Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Seminars in Immunology xxx (xxxx) xxx–xxx

Contents lists available at ScienceDirect

Seminars in Immunology
journal homepage: www.elsevier.com/locate/ysmim

Chemokines sound the alarmin: The role of atypical chemokine in


inflammation and cancer

Elena Monica Borronia,b, Benedetta Savinoa,b, Raffaella Bonecchia,c, Massimo Locatia,b,
a
Humanitas Clinical and Research Center, Via Manzoni 113, I-20089 Rozzano, Italy
b
Department of Medical Biotechnologies and Translational Medicine, Università degli Studi di Milano, Via fratelli Cervi, I-20090 Segrate, Italy
c
Department of Biomedical Sciences, Humanitas University, via Rita Levi Montalcini 4, I-20089 Pieve Emanuele, Italy

A R T I C LE I N FO A B S T R A C T

Keywords: As main drivers of leukocyte recruitment during inflammatory reactions, chemokines act as mediatrs of alarmins
Chemokines in priming host defense responses after tissue exposure to toxic or infectious agents, immunomediated damage,
Scavenger receptors and in inflammation-driven tumors. Chemokines can therefore be considered alarm signals generated by tissues
Atypical chemokine receptors in a broad number of conditions, and mechanisms controlling chemokines biological activities are therefore key
Cancer related inflammation
to regulate tissue reactions induced by alarmins. By transporting, presenting or scavenging different sets of
chemokines, atypical chemokine receptors represent an emerign subfamily of chemokine receptors which op-
erates in tissues as chemokine gatekeepers in order to establish and shape their gradients and coordinate leu-
kocyte recruitment.

1. Introduction receptors. The chemokine receptors family includes 10 receptors re-


cognizing CC chemokines (CCR1 to 10) and 6 recognizing CXC che-
When exposed to different types of agents perturbing homeostasis, mokines (CXCR1 to 6), plus 2 receptors (XCR1 and CX3CR1) re-
vascularized tissues react by activating a stereotyped process called cognizing the only known C and CX3C chemokine known, respectively
inflammation. Vascular permeability and adhesiveness are increased [2]. These 18 receptors represent a distinct subfamily of the G protein-
allowing circulating leukocytes to access the subendothelial space and coupled receptors (GPCR) superfamily and all share the ability to
finally their entrance and activation in the inflamed tissue. A number of trigger a Gαi-mediated signaling pathway inducing directional cell
soluble mediators are implicated in the tight coordination of this pro- movement along the ligand gradient [3].
cess. Among these, a large family of chemoattractant cytokines name Being chemokines direct mediators of leukocyte infiltration during
chemokines (from chemotactic cytokines) for their main role in selecting inflammatory reactions, mechanisms controlling their biological activ-
the appropriate leukocyte subset to be recruited and in directing them ities are key to regulate tissue reactions induced by alarmins. A first
in their directional migration in the tissue, finally resulting in their mechanism is represented by the tight control of chemokine expression,
accumulation and activation in response to the causing agent [1]. particularly for “inflammatory” chemokines, which are only produced
The chemokine system is a complex network including over 50 li- in response to inflammatory and immune stimuli, while “homeostatic”
gands and 25 receptors. Beyond their chemotactic activity, ligands chemokines control leukocyte homing and lymphocyte recirculation in
share a conserved protein structure called “chemokine scaffold”, which normal conditions [4]. Chemokines are also targets of post-translational
is based on two cysteine residues located at the far end of the amino- modifications that influence their functional properties, including pro-
terminus of the protein. The relative positioning of these cysteine re- cessing at the amino- and carboxyl-termini by proteases and mod-
sidues identifies two main subfamilies (CC and CXC chemokines, which ifications such as nitration and citrullination [5,6]. We will here focus
have the cysteine residues adjacent or separated by a single intervening on the non-redundant role in the control of the chemokine system
amino acid, respectively) and two minor subfamilies (C and CX3C, with played by a distinct group of “atypical” chemokine receptors [7], which
have a single cysteine residue in the amino-terminus or with three re- control the chemokine system by shaping chemokine distribution
sidues separating the cysteine tandem, respectively). This structure- (concentration and gradient) in tissues by means of clearance and
based chemokine classification is reflected in the classification of their transport mechanisms [8].


Corresponding author at: Humanitas Clinical and Research Center, Via Manzoni 113, 20089 Rozzano, Italy.
E-mail addresses: elena.borroni@unimi.it (E.M. Borroni), benedetta.savino@humanitasresearch.it (B. Savino),
raffaella.bonecchi@humanitasresearch.it (R. Bonecchi), massimo.locati@humanitasresearch.it (M. Locati).

https://doi.org/10.1016/j.smim.2018.10.005
Received 23 August 2018; Accepted 8 October 2018
1044-5323/ © 2018 Published by Elsevier Ltd.

Please cite this article as: Borroni, E.M., Seminars in Immunology, https://doi.org/10.1016/j.smim.2018.10.005
E.M. Borroni et al. Seminars in Immunology xxx (xxxx) xxx–xxx

Fig. 1. ACKRs: tissue distribution, ligand specificity, and biological functions. ACKRs recognize chemokines and control their distribution in tissues, thus operating as
a prominent regulatory mechanisms of the chemokine system. Chemokines are color-coded as pro-inflammatory (red), homeostatic (green) and those with mixed
function (yellow). Non-chemokine ligands and reported in grey.

2. Structural and functional properties of ACKRs 2.2. Ligands

As mentioned, conventional chemokine receptors represent a well- Consistent with their regulatory activity on a highly promiscuous
characterized GPCR subfamily able to directly induce leukocyte mi- system, all ACKRs recognizes more than one ligand, and some show a
gration. More recently, a distinct subfamily of chemokine receptors remarkably broad number of ligands, such as in case of ACKR1 and
involved in the fine-tuning chemokine-based responses in both ACKR2 [33]. Again at variance with conventional chemokine receptors,
homeostatic and inflammatory contexts has been recognized. These are ACKRs binding specificity may not be restricted by the structural
referred to as “atypical” chemokine receptors (ACKRs) as specific properties of the ligand, so that some ACKRs are selective for CC or CXC
structural determinants cause either their apparent inability to signal or chemokines (ACKR2 and ACKR3, which are restricted to CC and CXC
their ability to use alternative signalling pathways to those seen with chemokines, respectively), but other promiscuous receptors cross this
the conventional receptors [8,9]. This receptor subfamily includes at boundary and recognize ligands of different subfamilies (ACKR1 and
present four accepted members (ACKR1, formerly known as Duffy ACKR4). Finally, ACKRs also recognize viral chemokines, as for vCCL2/
Antigen for Chemokines or DARC [10]; ACKR2, formerly known as D6 vMIP-II which acts as a ligand for ACKR3 [34], and in some cases
[11,12]; ACKR3, formerly known as CXCR7 [13]; ACKR4, formerly function as pathogen coreceptors (ACKR2 as HIV-1/HIV-2 coreceptor
known as CCRL1 or CCXCKR [14]), plus a fifth member for which on astrocytes [35]; ACKR1 as cell receptor for Plasmodium vivax and
evidence are still not conclusive (ACKR5) [2] (Fig. 1). knowlesi [36,37]). Besides chemokines, ACKR recognition of chemotatic
peptides has also been observed. In particular, ACKR3 binds the
pleiotropic cytokine with chemokine-like functions Macrophage Mi-
2.1. Expression pattern gration-Inhibitory Factor [38], as well as adrenomedullin and the in-
termediate opioid peptide BAM22 [39,40], and ACKR5 binds the che-
Though ACKRs have been reported in some leukocyte subsets, in- moattractant chemerin [41]. This evidence indicates that ACKRs exert
cluding hematopoietic progenitors, B and T lymphocytes and alveolar control on leukocyte chemoattraction beyond the chemokine system
macrophages, as well as erythrocytes, differently from their conven- borders.
tional counterparts ACKRs are usually poorly expressed within the he-
matopoietic compartment. Conversely, ACKRs are usually expressed at 2.3. Structure
barrier tissues (i.e. skin, lung, gut and placenta; ACKR2 [15]) and on
stromal cells (ACKR4 [16];), as well as vascular (ACKR1 [17], ACKR3 As a consequence of the growing interest on the role of ACKRs in a
[18], ACKR5 [19]) and lymphatic (ACKR2 [20], ACKR4 [21]) en- variety of in vivo contexts, information on the structure/function re-
dothelial cells. In some cases they may display distinct functions de- lationship regulating their chemokine binding properties has recently
pending upon their expression profile, as exemplified by ACKR1 that became available. Regions involved in ligand binding in conventional
acts as a chemokine sink on erythrocytes and as a chemokine trans- chemokine receptors include the N terminus and, in particular, a sul-
porter on endothelial cells [22], but in general their expression profile fated tyrosine motif in this domain [42]. This structural determinant is
is well in line with key function of chemokine gradient remodelers. also maintained within the ACKR subfamily and appears to be equally
Emerging evidence clearly indicates that, besides tumor micro- relevant as, for example, a conserved sulfated tyrosine residue in the N-
environment cells, ACKRs expression also occurs directly on cancer terminus of ACKR2 is essential for ligand internalization. Of note, a
cells. As examples, ACKR2 expression has been reported on Kaposi sulfated peptide derived from this region is capable of binding in-
sarcoma spindle cells, breast and cervical carcinoma [23–25], and flammatory chemokines and to inhibit their interaction with their
ACKR5 is expressed on human high grade glioblastoma and breast cognate conventional receptors [43], opening new perspectives into
cancer cells [26–28]. Similarly, ACKRs expression has also been re- pharmacological targeting of the chemokine system mutated by ACKRs.
ported in hematological malignancies, such as diffuse large B cell While ACKRs seem to share with conventional receptors structural
lymphoma and chronic lymphocytic leukemia cells [29,30], and acute determinants for ligand binding, they show alterations in highly con-
lymphoid and myeloid leukemia cells [31,32]. served structural determinants in TM domains involved in “micro-

2
E.M. Borroni et al. Seminars in Immunology xxx (xxxx) xxx–xxx

switch” elements required for chemokine receptors activation [44].


According to this, ACKRs are actually classified as “not DRY” receptors,
as they either lack the DRY in TM3 (ACKR1, ACKR5) or exhibit sig-
nificant modifications in the DRYLAIV consensus (ACKR2, ACKR3,
ACKR4) [45]. ACKRs also display modifications on an highly conserved
site located in TM3/ILC2 which is important for the selectivity of re-
ceptor/G protein interaction and the efficiency of G protein activation
[45,46]. These structural variants account for the functional differences
in the signaling pathways of the two chemokine receptor families [46].
Besides TM regions, ACKRs also display an intracytoplasmatic C-
terminal domain particularly rich of Ser/Thr clusters, involved in direct
physical interaction with β-arrestins, which critically impact on their
constitutive and ligand-induced β-arrestins coupling properties [47].
Interestingly, although the role of Ser/Thr residues phosphorylation is
still debated [48,49], these residues are well conserved across species,
further supporting their potential role in regulating ACKRs signaling
properties and biological functions [45].

2.4. Signaling and trafficking properties

GPCR are currently thought to signal through G protein and β-ar- Fig. 2. Role of ACKR as regulators of tissue response to alarmins. Chemokines
can be viewed as alarm signals generated in response to alarmins released by
restin modules in balanced fashion [50]. Under specific conditions
stressed tissues. By acting at conventional chemokine receptors expressed on
several GPCR, including chemokine receptors, may adopt different G
different leukocytes subsets, chemokine coordinate their recruitment and acti-
protein or β-arrestin signaling properties and operate as biased re- vation and promote inflammation. When expressed on the same leukocyte
ceptors [50,51]. A common feature of ACKRs is their intrinsic inability target, atypical chemokine receptors may control its activation by interfering
to promote cell migration as a consequence of their negligible signaling with conventional chemokine receptors’ signalling. Atypical chemokine re-
activity via G protein-dependent signaling pathways [52–55]. While ceptors expressed on endothelial cells and other non-leukocyte cells also tune
structure/activity properties responsible for impaired G protein cou- chemokine concentrations by transporting or degrading them.
pling and signaling properties of ACKRs deserve further investigation,
growing evidence clearly indicates that their ability to interact with β- ACKRs ligands. In fact, differently from transporter and presenter,
arrestins is relevant for their signaling activities [56]. Therefore, ACKR scavenger ACKRs exhibit constitutive trafficking as a mechanism to
may be viewed as the prototype of constitutive β-arrestin-biased GPCR immediately cope with changes in chemokine needs within the tissue.
[45]. With the exception of ACKR1, whose ability to activate the β- Interestingly, constitutive trafficking correlates with ACKRs cellular
arrestin module is not defined, ligand-induced association with β-ar- localization, as scavenger ACKRs are mainly located in the cytoplasm
restins has been described for all ACKRs. Each ACKR however shows a shuttling within recycling endosomes [65,66], whereas transporter and
unique pattern of recruited β-arrestins, which can be restricted (ACKR2: presenter are preferentially distributed on the plasma membrane
β-arrestin1; ACKR3: β-arrestin2) or promiscuous (ACKR4 and ACKR5) [22,41]. Ligand stimulation promotes mechanisms to increase ACKRs
[33]. With the remarkable exception of ACKR4, β-arrestin coupling has activity by improving receptor internalization for transporters [67] and
long been considered mandatory for trafficking and scavenger function accelerating recycling routes of scavengers [59,65,68].
of all ACKRs [57–59]. Similarly to other chemoattractant receptors Under specific conditions ACKRs are also co-expressed on leuko-
[60], a role for β-arrestins in chemokine degradation by modulation of cytes together with their conventional counterparts, and in these cases
the intracellular receptor transport has been demonstrated for ACKR2 they fine tune the biological response to chemokines not only seques-
[57] and ACKR3 [58,61], though recent evidence implies β-arrestin tering the ligand but also interfering with the activation of conventional
recruitment to be overlapped but functionally unrelated to the chemokine receptors [29,69–72]. The ability to interfere with conven-
scavenging activity of ACKR3 [62]. tional chemokine receptors signaling represents an additional me-
chanism used by all ACKRs to exert their regulatory activity. Interest-
2.5. Mechanism of action ingly, this effect goes beyond direct competition with cognate ligands or
ability to form oligomers with conventional counterparts, as it has been
ACKRs exert their immune-modulatory functions by acting as clearly demonstrated that some ACKRs (ACKR2 and ACKR4) impair
“checkpoints” of chemokine availability, shaping the chemokine gra- signaling activity of non-cognate conventional chemokine receptors
dient in a spatiotemporal and context-dependent manner, implying a mainly by sequestering β-arrestins and affecting the efficiency of its
key role for this receptor subfamily in the control of tissue reactions to coupling to conventional receptors [72,73].
injury and alarmins activity (Fig. 2). Depending on the mechanism used Finally, increasing evidences indicate that ACKRs functions goes
to fulfill this function, ACKRs can be divided into three main functional beyond mere chemokine regulatory activity. ACKR3 and ACKR5 re-
categories: scavengers, transporters, and presenters [33,63]. Chemo- cognize the no-chemokine ligands, and ACKR3 also acts as a mitogenic
kine scavenging represents the leading mechanism as it is shared by signal in cancer cells through a mechanism involving constitutive β-
most of ACKRs (ACKR2, ACKR3, ACKR4), while chemokine transport arrestin-independent Src-dependent transactivation of EGFR [74,75].
and presentation are restricted to ACKR1 and ACKR5, respectively.
Noteworthy, when expressed on erythrocytes ACKR1 also can act as a
chemokine sink/reservoir. A scavenger activity has been originally 3. Biological functions of ACKR
hypothesized for ACKR5, but a recent publication clearly demonstrated
that this receptor is devoid of ligand scavenging properties and ex- In the last decade accumulating evidence has clearly established a
clusively functions as an anchoring protein on the surface of endothelial key role of ACKRs as regulators of immune and inflammatory responses,
cells [64]. exerted by their ability to scavenge, transport, or store chemokines
ACKRs ability to fine tune the chemokine gradient relies on their [8,9,76]. As mentioned, they also regulate the activity of conventional
unique trafficking properties, which dictate the intracellular fate of chemokine receptors with which they share the ligands by forming

3
E.M. Borroni et al. Seminars in Immunology xxx (xxxx) xxx–xxx

heterodimers or modulating their expression levels and signaling ac- control excessive Th17 responses that can lead to immunopathology
tivity. Subsequently, the use of gene-targeted mice and the study of [98].
human ACKR genetic variants has also revealed their relevance in
tumor biology. Initially thought to be directly related to their role as 3.2. Role of ACKRs in hematopoiesis
negative regulators of inflammation, new and unexpected ACKR func-
tions have subsequently emerged. Hematopoiesis, which takes place in the bone marrow and in sec-
ondary lymphoid organs throughout the adult life, is the process by
3.1. Role of ACKRs in the regulation of immune responses which mature blood cells differentiate from a common hematopoietic
stem cells (HSC) [99]. In hematopoietic organs, HSC reside inside areas
In vivo models of acute and chronic inflammation using full Ackr1−/ called “niches” that are formed by mesenchymal cells and endothelial

mice have originally indicated a prominent proinflammatory role vessels [100]. It is known from a long time that these stromal cells
exerted by ACKR1. Indeed Ackr1−/− mice display reduced neutrophil produce cytokines that act on HSC and thus have a major impact on
recruitment in acute lung and kidney injury models, which results in hematopoiesis [101]. Chemokines and their receptors are also im-
tissue protection [77–80]. Ackr1−/− mice are also partially protected portant players in this process [102,103]. In particular, CXCL12 is
from atheroma development [81], but when subjected to high fat diet fundamental for HSC homeostasis [104], and several inflammatory
show increased adipose tissue inflammation and weight gain, despite chemokines have myelosuppressive activity and control mobilization of
having lower levels of circulating CCL2 [82]. These divergences could mature and immature leukocytes [105].
be reconciled by the fact that ACKR1 prominent function depends upon Considering their role as regulators of chemokine bioavailability, it
its expression profile. Indeed, when expressed by erythrocytes ACKR1 is not surprisingly that also ACKRs have a role in hematopoiesis. In
functions as a “sink” for inflammatory chemokines, thus limiting ex- particular, several data indicate that ACKR1 and ACKR2 are central
cessive leukocyte extravasation. Consistent with this, individuals of controllers of myeloid differentiation [106,107]. Healthy individuals of
African origin who lack ACKR1 expression on erythrocytes (referred to African ancestry bearing a specific ACKR1 variant are neutropenic [83].
as “Duffy-null”) have higher concentrations of circulating chemokines In this context, Duchene and colleagues have recently found that
[73,83]. On the contrary, when expressed on endothelial cells ACKR1 ACKR1 is expressed by bone marrow nucleated erythroid cells. Using a
supports chemokine internalization and transcytosis and presents in- conditional mouse model, they demonstrated that the expression of
flammatory chemokines on the luminal surface of vessels [22], thus ACKR1 in these cells is required for their direct interaction with HSC
promoting inflammation. and for neutrophil differentiation [108]. The role of ACKR1 in myeloid
Differently from ACKR1, ACKR2 has been reproducibly shown to differentiation has been recently confirmed in a GWAS study [109].
limit inflammation and promote its resolution. Indeed, the impaired Similarly to ACKR1, a role for ACKR2 in hematopoiesis has been
clearance of inflammatory chemokines observed in Ackr2−/− mice suggested by the observation that a polymorphism in the ACKR2 gene is
results in increased inflammation in response to infectious agents [84], linked to altered number of circulating monocytes [110], and direct
ischaemic damage [85], and in several inflammatory models related to evidence was then provided by studies performed with ACKR2−/− mice
various tissues, including skin, gut, lung, kidney, joints, and placenta that have detected increased number of circulating inflammatory
[86–89]. ACKR2 also limits the induction of adaptive immune re- monocytes [91] and increased myeloid differentiation of HSC [111].
sponses, as Ackr2−/− mice are resistant to the induction of experi- Data on the role of other ACKRs on hematopoiesis are scanty.
mental autoimmune encephalomyelitis [90], and are protected from ACKR4 was found to be a regulator of B cell differentiation, as it limits
GVHD [91]. The mechanism of the protective effect of ACKR2 in migration of B cells in the splenic interfollicular zones and, as a con-
adaptive immunity is still unclear. ACKR2 deficiency does not suppress sequence, it reduces the numbers of activated B cells [112]. Being a
autoreactive T-cell priming but enhances T-cell polarization toward high affinity receptor for CXCL12 [113], ACKR3 also likely has a role in
Th17 cells [92]. Ackr2−/− mice are also protected from renal in- hematopoiesis, though at present no data are available.
flammation and renal fibrosis in a model of diabetic nephropathy [87]
and from bleomycin-induced lung fibrosis [93]. In this latter model the 3.3. Role of ACKRs in tumor biology
mechanism of protection was correlated with increased IFNγ-producing
γδT cell influx and reduced Th17 response. Chemokines are key mediators of chronic inflammatory processes,
Opposite to ACKR2, evidence indicates a proinflammatory role for and their expression is often regulated by oncogenic pathways and
ACKR3. Endothelial cells increase ACKR3 expression during in- transcription factors deregulated in the pathogenesis of cancer, being
flammatory reactions, and lymphocytes purified from inflammatory therefore important players in both pathways linking inflammation to
bowel disease patients show enhanced ACKR3 expression [94]. ACKR3 cancer [114]. Chemokines control several aspects of tumor biology,
is also expressed by macrophages in the atherosclerotic plaque, where it including immune infiltrate at the primary tumor site, the angiogenesis
was associated with a pro-inflammatory phenotype that included pro- process, cancer cell proliferation, and migration to metastatic sites
duction of inflammatory chemokines and phagocytic activity [95]. Si- [114]. Both preclinical observations obtained in ACKR gene‐targeted
milarly, ACKR3 expressed by rheumatoid arthritis synovium promotes mice and clinical data provide evidence that the regulation exerted by
the inflammatory process by increasing angiogenesis [96] and is also ACKRs on the chemokine system has an important role in cancer
expressed by brain microvascular endothelial cells during experimental biology [115].
inflammatory conditions, such as permanent middle cerebral artery The role of ACKR1 has been evaluated in different tumor models.
occlusion and experimental autoimmune encephalomyelitis, where it ACKR1 overexpression in breast and lung cancer cell lines inhibited
favors leukocyte extravasation by enhancing leukocyte adhesion to the tumor angiogenesis and metastasis [116,117], and transgenic mice
endothelial surface [97]. Thus, ACKR3 promotes immune responses by overexpressing ACKR1 on endothelial cells displayed reduced growth
enhancing leukocyte extravasation and promoting leukocyte pro-in- and angiogenesis when injected with the human melanoma cell line
flammatory activities. Mela [118]. ACKR1 overexpression in pancreatic ductal adenocarci-
Though data on the role of ACKR4 in immune responses are scanty, noma cells also inhibited their proliferation by suppressing STAT3 ac-
this ACKR appears to be an important regulator of the adaptive immune tivation through the inhibition of CXCR2 signaling [119]. In addition,
response. Indeed, ACKR4 expression in lymph nodes is necessary for other studies indicated that ACKR1 expressed by endothelial cells can
creating a gradient of the CCR7 ligands, CCL19 and CCL21, in the inhibit metastasis, with a mechanism unrelated to its chemokine control
subcapsular sinus [14]. In addition, using ACKR4−/− mice, it was de- activity. Indeed, ACKR1 interaction with the tetraspanin CD82/KAI
monstrated that homeostatic chemokine clearance is necessary to expressed by tumor cells resulted in increased p21 levels, inducing their

4
E.M. Borroni et al. Seminars in Immunology xxx (xxxx) xxx–xxx

senescence and inhibiting lung metastasis [120,121]. Consistent with and colleagues have reported that ACKR3 inhibits breast tumor me-
its protective role, Ackr1−/− mice showed enhanced prostate cancer tastasis by decreasing CXCR4‐mediated effects, such as production of
growth correlated with increased levels of angiogenic CXC chemokines, metalloproteinase‐12 and matrix degradation [144], and ACKR3 ex-
such as CXCL8 and CXCL2 [122]. Despite this preclinical result how- pression was also correlated with a reduced metastatic phenotype in
ever, no correlation between the “Duffy-null” phenotype and the higher rhabdomyosarcomas [145]. Also the role of ACKR3 in tumor angio-
prostate cancer incidence reported in African American people was genesis is at present controversial. ACKR3 is expressed by endothelial
found [123]. progenitors and tumor endothelial cells and it has been reported to have
A protective role of ACKR1 in cancer is also suggested by data re- a proangiogenic role by inducing endothelial progenitor transen-
ferring to human tumors. In breast, thyroid, colorectal, laryngeal dothelial migration and survival [146,147]. However, an opposite role
squamous cell and pancreatic tumors, ACKR1 expression was positively for endothelial ACKR3 was found by the use of conditional knockout
correlated with a better outcome, even if the cell types expressing mice with selective depletion of the receptor in vascular endothelial
ACKR1 in these samples were not characterized [116,119,124–128]. In cells, as by decreasing CXCL12 plasma levels ACKR3 protected these
summary, evidence indicates ACKR1 as a negative regulator of tumor animals from lung metastasis in breast cancer after orthotopic injection
growth, with inhibitory effects on tumor angiogenesis and metastasis of the AT‐3 cell line [148]. ACKR3 expression by endothelial cells was
largely mediated by reducing angiogenetic chemokine levels. also correlated with a better prognosis in patients with glioblastoma
As mentioned, ACKR2 acts as a CC inflammatory chemokine sca- [149], and in a murine model of glioblastoma ACKR3 monoclonal an-
venger, inhibiting leukocytes recruitment and limiting inflammation. tibodies in combination with the chemotherapy agent temozolomide
Consistently with this, ACKR2 was found protective in different in- induced the killing of tumor and endothelial cells by NK and macro-
flammation-induced tumor models, such as the TPA/DMBA skin and phages and resulted in extended survival [150].
the AOM/DSS colon cancer models [129]. Conversely, despite its ex- These results indicate that ACKR3 has multiple roles in tumor
pression resulted in decreased inflammation, no difference was found biology, not restricted to its ability to regulate CXCL12 bioavailability
on tumor growth in the diethylnitrosamine-induced liver cancer and and CXCR4 signaling but also involving direct activation of intracellular
oral squam cancer models [130,131]. ACKR2 expression was found G protein–independent pathways that promote cell growth and sur-
downregulated in human cancer samples, such as colon adenocarci- vival. Despite its expression by tumor cells being correlated in most of
noma and Kaposi sarcoma [23,132]. In this latter tumor, ACKR2 the cases with an increase in tumor growth, contrasting results have
down‐regulation was directly dependent by the oncogenic pathway K- been reported on its role for metastatic spread, as it can protect from
Ras/B-Raf/MEK/MAPK and was particularly evident in more‐aggressive metastatic dissemination but also promote angiogenesis.
forms [23]. Similarly, expression of ACKR2, ACKR1, and ACKR4 was Few results have been published on the role of ACKR4 in cancer
correlated with a better outcome also in cervical squamous cell cancer biology. in vitro, ACKR4 overexpression inhibited proliferation of some
and gastric cancer [25,133]. Conversely, in the ApcMin/+ model of in- breast and hepatocellular carcinoma cell lines [151,152]. In vivo, beside
testinal tumor, where inflammatory chemokines are needed to establish inhibition of tumor growth, there is also evidence for reduced metas-
a mast cell-dependent process of CD8+ T-cell recruitment which med- tasis. ACKR4 overexpression in a colon cancer model reduced tumor
iate immune surveillance, ACKR2 plays a protumoral role. Similarly, in cell migration and matrigel invasion [153]. This result is in line with
breast cancer ACKR2 expression was found to be inversely correlated the protective role of ACKR4 in tumor growth and dissemination in
with lymph node metastasis and clinical stage and positively correlated human breast, hepatocellular, and colon cancer samples, in which
to disease‐free survival rate [124], and similar findings have been re- ACKR4 down‐regulation is correlated with worse outcome [151–153].
ported for human lung cancer [134]. Collectively, these results indicate Opposite results however have been found using the breast cancer cell
that the chemokine scavenger activity of ACKR2 results in protection line 4T1.2 overexpressing ACKR4, which displayed increase levels of
from cancer growth when recruited leukocytes sustain and enhance TGF‐β1, enhanced EMT, and increased lung metastasis [154].
tumor growth, while having an opposite effect in experimental settings
where a protective immune response is required. In line with this latter 4. Conclusions
scenario, we recently reported that Ackr2−/− mice are protected by
metastasis in breast cancer and melanoma models as the consequence of After infection or tissue damage, alarmins are release in the extra-
the release from the bone marrow of neutrophils with antimetastatic cellular milieu and prime host defense responses. A downstream ef-
activity in ACKR2−/− mice [111]. fector is represented by chemokines, which are recognized as the main
ACKR3 was found up‐regulated on several tumors and on tumor- drivers of leukocyte recruitment during inflammatory reactions.
associated endothelial cells [135], with a mechanism involving hy- Mechanisms controlling chemokines biological activities are therefore
poxia, DNA methylation of the tumor suppressor gene hypermethylated key to regulate tissue reactions induced by alarmins. By different me-
in cancer 1, and expression of microRNA‐430 and microRNA‐101 chanisms, including transport, presentation and scavenging, atypical
[136]. Different from the other ACKRs, ACKR3 clearly promotes tu- chemokine receptors shape chemokine gradients in tissues and are
morigenesis by increasing tumor cell proliferation and inhibiting emerging as a new family of chemokine gatekeepers in inflammation
apoptosis. In particular, in lung cancer TGF‐β1 increased ACKR3 ex- and cancer.
pression and this correlated with decreased survival rate [137], while in
breast and prostate cancers ACKR3 was reported to form heterodimers Funding sources
with EGFR and promote tumor cell proliferation in a ligand‐indepen-
dent way [74,138,139]. ACKR3 expression has also been correlated This work was supported by the The Italian Association for Cancer
with poor prognosis in renal cell carcinoma, where it promoted tumor Research (AIRC-IG 2016 #19014 to ML and #20269 to RB) and the
growth by activating the mTOR pathway [140]. Italian Ministry of Health (Ricerca Finalizzata GR-2013-02356521 to
While the role of ACKR3 in promoting cancer growth is well as- EMB and GR-2013-02356522 to BS).
sessed, the role of ACKR3 in the metastasis process is contrasting. This
receptor was found to promote metastasis in a breast cancer model References
[141] as well as in hepatocellular carcinoma through up‐regulation of
osteopontin [142], was required for bone marrow and brain invasion [1] I.F. Charo, R.M. Ransohoff, The many roles of chemokines and chemokine re-
and for local tumor growth in a disseminated in vivo lymphoma model ceptors in inflammation, N. Engl. J. Med. 354 (6) (2006) 610–621.
[2] F. Bachelerie, A. Ben-Baruch, A.M. Burkhardt, C. Combadiere, J.M. Farber,
[30], and was also reported to regulate CXCR4‐mediated transen- G.J. Graham, R. Horuk, A.H. Sparre-Ulrich, M. Locati, A.D. Luster, A. Mantovani,
dothelial migration of tumor cells [143]. On the contrary, Hernandez

5
E.M. Borroni et al. Seminars in Immunology xxx (xxxx) xxx–xxx

K. Matsushima, P.M. Murphy, R. Nibbs, H. Nomiyama, C.A. Power, A.E. Proudfoot, squamous cell cancer, Gynecol. Oncol. 130 (1) (2013) 181–187.
M.M. Rosenkilde, A. Rot, S. Sozzani, M. Thelen, O. Yoshie, A. Zlotnik, [26] F. Yin, Z. Xu, Z. Wang, H. Yao, Z. Shen, F. Yu, Y. Tang, D. Fu, S. Lin, G. Lu,
International Union of basic and clinical pharmacology. [corrected]. LXXXIX. H.F. Kung, W.S. Poon, Y. Huang, M.C. Lin, Elevated chemokine CC-motif receptor-
Update on the extended family of chemokine receptors and introducing a new like 2 (CCRL2) promotes cell migration and invasion in glioblastoma, Biochem.
nomenclature for atypical chemokine receptors, Pharmacol. Rev. 66 (1) (2014) Biophys. Res. Commun. 429 (3-4) (2012) 168–172.
1–79. [27] P. Sarmadi, G. Tunali, G. Esendagli-Yilmaz, K.B. Yilmaz, G. Esendagli, CRAM-a
[3] M. Thelen, Dancing to the tune of chemokines, Nat. Immunol. 2 (2) (2001) indicates IFN-gamma-associated inflammatory response in breast cancer, Mol.
129–134. Immunol. 68 (2 Pt C) (2015) 692–698.
[4] A. Mantovani, The chemokine system: redundancy for robust outputs, Immunol. [28] L.P. Wang, J. Cao, J. Zhang, B.Y. Wang, X.C. Hu, Z.M. Shao, Z.H. Wang, Z.L. Ou,
Today 20 (6) (1999) 254–257. The human chemokine receptor CCRL2 suppresses chemotaxis and invasion by
[5] P. Proost, T. Loos, A. Mortier, E. Schutyser, M. Gouwy, S. Noppen, C. Dillen, blocking CCL2-induced phosphorylation of p38 MAPK in human breast cancer
I. Ronsse, R. Conings, S. Struyf, G. Opdenakker, P.C. Maudgal, J. Van Damme, cells, Med. Oncol. 32 (11) (2015) 254.
Citrullination of CXCL8 by peptidylarginine deiminase alters receptor usage, [29] J. Catusse, M. Leick, M. Groch, D.J. Clark, M.V. Buchner, K. Zirlik, M. Burger, Role
prevents proteolysis, and dampens tissue inflammation, J. Exp. Med. 205 (9) of the atypical chemoattractant receptor CRAM in regulating CCL19 induced CCR7
(2008) 2085–2097. responses in B-cell chronic lymphocytic leukemia, Mol. Cancer 9 (2010) 297.
[6] B. Molon, S. Ugel, F. Del Pozzo, C. Soldani, S. Zilio, D. Avella, A. De Palma, [30] V. Puddinu, S. Casella, E. Radice, S. Thelen, S. Dirnhofer, F. Bertoni, M. Thelen,
P. Mauri, A. Monegal, M. Rescigno, B. Savino, P. Colombo, N. Jonjic, S. Pecanic, ACKR3 expression on diffuse large B cell lymphoma is required for tumor
L. Lazzarato, R. Fruttero, A. Gasco, V. Bronte, A. Viola, Chemokine nitration spreading and tissue infiltration, Oncotarget 8 (49) (2017) 85068–85084.
prevents intratumoral infiltration of antigen-specific T cells, J. Exp. Med. 208 (10) [31] R.C.C. Melo, A.L. Longhini, C.L. Bigarella, M.O. Baratti, F. Traina, P. Favaro, P. de
(2011) 1949–1962. Melo Campos, S.T. Saad, CXCR7 is highly expressed in acute lymphoblastic leu-
[7] F. Bachelerie, G.J. Graham, M. Locati, A. Mantovani, P.M. Murphy, R. Nibbs, kemia and potentiates CXCR4 response to CXCL12, PLoS One 9 (1) (2014) e85926.
A. Rot, S. Sozzani, M. Thelen, New nomenclature for atypical chemokine re- [32] H.Y. Kim, S.Y. Lee, D.Y. Kim, J.Y. Moon, Y.S. Choi, I.C. Song, H.J. Lee, H.J. Yun,
ceptors, Nat. Immunol. 15 (3) (2014) 207–208. S. Kim, D.Y. Jo, Expression and functional roles of the chemokine receptor CXCR7
[8] A. Mantovani, R. Bonecchi, M. Locati, Tuning inflammation and immunity by in acute myeloid leukemia cells, Blood Res. 50 (4) (2015) 218–226.
chemokine sequestration: decoys and more, Nat. Rev. Immunol. 6 (12) (2006) [33] A. Vacchini, M. Locati, E.M. Borroni, Overview and potential unifying themes of
907–918. the atypical chemokine receptor family, J. Leukoc. Biol. 99 (6) (2016) 883–892.
[9] R.J. Nibbs, G.J. Graham, Immune regulation by atypical chemokine receptors, [34] M. Szpakowska, N. Dupuis, A. Baragli, M. Counson, J. Hanson, J. Piette,
nature reviews, Immunology 13 (11) (2013) 815–829. A. Chevigne, Human herpesvirus 8-encoded chemokine vCCL2/vMIP-II is an
[10] I. Novitzky-Basso, A. Rot, Duffy antigen receptor for chemokines and its involve- agonist of the atypical chemokine receptor ACKR3/CXCR7, Biochem. Pharmacol.
ment in patterning and control of inflammatory chemokines, Front. Immunol. 3 114 (2016) 14–21.
(2012) 266. [35] S.J. Neil, M.M. Aasa-Chapman, P.R. Clapham, R.J. Nibbs, A. McKnight, R.A. Weiss,
[11] A.M. Fra, M. Locati, K. Otero, M. Sironi, P. Signorelli, M.L. Massardi, M. Gobbi, The promiscuous CC chemokine receptor D6 is a functional coreceptor for primary
A. Vecchi, S. Sozzani, A. Mantovani, Cutting edge: scavenging of inflammatory CC isolates of human immunodeficiency virus type 1 (HIV-1) and HIV-2 on astrocytes,
chemokines by the promiscuous putatively silent chemokine receptor D6, J. J. Virol. 79 (15) (2005) 9618–9624.
Immunol. 170 (5) (2003) 2279–2282. [36] L.H. Miller, S.J. Mason, D.F. Clyde, M.H. McGinniss, The resistance factor to
[12] G.J. Graham, D6/Ackr2, Front. Immunol. 6 (2015) 280. plasmodium vivax in blacks. The Duffy-blood-group genotype, FyFy, N. Engl. J.
[13] L. Sanchez-Martin, P. Sanchez-Mateos, C. Cabanas, CXCR7 impact on CXCL12 Med. 295 (6) (1976) 302–304.
biology and disease, Trends Mol. Med. 19 (1) (2013) 12–22. [37] L.H. Miller, S.J. Mason, J.A. Dvorak, M.H. McGinniss, I.K. Rothman, Erythrocyte
[14] M.H. Ulvmar, K. Werth, A. Braun, P. Kelay, E. Hub, K. Eller, L. Chan, B. Lucas, receptors for (Plasmodium knowlesi) malaria: Duffy blood group determinants,
I. Novitzky-Basso, K. Nakamura, T. Rulicke, R.J. Nibbs, T. Worbs, R. Forster, Science 189 (4202) (1975) 561–563.
A. Rot, The atypical chemokine receptor CCRL1 shapes functional CCL21 gradients [38] S. Alampour-Rajabi, O.El Bounkari, A. Rot, G. Muller-Newen, F. Bachelerie,
in lymph nodes, Nat. Immunol. 15 (7) (2014) 623–630. M. Gawaz, C. Weber, A. Schober, J. Bernhagen, MIF interacts with CXCR7 to
[15] Y.Martinez de la Torre, C. Buracchi, E.M. Borroni, J. Dupor, R. Bonecchi, promote receptor internalization, ERK1/2 and ZAP-70 signaling, and lymphocyte
M. Nebuloni, F. Pasqualini, A. Doni, E. Lauri, C. Agostinis, R. Bulla, D.N. Cook, chemotaxis, FASEB J. 29 (11) (2015) 4497–4511.
B. Haribabu, P. Meroni, D. Rukavina, L. Vago, F. Tedesco, A. Vecchi, S.A. Lira, [39] K.R. Klein, N.O. Karpinich, S.T. Espenschied, H.H. Willcockson, W.P. Dunworth,
M. Locati, A. Mantovani, Protection against inflammation- and autoantibody- S.L. Hoopes, E.J. Kushner, V.L. Bautch, K.M. Caron, Decoy receptor CXCR7
caused fetal loss by the chemokine decoy receptor D6, Proc. Natl. Acad. Sci. U. S A. modulates adrenomedullin-mediated cardiac and lymphatic vascular develop-
104 (7) (2007) 2319–2324. ment, Dev. Cell 30 (5) (2014) 528–540.
[16] S.A. Bryce, R.A. Wilson, E.M. Tiplady, D.L. Asquith, S.K. Bromley, A.D. Luster, [40] Y. Ikeda, H. Kumagai, A. Skach, M. Sato, M. Yanagisawa, Modulation of circadian
G.J. Graham, R.J. Nibbs, ACKR4 on stromal cells scavenges CCL19 to enable glucocorticoid oscillation via adrenal opioid-CXCR7 signaling alters emotional
CCR7-dependent trafficking of APCs from inflamed skin to lymph nodes, J. behavior, Cell 155 (6) (2013) 1323–1336.
Immunol. 196 (8) (2016) 3341–3353. [41] B.A. Zabel, S. Nakae, L. Zuniga, J.Y. Kim, T. Ohyama, C. Alt, J. Pan, H. Suto,
[17] M. Kashiwazaki, T. Tanaka, H. Kanda, Y. Ebisuno, D. Izawa, N. Fukuma, D. Soler, S.J. Allen, T.M. Handel, C.H. Song, S.J. Galli, E.C. Butcher, Mast cell-
N. Akimitsu, K. Sekimizu, M. Monden, M. Miyasaka, A high endothelial venule- expressed orphan receptor CCRL2 binds chemerin and is required for optimal
expressing promiscuous chemokine receptor DARC can bind inflammatory, but not induction of IgE-mediated passive cutaneous anaphylaxis, J. Exp. Med. 205 (10)
lymphoid, chemokines and is dispensable for lymphocyte homing under physio- (2008) 2207–2220.
logical conditions, Int. Immunol. 15 (10) (2003) 1219–1227. [42] J.P. Ludeman, M.J. Stone, The structural role of receptor tyrosine sulfation in
[18] R.D. Berahovich, B.A. Zabel, S. Lewen, M.J. Walters, K. Ebsworth, Y. Wang, chemokine recognition, Br. J. Pharmacol. 171 (5) (2014) 1167–1179.
J.C. Jaen, T.J. Schall, Endothelial expression of CXCR7 and the regulation of [43] K.D. Hewit, A. Fraser, R.J. Nibbs, G.J. Graham, The N-terminal region of the
systemic CXCL12 levels, Immunology 141 (1) (2014) 111–122. atypical chemokine receptor ACKR2 is a key determinant of ligand binding, J.
[19] J. Monnier, S. Lewen, E. O’Hara, K. Huang, H. Tu, E.C. Butcher, B.A. Zabel, Biol. Chem. 289 (18) (2014) 12330–12342.
Expression, regulation, and function of atypical chemerin receptor CCRL2 on en- [44] R. Nygaard, T.M. Frimurer, B. Holst, M.M. Rosenkilde, T.W. Schwartz, Ligand
dothelial cells, J. Immunol. 189 (2) (2012) 956–967. binding and micro-switches in 7TM receptor structures, Trends Pharmacol. Sci. 30
[20] R.J. Nibbs, E. Kriehuber, P.D. Ponath, D. Parent, S. Qin, J.D. Campbell, (5) (2009) 249–259.
A. Henderson, D. Kerjaschki, D. Maurer, G.J. Graham, A. Rot, The beta-chemokine [45] C. Cancellieri, A. Vacchini, M. Locati, R. Bonecchi, E.M. Borroni, Atypical che-
receptor D6 is expressed by lymphatic endothelium and a subset of vascular tu- mokine receptors: from silence to sound, Biochem. Soc. Trans. 41 (1) (2013)
mors, Am. J. Pathol. 158 (3) (2001) 867–877. 231–236.
[21] D. Malhotra, A.L. Fletcher, J. Astarita, V. Lukacs-Kornek, P. Tayalia, S.F. Gonzalez, [46] H. Daiyasu, W. Nemoto, H. Toh, Evolutionary analysis of functional divergence
K.G. Elpek, S.K. Chang, K. Knoblich, M.E. Hemler, M.B. Brenner, M.C. Carroll, among chemokine receptors, decoy receptors, and viral receptors, Front.
D.J. Mooney, S.J. Turley, Immunological genome project, transcriptional profiling Microbiol. 3 (2012) 264.
of stroma from inflamed and resting lymph nodes defines immunological hall- [47] F. Huttenrauch, A. Nitzki, F.T. Lin, S. Honing, M. Oppermann, Beta-arrestin
marks, Nat. Immunol. 13 (5) (2012) 499–510. binding to CC chemokine receptor 5 requires multiple C-terminal receptor phos-
[22] M. Pruenster, L. Mudde, P. Bombosi, S. Dimitrova, M. Zsak, J. Middleton, phorylation sites and involves a conserved Asp-Arg-Tyr sequence motif, J. Biol.
A. Richmond, G.J. Graham, S. Segerer, R.J. Nibbs, A. Rot, The Duffy antigen re- Chem. 277 (34) (2002) 30769–30777.
ceptor for chemokines transports chemokines and supports their promigratory [48] E. Galliera, V.R. Jala, J.O. Trent, R. Bonecchi, P. Signorelli, R.J. Lefkowitz,
activity, Nat. Immunol. 10 (1) (2009) 101–108. A. Mantovani, M. Locati, B. Haribabu, Beta-arrestin-dependent constitutive in-
[23] B. Savino, N. Caronni, A. Anselmo, F. Pasqualini, E.M. Borroni, G. Basso, ternalization of the human chemokine decoy receptor D6, J. Biol. Chem. 279 (24)
G. Celesti, L. Laghi, A. Tourlaki, V. Boneschi, L. Brambilla, M. Nebuloni, G. Vago, (2004) 25590–25597.
A. Mantovani, M. Locati, R. Bonecchi, ERK-dependent downregulation of the [49] C.V. McCulloch, V. Morrow, S. Milasta, I. Comerford, G. Milligan, G.J. Graham,
atypical chemokine receptor D6 drives tumor aggressiveness in Kaposi sarcoma, N.W. Isaacs, R.J. Nibbs, Multiple roles for the C-terminal tail of the chemokine
Cancer Immunol. Res. 2 (7) (2014) 679–689. scavenger D6, J. Biol. Chem. 283 (12) (2008) 7972–7982.
[24] A.L. Chew, W.Y. Tan, B.Y. Khoo, Potential combinatorial effects of recombinant [50] S.K. Shenoy, R.J. Lefkowitz, Beta-arrestin-mediated receptor trafficking and signal
atypical chemokine receptors in breast cancer cell invasion: a research perspec- transduction, Trends Pharmacol. Sci. 32 (9) (2011) 521–533.
tive, Biomed. Rep. 1 (2) (2013) 185–192. [51] J. Corbisier, C. Gales, A. Huszagh, M. Parmentier, J.Y. Springael, Biased signaling
[25] T. Hou, D. Liang, L. Xu, X. Huang, Y. Huang, Y. Zhang, Atypical chemokine re- at chemokine receptors, J. Biol. Chem. 290 (15) (2015) 9542–9554.
ceptors predict lymph node metastasis and prognosis in patients with cervical [52] V. Odemis, J. Lipfert, R. Kraft, P. Hajek, G. Abraham, K. Hattermann, R. Mentlein,

6
E.M. Borroni et al. Seminars in Immunology xxx (xxxx) xxx–xxx

J. Engele, The presumed atypical chemokine receptor CXCR7 signals through G(i/ [78] J.S. Lee, C.W. Frevert, M.M. Wurfel, S.C. Peiper, V.A. Wong, K.K. Ballman,
o) proteins in primary rodent astrocytes and human glioma cells, Glia 60 (3) J.T. Ruzinski, J.S. Rhim, T.R. Martin, R.B. Goodman, Duffy antigen facilitates
(2012) 372–381. movement of chemokine across the endothelium in vitro and promotes neutrophil
[53] A. Levoye, K. Balabanian, F. Baleux, F. Bachelerie, B. Lagane, CXCR7 hetero- transmigration in vitro and in vivo, J. Immunol. 170 (10) (2003) 5244–5251.
dimerizes with CXCR4 and regulates CXCL12-mediated G protein signaling, Blood [79] A. Zarbock, J. Bishop, H. Muller, M. Schmolke, K. Buschmann, H.Van Aken,
113 (24) (2009) 6085–6093. K. Singbartl, Chemokine homeostasis vs. Chemokine presentation during severe
[54] A.O. Watts, F. Verkaar, M.M. van der Lee, C.A. Timmerman, M. Kuijer, J. van acute lung injury: the other side of the Duffy antigen receptor for chemokines, Am.
Offenbeek, L.H. van Lith, M.J. Smit, R. Leurs, G.J. Zaman, H.F. Vischer, Beta- J. Physiol. Lung cell. Mol. Physiol. 298 (3) (2010) L462–71.
arrestin recruitment and G protein signaling by the atypical human chemokine [80] A. Zarbock, M. Schmolke, S.G. Bockhorn, M. Scharte, K. Buschmann, K. Ley,
decoy receptor CCX-CKR, J. Biol. Chem. 288 (10) (2013) 7169–7181. K. Singbartl, The Duffy antigen receptor for chemokines in acute renal failure: a
[55] O. De Henau, G.N. Degroot, V. Imbault, V. Robert, C. De Poorter, S. McHeik, facilitator of renal chemokine presentation, Crit. Care Med. 35 (9) (2007)
C. Gales, M. Parmentier, J.Y. Springael, Signaling properties of chemerin receptors 2156–2163.
CMKLR1, GPR1 and CCRL2, PLoS One 11 (10) (2016) e0164179. [81] W. Wan, Q. Liu, M.S. Lionakis, A.P. Marino, S.A. Anderson, M. Swamydas,
[56] V.V. Gurevich, E.V. Gurevich, Structural determinants of arrestin functions, Prog. P.M. Murphy, Atypical chemokine receptor 1 deficiency reduces atherogenesis in
Mol. Biol. Trans. Sci. 118 (2013) 57–92. ApoE-knockout mice, Cardiovasc. Res. 106 (3) (2015) 478–487.
[57] E.M. Borroni, C. Cancellieri, A. Vacchini, Y. Benureau, B. Lagane, F. Bachelerie, [82] T.W. Benson, D.S. Weintraub, M. Crowe, N.K.H. Yiew, O. Popoola, A. Pillai,
F. Arenzana-Seisdedos, K. Mizuno, A. Mantovani, R. Bonecchi, M. Locati, Beta- J. Joseph, K. Archer, C. Greenway, T.K. Chatterjee, J. Mintz, D.W. Stepp,
arrestin-dependent activation of the cofilin pathway is required for the scavenging B.K. Stansfield, W. Chen, J. Brittain, V.Y. Bogdanov, Y. Gao, J.G. Wilson, Y. Tang,
activity of the atypical chemokine receptor D6, Sci. Signal. 6 (273) (2013) S1–S3 H.W. Kim, N.L. Weintraub, Deletion of the Duffy antigen receptor for chemokines
ra30 1-11. (DARC) promotes insulin resistance and adipose tissue inflammation during high
[58] S. Rajagopal, J. Kim, S. Ahn, S. Craig, C.M. Lam, N.P. Gerard, C. Gerard, fat feeding, Mol. Cell. Endocrinol. 473 (2018) 79–88.
R.J. Lefkowitz, Beta-arrestin- but not G protein-mediated signaling by the "decoy" [83] R.E. Howes, A.P. Patil, F.B. Piel, O.A. Nyangiri, C.W. Kabaria, P.W. Gething,
receptor CXCR7, Proc. Natl. Acad. Sci. U. S. A. 107 (2) (2010) 628–632. P.A. Zimmerman, C. Barnadas, C.M. Beall, A. Gebremedhin, D. Menard,
[59] I. Comerford, S. Milasta, V. Morrow, G. Milligan, R. Nibbs, The chemokine re- T.N. Williams, D.J. Weatherall, S.I. Hay, The global distribution of the Duffy blood
ceptor CCX-CKR mediates effective scavenging of CCL19 in vitro, Eur. J. Immunol. group, Nat. Commun. 2 (2011) 266.
36 (7) (2006) 1904–1916. [84] D. Di Liberto, M. Locati, N. Caccamo, A. Vecchi, S. Meraviglia, A. Salerno,
[60] B.M. Wagener, N.A. Marjon, E.R. Prossnitz, Regulation of N-formyl peptide re- G. Sireci, M. Nebuloni, N. Caceres, P.J. Cardona, F. Dieli, A. Mantovani, Role of the
ceptor signaling and trafficking by arrestin-src kinase interaction, PLoS One 11 (1) chemokine decoy receptor D6 in balancing inflammation, immune activation, and
(2016) e0147442. antimicrobial resistance in mycobacterium tuberculosis infection, J. Exp. Med.
[61] M. Canals, D.J. Scholten, S. de Munnik, M.K. Han, M.J. Smit, R. Leurs, 205 (9) (2008) 2075–2084.
Ubiquitination of CXCR7 controls receptor trafficking, PLoS One 7 (3) (2012) [85] C. Cochain, C. Auvynet, L. Poupel, J. Vilar, E. Dumeau, A. Richart, A. Recalde,
e34192. Y. Zouggari, K.Y. Yin, P. Bruneval, G. Renault, C. Marchiol, P. Bonnin, B. Levy,
[62] B. Benredjem, M. Girard, D. Rhainds, G. St-Onge, N. Heveker, Mutational analysis R. Bonecchi, M. Locati, C. Combadiere, J.S. Silvestre, The chemokine decoy re-
of atypical chemokine receptor 3 (ACKR3/CXCR7) interaction with its chemokine ceptor D6 prevents excessive inflammation and adverse ventricular remodeling
ligands CXCL11 and CXCL12, J. Biol. Chem. 292 (1) (2017) 31–42. after myocardial infarction, Arterioscler. Thromb. Vasc. Biol. 32 (9) (2012)
[63] M.H. Ulvmar, E. Hub, A. Rot, Atypical chemokine receptors, Exp. Cell. Res. 317 (5) 2206–2213.
(2011) 556–568. [86] G.J. Graham, M. Locati, Regulation of the immune and inflammatory responses by
[64] C. Mazzotti, V. Gagliostro, D. Bosisio, A. Del Prete, L. Tiberio, M. Thelen, the’ atypical’ chemokine receptor D6, J. Pathol. 229 (2) (2013) 168–175.
S. Sozzani, The atypical receptor CCRL2 (C-C chemokine receptor-like 2) does not [87] A. Bideak, A. Blaut, J.M. Hoppe, M.B. Muller, G. Federico, N. Eltrich, H.J. Grone,
act as a decoy receptor in endothelial cells, Front. Immunol. 8 (2017) 1233. M. Locati, V. Vielhauer, The atypical chemokine receptor 2 limits renal in-
[65] R. Bonecchi, E.M. Borroni, A. Anselmo, A. Doni, B. Savino, M. Mirolo, M. Fabbri, flammation and fibrosis in murine progressive immune complex glomerulone-
V.R. Jala, B. Haribabu, A. Mantovani, M. Locati, Regulation of D6 chemokine phritis, Kidney Int. 93 (4) (2018) 826–841.
scavenging activity by ligand- and Rab11-dependent surface up-regulation, Blood [88] K. Eller, A.R. Rosenkranz, Atypical chemokine receptors-"chemokine PACMANs"
112 (3) (2008) 493–503. as new therapeutic targets in glomerulonephritis, Kidney Int. 93 (4) (2018)
[66] F. Hoffmann, W. Muller, D. Schutz, M.E. Penfold, Y.H. Wong, S. Schulz, R. Stumm, 774–775.
Rapid uptake and degradation of CXCL12 depend on CXCR7 carboxyl-terminal [89] H.M. Baldwin, M.D. Singh, V. Codullo, V. King, H. Wilson, I. McInnes,
serine/threonine residues, J. Biol. Chem. 287 (34) (2012) 28362–28377. G.J. Graham, Elevated ACKR2 expression is a common feature of inflammatory
[67] Y. Zhao, N.S. Mangalmurti, Z. Xiong, B. Prakash, F. Guo, D.B. Stolz, J.S. Lee, Duffy arthropathies, Rheumatology (Oxford) 56 (9) (2017) 1607–1617.
antigen receptor for chemokines mediates chemokine endocytosis through a [90] L. Liu, G.J. Graham, A. Damodaran, T. Hu, S.A. Lira, M. Sasse, C. Canasto-
macropinocytosis-like process in endothelial cells, PLoS One 6 (12) (2011) Chibuque, D.N. Cook, R.M. Ransohoff, Cutting edge: the silent chemokine receptor
e29624. D6 is required for generating T cell responses that mediate experimental auto-
[68] K.E. Luker, J.M. Steele, L.A. Mihalko, P. Ray, G.D. Luker, Constitutive and che- immune encephalomyelitis, J. Immunol. 177 (1) (2006) 17–21.
mokine-dependent internalization and recycling of CXCR7 in breast cancer cells to [91] B. Savino, M.G. Castor, N. Caronni, A. Sarukhan, A. Anselmo, C. Buracchi,
degrade chemokine ligands, Oncogene 29 (32) (2010) 4599–4610. F. Benvenuti, V. Pinho, M.M. Teixeira, A. Mantovani, M. Locati, R. Bonecchi,
[69] A. Chakera, R.M. Seeber, A.E. John, K.A. Eidne, D.R. Greaves, The duffy antigen/ Control of murine Ly6C(high) monocyte traffic and immunosuppressive activities
receptor for chemokines exists in an oligomeric form in living cells and func- by atypical chemokine receptor D6, Blood 119 (22) (2012) 5250–5260.
tionally antagonizes CCR5 signaling through hetero-oligomerization, Mol. [92] C.A. Hansell, L.M. MacLellan, R.S. Oldham, J. Doonan, K.J. Chapple,
Pharmacol. 73 (5) (2008) 1362–1370. E.J. Anderson, C. Linington, I.B. McInnes, R.J. Nibbs, C.S. Goodyear, The atypical
[70] R. Bonecchi, M. Locati, E. Galliera, M. Vulcano, M. Sironi, A.M. Fra, M. Gobbi, chemokine receptor ACKR2 suppresses Th17 responses to protein autoantigens,
A. Vecchi, S. Sozzani, B. Haribabu, J. Van Damme, A. Mantovani, Differential Immunol. Cell Biol. 93 (2) (2015) 167–176.
recognition and scavenging of native and truncated macrophage-derived chemo- [93] R.C. Russo, B. Savino, M. Mirolo, C. Buracchi, G. Germano, A. Anselmo,
kine (macrophage-derived chemokine/CC chemokine ligand 22) by the D6 decoy L. Zammataro, F. Pasqualini, A. Mantovani, M. Locati, M.M. Teixeira, The atypical
receptor, J. Immunol. 172 (8) (2004) 4972–4976. chemokine receptor ACKR2 drives pulmonary fibrosis by tuning influx of
[71] A.K. Singh, R.K. Arya, A.K. Trivedi, S. Sanyal, R. Baral, O. Dormond, D.M. Briscoe, CCR2(+) and CCR5(+) IFNgamma-producing gammadeltaT cells in mice,
D. Datta, Chemokine receptor trio: CXCR3, CXCR4 and CXCR7 crosstalk via American journal of physiology, Lung Cell. Mol. Physiol. 314 (6) (2018)
CXCL11 and CXCL12, Cytok. Growth Factor rev. 24 (1) (2013) 41–49. L1010–L1025.
[72] J. Vinet, M. van Zwam, I.M. Dijkstra, N. Brouwer, H.R. van Weering, A. Watts, [94] L. Werner, H. Elad, E. Brazowski, H. Tulchinsky, S. Vigodman, U. Kopylov,
M. Meijer, M.R. Fokkens, V. Kannan, D. Verzijl, H.F. Vischer, M.J. Smit, R. Leurs, Z. Halpern, H. Guzner-Gur, I. Dotan, Reciprocal regulation of CXCR4 and CXCR7
K. Biber, H.W. Boddeke, Inhibition of CXCR3-mediated chemotaxis by the human in intestinal mucosal homeostasis and inflammatory bowel disease, J. Leukoc.
chemokine receptor-like protein CCX-CKR, Br. J. Pharmacol. 168 (6) (2013) Biol. 90 (3) (2011) 583–590.
1375–1387. [95] M. Chatterjee, S.N. von Ungern-Sternberg, P. Seizer, F. Schlegel, M. Buttcher,
[73] C.A. Hansell, C.E. Hurson, R.J. Nibbs, DARC and D6: silent partners in chemokine N.A. Sindhu, S. Muller, A. Mack, M. Gawaz, Platelet-derived CXCL12 regulates
regulation? Immunol. Cell Biol. 89 (2) (2011) 197–206. monocyte function, survival, differentiation into macrophages and foam cells
[74] N. Salazar, D. Munoz, G. Kallifatidis, R.K. Singh, M. Jorda, B.L. Lokeshwar, The through differential involvement of CXCR4-CXCR7, Cell Death Dis. 6 (2015)
chemokine receptor CXCR7 interacts with EGFR to promote breast cancer cell e1989.
proliferation, Mol. Cancer 13 (2014) 198. [96] K. Watanabe, M.E. Penfold, A. Matsuda, N. Ohyanagi, K. Kaneko, Y. Miyabe,
[75] G. Kallifatidis, D. Munoz, R.K. Singh, N. Salazar, J.J. Hoy, B.L. Lokeshwar, Beta- K. Matsumoto, T.J. Schall, N. Miyasaka, T. Nanki, Pathogenic role of CXCR7 in
arrestin-2 counters CXCR7-mediated EGFR transactivation and proliferation, Mol. rheumatoid arthritis, Arthritis Rheum. 62 (11) (2010) 3211–3220.
Cancer Res. MCR 14 (5) (2016) 493–503. [97] L. Cruz-Orengo, D.W. Holman, D. Dorsey, L. Zhou, P. Zhang, M. Wright,
[76] R. Bonecchi, G.J. Graham, Atypical chemokine receptors and their roles in the E.E. McCandless, J.R. Patel, G.D. Luker, D.R. Littman, J.H. Russell, R.S. Klein,
Resolution of the inflammatory response, Front. Immunol. 7 (2016) 224. CXCR7 influences leukocyte entry into the CNS parenchyma by controlling ab-
[77] J.S. Lee, M.M. Wurfel, G. Matute-Bello, C.W. Frevert, M.R. Rosengart, luminal CXCL12 abundance during autoimmunity, J. Exp. Med. 208 (2) (2011)
M. Ranganathan, V.W. Wong, T. Holden, S. Sutlief, A. Richmond, S. Peiper, 327–339.
T.R. Martin, The Duffy antigen modifies systemic and local tissue chemokine re- [98] I. Comerford, R.J. Nibbs, W. Litchfield, M. Bunting, Y. Harata-Lee, S. Haylock-
sponses following lipopolysaccharide stimulation, J. Immunol. 177 (11) (2006) Jacobs, S. Forrow, H. Korner, S.R. McColl, The atypical chemokine receptor CCX-
8086–8094. CKR scavenges homeostatic chemokines in circulation and tissues and suppresses

7
E.M. Borroni et al. Seminars in Immunology xxx (xxxx) xxx–xxx

Th17 responses, Blood 116 (20) (2010) 4130–4140. [122] H. Shen, R. Schuster, K.F. Stringer, S.E. Waltz, A.B. Lentsch, The Duffy antigen/
[99] M. Kondo, A.J. Wagers, M.G. Manz, S.S. Prohaska, D.C. Scherer, G.F. Beilhack, receptor for chemokines (DARC) regulates prostate tumor growth, FASEB J. 20 (1)
J.A. Shizuru, I.L. Weissman, Biology of hematopoietic stem cells and progenitors: (2006) 59–64.
implications for clinical application, Annu. Rev. Immunol. 21 (2003) 759–806. [123] B. Nemesure, S.Y. Wu, A. Hennis, M.C. Leske, Distribution of Duffy antigen re-
[100] G.M. Crane, E. Jeffery, S.J. Morrison, Adult haematopoietic stem cell niches, ceptor for chemokines (DARC) and risk of prostate cancer in Barbados, West
nature reviews, Immunology 17 (9) (2017) 573–590. Indies, J Immigr Minor Health 17 (3) (2015) 679–683.
[101] M.G. Manz, S. Boettcher, Emergency granulopoiesis, nature reviews, Immunology [124] X.H. Zeng, Z.L. Ou, K.D. Yu, L.Y. Feng, W.J. Yin, J. Li, Z.Z. Shen, Z.M. Shao,
14 (5) (2014) 302–314. Coexpression of atypical chemokine binders (ACBs) in breast cancer predicts
[102] H.E. Broxmeyer, Chemokines in hematopoiesis, Curr. Opin. Hematol. 15 (1) better outcomes, Breast Cancer Res. Treat. 125 (3) (2011) 715–727.
(2008) 49–58. [125] X.H. Zeng, Z.L. Ou, K.D. Yu, L.Y. Feng, W.J. Yin, J. Li, Z.Z. Shen, Z.M. Shao,
[103] O. Bonavita, V. Mollica Poeta, M. Massara, A. Mantovani, R. Bonecchi, Regulation Absence of multiple atypical chemokine binders (ACBs) and the presence of VEGF
of hematopoiesis by the chemokine system, Cytokine 109 (2018) 76–80. and MMP-9 predict axillary lymph node metastasis in early breast carcinomas,
[104] D. Karpova, H. Bonig, Concise review: CXCR4/CXCL12 signaling in immature Med. Oncol. 31 (9) (2014) 145.
hematopoiesis–lessons from pharmacological and genetic models, Stem Cells 33 [126] F.R. Latini, A.U. Bastos, C.P. Arnoni, J.G. Muniz, R.M. Person, W. Baleotti Jr,
(8) (2015) 2391–2399. J.A. Barreto, L. Castilho, J.M. Cerutti, DARC (Duffy) and BCAM (Lutheran) re-
[105] H.E. Broxmeyer, L.M. Pelus, C.H. Kim, G. Hangoc, S. Cooper, R. Hromas, duced expression in thyroid cancer, Blood Cells Mol. Dis. 50 (3) (2013) 161–165.
Synergistic inhibition in vivo of bone marrow myeloid progenitors by myelosup- [127] S. Zhou, M. Liu, Y. Hu, W. An, X. Liang, W. Yu, F. Piao, Expression of Duffy antigen
pressive chemokines and chemokine-accelerated recovery of progenitors after receptor for chemokines (DARC) is down-regulated in colorectal cancer, J. Recept.
treatment of mice with Ara-C, Exp. Hematol. 34 (8) (2006) 1069–1077. Signal Transduct. Res. 35 (5) (2015) 462–467.
[106] M. Permanyer, B. Bosnjak, R. Forster, Dual role for atypical chemokine receptor 1 [128] G. Sun, Y. Wang, Y. Zhu, C. Huang, Q. Ji, Duffy antigen receptor for chemokines in
in myeloid cell hematopoiesis and distribution, Cell Mol. Immunol. 15 (4) (2018) laryngeal squamous cell carcinoma as a negative regulator, Acta Otolaryngol. 131
399–401. (2) (2011) 197–203.
[107] M. Locati, A. Mantovani, R. Bonecchi, Atypical matters in myeloid differentiation, [129] S. Vetrano, E.M. Borroni, A. Sarukhan, B. Savino, R. Bonecchi, C. Correale,
Nat. Immunol. 18 (7) (2017) 711–712. V. Arena, M. Fantini, M. Roncalli, A. Malesci, A. Mantovani, M. Locati, S. Danese,
[108] J. Duchene, I. Novitzky-Basso, A. Thiriot, M. Casanova-Acebes, M. Bianchini, The lymphatic system controls intestinal inflammation and inflammation-asso-
S.L. Etheridge, E. Hub, K. Nitz, K. Artinger, K. Eller, J. Caamano, T. Rulicke, ciated colon cancer through the chemokine decoy receptor D6, Gut 59 (2) (2010)
P. Moss, R.T.A. Megens, U.H. von Andrian, A. Hidalgo, C. Weber, A. Rot, Atypical 197–206.
chemokine receptor 1 on nucleated erythroid cells regulates hematopoiesis, Nat. [130] C. Schneider, A. Teufel, T. Yevsa, F. Staib, A. Hohmeyer, G. Walenda,
Immunol. 18 (7) (2017) 753–761. H.W. Zimmermann, M. Vucur, S. Huss, N. Gassler, H.E. Wasmuth, S.A. Lira,
[109] B.A. Charles, M.M. Hsieh, A.A. Adeyemo, D. Shriner, E. Ramos, K. Chin, L. Zender, T. Luedde, C. Trautwein, F. Tacke, Adaptive immunity suppresses for-
K. Srivastava, N.A. Zakai, M. Cushman, L.A. McClure, V. Howard, W.A. Flegel, mation and progression of diethylnitrosamine-induced liver cancer, Gut 61 (12)
C.N. Rotimi, G.P. Rodgers, Analyses of genome wide association data, cytokines, (2012) 1733–1743.
and gene expression in African-Americans with benign ethnic neutropenia, PLoS [131] J.M. da Silva, T.P.M. Dos Santos, A.M. Saraiva, A.L. Fernandes de Oliveira,
One 13 (3) (2018) e0194400. G.P. Garlet, A.C. Batista, R.A. de Mesquita, R.C. Russo, T.A. da Silva, Role of
[110] D.R. Crosslin, A. McDavid, N. Weston, X. Zheng, E. Hart, M. de Andrade, I.J. Kullo, atypical chemokine receptor ACKR2 in experimental oral squamous cell carcino-
C.A. McCarty, K.F. Doheny, E. Pugh, A. Kho, M.G. Hayes, M.D. Ritchie, A. Saip, genesis, Cytokine (2018).
D.C. Crawford, P.K. Crane, K. Newton, D.S. Carrell, C.J. Gallego, M.A. Nalls, R. Li, [132] V. Langenes, H. Svensson, L. Borjesson, B. Gustavsson, M. Bemark, A. Sjoling,
D.B. Mirel, A. Crenshaw, D.J. Couper, T. Tanaka, F.J. van Rooij, M.H. Chen, M. Quiding-Jarbrink, Expression of the chemokine decoy receptor D6 is decreased
A.V. Smith, N.A. Zakai, Q. Yango, M. Garcia, Y. Liu, T. Lumley, A.R. Folsom, in colon adenocarcinomas, Cancer Immunol. Immunother. 62 (11) (2013)
A.P. Reiner, J.F. Felix, A. Dehghan, J.G. Wilson, J.C. Bis, C.S. Fox, N.L. Glazer, 1687–1695.
L.A. Cupples, J. Coresh, G. Eiriksdottir, V. Gudnason, S. Bandinelli, T.M. Frayling, [133] Z. Zhu, Z. Sun, Z. Wang, P. Guo, X. Zheng, H. Xu, Prognostic impact of atypical
A. Chakravarti, C.M. van Duijn, D. Melzer, D. Levy, E. Boerwinkle, A.B. Singleton, chemokine receptor expression in patients with gastric cancer, J. Surg. Res. 183
D.G. Hernandez, D.L. Longo, J.C. Witteman, B.M. Psaty, L. Ferrucci, T.B. Harris, (1) (2013) 177–183.
C.J. O’Donnell, S.K. Ganesh, C.H.W. Group, E.B. Larson, C.S. Carlson, G.P. Jarvik, [134] F.Y. Wu, J. Fan, L. Tang, Y.M. Zhao, C.C. Zhou, Atypical chemokine receptor D6
R. electronic Medical, N. Genomics, Genetic variation associated with circulating inhibits human non-small cell lung cancer growth by sequestration of chemokines,
monocyte count in the eMERGE network, Hum. Mol. Genet. 22 (10) (2013) Oncol. Lett 6 (1) (2013) 91–95.
2119–2127. [135] J.M. Burns, B.C. Summers, Y. Wang, A. Melikian, R. Berahovich, Z. Miao,
[111] M. Massara, O. Bonavita, B. Savino, N. Caronni, V. Mollica Poeta, M. Sironi, M.E. Penfold, M.J. Sunshine, D.R. Littman, C.J. Kuo, K. Wei, B.E. McMaster,
E. Setten, C. Recordati, L. Crisafulli, F. Ficara, A. Mantovani, M. Locati, K. Wright, M.C. Howard, T.J. Schall, A novel chemokine receptor for SDF-1 and I-
R. Bonecchi, ACKR2 in hematopoietic precursors as a checkpoint of neutrophil TAC involved in cell survival, cell adhesion, and tumor development, J. Exp. Med.
release and anti-metastatic activity, Nat. Commun. 9 (1) (2018) 676. 203 (9) (2006) 2201–2213.
[112] E.E. Kara, C.R. Bastow, D.R. McKenzie, C.E. Gregor, K.A. Fenix, R. Babb, [136] C. Freitas, A. Desnoyer, F. Meuris, F. Bachelerie, K. Balabanian, V. Machelon, The
T.S. Norton, D. Zotos, L.B. Rodda, J.R. Hermes, K. Bourne, D.S. Gilchrist, relevance of the chemokine receptor ACKR3/CXCR7 on CXCL12-mediated effects
R.J. Nibbs, M. Alsharifi, C.G. Vinuesa, D.M. Tarlinton, R. Brink, G.R. Hill, in cancers with a focus on virus-related cancers, Cytok. Growth Factor Rev. 25 (3)
J.G. Cyster, I. Comerford, S.R. McColl, Atypical chemokine receptor 4 shapes ac- (2014) 307–316.
tivated B cell fate, J. Exp. Med. 215 (3) (2018) 801–813. [137] Y.C. Wu, S.J. Tang, G.H. Sun, K.H. Sun, CXCR7 mediates TGFbeta1-promoted EMT
[113] U. Naumann, E. Cameroni, M. Pruenster, H. Mahabaleshwar, E. Raz, H.G. Zerwes, and tumor-initiating features in lung cancer, Oncogene 35 (16) (2016)
A. Rot, M. Thelen, CXCR7 functions as a scavenger for CXCL12 and CXCL11, PLoS 2123–2132.
One 5 (2) (2010) e9175. [138] J. Wang, Y. Shiozawa, J. Wang, Y. Wang, Y. Jung, K.J. Pienta, R. Mehra, R. Loberg,
[114] A. Mantovani, B. Savino, M. Locati, L. Zammataro, P. Allavena, R. Bonecchi, The R.S. Taichman, The role of CXCR7/RDC1 as a chemokine receptor for CXCL12/
chemokine system in cancer biology and therapy, Cytok. Growth Factor Rev. 21 SDF-1 in prostate cancer, J. Biol. Chem. 283 (7) (2008) 4283–4294.
(1) (2010) 27–39. [139] R.K. Singh, B.L. Lokeshwar, The IL-8-regulated chemokine receptor CXCR7 sti-
[115] M. Massara, O. Bonavita, A. Mantovani, M. Locati, R. Bonecchi, Atypical chemo- mulates EGFR signaling to promote prostate cancer growth, Cancer Res. 71 (9)
kine receptors in cancer: friends or foes? J. Leukoc. Biol. 99 (6) (2016) 927–933. (2011) 3268–3277.
[116] J. Wang, Z.L. Ou, Y.F. Hou, J.M. Luo, Z.Z. Shen, J. Ding, Z.M. Shao, Enhanced [140] C. Ierano, S. Santagata, M. Napolitano, F. Guardia, A. Grimaldi, E. Antignani,
expression of Duffy antigen receptor for chemokines by breast cancer cells at- G. Botti, C. Consales, A. Riccio, M. Nanayakkara, M.V. Barone, M. Caraglia,
tenuates growth and metastasis potential, Oncogene 25 (54) (2006) 7201–7211. S. Scala, CXCR4 and CXCR7 transduce through mTOR in human renal cancer cells,
[117] C.L. Addison, J.A. Belperio, M.D. Burdick, R.M. Strieter, Overexpression of the Cell Death Dis. 5 (2014) e1310.
duffy antigen receptor for chemokines (DARC) by NSCLC tumor cells results in [141] Z. Miao, K.E. Luker, B.C. Summers, R. Berahovich, M.S. Bhojani, A. Rehemtulla,
increased tumor necrosis, BMC Cancer 4 (2004) 28. C.G. Kleer, J.J. Essner, A. Nasevicius, G.D. Luker, M.C. Howard, T.J. Schall, CXCR7
[118] L.W. Horton, Y. Yu, S. Zaja-Milatovic, R.M. Strieter, A. Richmond, Opposing roles (RDC1) promotes breast and lung tumor growth in vivo and is expressed on tumor-
of murine duffy antigen receptor for chemokine and murine CXC chemokine re- associated vasculature, Proc. Natl. Acad. Sci. U. S. A. 104 (40) (2007)
ceptor-2 receptors in murine melanoma tumor growth, Cancer Res. 67 (20) (2007) 15735–15740.
9791–9799. [142] T.C. Xue, R.X. Chen, Z.G. Ren, J.H. Zou, Z.Y. Tang, S.L. Ye, Transmembrane re-
[119] S. Maeda, S. Kuboki, H. Nojima, H. Shimizu, H. Yoshitomi, K. Furukawa, ceptor CXCR7 increases the risk of extrahepatic metastasis of relatively well-dif-
M. Miyazaki, M. Ohtsuka, Duffy antigen receptor for chemokines (DARC) ex- ferentiated hepatocellular carcinoma through upregulation of osteopontin, Oncol.
pressing in cancer cells inhibits tumor progression by suppressing CXCR2 signaling Rep. 30 (1) (2013) 105–110.
in human pancreatic ductal adenocarcinoma, Cytokine 95 (2017) 12–21. [143] B.A. Zabel, S. Lewen, R.D. Berahovich, J.C. Jaen, T.J. Schall, The novel chemokine
[120] S. Bandyopadhyay, R. Zhan, A. Chaudhuri, M. Watabe, S.K. Pai, S. Hirota, receptor CXCR7 regulates trans-endothelial migration of cancer cells, Mol. Cancer
S. Hosobe, T. Tsukada, K. Miura, Y. Takano, K. Saito, M.E. Pauza, S. Hayashi, 10 (2011) 73.
Y. Wang, S. Mohinta, T. Mashimo, M. Iiizumi, E. Furuta, K. Watabe, Interaction of [144] L. Hernandez, M.A. Magalhaes, S.J. Coniglio, J.S. Condeelis, J.E. Segall, Opposing
KAI1 on tumor cells with DARC on vascular endothelium leads to metastasis roles of CXCR4 and CXCR7 in breast cancer metastasis, Breast Cancer Res. 13 (6)
suppression, Nat. Med. 12 (8) (2006) 933–938. (2011) R128.
[121] P. Khanna, C.Y. Chung, R.I. Neves, G.P. Robertson, C. Dong, CD82/KAI expression [145] K. Grymula, M. Tarnowski, M. Wysoczynski, J. Drukala, F.G. Barr, J. Ratajczak,
prevents IL-8-mediated endothelial gap formation in late-stage melanomas, M. Kucia, M.Z. Ratajczak, Overlapping and distinct role of CXCR7-SDF-1/ITAC and
Oncogene 33 (22) (2014) 2898–2908. CXCR4-SDF-1 axes in regulating metastatic behavior of human

8
E.M. Borroni et al. Seminars in Immunology xxx (xxxx) xxx–xxx

rhabdomyosarcomas, Int. J. Cancer 127 (11) (2010) 2554–2568. combination with TMZ activates immune responses and extends survival in mouse
[146] X. Dai, Y. Tan, S. Cai, X. Xiong, L. Wang, Q. Ye, X. Yan, K. Ma, L. Cai, The role of GBM models, Mol. Ther. 26 (5) (2018) 1354–1365.
CXCR7 on the adhesion, proliferation and angiogenesis of endothelial progenitor [151] L.Y. Feng, Z.L. Ou, F.Y. Wu, Z.Z. Shen, Z.M. Shao, Involvement of a novel che-
cells, J. Cell Mol. Med. 15 (6) (2011) 1299–1309. mokine decoy receptor CCX-CKR in breast cancer growth, metastasis and patient
[147] N. Maishi, N. Ohga, Y. Hida, K. Akiyama, K. Kitayama, T. Osawa, Y. Onodera, survival, Clin. Cancer Res. 15 (9) (2009) 2962–2970.
N. Shinohara, K. Nonomura, M. Shindoh, K. Hida, CXCR7: a novel tumor en- [152] J.Y. Shi, L.X. Yang, Z.C. Wang, L.Y. Wang, J. Zhou, X.Y. Wang, G.M. Shi, Z.B. Ding,
dothelial marker in renal cell carcinoma, Pathol. Int. 62 (5) (2012) 309–317. A.W. Ke, Z. Dai, S.J. Qiu, Q.Q. Tang, Q. Gao, J. Fan, CC chemokine receptor-like 1
[148] A.C. Stacer, J. Fenner, S.P. Cavnar, A. Xiao, S. Zhao, S.L. Chang, A. Salomonnson, functions as a tumour suppressor by impairing CCR7-related chemotaxis in he-
K.E. Luker, G.D. Luker, Endothelial CXCR7 regulates breast cancer metastasis, patocellular carcinoma, J. Pathol. 235 (4) (2015) 546–558.
Oncogene 35 (13) (2016) 1716–1724. [153] Y. Zhu, W. Tang, Y. Liu, G. Wang, Z. Liang, L. Cui, CCX-CKR expression in col-
[149] P. Birner, A. Tchorbanov, S. Natchev, J. Tuettenberg, M. Guentchev, The che- orectal cancer and patient survival, Int. J. Biol. Markers 29 (1) (2014) e40–8.
mokine receptor CXCR7 influences prognosis in human glioma in an IDH1-de- [154] Y. Harata-Lee, M.E. Turvey, J.A. Brazzatti, C.E. Gregor, M.P. Brown, M.J. Smyth,
pendent manner, J. Clin. Pathol. 68 (10) (2015) 830–834. I. Comerford, S.R. McColl, The atypical chemokine receptor CCX-CKR regulates
[150] N. Salazar, J.C. Carlson, K. Huang, Y. Zheng, C. Oderup, J. Gross, A.D. Jang, metastasis of mammary carcinoma via an effect on EMT, Immunol. Cell Biol. 92
T.M. Burke, S. Lewen, A. Scholz, S. Huang, L. Nease, J. Kosek, M. Mittelbronn, (10) (2014) 815–824.
E.C. Butcher, H. Tu, B.A. Zabel, A chimeric antibody against ACKR3/CXCR7 in

You might also like