Вы находитесь на странице: 1из 11

Molecular Immunology 47 (2010) 1650–1660

Contents lists available at ScienceDirect

Molecular Immunology
journal homepage: www.elsevier.com/locate/molimm

Review

Scavenger receptor CD163, a Jack-of-all-trades and potential


target for cell-directed therapy
Hanne Van Gorp, Peter L. Delputte 1 , Hans J. Nauwynck ∗,1
Department of Virology, Parasitology and Immunology, Faculty of Veterinary Medicine, Ghent University, Salisburylaan 133,
B-9820 Merelbeke, Belgium

a r t i c l e i n f o a b s t r a c t

Article history: Scavenger receptor CD163 contains nine scavenger receptor cysteine-rich (SRCR) domains and because of
Received 19 January 2010 the presence of this ancient and highly conserved protein motif, CD163 belongs to the SRCR superfamily.
Received in revised form 9 February 2010 Expression of CD163 is restricted to cells of the monocyte/macrophage lineage and is tightly regulated,
Accepted 14 February 2010
with a general tendency of anti-inflammatory signals to induce CD163 synthesis, while pro-inflammatory
Available online 17 March 2010
signals rather seem to downregulate CD163 expression. The first-identified and most-studied function
of CD163 is related to its capacity to bind and internalize haemoglobin–haptoglobin (HbHp) complexes.
Keywords:
Later on, its functional repertoire was expanded, with the identification of CD163 as an erythroblast
Scavenger receptor
CD163
adhesion receptor, a receptor for tumour necrosis factor-like weak inducer of apoptosis (TWEAK), as
Macrophages well as a receptor for distinct pathogens encompassing bacteria and viruses. Interaction of one of these
Ectodomain shedding ligands with CD163 might result in receptor-mediated endocytosis, but might as well trigger a signalling
Targeting cascade leading to the secretion of signalling molecules, which implicates that CD163 also acts as an
immunomodulator. Not only the membrane-bound form of CD163 has an immunomodulating capacity,
but also soluble CD163, which is generated via ectodomain shedding, is able to exert anti-inflammatory
effects. Furthermore, the concentration of this soluble protein is significantly increased under specific
pathological conditions, making it a useful marker protein for certain diseases. Finally, its restricted
expression pattern and potential to internalize make CD163 an attractive candidate as gateway for cell-
directed (immuno)therapy. This review aims to summarize current knowledge on CD163’s biology and its
different biological functions beyond HbHp scavenging, thereby mainly focussing on the more recently
discovered ones. Furthermore, current data supporting the capacity of CD163 to serve as a diagnostic
marker in certain diseases and its potential as a target molecule for cell-directed therapy are surveyed.
© 2010 Elsevier Ltd. All rights reserved.

1. Introduction 1996), 2A10 (Sanchez et al., 1999), along with p155 (Morganelli and
Guyre, 1988), M130 (Ritter et al., 1999), and haemoglobin scavenger
Macrophages and their progenitor cells, monocytes, are key receptor (HbSR) (Kristiansen et al., 2001).
players in the immune system where they fulfil a myriad of func- Only in 2001 however, a function could be attributed to CD163
tions aided by a vast repertoire of surface molecules. Scavenger when this glycoprotein was identified as internalization recep-
receptor CD163 is one of them and was initially identified as RM3/1 tor for haemoglobin–haptoglobin (HbHp) complexes (Kristiansen
in 1987 (Zwadlo et al., 1987) before receiving its CD number in 1996 et al., 2001). This function of CD163 has been the subject of
(Kishimoto et al., 1997). Different antibodies were raised against extensive investigation, as discussed in several reviews (Ascenzi
CD163 prior to its cloning and proper characterization, resulting in a et al., 2005; Fabriek et al., 2005; Graversen et al., 2002; Nielsen
series of different names for the same protein. Besides its most com- and Moestrup, 2009; Nielsen et al., 2010; Onofre et al., 2009;
monly used name CD163, this scavenger receptor is also known as Schaer et al., 2007; Zuwala-Jagiello, 2006). Meanwhile however,
Ki-M8 (Radzun et al., 1987), Ber-MAC3 (Backe et al., 1991), GHI/61 also other CD163 ligands were described, indicating that CD163
and SM4 (Pulford et al., 1992), RM3/1 (Zwadlo et al., 1987), ED2 anti- is a multifunctional receptor involved in receptor-mediated endo-
gen (Dijkstra et al., 1985), AM-3K (Komohara et al., 2006; Zeng et al., cytosis and in signalling pathways upon interaction with diverse
ligands. This review aims to summarize current knowledge on
CD163’s biology and its different biological functions beyond HbHp
∗ Corresponding author. Tel.: +32 9 264 73 66; fax: +32 9 264 74 95. scavenging, thereby mainly focussing on the more recently dis-
E-mail address: Hans.Nauwynck@Ugent.be (H.J. Nauwynck). covered ones. Furthermore, current data supporting the capacity
1
These authors share senior authorship. of CD163 to serve as a diagnostic marker in certain diseases and

0161-5890/$ – see front matter © 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.molimm.2010.02.008
H. Van Gorp et al. / Molecular Immunology 47 (2010) 1650–1660 1651

its potential as a target molecule for cell-directed therapy are gp-340/SAG/DMBT1 (Holmskov et al., 1997; Ligtenberg et al., 2007;
surveyed. Mollenhauer et al., 1997). Members are characterized by the pres-
ence of additional domains other than the SRCR domain, such as
epithelial growth factor (EGF) domain, zona pellucida (ZP) domain,
2. Classification: scavenger receptor cysteine-rich fibronectin domain, etc. The third subgroup contains CD163 (Law
superfamily group B et al., 1993) together with CD163-L1 (Gronlund et al., 2000), WC1
(Mackay et al., 1986; Wijngaard et al., 1992) and SCART (Table 1)
Scavenger receptor (SR) biology originated in 1979, when (Holm et al., 2009; Kisielow et al., 2008). They are characterized by
Brown and Goldstein explored the uptake of modified low-density the presence of a long-range repeat of 5 consecutive SRCR domains
lipoprotein (LDL) by macrophages. At present, the term SR defines a with a small linker domain between the second and third SRCR
vast number of glycoproteins involved in the recognition of polyan- domain. This long-range repeat is referred to as [b-c-d-e-d] cassette
ionic structures of either endogenous (e.g. oxidized or acetylated (Fig. 1). It has been suggested that CD163 emerged from CD163-L1
LDL) or exogenous (e.g. bacterial lipopolysaccharides (LPS)) ori- by gene duplication, and that during this process 3 of the first 6
gin. Although functionally related, SR are structurally quite diverse, SRCR domains of CD163-L1 were lost (Gronlund et al., 2000; Stover
containing, among others, collagenous, C-type lectin, leucine-rich et al., 2000).
repeat or scavenger receptor cysteine-rich (SRCR) domains. Based
on the multidomain structure, the SR family can be classified into
eight different classes (A–H) (Areschoug and Gordon, 2009; Krieger,
1997; Krieger and Herz, 1994; Murphy et al., 2005; Pluddemann et 3. Structure
al., 2007; Taylor et al., 2005).
The first SR to be cloned was the class A scavenger receptor SR-AI CD163 is a type I transmembrane protein composed of 9 extra-
that is composed of several distinct domains including a C-terminal cellular consecutive SRCR domains type B, with only SRCR 6 and 7
SRCR domain (Freeman et al., 1990). At present, this domain has separated by a proline-serine-threonine rich (PST) polypeptide of
been found in more than 25 proteins, including CD163, and its pres- approximately 35 amino acids (Fig. 1). Following the SRCR domains,
ence is the molecular signature of the so-called SRCR superfamily a short PST linker domain connects SRCR 9 with a transmembrane
(SRCR-SF) as nicely reviewed by Sarrias et al. (2004). Sequences domain and an intracellular cytoplasmic tail. This intracellular tail
containing the SRCR domain have been identified in representatives is subject to alternative splicing resulting in various CD163 isoforms
of diverse animal phyla, ranging from marine sponges (Blumbach that differ in the length of their cytoplasmic tail (Law et al., 1993;
et al., 1998; Pancer et al., 1997) and insects (Gorman et al., 2000), Ritter et al., 1999). The most abundant form has a tail of 49 amino
to avian (Gobel et al., 1996; Iwasaki et al., 2001; Koskinen et al., acids, while the less abundant forms have a tail of 84 or 89 amino
2001), amphibian (Goldberger et al., 1987; Yamada et al., 2000), and acids. The first 42 amino acids after the membrane-spanning seg-
various mammalian species (Aruffo et al., 1991; Holm et al., 2009; ment are in common for the three isoforms and this region contains
Mackay et al., 1986; Nunes et al., 1995; Polfliet et al., 2006; Takito consensus sequences for phosphorylation with protein kinase C and
et al., 1996). Notwithstanding the structural relationship due to the creatine kinase. Additional phosphorylation sites are present in the
presence of this ancient and highly conserved protein motif, there remaining part of the cytoplasmic tail of the isoforms with 84 amino
is, as yet, no unifying biological function for the SRCR-SF (Sarrias et acids. Furthermore, the common membrane-proximal region of the
al., 2004). cytoplasmic tail includes a hydrophobic internalization motif of
The SRCR domain is an extracellular domain consisting of the type Yxx␾, where ␾ represents a bulky hydrophobic residue
100–110 amino acid residues (Sarrias et al., 2004). The SRCR-SF can (Nielsen et al., 2006). The observed mass of human CD163 under
be divided into two groups (A and B), depending on the number of non-reducing conditions is 110 kDa, and 130 kDa under reducing
cysteine residues present in the SRCR domains, and, consequently, conditions (Pulford et al., 1992). CD163 is extensively glycosylated,
on the disulphide-bond pattern established. Another distinguishing predominantly with N-linked glycans as shown by a reduction in
feature is the number of exons coding for each SRCR domain. Group molecular weight after endoglycosidase F treatment (Fabriek et al.,
A SRCR domains contain 6 cysteine residues and are encoded by two 2007b; Hogger et al., 1998b).
exons, whereas those of group B contain 6 or 8 cysteine residues Recently, the crystallographic structure for a group B SRCR-SF
and are encoded by a single exon. Protein alignment and studies member was determined. More precisely, the N-terminal (Garza-
on the cysteine disulphide linkage pattern of several members of Garcia et al., 2008) and the C-terminal (Rodamilans et al., 2007a,b)
the SRCR-SF have shown that the relative position of cysteines and SRCR domains of human CD5 were analyzed (Fig. 1). SRCR domains
their disulphide bond pattern are well conserved within each SRCR have a compact fold consisting of a curved six/seven-stranded ␤-
domain. Thus, according to the usual nomenclature for number- sheet cradling an ␣-helix. The structure core is formed by the
ing cysteines 1–8 of group B domains, the disulphide pattern of association of helix ␣1 with a curved four-stranded antiparallel
an SRCR domain is C1–C4, C2–C7, C3–C8, and C5–C6. Cysteines C1 ␤-sheet, which includes ␤1, ␤2, ␤4 and ␤7. The three additional
and C4 are absent in group A domains, though always present in ␤-strands (␤3, ␤5, and ␤6) form another antiparallel ␤-sheet at
group B domains. Some group B domains lack the C2–C7 bridge the bottom of the domain core, as depicted by Rodamilans et
(Resnick et al., 1994; Sarrias et al., 2004). CD163 is composed of al. (2007b). For comparison with other SRCR domains, the only
nine type B SRCR domains (SRCR1–9) all containing four disul- information available comes from M2BP (Hohenester et al., 1999),
phide bridges except for SRCR8, which lacks the C2–C7 bridge hepsin (Somoza et al., 2003) and MARCO (Ojala et al., 2007), which
(Fig. 1). all belong to the group A SRCR-SF. At the sequence level, the only
Members of the group B SRCR-SF are further divided into differ- consistent difference between group A and group B SRCR domains
ent subgroups based on their structure, sequence homology, and is the C1–C4 disulphide bond present in group B and absent in group
extracellular domain organization (Gronlund et al., 2000; Sarrias A. Based on the 3D structures available, this disulphide bond does
et al., 2004). The first and most extensively studied subgroup com- not appear to have a major role in shaping the SRCR fold. Main dif-
prises CD5 (Jones et al., 1986), CD6 (Aruffo et al., 1991) and Sp␣ ferences are observed at the connecting loops, though overall type
(Gebe et al., 1997), which all contain three SRCR domains. The A and B SRCR domains share a similar fold with most of the sec-
second subgroup presents SRCR domains that are part of a mul- ondary elements conserved (Garza-Garcia et al., 2008; Rodamilans
tidomain mosaic extracellular region, and includes, among others, et al., 2007b).
1652 H. Van Gorp et al. / Molecular Immunology 47 (2010) 1650–1660

Fig. 1. Schematic representation of SRCR-SF members belonging to the CD163 subgroup, and sequence alignment of the 9 SRCR domains of CD163 and the third SRCR
domain of CD5. (A) CD163 and its closest relatives are type I membrane proteins composed of a variable number of scavenger receptor cysteine-rich (SRCR) domains, and
proline-serine-threonine rich (PST) domains. A common characteristic of these 4 proteins is the presence of a long-range repeat of 5 consecutive SRCR domains with a small
PST linker molecule separating the second and the third SRCR domain. This repeat is referred to as [b-c-d-e-d] cassette. WC1 is represented by the bovine homologue. (B)
CD163 contains 9 type B SRCR domains (SRCR1–9). Type B SRCR domains are characterized by the presence of 6 or 8 cysteine residues, highlighted in grey, resulting in a
typical disulphide linkage pattern appointed as C1–C4, C2–C7, C3–C8, and C5–C6. All 9 SRCR domains contain 4 disulphide bonds, except for SRCR8 that lacks the C2–C7
linkage. The crystal structure was determined for the third SRCR domain of human CD5 (Rodamilans et al., 2007b), which can be used to predict the potential secondary
structure of the CD163 SRCR domains, which are all type B domains as is the case for the CD5 SRCR domains. Furthermore, the ligand-binding pocket (LBP) responsible for
the interaction between several members of the SRCR-SF and their ligands is indicated.

4. Expression and regulation et al., 2005; Sanchez et al., 1999). CD163-positive macrophages
are also found during the healing phase of acute inflammation,
To date, CD163 has been described in human (Law et al., 1993), in chronic inflammation and in wound-healing tissue, whereas
pig (Sanchez et al., 1999), mouse (Schaer et al., 2001), and rat freshly infiltrated macrophages are CD163-negative (Verschure et
(Polfliet et al., 2006), but also dog, cattle and monkey homologues al., 1989; Zwadlo et al., 1987).
have been identified (Calvert et al., 2007; Sopp et al., 2007; Zwadlo- The expression of CD163 is regulated by a variety of factors
Klarwasser et al., 1992). Expression of CD163 is restricted to cells (Fig. 3). Genomic analysis of the CD163 promoter region revealed
of the monocyte/macrophage lineage. In tissues, CD163 staining several binding sites for transcription factors in the 5 flanking
is predominantly observed on resident tissue macrophages such region. These transcription factors have been shown to play an
as red pulp macrophages in the spleen, Kupffer cells in the liver, important role in myeloid-specific gene expression and differen-
and interstitial and alveolar macrophages in the lungs. In addi- tiation. In addition, three putative glucocorticoid receptor binding
tion, the CD163 antigen is present on perifollicular and medullary sites were identified (Ritter et al., 1999). Indeed, in vitro treatment
macrophages in lymph nodes, perifollicular macrophages in the of monocytes with the glucocorticoid dexamethasone strongly
tonsils, medullary and cortical macrophages in the thymus, and induced CD163 expression (Hogger et al., 1998a,b; Morganelli
perivascular and meningeal macrophages in the central ner- and Guyre, 1988; Van den Heuvel et al., 1999). Even more, in
vous system (Van den Heuvel et al., 1999). These mature tissue vivo administration of glucocorticoids in human volunteers also
macrophages show a substantially higher expression compared to resulted in an increase of the CD163-positive monocyte population
blood monocytes indicating that CD163 is a differentiation marker (Zwadlo-Klarwasser et al., 1990). Another anti-inflammatory medi-
of the macrophage lineage with increased expression along the ator, the cytokine IL-10, also strongly upregulated CD163 mRNA
macrophage differentiation pathway (Backe et al., 1991; Chamorro in monocytes and macrophages, as did the acute phase mediator

Table 1
Group B SRCR-SF members belonging to the CD163 subgroup.

Name Alternative names Expression Form Function References

CD163 M130, RM3/1, haemoglobin Monocytes/macrophages Membrane, soluble HbHp receptor, erythroblast Law et al. (1993) and
scavenger receptor (HbSR), adhesion receptor, virus Kristiansen et al.
p155, etc. receptor, inflammation, (2001)
immunity and host defence
CD163-L1 M160, CD163b Monocytes/macrophages Membrane Inflammation Gronlund et al. (2000)
WC1 T19 ␥␦ T lymphocytes Membrane Reversible G0/G1 cell cycle Mackay et al. (1986)
arrest and Wijngaard et al.
(1992)
SCART ␥␦ T lymphocytes Membrane Kisielow et al. (2008)
and Holm et al. (2009)
H. Van Gorp et al. / Molecular Immunology 47 (2010) 1650–1660 1653

Fig. 2. Schematic representation of the functional portfolio of CD163. (A) CD163 functions as a scavenger receptor for haemoglobin–haptoglobin (HbHp) complexes. Upon
HbHp complex formation, a neo-epitope is exposed, which interacts with the third SRCR domain of CD163, resulting in signalling events and endocytosis of the receptor–ligand
complex. HbHp-induced signalling triggers production of IL-10, which in turn upregulates production of HO-1 and CD163. Upon endocytosis, the receptor–ligand complex
enters early endosomes where HbHp complexes are released from CD163. The receptor then recycles to the cell surface, while HbHp complexes continue through the
endocytic pathway to end up in lysosomes where the protein moieties of the ligands are degraded. Haem is converted to the overall anti-inflammatory molecules CO, Fe2+ ,
and bilirubin by means of HO-1 and biliverdin reductase in the cytosol. (B) CD163 functions as an erythroblast adhesion receptor, thereby linking developing erythroblasts
to a macrophage. A thirteen amino acid motif in the second SRCR domain of CD163 mediates this interaction. (C) CD163 plays a role as receptor for the cytokine TNF-like
weak inducer of apoptosis (TWEAK). Forty-four putative interaction sites were determined, spread across the SRCR domains, excluding SRCR5. (D) CD163 acts as an innate
immune sensor for bacteria, and upon binding of bacteria to CD163 the production of pro-inflammatory cytokines is induced. The thirteen amino acid motif in SRCR2,
which is responsible for the interaction with erythroblasts, is also involved in the interaction between CD163 and bacteria. (E) CD163 is used by ASFV as an attachment
and internalization receptor to enter macrophages. In addition, CD163 is involved in entry of PRRSV. The virus and CD163 however do not interact at the cell surface, but
only intracellularly in early endosomes where CD163 facilitates PRRSV uncoating. CD163 SRCR domains are coloured in blue, with the domains responsible for interaction
with a specific ligand indicated in purple. Abbreviations: ASFV, African Swine Fever virus; BR, biliverdin reductase; EE, early endosome; HO-1, haem oxygenase-1; LE, late
endosome; Ly, lysosome; PRRSV, porcine reproductive and respiratory syndrome virus.

IL-6 (Buechler et al., 2000; Sulahian et al., 2000). Among 19 iden- differentiated to dendritic cells by granulocyte macrophage-colony
tified IL-10-inducible genes in human monocytes, CD163 in fact stimulating factor (GM-CSF) and IL-4, CD163 mRNA transcrip-
displayed the strongest upregulation (Williams et al., 2002). This is tion and protein expression were suppressed (Buechler et al.,
however not a feature common to all anti-inflammatory cytokines, 2000). However, low levels of CD163 expression were reported
since IL-4 and IL-13 were not able to increase CD163 expres- for monocyte-derived dendritic cells (Sulahian et al., 2000) and
sion (Buechler et al., 2000; Sulahian et al., 2000; Van den Heuvel for a small fraction of dendritic cells isolated from the blood
et al., 1999). Interestingly, in contrast to the anti-inflammatory (Maniecki et al., 2006), but the functional significance of this is not
mediators, pro-inflammatory mediators, like IFN-␥, LPS and TNF- yet clear.
␣, have been reported to suppress CD163 expression (Buechler
et al., 2000). The reduction observed upon LPS treatment how- 5. Functional role
ever, is due to ectodomain shedding (see ‘Soluble CD163 ) and is
observed already at 1 h post treatment, while starting from 24 h 5.1. CD163 as a receptor for haemoglobin clearance
an increase in CD163 surface re-expression is observed (Hintz et
al., 2002; Van den Heuvel et al., 1999; Weaver et al., 2006). It is The most extensively studied function of CD163 is related to the
of interest to note that activation of cell surface TLRs, like TLR4 binding of haemoglobin–haptoglobin (HbHp) complexes, as dis-
activation by LPS, results in increased production of IL-6 and IL-10 covered in 2001 by Kristiansen and co-workers (Fig. 2). Release
followed by upregulation of CD163 expression (Weaver et al., 2006, of Hb into plasma is a physiological phenomenon associated with
2007). So, CD163 expression is tightly regulated by pro- and anti- intravascular haemolysis occurring during destruction of senes-
inflammatory mediators among others suggesting a role for CD163 cent or malformed red blood cells. Efficient removal of free Hb
in immune responses. Consistent with the expression of CD163 on is essential for health because of the oxidative and toxic prop-
mature macrophages in vivo, in vitro differentiation of monocytes erties of the iron-containing haem present in Hb (Ascenzi et al.,
to macrophages by macrophage-colony stimulating factor (M-CSF) 2005; Nielsen et al., 2010). Upon release in the circulation, free
strongly induces CD163 mRNA and protein expression (Buechler Hb binds to the plasma glycoprotein Hp resulting in one of the
et al., 2000). Intriguingly, when the chemokine CXCL4 was used to strongest non-covalent interactions observed in plasma (Hwang
differentiate monocytes into macrophages, CD163 was downreg- and Greer, 1980). Binding of Hb to Hp leads to the exposure of
ulated (Gleissner et al., 2010). In addition, when monocytes were a neo-epitope interacting with high affinity with the third SRCR
1654 H. Van Gorp et al. / Molecular Immunology 47 (2010) 1650–1660

domain of CD163 in a calcium-dependent manner (Kristiansen 5.4. CD163 as a pathogen receptor


et al., 2001; Madsen et al., 2004). Upon binding to CD163, the
HbHp complexes are internalized by means of the Yxx␾-based The most recent discovery concerning the functional portfo-
internalization motif common to all three cytoplasmic tail variants lio of CD163 proposes a role for this molecule as innate immune
(Nielsen et al., 2006). Among the three variants, the short tail vari- sensor for bacteria. CD163 has been shown to bind both gram-
ant exhibits the highest capacity for ligand endocytosis, probably positive and gram-negative bacteria, and the CD163 motif involved
due to its more pronounced surface expression compared with the in this interaction has been shown to be the same as the previ-
long tail variants, which are most abundant in intracellular com- ously identified motif involved in erythroblast binding. Expression
partments (Backe et al., 1991; Nielsen et al., 2006; Van den Heuvel of CD163 in monocytic cells promoted bacteria-induced produc-
et al., 1999). Apart from ligand-induced internalization, CD163 also tion of pro-inflammatory cytokines, like TNF-␣. CD163 is suggested
undergoes constitutive and ligand-independent internalization. to act as an innate immune sensor for bacteria and inducer of
Upon internalization, cargo is delivered to early endosomes and local immunity, rather than as a phagocytic receptor (Fabriek et
CD163 recycles to the plasma membrane for a new round of endo- al., 2009). Interestingly, several other members of the SRCR-SF like
cytosis (Schaer et al., 2006). Once released, the haem subunit of Hb SR-AI (Hampton et al., 1991), gp-340 (Bikker et al., 2002, 2004),
is degraded by the rate-limiting haem oxygenase-1 (HO-1) enzyme, Sp␣ (Sarrias et al., 2005) and CD6 (Sarrias et al., 2007) have previ-
yielding biliverdin, free iron, and the carbon monoxide (CO) ously been demonstrated to mediate bacterial binding. It remains
molecule (Nielsen et al., 2010; Philippidis et al., 2004; Schaer et al., however to be analyzed whether pathogen scavenging is a general
2008). property shared by all members of the SRCR-SF or only by a selected
group of members.
Several pathogens have evolved mechanisms to evade recep-
5.2. CD163 as an erythroblast adhesion receptor tor recognition or to exploit these receptors for their own benefit,
mainly to enter their host cell. Two viruses, namely the African
Already in 1996, the ED2 antigen was reported to mediate ery- swine fever virus (ASFV) and the porcine reproductive and res-
throblast binding (Barbe et al., 1996). Though, only in 2007 when piratory syndrome virus (PRRSV) have been shown to (mis)use
the ED2 antigen was identified as the rat CD163 surface glyco- CD163 during entry into their target cells belonging to the mono-
protein, the role of CD163 as erythroblast adhesion molecule was cyte/macrophage lineage (Sanchez-Torres et al., 2003; Van Gorp
further explored. CD163 was shown to directly interact with ery- et al., 2008). For both viruses, expression of CD163 was shown
throblastic cells and a thirteen amino acid motif in the second SRCR to correlate with susceptibility to virus infection (Patton et al.,
domain of CD163 was identified to mediate this binding. Interac- 2009; Sanchez-Torres et al., 2003). The exact role of CD163 during
tion of this CD163 motif with erythroblasts promotes the growth infection of both viruses however seems to differ since CD163 is
and/or survival of these cells (Fabriek et al., 2007b). Apart from proposed to function as an attachment and internalization recep-
CD163, four other erythroblast adhesion molecules are known on tor for ASFV (Sanchez-Torres et al., 2003), while for PRRSV data
macrophages (Chasis, 2006; Koury, 2007): vascular cell adhesion sustain a role for CD163 during virus uncoating (Van Gorp et al.,
molecule-1 (VCAM-1) (Sadahira et al., 1995), av integrin (Lee et 2008, 2009). Analysis of the interaction between PRRSV and CD163
al., 2006), erythroblast–macrophage protein (EMP) (Hanspal and revealed that the cytoplasmic tail containing the internalization
Hanspal, 1994), and sialoadhesin (Rughetti et al., 2003). These motif is dispensable for PRRSV infection, as are the 4 N-terminal
adhesion molecules mediate adherence of developing erythrob- SRCR domains. The essential domains are more centrally located
lasts to a central macrophage forming a functional unit termed with a key role for SRCR domain 5, but also other SRCR domains
erythroblastic island. The interactions enhance erythroblast sur- and the 2 PST domains are required (Van Gorp et al., 2010). So far,
vival, proliferation and differentiation during the terminal stages no biological role has been attributed to these domains as other lig-
of erythropoiesis (Chasis, 2006). ands were shown to interact with SRCR domain 2 or 3 (Fabriek et al.,
2007b, 2009; Madsen et al., 2004). The interaction between CD163
and bacteria or erythroblasts could be pinpointed to an 11–13
5.3. CD163 as a TWEAK receptor amino acid motif within the second SRCR domain. This motif was
previously localized to a putative ligand-binding pocket of the SRCR
CD163 has recently been proposed as receptor for tumour necro- domain of M2BP and has been shown to mediate ligand-binding for
sis factor (TNF)-like weak inducer of apoptosis (TWEAK), a secreted several other members of the SRCR-SF, including gp-340, MARCO,
cytokine belonging to the TNF superfamily (Bover et al., 2007; and thus also CD163 (Bikker et al., 2004; Elomaa et al., 1998). The
Chicheportiche et al., 1997). Also fibroblast growth factor inducible potential of this motif to mediate cell and pathogen recognition is
14 (Fn14/TweakR) has been described as TWEAK-binding molecule further supported by the prediction that this motif is localized in a
(Wiley et al., 2001). Though, cells lacking this receptor were short loop extending from the surface of the SRCR domain (Fabriek
still TWEAK-sensitive, suggesting the existence of an alternative et al., 2009; Rodamilans et al., 2007b). Whether or not this motif
TWEAK receptor (Polek et al., 2003). Using a random combinato- is involved in the interaction between CD163 and ASFV or PRRSV
rial peptide library, CD163 was identified as a new TWEAK-binding remains to be discovered. So far, only one other member of the
protein. Peptide sequences obtained from the screening revealed a SRCR-SF has been related to virus infection, namely gp-340. The
total of 44 putative interaction sites between TWEAK and CD163, N-terminal SRCR domain of gp-340 has been implicated in bind-
all located in eight of the nine CD163 SRCR domains. Further- ing with HIV viral glycoprotein gp-120 (Wu et al., 2006). Yet, the
more, HbHp complexes and antibodies recognizing different SRCR functional relevance of this interaction is still a matter of debate
domains of CD163 were all able to inhibit TWEAK binding to CD163. since inhibition (Wu et al., 2006) as well as promotion (Cannon et
Together, these data suggest multiple interaction sites between al., 2008; Stoddard et al., 2007) of HIV infection has been described.
TWEAK and CD163, including the HbHp recognition site. Upon
binding to CD163-expressing macrophages, TWEAK is internalized 5.5. CD163 as immunomodulator
and degraded. CD163 is proposed as scavenger receptor for TWEAK,
preventing TWEAK from exerting its biological functions by seques- Different populations of activated macrophages can be dis-
tering it from the environment (Bover et al., 2007; Moreno et al., tinguished based on activation signals, secretory products, and
2009). biological markers. Whereas classical activation of macrophages
H. Van Gorp et al. / Molecular Immunology 47 (2010) 1650–1660 1655

Fig. 3. Regulation of CD163 expression and shedding. (A) CD163 expression is tightly regulated. Overall, anti-inflammatory signals seem to induce CD163 synthesis, while pro-
inflammatory signals seem to downregulate CD163 expression. Differentiation into macrophages via M-CSF results in higher CD163 expression levels, while differentiation
into dendritic cells via GM-CSF and IL-4 results in lower CD163 expression levels. (B) Several signals are able to induce protease-dependent cleavage of the extracellular part
of CD163, which results in the release of soluble CD163 (sCD163). Treatment with LPS results in shedding of sCD163, and consequently initial downregulation of cell surface
CD163. Over time however, CD163 cell surface expression is restored to similar or even higher levels compared to the situation prior to LPS treatment. The soluble form of
CD163 has anti-inflammatory properties and is used as a marker protein for certain pathological conditions.

by IFN-␥ and TNF-␣ results in decreased expression of CD163 faced immunomodulator, able to stimulate as well as suppress the
(Buechler et al., 2000), elevated levels of CD163 expression are immune response.
induced by glucocorticoids, IL-6, and IL-10 (Buechler et al., 2000;
Hogger et al., 1998a; Sulahian et al., 2000) resulting in a distinct 6. Soluble CD163 (sCD163)
population of cells called the ‘alternatively activated macrophages’.
The anti-inflammatory IL-10 is one of the typical cytokines pro- CD163 is not only present as a membrane-bound form, but
duced by alternatively activated macrophages, which fail to make can also be detected in high concentrations as soluble protein in
NO, are not efficient in antigen presentation and inhibit T lympho- plasma (Fig. 3) (Droste et al., 1999; Moller et al., 2002c). Rather
cyte proliferation. They are however implicated in wound-healing, than being the product of alternative splicing, soluble CD163
angiogenesis, and protection of the host from an overwhelming (sCD163) is actively shed from the plasma membrane by metal-
inflammatory response. Thus, alternatively activated macrophages loproteinases, though the responsible metalloproteinase and the
tend to serve more a regulatory and recovery function than cleavage site remain elusive so far (Droste et al., 1999; Hintz et
the effector killing functions associated with classically activated al., 2002; Matsushita et al., 2002; Moller et al., 2009). The result-
macrophages (Mosser, 2003). Early immunohistochemical studies ing soluble protein displays an electrophoretic mobility similar to
support the involvement of CD163-positive macrophages in the that of a recombinant soluble protein composed of all nine SRCR
late downregulatory phase of acute inflammation (Zwadlo et al., domains, suggesting that CD163 is cleaved near the plasma mem-
1987) and in chronic inflammation (Topoll et al., 1988). brane (Moller et al., 2002c). Later on, mass spectrometry was used
CD163 is well known as receptor for HbHp complexes. Inter- showing that sCD163 indeed covers more than 94% of the extracel-
estingly, the metabolites released upon HbHp degradation, more lular part of CD163 (Moller et al., 2009). In vitro, shedding of CD163
precisely bilirubin, free iron, and CO, all have strong anti-oxidative is induced in response to phorbol 13-myristate 12-acetate treat-
and anti-inflammatory effects (Otterbein et al., 2003). A key ment (PMA; also known as 12-O-tetradecanoylphorbol-13-acetate
molecule in this process is the HO-1 enzyme converting haem (TPA)) and acts via a protein kinase C-dependent mechanism
into its metabolites (Soares and Bach, 2009). In addition, bind- (Droste et al., 1999). Phorbol esters are well known to induce
ing of HbHp to CD163 elicits a direct anti-inflammatory effect ectodomain shedding of various transmembrane molecules (Dello
via the secretion of IL-10 (Philippidis et al., 2004), which in turn Sbarba and Rovida, 2002), though they are nonphysiological com-
upregulates CD163 expression thereby amplifying HbHp scaveng- pounds that cannot account for the activation of shedding in vivo.
ing via a positive feedback loop (Buechler et al., 2000; Sulahian et Furthermore, endotoxins, like LPS, are potent inducers of CD163
al., 2000). Furthermore, binding of HbHp also induces expression shedding both in vitro and in vivo, and this via activation of cell
of the rate-limiting enzyme HO-1 via an IL-10 autocrine mecha- surface TLRs (Hintz et al., 2002; Weaver et al., 2006). Interestingly,
nism (Philippidis et al., 2004). CD163-positive macrophages thus induced shedding of CD163 is followed by recovery and induction
efficiently metabolize toxic haem via CD163 binding and HO-1 of surface CD163 to similar or even higher levels than observed in
degradation, and meanwhile exert anti-inflammatory effects via untreated controls (Hintz et al., 2002; Pulford et al., 1992; Weaver
the release of IL-10 and haem metabolites. et al., 2007). Another natural stimulus triggering CD163 shedding is
Besides mediating anti-inflammatory effects, CD163 signalling found in immune complexes cross-linking Fc␥ receptors (Sulahian
has also been implicated in the production of pro-inflammatory et al., 2004). Furthermore, oxidative stress mediators, often present
cytokines. Cross-linking of CD163 with CD163-specific monoclonal under inflammatory conditions, were also identified as stimuli for
antibodies induced secretion of NO, IL-1␤, IL-6, and TNF-␣ (Polfliet CD163 ectodomain shedding (Timmermann and Hogger, 2005).
et al., 2006; Van den Heuvel et al., 1999). IL-1␤ and TNF-␣ are pro- Ectodomain shedding is observed for a wide variety of molecules
inflammatory cytokines, while NO and IL-6 exert both pro- and including adhesion molecules (e.g. ICAM-1), mediators of apopto-
anti-inflammatory effects. Furthermore, also bacteria elicit pro- sis (e.g. Fas ligand), receptors (e.g. CD14), cytokines (e.g. TNF-␣),
duction of pro-inflammatory cytokines upon binding with CD163 growth factors (e.g. EGF), etc. (Cauwe et al., 2007; Garton et
(Fabriek et al., 2009). All together, CD163 appears as a two- al., 2006). So far, only for one other member of the SRCR-SF,
1656 H. Van Gorp et al. / Molecular Immunology 47 (2010) 1650–1660

more precisely CD5, the process of ectodomain shedding has of macrophages and monocytes and expression is even upregu-
been described (Calvo et al., 1999). Ectodomain shedding repre- lated under inflammatory conditions like in rheumathoid arthritis.
sents a post-translational mechanism regulating cellular responses In addition to treatment of inflammatory diseases, other appli-
and interactions at multiple levels, like the release of cytokines cations are apparent including treatment of infectious diseases
expressed as pro-forms generating soluble mediators, or reduction in which macrophages function as hiding place for intracellu-
of the density of a receptor or adhesion molecule and production of lar pathogens like Mycobacterium tuberculosis (McKinney et al.,
soluble competitors for their own ligands. For CD163 however, little 2000). Furthermore, macrophages have the capacity to function
is known about the biological functions of its soluble form. Bind- as antigen-presenting cells, and have been found associated with
ing of sCD163 to HbHp complexes has been suggested to reduce tumours (Allavena et al., 2008), providing additional applications.
cellular iron uptake, thereby limiting pathogen access to free Hb CD163 also has an immunomodulating potential that can be tar-
and inhibiting growth of intracellular pathogens (Madsen et al., geted, thereby either stimulating or repressing the immune system
2004; Weaver et al., 2006). A recent study however showed that depending on the ligand and the cellular micro-environment. Not
HbHp complexes preferentially bind to the membrane-bound form only a multitude of applications can be imagined, also several
of CD163, and that sCD163 is normally not in complex with HbHp options are available to specifically target CD163, including the dif-
complexes in the plasma. This is presumably due to the di- or multi- ferent CD163 ligands and CD163-specific antibodies. Since CD163
valent nature of HbHp complexes in terms of CD163 binding (Moller undergoes constitutive internalization, it is to be expected that any
et al., 2009). Furthermore, a potential direct anti-inflammatory substance bound to it will be transferred across the plasma mem-
effect of sCD163 was described, showing inhibition of T lympho- brane. Concerns about the efficiency of CD163-directed therapy
cyte activation and proliferation upon interaction with sCD163 might however arise because of the existence of a soluble form of
(Hogger and Sorg, 2001). This phenomenon could not be observed CD163. The soluble form is naturally present at high concentrations
for membrane-bound CD163, clearly showing distinct functions for in serum and levels are even significantly upregulated under vari-
the soluble and membrane-bound form of CD163 (Frings et al., ous pathological conditions (Droste et al., 1999; Moller et al., 2002a,
2002). Interaction between sCD163 and T lymphocytes is medi- 2004; Schaer et al., 2005). This soluble protein might capture the
ated by non-muscle myosin heavy chain in T lymphocytes, though CD163-specific therapeutics before they can reach the macrophage,
further research is needed to clarify the consequences of this inter- as shown for CD97 (de Groot et al., 2009), thereby interfering with
action (Timmermann et al., 2004). efficient delivery and increasing off-target side effects. This knowl-
The concentration of sCD163 in serum or plasma from healthy edge emphasizes the importance of careful examination of the
individuals is relatively high, with an average of 2 mg/l, and can expression profile of the membrane-bound as well as the solu-
readily be measured using an ELISA (Moller et al., 2002b,c, 2003; ble receptor, as exemplified by a study exploring the potential of
Perez et al., 2008; Sulahian et al., 2001). Intriguingly, sCD163 lev- CD163 as target for patients with acute myeloid leukemia (AML).
els are increased many-fold during various pathological conditions It was shown that there is no increase in sCD163 concentration
and its use as marker protein monitoring macrophage activity in patients whose leukemic blasts displayed high surface CD163
is explored for an increasing number of diseases. Gaucher’s dis- expression, supporting the potential of CD163 as target molecule
ease is an inherited lysosomal storage disorder characterized by in the treatment of AML (Bachli et al., 2006). Furthermore, even
excessive accumulation of macrophages and patients exhibit highly if there is sCD163, this does not necessarily means that sCD163
increased levels of sCD163. Besides having a diagnostic value in serves as a sink for CD163-specific therapeutics. At least for HbHp
screening for Gaucher’s disease, sCD163 also correlates with clini- complexes it has been shown that they preferentially interact with
cal manifestations and might be useful in monitoring disease course membrane-bound CD163 instead of sCD163 (Moller et al., 2009).
and response to treatment (Moller et al., 2004). Overwhelming All together, further research is imperative to circumvent poten-
macrophage activity is also observed in patients with the reactive tial pitfalls and to fully explore the potential of CD163 as target for
haemophagocytic syndrome (RHS), though differentiation from cell-directed therapy.
other systemic inflammatory conditions long proved a challeng-
ing task since none of the used markers, like TNF-␣ and ferritin, 8. Concluding Remarks
is a specific indicator of macrophage activity. Recently, sCD163
was suggested as a potentially useful marker for RHS due to its Scavenger receptor CD163 manifests itself as a multifunctional
macrophage-specific expression pattern and highly increased lev- molecule fulfilling essential homeostatic functions in diverse bio-
els, up to 20 times, in RHS patients. Furthermore, a close correlation logical processes, with multiple ligands that are able to engage
was revealed between sCD163 levels and clinical disease activity CD163. The outcome of these interactions however seems to be
(Schaer et al., 2005). sCD163 has further been revealed as a promis- ligand-specific and diverse pathways might be set in motion.
ing marker molecule for macrophage activation in diseases and Binding of HbHp complexes not only triggers receptor-mediated
pathological conditions such as sepsis (Moller et al., 2002c, 2006), endocytosis, but also activates a signalling cascade resulting in
liver disease (Hiraoka et al., 2005a,b; Moller et al., 2007), autoim- the production of anti-inflammatory molecules. Other ligands
mune disorders (Baeten et al., 2004; Bleesing et al., 2007; De Rycke however, like bacteria, elicit a pro-inflammatory instead of an anti-
et al., 2005), multiple sclerosis (Fabriek et al., 2007a), and malaria inflammatory response. Furthermore, bacteria are not internalized
(Kusi et al., 2008). upon interaction with CD163, in contrast to HbHp complexes and
TWEAK. Currently however, little is known about the regulatory
mechanisms influencing the outcome of the ligand-CD163 inter-
7. CD163 as target molecule for cell-directed actions. What is known so far is that different CD163 domains
(immuno)therapy are implicated in the interaction between CD163 and HbHp com-
plexes or bacteria. To what extent these interacting SRCR domains
Apart from being a useful tool in the diagnosis of certain dis- of CD163 are a determining factor in the outcome of the interaction
eases, CD163 also appears as an interesting candidate for medical remains elusive, as are the signalling pathways triggered. There-
therapy. Its restricted expression pattern and potential to func- fore, more profound knowledge on the events following ligand
tion as an endocytic receptor make CD163 an attractive target binding to CD163 is required, not only to further our understanding
molecule for cell-directed therapy as already posted by Graversen of fundamental CD163 biology, but also because it is a prerequisite
et al. (2002). CD163 is selectively expressed on subpopulations in the development of CD163-specific therapeutics.
H. Van Gorp et al. / Molecular Immunology 47 (2010) 1650–1660 1657

CD163 is reported to function as an innate immune sensor for Blumbach, B., Pancer, Z., Diehl-Seifert, B., Steffen, R., Munkner, J., Muller, I., Muller,
bacteria, thereby protecting the host. For other pathogens however, W.E., 1998. The putative sponge aggregation receptor. Isolation and characteri-
zation of a molecule composed of scavenger receptor cysteine-rich domains and
like the porcine viruses ASFV and PRRSV, CD163 serves as a por- short consensus repeats. J. Cell Sci. 111, 2635–2644.
tal that allows infection of their target cells. Interestingly, CD163 Bover, L.C., Cardo-Vila, M., Kuniyasu, A., Sun, J., Rangel, R., Takeya, M., Aggarwal,
seems to contribute in a different manner during entry of both B.B., Arap, W., Pasqualini, R., 2007. A previously unrecognized protein–protein
interaction between TWEAK and CD163: potential biological implications. J.
viruses. For ASFV, CD163 is suggested to function as an attachment Immunol. 178, 8183–8194.
and internalization receptor, while for PRRSV CD163 is proposed Buechler, C., Ritter, M., Orso, E., Langmann, T., Klucken, J., Schmitz, G., 2000. Reg-
to act at a post-internalization stage, more precisely during virus ulation of scavenger receptor CD163 expression in human monocytes and
macrophages by pro- and antiinflammatory stimuli. J. Leukoc. Biol. 67, 97–
disassembly. Indicative of this new role of CD163 is that the cyto- 103.
plasmic tail of CD163 containing the internalization motif is not Calvert, J.G., Slade, D.E., Shields, S.L., Jolie, R., Mannan, R.M., Ankenbauer, R.G., Welch,
required for PRRSV infection, while it is essential for uptake of S.K., 2007. CD163 expression confers susceptibility to porcine reproductive and
respiratory syndrome viruses. J. Virol. 81, 7371–7379.
HbHp complexes, sustaining the functional flexibility of CD163.
Calvo, J., Places, L., Espinosa, G., Padilla, O., Vila, J.M., Villamor, N., Ingelmo, M., Gallart,
Furthermore, the CD163 domains essential for PRRSV infection are T., Vives, J., Font, J., Lozano, F., 1999. Identification of a natural soluble form of
different from the ones implicated in the interactions with other human CD5. Tissue Antigens 54, 128–137.
ligands and are more centrally located. Still, the precise molecular Cannon, G., Yi, Y., Ni, H., Stoddard, E., Scales, D.A., Van Ryk, D.I., Chaiken, I., Malamud,
D., Weissman, D., 2008. HIV envelope binding by macrophage-expressed gp340
mechanism underlying the CD163-dependent uncoating of PRRSV promotes HIV-1 infection. J. Immunol. 181, 2065–2070.
remains to be elucidated. Intriguingly, the CD163 ancestor, CD163- Cauwe, B., Van den Steen, P.E., Opdenakker, G., 2007. The biochemical, biological,
L1, is not able to replace CD163 functionality during PRRSV entry and pathological kaleidoscope of cell surface substrates processed by matrix
metalloproteinases. Crit. Rev. Biochem. Mol. Biol. 42, 113–185.
(Van Gorp et al., 2010), although CD163 originates from CD163-L1 Chamorro, S., Revilla, C., Alvarez, B., Alonso, F., Ezquerra, A., Dominguez, J., 2005.
and both proteins are structurally much alike. Therefore, it seems Phenotypic and functional heterogeneity of porcine blood monocytes and its
intriguing to explore the functional repertoire of CD163-L1 as well relation with maturation. Immunology 114, 63–71.
Chasis, J.A., 2006. Erythroblastic islands: specialized microenvironmental niches for
as the potential functional redundancy between CD163 and CD163- erythropoiesis. Curr. Opin. Hematol. 13, 137–141.
L1 to accomplish a holistic picture of these closely related scavenger Chicheportiche, Y., Bourdon, P.R., Xu, H., Hsu, Y.M., Scott, H., Hession, C., Garcia, I.,
receptors. Browning, J.L., 1997. TWEAK, a new secreted ligand in the tumor necrosis factor
family that weakly induces apoptosis. J. Biol. Chem. 272, 32401–32410.
de Groot, D.M., Vogel, G., Dulos, J., Teeuwen, L., Stebbins, K., Hamann, J., Owens,
Acknowledgements A.T.B.M., van Eenennaam, H., Bos, E., Boots, A.M., 2009. Therapeutic antibody tar-
geting of CD97 in experimental arthritis: the role of antigen expression, shedding
The authors would like to thank Wander Van Breedam for fruit- and internalization on the pharmacokinetics of anti-CD97 monoclonal antibody
1B2. J. Immunol. 183, 4127–4134.
ful discussions and critical reading of the manuscript. This work was
De Rycke, L., Baeten, D., Foell, D., Kruithof, E., Veys, E.M., Roth, J., De Keyser, F., 2005.
supported by the Industrial Research Fund (IOF) of Ghent Univer- Differential expression and response to anti-TNFalpha treatment of infiltrating
sity and by the European Union (Seventh Framework Programme, versus resident tissue macrophage subsets in autoimmune arthritis. J. Pathol.
Project No. 245141, coordinated by Dr. H. Nauwynck). 206, 17–27.
Dello Sbarba, P., Rovida, E., 2002. Transmodulation of cell surface regulatory
The authors declare that they have no conflict of interest. molecules via ectodomain shedding. Biol. Chem. 383, 69–83.
Dijkstra, C.D., Dopp, E.A., Joling, P., Kraal, G., 1985. The heterogeneity of mononu-
References clear phagocytes in lymphoid organs: distinct macrophage subpopulations in
the rat recognized by monoclonal antibodies ED1, ED2 and ED3. Immunology
Allavena, P., Sica, A., Garlanda, C., Mantovani, A., 2008. The Yin-Yang of tumor- 54, 589–599.
associated macrophages in neoplastic progression and immune surveillance. Droste, A., Sorg, C., Hogger, P., 1999. Shedding of CD163, a novel regulatory mech-
Immunol. Rev. 222, 155–161. anism for a member of the scavenger receptor cysteine-rich family. Biochem.
Areschoug, T., Gordon, S., 2009. Scavenger receptors: role in innate immunity and Biophys. Res. Commun. 256, 110–113.
microbial pathogenesis. Cell. Microbiol. 11, 1160–1169. Elomaa, O., Sankala, M., Pikkarainen, T., Bergmann, U., Tuuttila, A., Raatikainen-
Aruffo, A., Melnick, M.B., Linsley, P.S., Seed, B., 1991. The lymphocyte glycoprotein Ahokas, A., Sariola, H., Tryggvason, K., 1998. Structure of the human macrophage
CD6 contains a repeated domain structure characteristic of a new family of cell MARCO receptor and characterization of its bacteria-binding region. J. Biol.
surface and secreted proteins. J. Exp. Med. 174, 949–952. Chem. 273, 4530–4538.
Ascenzi, P., Bocedi, A., Visca, P., Altruda, F., Tolosano, E., Beringhelli, T., Fasano, M., Fabriek, B.O., Dijkstra, C.D., van den Berg, T.K., 2005. The macrophage scavenger
2005. Hemoglobin and heme scavenging. IUBMB Life 57, 749–759. receptor CD163. Immunobiology 210, 153–160.
Bachli, E.B., Schaer, D.J., Walter, R.B., Fehr, J., Schoedon, G., 2006. Functional expres- Fabriek, B.O., Moller, H.J., Vloet, R.P., van Winsen, L.M., Hanemaaijer, R., Teunis-
sion of the CD163 scavenger receptor on acute myeloid leukemia cells of sen, C.E., Uitdehaag, B.M., van den Berg, T.K., Dijkstra, C.D., 2007a. Proteolytic
monocytic lineage. J. Leukoc. Biol. 79, 312–318. shedding of the macrophage scavenger receptor CD163 in multiple sclerosis. J.
Backe, E., Schwarting, R., Gerdes, J., Ernst, M., Stein, H., 1991. Ber-MAC3: new mon- Neuroimmunol. 187, 179–186.
oclonal antibody that defines human monocyte/macrophage differentiation Fabriek, B.O., Polfliet, M.M., Vloet, R.P., van der Schors, R.C., Ligtenberg, A.J., Weaver,
antigen. J. Clin. Pathol. 44, 936–945. L.K., Geest, C., Matsuno, K., Moestrup, S.K., Dijkstra, C.D., van den Berg, T.K.,
Baeten, D., Moller, H.J., Delanghe, J., Veys, E.M., Moestrup, S.K., De Keyser, F., 2004. 2007b. The macrophage CD163 surface glycoprotein is an erythroblast adhesion
Association of CD163+ macrophages and local production of soluble CD163 receptor. Blood 109, 5223–5229.
with decreased lymphocyte activation in spondylarthropathy synovitis. Arthri- Fabriek, B.O., van Bruggen, R., Deng, D.M., Ligtenberg, A.J., Nazmi, K., Schornagel, K.,
tis Rheum. 50, 1611–1623. Vloet, R.P., Dijkstra, C.D., van den Berg, T.K., 2009. The macrophage scavenger
Barbe, E., Huitinga, I., Dopp, E.A., Bauer, J., Dijkstra, C.D., 1996. A novel bone marrow receptor CD163 functions as an innate immune sensor for bacteria. Blood 113,
frozen section assay for studying hematopoietic interactions in situ: the role 887–892.
of stromal bone marrow macrophages in erythroblast binding. J. Cell Sci. 109, Freeman, M., Ashkenas, J., Rees, D.J., Kingsley, D.M., Copeland, N.G., Jenkins, N.A.,
2937–2945. Krieger, M., 1990. An ancient, highly conserved family of cysteine-rich protein
Bikker, F.J., Ligtenberg, A.J., End, C., Renner, M., Blaich, S., Lyer, S., Wittig, R., domains revealed by cloning type I and type II murine macrophage scavenger
van’t Hof, W., Veerman, E.C., Nazmi, K., de Blieck-Hogervorst, J.M., Kioschis, receptors. Proc. Natl. Acad. Sci. U.S.A. 87, 8810–8814.
P., Nieuw Amerongen, A.V., Poustka, A., Mollenhauer, J., 2004. Bacteria binding Frings, W., Dreier, J., Sorg, C., 2002. Only the soluble form of the scavenger recep-
by DMBT1/SAG/gp-340 is confined to the VEVLXXXXW motif in its scavenger tor CD163 acts inhibitory on phorbol ester-activated T-lymphocytes, whereas
receptor cysteine-rich domains. J. Biol. Chem. 279, 47699–47703. membrane-bound protein has no effect. FEBS Lett. 526, 93–96.
Bikker, F.J., Ligtenberg, A.J., Nazmi, K., Veerman, E.C., van’t Hof, W., Bolscher, J.G., Garton, K.J., Gough, P.J., Raines, E.W., 2006. Emerging roles for ectodomain shed-
Poustka, A., Nieuw Amerongen, A.V., Mollenhauer, J., 2002. Identification of ding in the regulation of inflammatory responses. J. Leukoc. Biol. 79, 1105–
the bacteria-binding peptide domain on salivary agglutinin (gp-340/DMBT1), 1116.
a member of the scavenger receptor cysteine-rich superfamily. J. Biol. Chem. Garza-Garcia, A., Esposito, D., Rieping, W., Harris, R., Briggs, C., Brown, M.H., Driscoll,
277, 32109–32115. P.C., 2008. Three-dimensional solution structure and conformational plasticity
Bleesing, J., Prada, A., Siegel, D.M., Villanueva, J., Olson, J., Ilowite, N.T., Brunner, of the N-terminal scavenger receptor cysteine-rich domain of human CD5. J.
H.I., Griffin, T., Graham, T.B., Sherry, D.D., Passo, M.H., Ramanan, A.V., Filipovich, Mol. Biol. 378, 129–144.
A., Grom, A.A., 2007. The diagnostic significance of soluble CD163 and solu- Gebe, J.A., Kiener, P.A., Ring, H.Z., Li, X., Francke, U., Aruffo, A., 1997. Molecular
ble interleukin-2 receptor alpha-chain in macrophage activation syndrome and cloning, mapping to human chromosome 1 q21–q23, and cell binding character-
untreated new-onset systemic juvenile idiopathic arthritis. Arthritis Rheum. 56, istics of Spalpha, a new member of the scavenger receptor cysteine-rich (SRCR)
965–971. family of proteins. J. Biol. Chem. 272, 6151–6158.
1658 H. Van Gorp et al. / Molecular Immunology 47 (2010) 1650–1660

Gleissner, C.A., Shaked, I., Erbel, C., Bockler, D., Katus, H.A., Ley, K., 2010. CXCL4 Kusi, K.A., Gyan, B.A., Goka, B.Q., Dodoo, D., Obeng-Adjei, G., Troye-Blomberg, M.,
downregulates the atheroprotective hemoglobin receptor CD163 in human Akanmori, B.D., Adjimani, J.P., 2008. Levels of soluble CD163 and severity of
macrophages. Circ. Res. 106, 203–211. malaria in children in Ghana. Clin. Vaccine Immunol. 15, 1456–1460.
Gobel, T.W., Chen, C.H., Cooper, M.D., 1996. Expression of an avian CD6 candidate is Law, S.K., Micklem, K.J., Shaw, J.M., Zhang, X.P., Dong, Y., Willis, A.C., Mason, D.Y.,
restricted to alpha beta T cells, splenic CD8+ gamma delta T cells and embryonic 1993. A new macrophage differentiation antigen which is a member of the
natural killer cells. Eur. J. Immunol. 26, 1743–1747. scavenger receptor superfamily. Eur. J. Immunol. 23, 2320–2325.
Goldberger, G., Bruns, G.A., Rits, M., Edge, M.D., Kwiatkowski, D.J., 1987. Human com- Lee, G., Lo, A., Short, S.A., Mankelow, T.J., Spring, F., Parsons, S.F., Yazdanbakhsh,
plement factor I: analysis of cDNA-derived primary structure and assignment of K., Mohandas, N., Anstee, D.J., Chasis, J.A., 2006. Targeted gene deletion demon-
its gene to chromosome 4. J. Biol. Chem. 262, 10065–10071. strates that the cell adhesion molecule ICAM-4 is critical for erythroblastic island
Gorman, M.J., Andreeva, O.V., Paskewitz, S.M., 2000. Sp22D: a multidomain serine formation. Blood 108, 2064–2071.
protease with a putative role in insect immunity. Gene 251, 9–17. Ligtenberg, A.J., Veerman, E.C., Nieuw Amerongen, A.V., Mollenhauer, J., 2007.
Graversen, J.H., Madsen, M., Moestrup, S.K., 2002. CD163: a signal receptor scaveng- Salivary agglutinin/glycoprotein-340/DMBT1: a single molecule with variable
ing haptoglobin–hemoglobin complexes from plasma. Int. J. Biochem. Cell Biol. composition and with different functions in infection, inflammation and cancer.
34, 309–314. Biol. Chem. 388, 1275–1289.
Gronlund, J., Vitved, L., Lausen, M., Skjodt, K., Holmskov, U., 2000. Cloning of a Mackay, C.R., Maddox, J.F., Brandon, M.R., 1986. Three distinct subpopulations of
novel scavenger receptor cysteine-rich type I transmembrane molecule (M160) sheep T lymphocytes. Eur. J. Immunol. 16, 19–25.
expressed by human macrophages. J. Immunol. 165, 6406–6415. Madsen, M., Moller, H.J., Nielsen, M.J., Jacobsen, C., Graversen, J.H., van den Berg, T.,
Hampton, R.Y., Golenbock, D.T., Penman, M., Krieger, M., Raetz, C.R., 1991. Recog- Moestrup, K., 2004. Molecular characterization of the haptoglobin.hemoglobin
nition and plasma clearance of endotoxin by scavenger receptors. Nature 352, receptor CD163. Ligand binding properties of the scavenger receptor cysteine-
342–344. rich domain region. J. Biol. Chem. 279, 51561–51567.
Hanspal, M., Hanspal, J.S., 1994. The association of erythroblasts with macrophages Maniecki, M.B., Moller, H.J., Moestrup, S.K., Moller, B.K., 2006. CD163 positive subsets
promotes erythroid proliferation and maturation: a 30-kD heparin-binding pro- of blood dendritic cells: the scavenging macrophage receptors CD163 and CD91
tein is involved in this contact. Blood 84, 3494–3504. are coexpressed on human dendritic cells and monocytes. Immunobiology 211,
Hintz, K.A., Rassias, A.J., Wardwell, K., Moss, M.L., Morganelli, P.M., Pioli, P.A., 407–417.
Givan, A.L., Wallace, P.K., Yeager, M.P., Guyre, P.M., 2002. Endotoxin induces Matsushita, N., Kashiwagi, M., Wait, R., Nagayoshi, R., Nakamura, M., Matsuda, T.,
rapid metalloproteinase-mediated shedding followed by up-regulation of the Hogger, P., Guyre, P.M., Nagase, H., Matsuyama, T., 2002. Elevated levels of
monocyte hemoglobin scavenger receptor CD163. J. Leukoc. Biol. 72, 711– soluble CD163 in sera and fluids from rheumatoid arthritis patients and inhi-
717. bition of the shedding of CD163 by TIMP-3. Clin. Exp. Immunol. 130, 156–
Hiraoka, A., Horiike, N., Akbar, S.M., Michitaka, K., Matsuyama, T., Onji, M., 2005a. 161.
Expression of CD163 in the liver of patients with viral hepatitis. Pathol. Res. McKinney, J.D., Honer zu Bentrup, K., Munoz-Elias, E.J., Miczak, A., Chen, B., Chan,
Pract. 201, 379–384. W.T., Swenson, D., Sacchettini, J.C., Jacobs Jr., W.R., Russell, D.G., 2000. Persis-
Hiraoka, A., Horiike, N., Akbar, S.M., Michitaka, K., Matsuyama, T., Onji, M., 2005b. tence of Mycobacterium tuberculosis in macrophages and mice requires the
Soluble CD163 in patients with liver diseases: very high levels of soluble glyoxylate shunt enzyme isocitrate lyase. Nature 406, 735–738.
CD163 in patients with fulminant hepatic failure. J. Gastroenterol. 40, 52– Mollenhauer, J., Wiemann, S., Scheurlen, W., Korn, B., Hayashi, Y., Wilgenbus, K.K.,
56. von Deimling, A., Poustka, A., 1997. DMBT1, a new member of the SRCR super-
Hogger, P., Dreier, J., Droste, A., Buck, F., Sorg, C., 1998a. Identification of the integral family, on chromosome 10q25.3–26.1 is deleted in malignant brain tumours.
membrane protein RM3/1 on human monocytes as a glucocorticoid-inducible Nat. Genet. 17, 32–39.
member of the scavenger receptor cysteine-rich family (CD163). J. Immunol. Moller, H.J., Aerts, H., Gronbaek, H., Peterslund, N.A., Hyltoft Petersen, P., Hornung,
161, 1883–1890. N., Rejnmark, L., Jabbarpour, E., Moestrup, S.K., 2002a. Soluble CD163: a marker
Hogger, P., Erpenstein, U., Rohdewald, P., Sorg, C., 1998b. Biochemical charac- molecule for monocyte/macrophage activity in disease. Scand. J. Clin. Lab. Invest.
terization of a glucocorticoid-induced membrane protein (RM3/1) in human Suppl. 237, 29–33.
monocytes and its application as model system for ranking glucocorticoid Moller, H.J., de Fost, M., Aerts, H., Hollak, C., Moestrup, S.K., 2004. Plasma level of the
potency. Pharm. Res. 15, 296–302. macrophage-derived soluble CD163 is increased and positively correlates with
Hogger, P., Sorg, C., 2001. Soluble CD163 inhibits phorbol ester-induced lymphocyte severity in Gaucher’s disease. Eur. J. Haematol. 72, 135–139.
proliferation. Biochem. Biophys. Res. Commun. 288, 841–843. Moller, H.J., Gronbaek, H., Schiodt, F.V., Holland-Fischer, P., Schilsky, M., Munoz,
Hohenester, E., Sasaki, T., Timpl, R., 1999. Crystal structure of a scavenger receptor S., Hassanein, T., Lee, W.M., 2007. Soluble CD163 from activated macrophages
cysteine-rich domain sheds light on an ancient superfamily. Nat. Struct. Biol. 6, predicts mortality in acute liver failure. J. Hepatol. 47, 671–676.
228–232. Moller, H.J., Hald, K., Moestrup, S.K., 2002b. Characterization of an enzyme-linked
Holm, D., Fink, D.R., Gronlund, J., Hansen, S., Holmskov, U., 2009. Cloning and immunosorbent assay for soluble CD163. Scand. J. Clin. Lab. Invest. 62, 293–299.
characterization of SCART1, a novel scavenger receptor cysteine-rich type I Moller, H.J., Moestrup, S.K., Weis, N., Wejse, C., Nielsen, H., Pedersen, S.S., Attermann,
transmembrane molecule. Mol. Immunol. 46, 1663–1672. J., Nexo, E., Kronborg, G., 2006. Macrophage serum markers in pneumococ-
Holmskov, U., Lawson, P., Teisner, B., Tornoe, I., Willis, A.C., Morgan, C., Koch, C., Reid, cal bacteremia: prediction of survival by soluble CD163. Crit. Care Med. 34,
K.B., 1997. Isolation and characterization of a new member of the scavenger 2561–2566.
receptor superfamily, glycoprotein-340 (gp-340), as a lung surfactant protein-D Moller, H.J., Nielsen, M.J., Maniecki, M.B., Madsen, M., Moestrup, S.K., 2009. Soluble
binding molecule. J. Biol. Chem. 272, 13743–13749. macrophage-derived CD163: a homogenous ectodomain protein with a disso-
Hwang, P.K., Greer, J., 1980. Interaction between hemoglobin subunits in the ciable haptoglobin–hemoglobin binding. Immunobiology, 1003, Epub ahead of
hemoglobin–haptoglobin complex. J. Biol. Chem. 255, 3038–3041. print. 10.1016/j.imbio.2009.1005.
Iwasaki, K., Morimatsu, M., Inanami, O., Uchida, E., Syuto, B., Kuwabara, M., Niiyama, Moller, H.J., Petersen, P.H., Rejnmark, L., Moestrup, S.K., 2003. Biological variation of
M., 2001. Isolation, characterization, and cDNA cloning of chicken turpentine- soluble CD163. Scand. J. Clin. Lab. Invest. 63, 15–21.
induced protein, a new member of the scavenger receptor cysteine-rich (SRCR) Moller, H.J., Peterslund, N.A., Graversen, J.H., Moestrup, S.K., 2002c. Identification
family of proteins. J. Biol. Chem. 276, 9400–9405. of the hemoglobin scavenger receptor/CD163 as a natural soluble protein in
Jones, N.H., Clabby, M.L., Dialynas, D.P., Huang, H.J., Herzenberg, L.A., Strominger, J.L., plasma. Blood 99, 378–380.
1986. Isolation of complementary DNA clones encoding the human lymphocyte Moreno, J.A., Munoz-Garcia, B., Martin-Ventura, J.L., Madrigal-Matute, J., Orbe, J.,
glycoprotein T1/Leu-1. Nature 323, 346–349. Paramo, J.A., Ortega, L., Egido, J., Blanco-Colio, L.M., 2009. The CD163-expressing
Kishimoto, T., Goyert, S., Kikutani, H., Mason, D., Miyasaka, M., Moretta, L., Ohno, macrophages recognize and internalize TWEAK potential consequences in
T., Okumura, K., Shaw, S., Springer, T.A., Sugamura, K., Sugawara, H., von dem atherosclerosis. Atherosclerosis 207, 103–110.
Borne, A.E., Zola, H., 1997. CD antigens 1996. Blood 89, 3502. Morganelli, P.M., Guyre, P.M., 1988. IFN-gamma plus glucocorticoids stimulate the
Kisielow, J., Kopf, M., Karjalainen, K., 2008. SCART scavenger receptors identify a expression of a newly identified human mononuclear phagocyte-specific anti-
novel subset of adult gammadelta T cells. J. Immunol. 181, 1710–1716. gen. J. Immunol. 140, 2296–2304.
Komohara, Y., Hirahara, J., Horikawa, T., Kawamura, K., Kiyota, E., Sakashita, N., Araki, Mosser, D.M., 2003. The many faces of macrophage activation. J. Leukoc. Biol. 73,
N., Takeya, M., 2006. AM-3K, an anti-macrophage antibody, recognizes CD163, 209–212.
a molecule associated with an anti-inflammatory macrophage phenotype. J. Murphy, J.E., Tedbury, P.R., Homer-Vanniasinkam, S., Walker, J.H., Ponnambalam,
Histochem. Cytochem. 54, 763–771. S., 2005. Biochemistry and cell biology of mammalian scavenger receptors.
Koskinen, R., Salomonsen, J., Goodchild, M., Bumstead, N., Boyd, Y., Vainio, O., 2001. Atherosclerosis 182, 1–15.
Structure and chromosomal localization of chicken CD5. Scand. J. Immunol. 54, Nielsen, M.J., Madsen, M., Moller, H.J., Moestrup, S.K., 2006. The macrophage scav-
141–145. enger receptor CD163: endocytic properties of cytoplasmic tail variants. J.
Koury, M.J., 2007. Scavenger receptor helps erythroblasts stay on island. Blood 109, Leukoc. Biol. 79, 837–845.
5074–5075. Nielsen, M.J., Moestrup, S.K., 2009. Receptor targeting of hemoglobin mediated by
Krieger, M., 1997. The other side of scavenger receptors: pattern recognition for host the haptoglobins: roles beyond heme scavenging. Blood 114, 764–771.
defense. Curr. Opin. Lipidol. 8, 275–280. Nielsen, M.J., Moller, H.J., Moestrup, S.K., 2010. Hemoglobin and heme scavenger
Krieger, M., Herz, J., 1994. Structures and functions of multiligand lipoprotein recep- receptors. Antioxid. Redox Signal. 12, 261–273.
tors: macrophage scavenger receptors and LDL receptor-related protein (LRP). Nunes, D.P., Keates, A.C., Afdhal, N.H., Offner, G.D., 1995. Bovine gall-bladder mucin
Annu. Rev. Biochem. 63, 601–637. contains two distinct tandem repeating sequences: evidence for scavenger
Kristiansen, M., Graversen, J.H., Jacobsen, C., Sonne, O., Hoffman, H.J., Law, S.K., receptor cysteine-rich repeats. Biochem. J. 310, 41–48.
Moestrup, S.K., 2001. Identification of the haemoglobin scavenger receptor. Ojala, J.R., Pikkarainen, T., Tuuttila, A., Sandalova, T., Tryggvason, K., 2007. Crystal
Nature 409, 198–201. structure of the cysteine-rich domain of scavenger receptor MARCO reveals the
H. Van Gorp et al. / Molecular Immunology 47 (2010) 1650–1660 1659

presence of a basic and an acidic cluster that both contribute to ligand recogni- Schaer, D.J., Boretti, F.S., Hongegger, A., Poehler, D., Linnscheid, P., Staege, H., Muller,
tion. J. Biol. Chem. 282, 16654–16666. C., Schoedon, G., Schaffner, A., 2001. Molecular cloning and characterization of
Onofre, G., Kolackova, M., Jankovicova, K., Krejsek, J., 2009. Scavenger receptor CD163 the mouse CD163 homologue, a highly glucocorticoid-inducible member of the
and its biological functions. Acta Med. (Hradec Kralove) 52, 57–61. scavenger receptor cysteine-rich family. Immunogenetics 53, 170–177.
Otterbein, L.E., Soares, M.P., Yamashita, K., Bach, F.H., 2003. Heme oxygenase-1: Schaer, D.J., Schleiffenbaum, B., Kurrer, M., Imhof, A., Bachli, E., Fehr, J., Moller, H.J.,
unleashing the protective properties of heme. Trends Immunol. 24, 449–455. Moestrup, S.K., Schaffner, A., 2005. Soluble hemoglobin–haptoglobin scavenger
Pancer, Z., Munkner, J., Muller, I., Muller, W.E., 1997. A novel member of an ancient receptor CD163 as a lineage-specific marker in the reactive hemophagocytic
superfamily: sponge (Geodia cydonium, Porifera) putative protein that features syndrome. Eur. J. Haematol. 74, 6–10.
scavenger receptor cysteine-rich repeats. Gene 193, 211–218. Soares, M.P., Bach, F.H., 2009. Heme oxygenase-1: from biology to therapeutic poten-
Patton, J.B., Rowland, R.R., Yoo, D., Chang, K.O., 2009. Modulation of CD163 receptor tial. Trends Mol. Med. 15, 50–58.
expression and replication of porcine reproductive and respiratory syndrome Somoza, J.R., Ho, J.D., Luong, C., Ghate, M., Sprengeler, P.A., Mortara, K., Shrader, W.D.,
virus in porcine macrophages. Virus Res. 140, 161–171. Sperandio, D., Chan, H., McGrath, M.E., Katz, B.A., 2003. The structure of the extra-
Perez, C., Ortuno, E., Gomez, N., Garcia-Briones, M., Alvarez, B., Martinez de la Riva, P., cellular region of human hepsin reveals a serine protease domain and a novel
Alonso, F., Revilla, C., Dominguez, J., Ezquerra, A., 2008. Cloning and expression of scavenger receptor cysteine-rich (SRCR) domain. Structure 11, 1123–1131.
porcine CD163: its use for characterization of monoclonal antibodies to porcine Sopp, P., Werling, D., Baldwin, C., 2007. Cross-reactivity of mAbs to human CD anti-
CD163 and development of an ELISA to measure soluble CD163 in biological gens with cells from cattle. Vet. Immunol. Immunopathol. 119, 106–114.
fluids. Span. J. Agric. Res. 6, 59–72. Stoddard, E., Cannon, G., Ni, H., Kariko, K., Capodici, J., Malamud, D., Weissman, D.,
Philippidis, P., Mason, J.C., Evans, B.J., Nadra, I., Taylor, K.M., Haskard, D.O., Lan- 2007. gp340 expressed on human genital epithelia binds HIV-1 envelope protein
dis, R.C., 2004. Hemoglobin scavenger receptor CD163 mediates interleukin-10 and facilitates viral transmission. J. Immunol. 179, 3126–3132.
release and heme oxygenase-1 synthesis: antiinflammatory monocyte- Stover, C.M., Schleypen, J., Gronlund, J., Speicher, M.R., Schwaeble, W.J., Holmskov,
macrophage responses in vitro, in resolving skin blisters in vivo, and after U., 2000. Assignment of CD163B, the gene encoding M160, a novel scavenger
cardiopulmonary bypass surgery. Circ. Res. 94, 119–126. receptor, to human chromosome 12p13.3 by in situ hybridization and somatic
Pluddemann, A., Neyen, C., Gordon, S., 2007. Macrophage scavenger receptors and cell hybrid analysis. Cytogenet. Cell Genet. 90, 246–247.
host-derived ligands. Methods 43, 207–217. Sulahian, T.H., Hintz, K.A., Wardwell, K., Guyre, P.M., 2001. Development of an ELISA
Polek, T.C., Talpaz, M., Darnay, B.G., Spivak-Kroizman, T., 2003. TWEAK medi- to measure soluble CD163 in biological fluids. J. Immunol. Methods 252, 25–
ates signal transduction and differentiation of RAW264.7 cells in the absence 31.
of Fn14/TweakR. Evidence for a second TWEAK receptor. J. Biol. Chem. 278, Sulahian, T.H., Hogger, P., Wahner, A.E., Wardwell, K., Goulding, N.J., Sorg, C., Droste,
32317–32323. A., Stehling, M., Wallace, P.K., Morganelli, P.M., Guyre, P.M., 2000. Human mono-
Polfliet, M.M., Fabriek, B.O., Daniels, W.P., Dijkstra, C.D., van den Berg, T.K., 2006. The cytes express CD163, which is upregulated by IL-10 and identical to p155.
rat macrophage scavenger receptor CD163: expression, regulation and role in Cytokine 12, 1312–1321.
inflammatory mediator production. Immunobiology 211, 419–425. Sulahian, T.H., Pioli, P.A., Wardwell, K., Guyre, P.M., 2004. Cross-linking of FcgammaR
Pulford, K., Micklem, K., McCarthy, S., Cordell, J., Jones, M., Mason, D.Y., 1992. A triggers shedding of the hemoglobin–haptoglobin scavenger receptor CD163. J.
monocyte/macrophage antigen recognized by the four antibodies GHI/61, Ber- Leukoc. Biol. 76, 271–277.
MAC3, Ki-M8 and SM4. Immunology 75, 588–595. Takito, J., Hikita, C., Al-Awqati, Q., 1996. Hensin, a new collecting duct protein
Radzun, H.J., Kreipe, H., Bodewadt, S., Hansmann, M.L., Barth, J., Parwaresch, M.R., involved in the in vitro plasticity of intercalated cell polarity. J. Clin. Invest. 98,
1987. Ki-M8 monoclonal antibody reactive with an intracytoplasmic antigen of 2324–2331.
monocyte/macrophage lineage. Blood 69, 1320–1327. Taylor, P.R., Martinez-Pomares, L., Stacey, M., Lin, H.H., Brown, G.D., Gordon, S.,
Resnick, D., Pearson, A., Krieger, M., 1994. The SRCR superfamily: a family reminis- 2005. Macrophage receptors and immune recognition. Annu. Rev. Immunol. 23,
cent of the Ig superfamily. Trends Biochem. Sci. 19, 5–8. 901–944.
Ritter, M., Buechler, C., Langmann, T., Schmitz, G., 1999. Genomic organization and Timmermann, M., Buck, F., Sorg, C., Hogger, P., 2004. Interaction of soluble CD163
chromosomal localization of the human CD163 (M130) gene: a member of the with activated T lymphocytes involves its association with non-muscle myosin
scavenger receptor cysteine-rich superfamily. Biochem. Biophys. Res. Commun. heavy chain type A. Immunol. Cell Biol. 82, 479–487.
260, 466–474. Timmermann, M., Hogger, P., 2005. Oxidative stress and 8-iso-prostaglandin
Rodamilans, B., Ibanez, S., Bragado-Nilsson, E., Sarrias, M.R., Lozano, F., Blanco, F.J., F(2alpha) induce ectodomain shedding of CD163 and release of tumor necrosis
Montoya, G., 2007a. Expression, purification and crystallization of human CD5 factor-alpha from human monocytes. Free Radic. Biol. Med. 39, 98–107.
domain III, a nano-scale crystallization example. J. Struct. Biol. 159, 144–148. Topoll, H., Zwadlo, G., Lange, D.E., Sorg, C., 1988. Phenotypic dynamics of macrophage
Rodamilans, B., Munoz, I.G., Bragado-Nilsson, E., Sarrias, M.R., Padilla, O., Blanco, subsets during human gingivitis. J. Dent. Res. 67, 685.
F.J., Lozano, F., Montoya, G., 2007b. Crystal structure of the third extracellular Van den Heuvel, M.M., Tensen, C.P., van As, J.H., Van den Berg, T.K., Fluitsma, D.M.,
domain of CD5 reveals the fold of a group B scavenger cysteine-rich receptor Dijkstra, C.D., Dopp, E.A., Droste, A., Van Gaalen, F.A., Sorg, C., Hogger, P., Beelen,
domain. J. Biol. Chem. 282, 12669–12677. R.H., 1999. Regulation of CD163 on human macrophages: cross-linking of CD163
Rughetti, A., Biffoni, M., Pierelli, L., Rahimi, H., Bonanno, G., Barachini, S., Pellicciotta, induces signaling and activation. J. Leukoc. Biol. 66, 858–866.
I., Napoletano, C., Pescarmona, E., Del Nero, A., Pignoloni, P., Frati, L., Nuti, M., Van Gorp, H., Van Breedam, W., Delputte, P.L., Nauwynck, H.J., 2008. Sialoadhesin
2003. Regulated expression of MUC1 epithelial antigen in erythropoiesis. Br. J. and CD163 join forces during entry of the porcine reproductive and respiratory
Haematol. 120, 344–352. syndrome virus. J. Gen. Virol. 89, 2943–2953.
Sadahira, Y., Yoshino, T., Monobe, Y., 1995. Very late activation antigen 4-vascular Van Gorp, H., Van Breedam, W., Delputte, P.L., Nauwynck, H.J., 2009. The porcine
cell adhesion molecule 1 interaction is involved in the formation of erythrob- reproductive and respiratory syndrome virus requires trafficking through
lastic islands. J. Exp. Med. 181, 411–415. CD163 positive early endosomes, but not late endosomes, for productive infec-
Sanchez, C., Domenech, N., Vazquez, J., Alonso, F., Ezquerra, A., Dominguez, J., tion. Arch. Virol. 154, 1939–1943.
1999. The porcine 2A10 antigen is homologous to human CD163 and related Van Gorp, H., Van Breedam, W., Van Doorsselaere, J., Delputte, P.L., Nauwynck,
to macrophage differentiation. J. Immunol. 162, 5230–5237. H.J., 2010. Identification of the CD163 protein domains involved in infection
Sanchez-Torres, C., Gomez-Puertas, P., Gomez-del-Moral, M., Alonso, F., Escrib- of the porcine reproductive and respiratory syndrome virus. J. Virol. 84, 3101–
ano, J.M., Ezquerra, A., Dominguez, J., 2003. Expression of porcine CD163 on 3105.
monocytes/macrophages correlates with permissiveness to African swine fever Verschure, P.J., Van Noorden, C.J., Dijkstra, C.D., 1989. Macrophages and dendritic
infection. Arch. Virol. 148, 2307–2323. cells during the early stages of antigen-induced arthritis in rats: immunohisto-
Sarrias, M.R., Farnos, M., Mota, R., Sanchez-Barbero, F., Ibanez, A., Gimferrer, I., Vera, chemical analysis of cryostat sections of the whole knee joint. Scand. J. Immunol.
J., Fenutria, R., Casals, C., Yelamos, J., Lozano, F., 2007. CD6 binds to pathogen- 29, 371–381.
associated molecular patterns and protects from LPS-induced septic shock. Proc. Weaver, L.K., Hintz-Goldstein, K.A., Pioli, P.A., Wardwell, K., Qureshi, N., Vogel, S.N.,
Natl. Acad. Sci. U.S.A. 104, 11724–11729. Guyre, P.M., 2006. Pivotal advance: activation of cell surface Toll-like receptors
Sarrias, M.R., Gronlund, J., Padilla, O., Madsen, J., Holmskov, U., Lozano, F., 2004. The causes shedding of the hemoglobin scavenger receptor CD163. J. Leukoc. Biol.
scavenger receptor cysteine-rich (SRCR) domain: an ancient and highly con- 80, 26–35.
served protein module of the innate immune system. Crit. Rev. Immunol. 24, Weaver, L.K., Pioli, P.A., Wardwell, K., Vogel, S.N., Guyre, P.M., 2007. Up-regulation
1–37. of human monocyte CD163 upon activation of cell-surface Toll-like receptors. J.
Sarrias, M.R., Rosello, S., Sanchez-Barbero, F., Sierra, J.M., Vila, J., Yelamos, J., Vives, Leukoc. Biol. 81, 663–671.
J., Casals, C., Lozano, F., 2005. A role for human Sp alpha as a pattern recognition Wijngaard, P.L., Metzelaar, M.J., MacHugh, N.D., Morrison, W.I., Clevers, H.C., 1992.
receptor. J. Biol. Chem. 280, 35391–35398. Molecular characterization of the WC1 antigen expressed specifically on bovine
Schaer, C.A., Schoedon, G., Imhof, A., Kurrer, M.O., Schaer, D.J., 2006. Constitutive CD4-CD8-gamma delta T lymphocytes. J. Immunol. 149, 3273–3277.
endocytosis of CD163 mediates hemoglobin-heme uptake and determines the Wiley, S.R., Cassiano, L., Lofton, T., Davis-Smith, T., Winkles, J.A., Lindner, V., Liu,
noninflammatory and protective transcriptional response of macrophages to H., Daniel, T.O., Smith, C.A., Fanslow, W.C., 2001. A novel TNF receptor family
hemoglobin. Circ. Res. 99, 943–950. member binds TWEAK and is implicated in angiogenesis. Immunity 15, 837–846.
Schaer, C.A., Vallelian, F., Imhof, A., Schoedon, G., Schaer, D.J., 2008. Heme carrier Williams, L., Jarai, G., Smith, A., Finan, P., 2002. IL-10 expression profiling in human
protein (HCP-1) spatially interacts with the CD163 hemoglobin uptake path- monocytes. J. Leukoc. Biol. 72, 800–809.
way and is a target of inflammatory macrophage activation. J. Leukoc. Biol. 83, Wu, Z., Lee, S., Abrams, W., Weissman, D., Malamud, D., 2006. The N-terminal SRCR-
325–333. SID domain of gp-340 interacts with HIV type 1 gp120 sequences and inhibits
Schaer, D.J., Alayash, A.I., Buehler, P.W., 2007. Gating the radical hemoglobin to viral infection. AIDS Res. Hum. Retrovir. 22, 508–515.
macrophages: the anti-inflammatory role of CD163, a scavenger receptor. Yamada, K., Takabatake, T., Takeshima, K., 2000. Isolation and characterization of
Antioxid. Redox Signal. 9, 991–999. three novel serine protease genes from Xenopus laevis. Gene 252, 209–216.
1660 H. Van Gorp et al. / Molecular Immunology 47 (2010) 1650–1660

Zeng, L., Takeya, M., Takahashi, K., 1996. AM-3K, a novel monoclonal antibody spe- Zwadlo-Klarwasser, G., Bent, S., Haubeck, H.D., Sorg, C., Schmutzler, W., 1990.
cific for tissue macrophages and its application to pathological investigation. J. Glucocorticoid-induced appearance of the macrophage subtype RM 3/1 in
Pathol. 178, 207–214. peripheral blood of man. Int. Arch. Allergy Appl. Immunol. 91, 175–180.
Zuwala-Jagiello, J., 2006. Haemoglobin scavenger receptor: function in relation to Zwadlo-Klarwasser, G., Neubert, R., Stahlmann, R., Schmutzler, W., 1992. Influence
disease. Acta Biochim. Pol. 53, 257–268. of dexamethasone on the RM 3/1-positive macrophages in the peripheral blood
Zwadlo, G., Voegeli, R., Osthoff, K.S., Sorg, C., 1987. A monoclonal antibody to a and tissues of a New World monkey (the marmoset Callithrix jacchus). Int. Arch.
novel differentiation antigen on human macrophages associated with the down- Allergy Immunol. 97, 178–180.
regulatory phase of the inflammatory process. Exp. Cell Biol. 55, 295–304.

Вам также может понравиться