells in the skin. Annu. Rev. Immunol. 35: 469499.
Kelley, L.A., Mezulis, S., Yates, C.M., Wass, M.N., and Sternberg, M.J.E.
(2015). ThePhyre2 web portal for protein modeling, prediction and
analysis. Nat. Protoc. 10: 845858.
Kennedy, A.J. and Davenport, A.P. (2018). International union of basic
and clinical pharmacology CIII: chemerin receptors CMKLR1
(Chemerin1) and GPR1 (Chemerin2) nomenclature,
pharmacology, and function. Pharmacol. Rev. 70: 174196.
Kennedy, A.J., Yang, P., Read, C., Kuc, R.E., Yang, L., Taylor, E.J.A.,
Taylor, C.W., Maguire, J.J., and Davenport, A.P. (2016). Chemerin
elicits potent constrictor actions via chemokine-like receptor 1
(CMKLR1), not G-protein-coupled receptor 1 (GPR1), in human and
rat vasculature. J. Am. Heart Assoc. 5: e004421.
Kiczmer, P.,Seńkowska, A.P., Kula,A., Dawidowicz, M., Strzelczyk, J.K.,
Zajdel, E.N., Walkiewicz, K., Waniczek, D., Ostrowska, Z., and
Świętochowska, E. (2020). Assessment of CMKLR1 level in
colorectal cancer andits correlation with angiogenic markers. Exp.
Mol. Pathol. 113: 104377.
Kim, J.Y., Xue, K., Cao, M., Wang, Q., Liu, J.-Y., Leader, A., Han, J.Y., and
Tsang, B.K. (2013). Chemerin suppresses ovarian follicular
development and its potential involvement in follicular arrest in
rats treated chronically with dihydrotestosterone. Endocrinology
154: 29122923.
Konieczny, A., Braun, D., Wiing, D., Bernhardt, G., and Keller, M.
(2020). Oligopeptides as neuropeptide Y Y4 receptor ligands:
identication of a high-afnity tetrapeptide agonist and a
hexapeptide antagonist. J. Med. Chem. 63: 81988215.
Kralisch, S., Weise, S., Sommer, G., Lipfert, J., Lossner, U., Bluher, M.,
Stumvoll, M., and Fasshauer, M. (2009). Interleukin-1βinduces
the novel adipokine chemerin in adipocytes in vitro. Regul. Pept.
154: 102106.
Krautbauer, S., Wanninger, J., Eisinger, K., Hader, Y., Beck, M.,
Kopp, A., Schmid, A., Weiss, T.S., Dorn, C., and Buechler, C.
(2013). Chemerin is highly expressed in hepatocytes and is
induced in non-alcoholic steatohepatitis liver. Exp. Mol. Pathol.
95: 199205.
Krishnamoorthy, S., Recchiuti, A., Chiang, N., Yacoubian, S., Lee, C.-H.,
Yang, R., Petasis, N.A., and Serhan, C.N.(2010). Resolvin D1 binds
human phagocytes with evidencefor proresolving receptors. Proc.
Natl. Acad. Sci. U.S.A. 107: 16601665.
Kukla, M., Zwirska-Korczala, K., Hartleb, M., Waluga, M., Chwist, A.,
Kajor, M., Ciupinska-Kajor, M., Berdowska, A., Wozniak-Grygiel,
E., and Buldak, R. (2010). Serum chemerin and vaspin in non-
alcoholic fatty liver disease. Scand. J. Gastroenterol. 45:
235242.
Kumar, J.D., Kandola, S., Tiszlavicz, L., Reisz, Z., Dockray, G.J., and
Varro, A. (2016). The role of chemerin and ChemR23 in
stimulating the invasion of squamous oesophageal cancer cells.
Br. J. Cancer 114: 11521159.
Kumar, V., LaJevic, M., Pandrala, M., Jacobo, S.A., Malhotra, S.V., and
Zabel, B.A. (2019). Novel CMKLR1 inhibitors for application in
demyelinating disease. Sci. Rep. 9: 7178.
Kunimoto, H., Kazama, K., Takai, M., Oda, M., Okada, M., and
Yamawaki, H. (2015). Chemerin promotes the proliferation and
migration of vascular smooth muscle and increases mouse blood
pressure. Am. J. Physiol. Heart Circ. Physiol. 309: H1017H1028.
Li, J.-J., Yin, H.-K., Guan, D.-X., Zhao, J.-S., Feng, Y.-X., Deng, Y.-Z.,
Wang, X., Li, N., Wang, X.-F., Cheng, S.-Q., et al. (2018). Chemerin
suppresses hepatocellular carcinoma metastasis through
CMKLR1-PTEN-Akt axis. Br. J. Cancer 118: 13371348.
Li, L., Huang, C., Zhang, X., Wang, J., Ma, P., Liu, Y., Xiao, T.,
Zabel, B.A., and Zhang, J.V. (2014a). Chemerin-derived peptide
C-20 suppressed gonadal steroidogenesis. Am. J. Reprod.
Immunol. 71: 265277.
Li, L., Ma, P., Huang, C., Liu, Y., Zhang, Y., Gao, C., Xiao, T., Ren, P.-G.,
Zabel, B.A., and Zhang, J.V. (2014b). Expression of chemerin and
its receptors in rat testes and its action on testosterone
secretion. J. Endocrinol. 220: 155163.
Li, X., Zhu, Q., Wang, W., Qi, J., He, Y., Wang, Y., Lu, Y., Wu, H., Ding, Y.,
and Sun, Y. (2019a). Elevated chemerin induces insulin
resistance in human granulosa-lutein cells from polycystic ovary
syndrome patients. Faseb. J. 33, https://doi.org/10.1096/fj.
201802829R.
Li, X.X., Lee, J.D., Kemper, C., and Woodruff, T.M. (2019b). The
complement receptor C5aR2: a powerful modulator of innate and
adaptive immunity. J. Immunol. 202: 33393348.
Lima, P.D.A., Nivet, A.-L., Wang, Q., Chen, Y.-A., Leader, A., Cheung, A.,
Tzeng, C.-R., and Tsang, B.K. (2018). Polycystic ovary syndrome:
possible involvement of androgen-induced, chemerin-mediated
ovarian recruitment of monocytes/macrophages. Biol. Reprod.
99: 838852.
Lora, V., Bonaguri, C., Gisondi, P., Sandei, F., Battistelli, L., Russo, A.,
Melegari, A., Trenti, T., Lippi, G., and Girolomoni, G. (2013).
Autoantibody induction and adipokine levels in patients with
psoriasis treated with iniximab. Immunol. Res. 56: 382389.
Luangsay, S., Wittamer, V., Bondue, B., de Henau, O., Rouger, L., Brait,
M., Franssen, J.-D., de Nadai, P., Huaux, F., and Parmentier, M.
(2009). Mouse ChemR23 is expressed in dendritic cell subsets
and macrophages, and mediates an anti-inammatory activity
of chemerin in a lung disease model. J. Immunol. 183:
64896499.
Marchese, A., Docherty, J., Nguyen, T., Heiber, M., Cheng, R., Heng, H.,
Tsui, L.-C., Shi, X., George, S., and ODowd, B. (1994). Cloning of
human genes encoding novel G protein-coupled receptors.
Genomics 23: 609618.
Martínez-Muñoz, L., Villares, R., Rodríguez-Fernández, J.L.,
Rodríguez-Frade, J.M., and Mellado, M. (2018). Remodeling our
concept of chemokine receptor function: from monomers to
oligomers. J. Leukoc. Biol. 104: 323331.
Mattern, A., Zellmann, T., and Beck-Sickinger, A.G. (2014).
Processing, signaling, and physiological function of chemerin.
IUBMB Life 66: 1926.
Mazzotti, C., Gagliostro, V., Bosisio, D., Del Prete, A., Tiberio, L.,
Thelen, M., and Sozzani, S. (2017). The atypical receptor CCRL2
(C-C chemokine receptor-like 2) does not act as a decoy receptor
in endothelial cells. Front. Immunol. 8: 1233.
Meder,W.,Wendland,M.,Busmann,A.,Kutzleb,C.,Spodsberg,N.,
John, H., Richter, R., Schleuder, D., Meyer, M., and Forssmann, W.G.
T.F. Fischer and A.G. Beck-Sickinger: Chemerin exploring a versatile adipokine 15
(2003). Characterization of human circulating TIG2 as a ligand for
the orphan receptor ChemR23. FEBS Lett. 555: 495499.
Migeotte, I., Franssen, J.-D., Goriely, S., Willems, F., and Parmentier,
M. (2002). Distribution and regulation of expression of the
putative human chemokine receptor HCR in leukocyte
populations. Eur. J. Immunol. 32: 494501.
Monnier, J., Lewén, S., OHara, E., Huang, K., Tu, H., Butcher, E.C., and
Zabel, B.A. (2012). Expression, regulation, and function of
atypical chemerin receptor CCRL2 on endothelial cells.
J. Immunol. 189: 956967.
Muruganandan, S., Roman, A.A., and Sinal, C.J. (2010). Role of
chemerin/CMKLR1 signaling in adipogenesis and
osteoblastogenesis of bone marrow stem cells. J. Bone Miner.
Res. 25: 222234.
Nagpal, S., Patel, S., Jacobe, H., DiSepio, D., Ghosn, C., Malhotra, M.,
Teng, M., Duvic, M., and Chandraratna, R.A. (1997). Tazarotene-
induced gene 2 (TIG2), a novel retinoid-responsive gene in skin.
J. Invest. Dermatol. 109: 9195.
Nakajima, H., Nakajima, K., Nagano, Y., Yamamoto, M., Tarutani, M.,
Takahashi, M., Takahashi, Y., and Sano, S. (2010). Circulating
level of chemerin is upregulated in psoriasis. J. Dermatol. Sci. 60:
4547.
Nakamura, N., Naruse, K., Kobayashi, Y., Miyabe, M., Saiki, T.,
Enomoto, A., Takahashi, M., and Matsubara, T. (2018). Chemerin
promotes angiogenesis in vivo. Phys. Rep. 6: e13962.
Nati, M., Haddad, D., Birkenfeld, A.L., Koch, C.A., Chavakis, T., and
Chatzigeorgiou, A. (2016). The role of immune cells in
metabolism-related liver inammation and development of
non-alcoholic steatohepatitis (NASH). Rev. Endocr. Metab.
Disord. 17: 2939.
Neuschwander-Tetri, B.A. (2017). Non-alcoholic fatty liver disease.
BMC Med. 15: 45.
Neves, K.B., Nguyen Dinh Cat, A., Alves-Lopes, R., Harvey, K.Y.,
Costa, R.M.d., Lobato, N.S., Montezano, A.C., Oliveira, A.M.d.,
Touyz, R.M., and Tostes, R.C. (2018). Chemerin receptor blockade
improves vascular function in diabetic obese mice via redox-
sensitive and Akt-dependent pathways. Am. J. Physiol. Heart
Circ. Physiol. 315: H1851H1860.
Ohira, T., Arita, M., Omori, K., Recchiuti, A., van Dyke, T.E., and
Serhan, C.N. (2010). Resolvin E1 receptor activation signals
phosphorylation and phagocytosis. J. Biol. Chem. 285:
34513461.
Oostendorp, J., Hylkema, M.N., Luinge, M., Geerlings, M., Meurs, H.,
Timens, W., Zagsma, J., Postma, D.S., Boddeke, H.W., and
Biber, K. (2004). Localization and enhanced mRNA expression of
the orphan chemokine receptor L-CCR in the lung in a murine
model of ovalbumin-induced airway inammation. J. Histochem.
Cytochem. 52: 401410.
Otero, K., Vecchi, A., Hirsch, E., Kearley, J., Vermi, W., Del Prete, A.,
Gonzalvo-Feo, S., Garlanda, C., Azzolino, O., Salogni, L., et al.
(2010). Nonredundant role of CCRL2 in lung dendritic cell
trafcking. Blood 116: 29422949.
Pachynski, R.K., Wang, P., Salazar, N., Zheng, Y., Nease, L., Rosalez, J.,
Leong, W.-I., Virdi, G., Rennier, K., Shin, W.J., et al. (2019).
Chemerin suppresses breast cancer growth by recruiting immune
effector cells into the tumor microenvironment. Front. Immunol.
10: 983.
Pachynski, R.K., Zabel, B.A., Kohrt, H.E., Tejeda, N.M., Monnier, J.,
Swanson, C.D., Holzer, A.K., Gentles, A.J., Sperinde, G.V.,
Edalati, A., et al. (2012). The chemoattractant chemerin
suppresses melanoma by recruiting natural killer cell antitumor
defenses. J. Exp. Med. 209: 14271435.
Pandey, S., Maharana, J., Li, X.X., Woodruff, T.M., and Shukla, A.K.
(2020). Emerging insights into the structure and function of
complement C5a receptors. Trends Biochem. Sci. 4: 693705.
Parlee, S.D., McNeil, J.O., Muruganandan, S., Sinal, C.J., and Goralski,
K.B. (2012). Elastase and tryptase govern TNFα-mediated
production of active chemerin by adipocytes. PLoS One 7:
e51072.
Parolini, S., Santoro, A., Marcenaro, E., Luini, W., Massardi, L.,
Facchetti, F., Communi, D., Parmentier, M., Majorana, A., Sironi,
M., et al. (2007). The role of chemerin in the colocalization of NK
and dendritic cell subsets into inamed tissues. Blood 109:
36253632.
Patel, L., Charlton, S.J., Chambers, J.K., and Macphee, C.H. (2001).
Expression and functional analysis of chemokine receptors in
human peripheral blood leukocyte populations. Cytokine 14:
2736.
Pazgier, M., Ericksen, B., Ling, M., Toth, E., Shi, J., Li, X., Galliher-
Beckley, A., Lan, L., Zou, G., Zhan, C., et al. (2013). Structural and
functional analysis of the pro-domain of human cathelicidin,
LL-37. Biochemistry 52: 15471558.
Peyrassol, X., Laeremans, T., Gouwy, M., Lahura, V., Debulpaep, M.,
van Damme, J., Steyaert, J., Parmentier, M., and Langer, I. (2016).
Development by genetic immunization of monovalent antibodies
(nanobodies) behaving as antagonists of the human ChemR23
receptor. J. Immunol. 196: 28932901.
Rawlings, N.D. and Barrett, A.J. (1990). Evolution of proteins of the
cystatin superfamily. J. Mol. Evol. 30: 6071.
Regan-Komito, D., Valaris, S., Kapellos, T.S., Recio, C., Taylor, L.,
Greaves, D.R., and Iqbal, A.J. (2017). Absence of the non-
signalling chemerin receptor CCRL2 exacerbates acute
inammatory responses in vivo. Front. Immunol. 8: 1621.
Reilly, S.M. and Saltiel, A.R. (2017). Adapting to obesity with
adipose tissue inammation. Nat. Rev. Endocrinol. 13:
633643.
Rennier, K., Shin, W.J., Krug, E., Virdi, G., and Pachynski, R.K. (2020).
Chemerin reactivates PTEN and suppresses PD-L1 in tumor cells
via modulation of a novel CMKLR1-mediated signaling cascade.
Clin. Cancer Res. 26: 50195035.
Reverchon, M., Cornuau, M., Ramé, C., Guerif, F., Royère, D., and
Dupont, J. (2012). Chemerin inhibits IGF-1-induced progesterone
and estradiol secretion in human granulosa cells. Hum. Reprod.
27: 17901800.
Reyes, N., Benedetti, I., Rebollo, J., Correa, O., and Geliebter, J. (2017).
Atypical chemokine receptor CCRL2 is overexpressed in prostate
cancer cells. J. Biomed. Res., https://doi.org/10.7555/JBR.32.
20170057 (Epub ahead of print).
Rocha, V.Z. and Libby, P. (2009). Obesity, inammation, and
atherosclerosis. Nat. Rev. Cardiol. 6: 399409.
Roh, S., Song,S.-H., Choi, K.-C.,Katoh, K., Wittamer,V., Parmentier, M.,
and Sasaki, S.(2007). Chemerina new adipokine that modulates
adipogenesis via its own receptor. Biochem. Biophys. Res.
Commun. 362: 10131018.
Rouger, L., Denis, G.R., Luangsay, S., and Parmentier, M. (2013).
ChemR23 knockout mice display mild obesity but no decit in
adipocyte differentiation. J. Endocrinol. 219: 279289.
Rourke, J. (2015). The chemerin receptor GPR1 signals through a RhoA/
ROCK pathway and contributes to glucose homeostasis in obese
mice, Ph.D. thesis, Halifax, Canada: Dalhousie University.
16 T.F. Fischer and A.G. Beck-Sickinger: Chemerin exploring a versatile adipokine
Rourke, J.L., Dranse, H.J., and Sinal, C.J. (2015). CMKLR1 and GPR1
mediate chemerin signaling through the RhoA/ROCK pathway.
Mol. Cell. Endocrinol. 417: 3651.
Rourke, J.L., Muruganandan, S., Dranse, H.J., McMullen, N.M., and
Sinal, C.J. (2014). Gpr1 is an active chemerin receptor inuencing
glucose homeostasis in obese mice. J. Endocrinol. 222: 201215.
Sablinski, T. (2019). A multicenter, double-masked, parallel-group,
vehicle-controlled study to assess the efcacy and safety of
RX-10045 nanomicellar ophthalmic solution for treatment of
ocular inammation and pain in subjects undergoing cataract
surgery. NCT02329743, ATR-45-1401, Available at:
<https://clinicaltrials.gov/ct2/show/NCT02329743>.
Saeedi, P., Petersohn, I., Salpea, P., Malanda, B., Karuranga, S.,
Unwin, N., Colagiuri, S., Guariguata, L., Motala, A.A.,
Ogurtsova, K., et al. (2019). Global and regional diabetes
prevalence estimates for 2019 and projections for 2030 and 2045:
results from the International Diabetes Federation Diabetes Atlas,
9th edition. Diabetes Res. Clin. Pract. 157: 107843.
Sarmadi, P., Tunali, G., Esendagli-Yilmaz, G., Yilmaz, K.B., and
Esendagli, G. (2015). CRAM-A indicates IFN-γ-associated
inammatory response in breast cancer. Mol. Immunol. 68:
692698.
Scheja, L. and Heeren, J. (2019). The endocrine function of adipose
tissues in health and cardiometabolic disease. Nat. Rev.
Endocrinol. 15: 507524.
Schioppa, T., Sozio, F., Barbazza, I., Scutera, S., Bosisio, D., Sozzani, S.,
and Del Prete, A. (2020). Molecular basis for CCRL2 regulation of
leukocyte migration. Front. Cell Dev. Biol. 8: 615031.
Schultz, S., Saalbach, A., Heiker, J.T., Meier,R., Ze llmann, T., Simon, J.C.,
and Beck-Sickinger, A.G. (2013). Proteolytic activation of
prochemerin by kallikrein 7 breaks an ionic linkage and results in
C-terminal rearrangement. Biochem. J. 452: 271280.
Scola, A.-M., Johswich, K.-O., Morgan, B.P., Klos, A., and Monk, P.N.
(2009). The human complement fragment receptor, C5L2, is a
recycling decoy receptor. Mol. Immunol. 46: 11491162.
Sell, H., Divoux, A., Poitou,C., Basdevant,A., Bouillot, J.-L., Bedossa, P.,
Tordjman,J., Eckel, J., and Clément, K. (2010).Chemerin correlates
with markersfor fatty liver in morbidlyobese patients and strongly
decreases after weight loss induced by bariatric surgery. J. Clin.
Endocrinol. Metab. 95: 28922896.
Seran, D.S., Allyn, B., Sassano, M.F., Timoshchenko, R.G., Mattox, D.,
Brozowski, J.M., Siderovski, D.P., Truong, Y.K., Esserman, D.,
Tarrant, T.K., et al. (2019). Chemerin-activated functions of
CMKLR1 are regulated by G protein-coupled receptor kinase 6
(GRK6) and β-arrestin 2 in inammatory macrophages. Mol.
Immunol. 106: 1221.
Shimamura, K., Matsuda, M., Miyamoto, Y., Yoshimoto, R., Seo, T.,
and Tokita, S. (2009). Identication of a stable chemerin analog
with potent activity toward ChemR23. Peptides 30: 15291538.
Shin, W.J., Zabel, B.A., and Pachynski, R.K. (2018). Mechanisms and
functions of chemerin in cancer: potential roles in therapeutic
intervention. Front. Immunol. 9: 2772.
Siciliano, S.J., Rollins, T.E., DeMartino, J., Konteatis, Z., Malkowitz, L.,
van Riper, G., Bondy, S., Rosen, H., and Springer, M.S. (1994).
Two-site binding of C5a by its receptor: an alternative binding
paradigm for G protein-coupled receptors. Proc. Natl. Acad. Sci.
U.S.A. 91: 12141218.
Sledzinski, T., Korczynska, J., Hallmann, A., Kaska, L., Proczko-
Markuszewska, M., Stefaniak, T., Sledzinski, M., and
Swierczynski, J. (2013). The increase of serum chemerin
concentration is mainly associated with the increase of body
mass index in obese, non-diabetic subjects. J. Endocrinol. Invest.
36: 428434.
Smolen, J.S., Aletaha, D., and McInnes, I.B. (2016). Rheumatoid
arthritis. Lancet 388: 20232038.
Sriram, K. and Insel, P.A. (2018). G protein-coupled receptors as
targets for approved drugs: how many targets and how many
drugs? Mol. Pharmacol. 93: 251258.
Stefanov, T., Blüher, M., Vekova, A., Bonova, I., Tzvetkov, S.,
Kurktschiev, D., and Temelkova-Kurktschiev, T. (2014).
Circulating chemerin decreases in response to a combined
strength and endurance training. Endocrine 45: 382391.
Taheri Chadorneshin, H., Cheragh-Birjandi, S., Goodarzy, S., and
Ahmadabadi, F. (2019). The impact of high intensity interval
training on serum chemerin, tumor necrosis factor-alpha and
insulin resistance in overweight women. Obes. Med. 14: 100101.
Takahashi, M., Takahashi, Y., Takahashi, K., Zolotaryov, F.N.,
Hong, K.S., Kitazawa, R., Iida, K., Okimura, Y., Kaji, H.,
Kitazawa, S., et al. (2008). Chemerin enhances insulin signaling
and potentiates insulin-stimulated glucose uptake in 3T3-L1
adipocytes. FEBS Lett. 582: 573578.
Tang, M., Huang, C., Wang, Y.-F., Ren, P.-G., Chen, L., Xiao, T.-X.,
Wang, B.-B., Pan, Y.-F., Tsang, B.K., Zabel, B.A., et al. (2016).
CMKLR1 deciency maintains ovarian steroid production in mice
treated chronically with dihydrotestosterone. Sci. Rep. 6: 21328.
Thiele, S. and Rosenkilde, M.M. (2014). Interaction of chemokines
with their receptorsfrom initial chemokine binding to receptor
activating steps. Curr. Med. Chem. 21: 35943614.
Tu, J., Yang, Y., Zhang, J., Lu, G., Ke, L., Tong, Z., Kasimu, M., Hu, D.,
Xu, Q., and Li, W. (2020). Regulatory effect of chemerin and
therapeutic efcacy of chemerin-9 in pancreatogenic diabetes
mellitus. Mol. Med. Rep. 21: 981988.
Tümmler, C., Snapkov, I., Wickström, M., Moens, U., Ljungblad, L.,
Maria Elfman, L.H., Winberg, J.-O., Kogner, P., Johnsen, J.I., and
Sveinbjørnsson, B. (2017). Inhibition of chemerin/CMKLR1 axis in
neuroblastoma cells reduces clonogenicity and cell viability
in vitro and impairs tumor growth in vivo. Oncotarget 8:
9513595151.
van der Vorst, E.P.C., Mandl, M., Müller, M., Neideck, C., Jansen, Y.,
Hristov, M., Gencer, S., Peters, L.J.F., Meiler, S., Feld, M., et al.
(2019). Hematopoietic ChemR23 (chemerin receptor 23) fuels
atherosclerosis by sustaining an M1 macrophage-phenotype and
guidance of plasmacytoid dendritic cells to murine lesions-brief
report. Arterioscler. Thromb. Vasc. Biol. 39: 685693.
Veale, D.J. and Fearon, U. (2018). The pathogenesis of psoriatic
arthritis. Lancet 391: 22732284.
Wang, L., Zhong, Y., Ding, Y., Shi, X., Huang, J., and Zhu, F. (2014).
Elevated serum chemerin in Chinese women with
hyperandrogenic PCOS. Gynecol. Endocrinol. 30: 746750.
Watts, S.W., Dorrance, A.M., Penfold, M.E., Rourke, J.L., Sinal, C.J.,
Seitz, B., Sullivan, T.J., Charvat, T.T., Thompson, J.M., Burnett, R.,
et al. (2013). Chemerin connects fat to arterial contraction.
Arterioscler. Thromb. Vasc. Biol. 33: 13201328.
Weigert,J.,Neumeier,M.,Wanninger,J.,Filarsky,M.,Bauer,S.,Wiest,R.,
Farkas, S., Scherer, M.N., Schäfer, A., Aslanidis, C., et al. (2010).
Systemic chemerin is related to inammation rather than obesity in
type 2 diabetes. Clin. Endocrinol. 72: 342348.
Weiß, E. and Kretschmer, D. (2018). Formyl-peptide receptors in
infection, inammation, and cancer. Trends Immunol. 39:
815829.
T.F. Fischer and A.G. Beck-Sickinger: Chemerin exploring a versatile adipokine 17
Wittamer, V., Bondue, B., Guillabert, A., Vassart, G., Parmentier, M.,
and Communi, D. (2005). Neutrophil-mediated maturation of
chemerin: a link between innate and adaptive immunity.
J. Immunol. 175: 487493.
Wittamer, V., Franssen, J.-D., Vulcano, M., Mirjolet, J.-F., Le Poul, E.,
Migeotte, I., Brezillon, S., Tyldesley, R., Blanpain, C., Detheux,M.,
et al. (2003). Specic recruitment of antigen-presenting cells by
chemerin, a novel processed ligand from human inammatory
uids. J. Exp. Med. 198: 977985.
Wittamer, V., Gregoire, F., Robberecht, P., Vassart, G., Communi, D.,
and Parmentier, M. (2004). The C-terminal nonapeptide of
mature chemerin activates the chemerin receptor with low
nanomolar potency. J. Biol. Chem. 279: 99569962.
Yamaguchi, Y., Du, X.-Y., Zhao, L., Morser, J., and Leung, L.L.K. (2011).
Proteolytic cleavage of chemerin protein is necessary for
activation to the active form, Chem157S, which functions as a
signaling molecule in glioblastoma. J. Biol. Chem. 286:
3951039519.
Yang, Y.-L., Ren, L.-R., Sun, L.-F., Huang, C., Xiao, T.-X., Wang, B.-B.,
Chen, J., Zabel, B.A., Ren, P., and Zhang, J.V. (2016). The role of
GPR1 signaling in mice corpus luteum. J. Endocrinol. 230:
5565.
Yang, Z., Han, S., Keller, M., Kaiser, A., Bender, B.J., Bosse, M.,
Burkert, K., Kögler, L.M., Wiing, D., Bernhardt, G., et al. (2018).
Structural basis of ligand binding modes at the neuropeptide Y Y1
receptor. Nature 556: 520524.
Yu, G. (2020). Using ggtree to visualize data on tree-like structures.
Curr. Protoc. Bioinf. 69: e96.
Zabel, B.A., Allen, S.J., Kulig, P., Allen, J.A., Cichy, J., Handel, T.M., and
Butcher, E.C. (2005). Chemerin activation by serine proteases of
the coagulation, brinolytic, and inammatory cascades. J. Biol.
Chem. 280: 3466134666.
Zabel, B.A., Nakae, S., Zúñiga, L., Kim, J.-Y., Ohyama, T., Alt, C., Pan, J.,
Suto, H., Soler, D., Allen, S.J., et al. (2008). Mast cell-expressed
orphan receptor CCRL2 binds chemerin and is required for
optimal induction of IgE-mediated passive cutaneous
anaphylaxis. J. Exp. Med. 205: 22072220.
Zabel, B.A., Zuniga, L., Ohyama, T., Allen, S.J., Cichy, J., Handel, T.M.,
and Butcher, E.C. (2006). Chemoattractants, extracellular
proteases, and the integrated host defense response. Exp.
Hematol. 34: 10211032.
Zanetti, M. (2004). Cathelicidins, multifunctional peptides of the
innate immunity. J. Leukoc. Biol. 75: 3948.
Zhao, L., Yamaguchi, Y., Ge, X., Robinson, W.H., Morser, J., and Leung,
L.L.K. (2018). Chemerin 156F, generated by chymase cleavage of
prochemerin, is elevated in joint uids of arthritis patients.
Arthritis Res. Ther. 20: 132.
Zheng, C., Chen, D., Zhang, Y., Bai, Y., Huang, S., Zheng, D., Liang, W.,
She, S., Peng, X., Wang, P., et al. (2018). FAM19A1 is a new ligand
for GPR1 that modulates neural stem-cell proliferation and
differentiation. Faseb. J. 32: 58745890.
Zhou, J., Liao, D., Zhang, S., Cheng, N., He, H., and Ye, R.D. (2014).
Chemerin C9 peptide induces receptor internalization through a
clathrin-independent pathway. Acta Pharmacol. Sin. 35:
653663.
Zhuang, Y., Liu, H., Zhou, X. E., Verma, R.K., de Waal, P.W., Jang, W.,
Xu, T.-H., Wang, L., Meng, X., Zhao, G., et al. (2020). Structure of
formylpeptide receptor 2-G
i
complex reveals insights into ligand
recognition and signaling. Nat. Commun. 11: 112.
Zimny, S., Pohl, R., Rein-Fischboeck, L., Haberl, E.M., Krautbauer, S.,
Weiss, T.S., and Buechler, C. (2017). Chemokine (CC-motif)
receptor-like 2 mRNA is expressed in hepatic stellate cells and is
positively associated with characteristics of non-alcoholic
steatohepatitis in mice and men. Exp. Mol. Pathol. 103: 18.
18 T.F. Fischer and A.G. Beck-Sickinger: Chemerin exploring a versatile adipokine
... Chemerin, also known as retinoic acid receptor responder protein 2 (RARRES2) or TIG2, is an adipokine and immunomodulator encoded by the RARRES2 gene [1][2][3][4][5][6]. This adipokine is synthesized as an inactive precursor that undergoes activation and a subsequent inactivation through a series of proteolytic cleavages [7,8]. The inactive circulating precursor, chemerin163S (chem163S) in humans, is converted into a form called chem158K, which possesses 5% of fully active chemerin [9]. ...
... Chemerin is a critical player in metabolic and immune regulation, particularly in diseases which provoke systemic inflammatory conditions [8,[52][53][54]. The proteolytic cleavage and regulatory patterns of chemerin are consistent across human and mouse models, with several studies indicating that variation in inflammatory responses between conditions such as rheumatoid arthritis and osteoarthritis is linked to the generation of different chemerin forms by proteolysis [11,12]. ...
Article
Full-text available
Chemerin acts as both a chemotactic agent and an adipokine that undergoes proteolytic cleavage, converting inactive precursors into their active forms before being subsequently inactivated. Elevated chemerin levels are linked to obesity and type 2 diabetes mellitus (T2D). This study aimed to elucidate the effects of T2D and obesity on chemerin levels by comparing plasma samples from individuals with a normal weight and T2D (BMI < 25; NWD group n = 22) with those from individuals who are overweight or obese and have T2D (BMI ≥ 25; OWD group n = 39). The total chemerin levels were similar in the NWD and OWD groups, suggesting that T2D may equalize the chemerin levels irrespective of obesity status. The cleavage of chemerin has been previously linked to myocardial infarction and stroke in NWD, with potential implications for inflammation and mortality. OWD plasma exhibited lower levels of cleaved chemerin than the NWD group, suggesting less inflammation in the OWD group. Here, we showed that the interaction between obesity and T2D leads to an equalization in the total chemerin levels. The cleaved chemerin levels and the associated inflammatory state, however, differ significantly, underscoring the complex relationship between chemerin, T2D, and obesity.
... Chemerin levels in blood have a positive correlation with BMI [35,36] but, to our knowledge, there have been no reports comparing the levels of different chemerin forms in individuals with diabetes and in those without diabetes with different degrees of insulin resistance. We hypothesized that, due to higher inflammation, chemerin levels would be higher and more chemerin activation would occur in individuals with diabetes than in those without diabetes. ...
... In patients with obesity, chemerin levels are increased and more activation of chemerin occurs [7,26,50,51]. Chemerin levels are also increased in patients with metabolic syndrome and both type 1 diabetes and T2D [35,[52][53][54]. Diabetic kidney disease was associated with higher levels of serum chemerin [55]. ...
Article
Full-text available
Chemerin is a chemokine/adipokine, regulating inflammation, adipogenesis and energy metabolism whose activity depends on successive proteolytic cleavages at its C-terminus. Chemerin levels and processing are correlated with insulin resistance. We hypothesized that chemerin processing would be higher in individuals with type 2 diabetes (T2D) and in those who are insulin resistant (IR). This hypothesis was tested by characterizing different chemerin forms by specific ELISA in the plasma of 18 participants with T2D and 116 without T2D who also had their insulin resistance measured by steady-state plasma glucose (SSPG) concentration during an insulin suppression test. This approach enabled us to analyze the association of chemerin levels with a direct measure of insulin resistance (SSPG concentration). Participants were divided into groups based on their degree of insulin resistance using SSPG concentration tertiles: insulin sensitive (IS, SSPG ≤ 91 mg/dL), intermediate IR (IM, SSPG 92–199 mg/dL), and IR (SSPG ≥ 200 mg/dL). Levels of different chemerin forms were highest in patients with T2D, second highest in individuals without T2D who were IR, and lowest in persons without T2D who were IM or IS. In the whole group, chemerin levels positively correlated with both degree of insulin resistance (SSPG concentration) and adiposity (BMI). Participants with T2D and those without T2D who were IR had the most proteolytic processing of chemerin, resulting in higher levels of both cleaved and degraded chemerin. This suggests that increased inflammation in individuals who have T2D or are IR causes more chemerin processing.
... Chemerin, an immunomodulatory adipokine, was initially identified as a gene for the retinoic acid-receptor responder 2 (RARRES2) in lesions of psoriatic skin (9,10). Further investigations have revealed its involvement in various other biological processes, such as systemic inflammation, angiogenesis, and oxidative stress (11,12), all of which are suggested to contribute to the development of PE (13). ...
Article
Full-text available
Chemerin is a multifunctional adipokine associated with systemic inflammation, angiogenesis, and oxidative stress. Emerging evidence suggests a potential link between chemerin and the pathogenesis of preeclampsia (PE). In this systematic review and meta-analysis, we aimed to evaluate the serum chemerin levels in women with PE. A systematic search was conducted across Medline, Web of Science, and Embase databases from inception until April 15, 2023, to identify studies comparing serum chemerin levels in pregnant women with and without PE. A random-effects model was employed to pool the results, accounting for heterogeneity. Thirteen datasets from ten observational studies, encompassing 832 women with PE and 1298 healthy pregnant women, were analyzed. The pooled findings indicated a statistically significant elevation in serum chemerin levels in women with PE compared to controls (mean difference [MD] = 89.56 ng/mL, 95% confidence interval [CI] 62.14 - 116.98; P < 0.001; I2 = 87%). The subgroup analysis revealed consistent findings across studies that measured chemerin levels before or after the diagnosis of PE, studies that did or did not match the body mass index (BMI), and studies with varying quality scores (P values for subgroup differences were all > 0.05). Compared to controls, women with severe PE exhibited a significantly greater increase in serum chemerin levels than those with mild PE (P value for subgroup difference = 0.007). Additionally, meta-regression analysis results suggested that the mean BMI of the included pregnant women might positively modify the difference in circulating chemerin levels between women with and without PE (coefficient = 8.92; P = 0.045). In conclusion, this meta-analysis suggests a positive correlation between elevated serum chemerin levels and PE diagnosis in comparison to pregnant women without the condition.
Article
Full-text available
The integration of exercise prescriptions into cancer adjuvant therapy presents challenges stemming from the ambiguity surrounding the precise mechanism through which exercise intervention mitigates the risk of hepatocellular carcinoma (HCC) mortality and recurrence. Elucidation of this specific mechanism has substantial social and clinical implications. In this study, tumor-bearing mice engaged in voluntary wheel running exhibited a notable decrease in tumor growth, exceeding 30%. Microarray analysis revealed an upregulation of cytokine-related pathways as a potential explanation for this effect. The inclusion of granulocyte-macrophage colony-stimulating factor (GM-CSF) was found to enhance tumor cell proliferation, while the absence of GM-CSF resulted in a marked inhibition of tumor cell growth. The findings suggest that exercise-induced serum from mice can impede the proliferation of mouse tumor cells, with the adipokine chemerin inhibiting the growth factor GM-CSF. Additionally, exercise was found to stimulate chemerin secretion by brown adipose tissue. Chemerin suppression led to a reduction in the inhibition of tumor cell proliferation. The results of this study suggest that exercise may stimulate the release of adipokines from brown adipose tissue, transport them through the blood to the distant tumor microenvironment, and downregulate GM-CSF expression, alleviating tumor immunosuppression in the tumor microenvironment, thereby inhibiting at HCC progression. These findings provide a theoretical basis for incorporating exercise prescription into cancer treatment.
Article
Peripheral artery disease (PAD) biomarkers have been studied in isolation; however, an algorithm that considers a protein panel to inform PAD prognosis may improve predictive accuracy. Biomarker-based prediction models were developed and evaluated using a model development (n = 270) and prospective validation cohort (n = 277). Plasma concentrations of 37 proteins were measured at baseline and the patients were followed for 2 years. The primary outcome was 2-year major adverse limb event (MALE; composite of vascular intervention or major amputation). Of the 37 proteins tested, 6 were differentially expressed in patients with vs. without PAD (ADAMTS13, ICAM-1, ANGPTL3, Alpha 1-microglobulin, GDF15, and endostatin). Using 10-fold cross-validation, we developed a random forest machine learning model that accurately predicts 2-year MALE in a prospective validation cohort of PAD patients using a 6-protein panel (AUROC 0.84). This algorithm can support PAD risk stratification, informing clinical decisions on further vascular evaluation and management.
Article
Full-text available
Major determinants of the biological background or reserve, such as age, biological sex, comorbidities (diabetes, hypertension, obesity, etc.), and medications (e.g., anticoagulants), are known to affect outcome after traumatic brain injury (TBI). With the unparalleled data richness of coronavirus disease 2019 (COVID-19; ∼375,000 and counting!) as well as the chronic form, long-COVID, also called post-acute sequelae SARS-CoV-2 infection (PASC), publications (∼30,000 and counting) covering virtually every aspect of the diseases, pathomechanisms, biomarkers, disease phases, symptomatology, etc., have provided a unique opportunity to better understand and appreciate the holistic nature of diseases, interconnectivity between organ systems, and importance of biological background in modifying disease trajectories and affecting outcomes. Such a holistic approach is badly needed to better understand TBI-induced conditions in their totality. Here, I briefly review what is known about long-COVID/PASC, its underlying—suspected—pathologies, the pathobiological changes induced by TBI, in other words, the TBI endophenotypes, discuss the intersection of long-COVID/PASC and TBI-induced pathobiologies, and how by considering some of the known factors affecting the person's biological background and the inclusion of mechanistic molecular biomarkers can help to improve the clinical management of TBI patients.
Article
Glucagon-like peptide-1 (GLP-1), secreted by intestinal L-cells, plays a pivotal role in the modulation of β-cell insulin secretion in a glucose-dependent manner, concurrently promoting β-cell survival and β-cell mass. Notably, GLP-1 has emerged as an effective second-line treatment for type 2 diabetes mellitus, gaining further prominence for its pronounced impact on body weight reduction, positioning it as a potent antiobesity agent. However, the mechanism by which GLP-1 improves obesity remains unclear. Some reports suggest that this mechanism may be associated with the regulation of adipokine synthesis within adipose tissue. Chemerin, a multifunctional adipokine and chemokine, has been identified as a pivotal player in adipocyte differentiation and the propagation of systemic inflammation, a hallmark of obesity. This review provides a comprehensive overview of the mechanisms by which GLP-1 and chemerin play crucial roles in obesity and obesity-related diseases. It discusses well-established aspects, such as their effects on food intake and glycolipid metabolism, as well as recent insights, including their influence on macrophage polarization and adipose tissue thermogenesis. GLP-1 has been shown to increase the population of anti-inflammatory M2 macrophages, promote brown adipose tissue thermogenesis, and induce the browning of white adipose tissue. In contrast, chemerin exhibits opposite effects in these processes. In addition, recent research findings have demonstrated the promising potential of GLP-1-based therapies in directly or indirectly regulating chemerin expression. In an intriguing reciprocal relationship, chemerin has also been newly identified as a negative regulator of GLP-1 in vivo. This review delineates the intricate interplay between GLP-1 and chemerin, unraveling their mutual inhibitory interactions. To the best of our knowledge, no previous reviews have focused on this specific topic, making this review particularly valuable in expanding our understanding of the endocrine mechanisms of obesity and providing potential strategies for the treatment of obesity and related diseases.
Article
Full-text available
Rheumatoid arthritis (RA) is a chronic inflammatory disease manifested by joint involvement, extra-articular manifestations, and general symptoms. Adipose tissue, previously perceived as an inert energy storage organ, has been recognised as a significant contributor to RA pathophysiology. Adipokines modulate immune responses, inflammation, and metabolic pathways in RA. Although most adipokines have a pro-inflammatory and aggravating effect on RA, some could counteract this pathological process. The coexistence of RA and sarcopenic obesity (SO) has gained attention due to its impact on disease severity and outcomes. Sarcopenic obesity further contributes to the inflammatory milieu and metabolic disturbances. Recent research has highlighted the intricate crosstalk between adipose tissue and skeletal muscle, suggesting potential interactions between these tissues in RA. This review summarizes the roles of adipokines in RA, particularly in inflammation, immune modulation, and joint destruction. In addition, it explores the emerging role of adipomyokines, specifically irisin and myostatin, in the pathogenesis of RA and their potential as therapeutic targets. We discuss the therapeutic implications of targeting adipokines and adipomyokines in RA management and highlight the challenges and future directions for research in this field.
Article
Chronic rhinosinusitis (CRS) is an inflammatory disease of paranasal sinuses. This study is formulated to explore the roles of pro‐inflammatory factors Chemerin and interleukin‐17 (IL‐17) in CRS. Patients suffering from CRS without/with nasal polyps (CRSsNP/CRSwNP), along with volunteers, were recruited. CRS rabbit models were constructed by Staphylococcus aureus infection and rabbits were injected with lentiviral vectors of short hairpin RNA‐targeting Chemerin (shChemerin), followed by micro‐computed tomography (CT) scan. Levels of Chemerin and IL‐17 were determined, and histopathological lesions were observed in subjects and CRS rabbits. Correlations between Chemerin/IL‐17 level and Lund‐Mackay/Lund‐Kennedy scores of subjects and the predictive value of Chemerin or IL‐17 for CRS were analyzed. In CRS patients and rabbits, inflammatory degrees and the level of Chemerin/IL‐17 were increased in pathological tissues or plasma, while Chemerin silencing alleviated CRS symptoms of CRS rabbits. Chemerin and IL‐17 were mainly located in the immune cells of pathological tissues and presented the positive correlation with Lund‐Mackay/Lund‐Kennedy score of CRS patients. Also, they showed high predictive value for CRS. Micro‐CT scan uncovered that CRS rabbits had increased bone remodeling, which was alleviated by Chemerin silencing. Collectively, Chemerin and IL‐17 are potential predictors and Chemerin silencing alleviates inflammatory response and bone remodeling in chronic rhinosinusitis.
Article
Full-text available
Chemerin is a small chemotactic protein and a modulator of the innate immune system. Its activity is mainly mediated by the chemokine-like receptor 1 (CMKLR1), a receptor expressed by natural killer cells, dendritic cells, and macrophages. Downregulation of chemerin is part of the immune evasion strategy exploited by several cancer types, including melanoma, breast cancer, and hepatocellular carcinoma. Administration of chemerin can potentially counteract these effects, but synthetically accessible, metabolically stable analogs are required. Other tumors display overexpression of CMKLR1, offering a potential entry point for targeted delivery of chemotherapeutics. Here, we present cyclic derivatives of the chemerin C-terminus (chemerin-9), the minimal activation sequence of chemerin. Chemerin-9 derivatives that were cyclized through positions four and nine retained activity while displaying full stability in blood plasma for more than 24 h. Therefore, these peptides could be used as a drug shuttle system to target cancer cells as demonstrated here by methotrexate conjugates.
Article
Full-text available
Proteins are essential to life, and understanding their structure can facilitate a mechanistic understanding of their function. Through an enormous experimental effort1–4, the structures of around 100,000 unique proteins have been determined5, but this represents a small fraction of the billions of known protein sequences6,7. Structural coverage is bottlenecked by the months to years of painstaking effort required to determine a single protein structure. Accurate computational approaches are needed to address this gap and to enable large-scale structural bioinformatics. Predicting the 3-D structure that a protein will adopt based solely on its amino acid sequence, the structure prediction component of the ‘protein folding problem’8, has been an important open research problem for more than 50 years9. Despite recent progress10–14, existing methods fall far short of atomic accuracy, especially when no homologous structure is available. Here we provide the first computational method that can regularly predict protein structures with atomic accuracy even where no similar structure is known. We validated an entirely redesigned version of our neural network-based model, AlphaFold, in the challenging 14th Critical Assessment of protein Structure Prediction (CASP14)15, demonstrating accuracy competitive with experiment in a majority of cases and greatly outperforming other methods. Underpinning the latest version of AlphaFold is a novel machine learning approach that incorporates physical and biological knowledge about protein structure, leveraging multi-sequence alignments, into the design of the deep learning algorithm.
Article
Full-text available
Tight regulation of cytokines is essential for the initiation and resolution of inflammation. Chemerin, a mediator of innate immunity, mainly acts on chemokine-like receptor 1 (CMKLR1) to induce the migration of macrophages and dendritic cells. The role of the second chemerin receptor, G protein-coupled receptor 1 (GPR1), is still unclear. Here we demonstrate that GPR1 shows ligand-induced arrestin3 recruitment and internalization. The chemerin C-terminus triggers this activation by folding into a loop structure, binding to aromatic residues in the extracellular loops of GPR1. While this overall binding mode is shared between GPR1 and CMKLR1, differences in their respective extracellular loop 2 allowed for the design of the first GPR1-selective peptide. However, our results suggest that ligand-induced arrestin recruitment is not the only mode of action of GPR1. This receptor also displays constitutive internalization, which allows GPR1 to internalize inactive peptides efficiently by an activation-independent pathway. Our results demonstrate that GPR1 takes a dual role in regulating chemerin activity: as a signaling receptor for arrestin-based signaling on one hand, and as a scavenging receptor with broader ligand specificity on the other. Graphic abstract
Article
Full-text available
CCRL2 is a seven-transmembrane domain receptor that belongs to the chemokine receptor family. At difference from other members of this family, CCRL2 does not promote chemotaxis and shares structural features with atypical chemokine receptors (ACKRs). However, CCRL2 also differs from ACKRs since it does not bind chemokines and is devoid of scavenging functions. The only commonly recognized CCRL2 ligand is chemerin, a non-chemokine chemotactic protein. CCRL2 is expressed both by leukocytes and non-hematopoietic cells. The genetic ablation of CCRL2 has been instrumental to elucidate the role of this receptor as positive or negative regulator of inflammation. CCRL2 modulates leukocyte migration by two main mechanisms. First, when CCRL2 is expressed by barrier cells, such endothelial, and epithelial cells, it acts as a presenting molecule, contributing to the formation of a non-soluble chemotactic gradient for leukocytes expressing CMKLR1, the functional chemerin receptor. This mechanism was shown to be crucial in the induction of NK cell-dependent immune surveillance in lung cancer progression and metastasis. Second, by forming heterocomplexes with other chemokine receptors. For instance, CCRL2/CXCR2 heterodimers were shown to regulate the activation of β2-integrins in mouse neutrophils. This mini-review summarizes the current understanding of CCRL2 biology, based on experimental evidence obtained by the genetic deletion of this receptor in in vivo experimental models. Further studies are required to highlight the complex functional role of CCRL2 in different organs and pathological conditions.
Article
Full-text available
Murine chemerin is C-terminally processed to the bioactive isoforms, muChem-156 and muChem-155, among which the longer variant protects from hepatocellular carcinoma (HCC). However, the role of muChem-155 is mostly unknown. Here, we aimed to compare the effects of these isoforms on the proliferation, migration and the secretome of the human hepatocyte cell lines HepG2 and Huh7 and the murine Hepa1-6 cell line. Therefore, huChem-157 and -156 were overexpressed in the human cells, and the respective murine variants, muChem-156 and -155, in the murine hepatocytes. Both chemerin isoforms produced by HepG2 and Hepa1-6 cells activated the chemerin receptors chemokine-like receptor 1 (CMKLR1) and G protein-coupled receptor 1 (GPR1). HuChem-157 was the active isoform in the Huh7 cell culture medium. The potencies of muChem-155 and muChem-156 to activate human GPR1 and mouse CMKLR1 were equivalent. Human CMKLR1 was most responsive to muChem-156. Chemerin variants showed no effect on cell viability and proliferation. Activation of the mitogen-activated protein kinases Erk1/2 and p38, and protein levels of the epithelial–mesenchymal transition marker, E-cadherin, were not regulated by the chemerin variants. Migration was reduced in HepG2 and Hepa1-6 cells by the longer isoform. Protective effects of chemerin in HCC include the modulation of cytokines but huChem-156 and huChem-157 overexpression did not change IL-8, CCL20 or osteopontin in the hepatocytes. The conditioned medium of the transfected hepatocytes failed to alter these soluble factors in the cell culture medium of peripheral blood mononuclear cells (PBMCs). Interestingly, the cell culture medium of Huh7 cells producing the inactive variant huChem-155 reduced CCL2 and IL-8 in PBMCs. To sum up, huChem-157 and muChem-156 inhibited hepatocyte migration and may protect from HCC metastasis. HuChem-155 was the only human isoform exerting anti-inflammatory effects on immune cells.
Article
Full-text available
Background: The role of adipokines in the development of atherosclerosis (AS) has received increasing attention in recent years. This study aimed to explore the effects of chemerin on the functions of human endothelial progenitor cells (EPCs) and to investigate its role in lipid accumulation in ApoE-knockout (ApoE-/-) mice. Methods: EPCs were cultured and treated with chemerin together with the specific p38 mitogen-activated protein kinase (MAPK) inhibitor SB 203580 in a time- and dose-dependent manner. Changes in migration, adhesion, proliferation and the apoptosis rate of EPCs were detected. ApoE-/- mice with high-fat diet-induced AS were treated with chemerin with or without SB 203580. Weights were recorded, lipid indicators were detected, and tissues sections were stained. Results: The data showed that chemerin enhanced the adhesion and migration abilities of EPCs, and reduced the apoptosis ratio and that this effect might be mediated through the p38 MAPK pathway. Additionally, chemerin increased the instability of plaques. Compared with the control group and the inhibitor group, ApoE-/- mice treated with chemerin protein had more serious arterial stenosis, higher lipid contents in plaques and decreased collagen. Lipid accumulation in the liver and kidney and inflammation in the hepatic portal area were enhanced by treatment with chemerin, and the size of adipocytes also increased after chemerin treatment. In conclusion, chemerin can enhance the adhesion and migration abilities of human EPCs and reduce the apoptosis ratio. In animals, chemerin can increase lipid accumulation in atherosclerotic plaques and exacerbate plaques instability. At the same time, chemerin can cause abnormal lipid accumulation in the livers and kidneys of model animals. After specifically blocking the p38 MAPK pathway, the effect of chemerin was reduced. Conclusions: In conclusion, this study showed that chemerin enhances the adhesion and migration abilities of EPCs and increases the instability of plaques and abnormal lipid accumulation in ApoE-/- mice. Furthermore, these effects might be mediated through the p38 MAPK pathway.
Article
Chemerin, an adipocyte-secreted adipokine, is hypothesized to participate in energy homeostasis and glucoregulation. However, the physiologic effect of endogenous chemerin on glucose metabolism is unclear. The present studies tested the hypotheses that chemerin deficiency alters whole-body glucose homeostasis following switches to high-fat diet. Adult, male chemerin knockout and C57BL/6J control wild type mice were studied. During the following 4 weeks, chow- or high-fat diet maintained chemerin knockout mice showed elevated fasting glucose levels and glucose intolerance as well as insulin intolerance. Chemerin deficiency impaired adaptation to glucose and insulin challenge, leading to increased glucose levels. Moreover, the mRNA and protein levels of GLUT4 and PGC-1α expression in both skeletal muscle and adipose tissue were significantly decreased in chemerin knockout mice relative to the wild type, respectively. Taken together, the results support the hypotheses that chemerin helps adapt glucose metabolism to changes in dietary fat and modulates glucose consumption in mice by activation of PGC-1α/GLUT4 axis. Chemerin may play a significant role in elevation of glucose uptake and insulin sensitivity to promote glucose clearance.
Article
The chemokine-like receptor 1 (CMKLR1) is a promising target for treating autoinflammatory diseases, cancer, and reproductive disorders. However, the interaction between CMKLR1 and its protein-ligand chemerin remains uncharacterized, and no drugs targeting this interaction have passed clinical trials. Here, we identify the binding mode of chemerin-9, the C-terminus of chemerin, at the receptor by combining complementary mutagenesis with structure-based modeling. Incorporating our experimental data, we present a detailed model of this binding site, including experimentally confirmed pairwise interactions for the most critical ligand residues: Chemerin-9 residue F8 binds to a hydrophobic pocket in CMKLR1 formed by the extracellular loop (ECL) 2, while F6 interacts with Y2.68, suggesting a turn-like structure. On the basis of this model, we created the first cyclic peptide with nanomolar activity, confirming the overall binding conformation. This constrained agonist mimics the loop conformation adopted by the natural ligand and can serve as a lead compound for future drug design.
Article
We previously reported 2-aminobenzoxazole analogue 1 as a potent ChemR23 inhibitor. The compound showed inhibitory activity against chemerin-induced calcium signaling through ChemR23 internalization in CAL-1 cells, which are cell lines of plasmacytoid dendric cells (pDCs). Furthermore, compound 2 inhibited chemotaxis of CAL-1 triggered by chemerin in vitro. However, we noted a difference in the ChemR23 response to our inhibitor between rodents and non-rodents in a previous study. To address this issue, we performed optimization of ChemR23 inhibitors using CAL-1 cells endogenously expressing human ChemR23 and conducted a pharmacokinetics study in cynomolgus monkeys. Various substituents at the 4-position of the benzoxazole ring exhibited potent in vitro bioactivity, while those at the 6-position were not tolerated. Among substituents, a carboxyl group was identified as key for improving the oral bioavailability in cynomolgus monkeys. Compound 38a with the acidic part changed from a tetrazole group to a 1,2,4-oxadiazol-5-one group to improve bioactivity and pharmacokinetic parameters exhibited inhibitory activity against chemerin-induced chemotaxis in vitro. In addition, we confirmed the ChemR23 internalization of pDCs by compound 38a orally administered to cynomolgus monkeys. These 2-aminobenzoxazole-based ChemR23 inhibitors may be useful as novel immunotherapeutic agents capable of suppressing the migration of pDCs, which are known to be major producers of type I interferons in the lesion area of certain autoimmune diseases, such as systemic lupus erythematosus and psoriasis.
Article
Within the family of neuropeptide Y (NPY) receptors, the Y4 receptor (Y4R) is unique as it prefers pancreatic polypeptide (PP) over NPY and peptide YY (PYY). Today, low molecular weight Y4R ligands are lacking, in particular antagonists. We synthesized a series of peptidic NPY Y4R ligands, derived from the hexapeptide acetyl-Arg-Tyr-Arg-Leu-Arg-Tyr-NH2 (1), reported to be a Y4R partial agonist with high affinity (pKi Y4R: 8.43). Peptide 1 was N-terminally extended as well as truncated, subjected to a D-amino acid scan, and Leu was replaced by different amino acids. Compounds were characterized by radioligand competition binding and functional studies (Cai²⁺ mobilization and β-arrestin 1/2 recruitment). N-terminal truncation of 1 resulted in a tetrapeptide (Arg-Leu-Arg-Tyr-NH2) being a Y4R partial agonist with retained Y4R affinity (pKi: 8.47). Remarkably, replacement of Leu in 1 and in derivatives of 1 by Trp turned Y4R agonism to antagonism, giving Y4R antagonists with pKi values ≤ 7.57.