CHEMOKINES have recently become the focus of intense interest and discussion. This has occurred, in part, because of an enlarging view of what chemokines do. As recently as 2 or 3 years ago a chemokine review would have begun with a discussion of the importance of chemotactic factors in controlling leukocyte function and trafficking, and would have pointed out that specificity for leukocyte subsets is what sets chemokines apart from other chemoattractants. For example, formylated peptides (eg, f-Met-Leu-Phe), complement fragments (eg, C5a), or arachidonic acid metabolites (eg, LTB4) attract neutrophils and monocytes with equal potency, whereas some chemokines (eg, interleukin-8 [IL-8]) attract neutrophils but have no discernible effects on monocytes. This discrimination has led to the frequently espoused proposition that chemokines are involved in the pathogenesis of diseases having characteristic infiltrates.

However, we now know that chemokines and their receptors are expressed by a wide variety of nonhematopoietic cells, and that chemokine function extends far beyond leukocyte physiology. Even within the world of leukocytes, the connections among chemokines, their receptors, and human immunodeficiency virus (HIV) infection broadens the previously narrow focus on chemokines as mere chemoattractants. Furthermore, a proliferation of animal models has more precisely defined the functions of chemokines in vivo. This review will attempt to describe what is understood about chemokines in light of recent discoveries.

Based on genomics efforts, it has been estimated that there may be as many as 40 to 50 human chemokines. The determination that a gene encodes a chemokine depends on structural and, to a lesser extent, genetic criteria.1 Chemokines are small, with molecular weights in the range of 8 to 12 kD, but there are exceptions which involve proteins comprised of multiple domains, one of which looks like a chemokine.2,3 Chemokine domains are defined by the presence of four cysteines in highly conserved positions (Fig 1). One major chemokine subfamily is called “CXC” because the two amino acids nearest the N-termini of these proteins are separated by a single amino acid. This is in contrast to the other major subfamily which is called “CC” because these two cysteines are adjacent.

Fig. 1.

Chemokine families. Chemokines are divided into families based on structural and genetic considerations. All chemokines are structurally similar, having at least three β-pleated sheets (indicated as β1-3) and a C-terminal α-helix. Most chemokines also have at least four cysteines in conserved positions. In the CXC chemokine family, the two cysteines nearest the N-termini of family members are separated by a single (and variable) amino acid. The genes encoding these proteins cluster at human chromosome 14q12-21 (except for SDF-1α whose gene maps to chromosome 1053). In the CC chemokine family, the two cysteines nearest the N-termini of these proteins are adjacent. Their genes cluster at 17q11.2-12 (except for MIP-3β, whose gene maps to chromosome 9117 and MIP-3α/LARC which maps to chromosome 2117a). Lymphotactin is a structurally related chemokine having only one cysteine near its N-terminus and is said to belong to the C chemokine family. The CX3C chemokine (also called “fractalkine” or “neurotactin”) has a typical chemokine-like structure at its N-terminus except for the placement of three amino acids between the first two cysteines. This chemokine domain occurs at the end of a long stalk which is heavily substituted with mucin-like carbohydrates. The protein is embedded in the membrane and has a short cytoplasmic domain.

Fig. 1.

Chemokine families. Chemokines are divided into families based on structural and genetic considerations. All chemokines are structurally similar, having at least three β-pleated sheets (indicated as β1-3) and a C-terminal α-helix. Most chemokines also have at least four cysteines in conserved positions. In the CXC chemokine family, the two cysteines nearest the N-termini of family members are separated by a single (and variable) amino acid. The genes encoding these proteins cluster at human chromosome 14q12-21 (except for SDF-1α whose gene maps to chromosome 1053). In the CC chemokine family, the two cysteines nearest the N-termini of these proteins are adjacent. Their genes cluster at 17q11.2-12 (except for MIP-3β, whose gene maps to chromosome 9117 and MIP-3α/LARC which maps to chromosome 2117a). Lymphotactin is a structurally related chemokine having only one cysteine near its N-terminus and is said to belong to the C chemokine family. The CX3C chemokine (also called “fractalkine” or “neurotactin”) has a typical chemokine-like structure at its N-terminus except for the placement of three amino acids between the first two cysteines. This chemokine domain occurs at the end of a long stalk which is heavily substituted with mucin-like carbohydrates. The protein is embedded in the membrane and has a short cytoplasmic domain.

Close modal

Many of the genes encoding chemokines have been mapped, and they cluster at specific loci. CC chemokine genes are grouped at 17q11.2-12 and CXC chemokine genes at 4q13 (although there are exceptions; see below). This suggests that chemokines arose by duplication and divergence from a primordial chemokine gene, with an early split into the two loci. Because mice appear to have fewer chemokines than humans (for example, there is no clear-cut murine homolog of IL-8), some of this development may have occurred relatively recently in evolutionary terms, but this has not been systematically examined. Similar clustering has been observed among genes encoding chemokine receptors.

One exception to the CC/CXC rule is lymphotactin, a potent attractant for T lymphocytes, but not monocytes.4 Although it is the right size to be a chemokine and it has several characteristic sequence signatures common to CC chemokines, it has only two cysteines. Nonetheless, these cysteines correspond to cysteines 2 and 4 of chemokines, and it has been suggested that lymphotactin belongs to a third chemokine subfamily denoted C, because of the lone cysteine in the N-terminal domain. Consistent with its new assignment, lymphotactin's gene maps to 1q. So far, lymphotactin is the only member of this putative family.

Another exception is a CX3C chemokine (also referred to as “fractalkine” or “neurotactin”) that is an integral membrane protein with a chemokine domain at its N-terminus.3,3a This domain differs from other chemokines by the presence of three amino acids intervening between the first two cysteines. Modeling studies suggest that the three-dimensional structure of chemokines (see below) can only accommodate 0, 1, or 3 amino acids between the first two cysteines, explaining the absence of a CX2C chemokine. Like lymphotactin, the gene encoding CX3C chemokine maps to a locus different from other chemokines, and so far it is the sole example of a chemokine with this structural motif.

Discussions of chemokine function tend to look like a collection of disconnected observations. It is extremely difficult to divine rules, particularly because the dose responses for given effects in vitro may not be relevant in vivo. For example, the fact that the ED50 for MCP-1's ability to attract monocytes is 10-fold lower than MCP-2's may imply that MCP-1 is a more potent chemoattractant.5 However, a 10-fold increase in MCP-2 expression in vivo would make it physiologically equipotent with MCP-1. Therefore, the following discussion merely catalogs chemokine properties without necessarily implying their physiologic importance in vivo.

ELR chemokines.The prototypic CXC chemokine is IL-8, which was purified by several groups as a monocyte-derived factor that attracts neutrophils, but not monocytes, in Boyden chamber assays.6-8 Several other CXC chemokines are also potent neutrophil chemoattractants (see Table 1), and structure/activity analyses show that this property depends on the presence of a three–amino acid motif, ELR (glutamate-leucine-arginine), between the N-terminus and the first cysteine.9,10 However, these amino acids must appear in positions close to the proteins' N-termini because platelet basic protein (PBP) and two of its N-terminally truncated derivatives, CTAPIII and β-thromboglobulin, have very weak neutrophil chemoattractant activity despite the presence of ELR. Only NAP-2, a further truncated product in which ELR appears close to the N-terminus, is an active PBP-derived neutrophil attractant.11 

Table 1.

CXC Chemokines

NameTarget Cells
ELR 
IL-8 Neutrophils, T lymphocytes, basophils, angiogenesis (?endothelial cells) 
GRO-α(MGSA) Neutrophils, melanoma cells, (?endothelial cells) 
GRO-β(MIP-2α) Neutrophils, (?endothelial cells) 
GRO-γ(MIP-2β) Neutrophils, (?endothelial cells) 
ENA-78 Neutrophils 
GCP-2 Neutrophils 
Platelet basic protein 
CTAP-III Fibroblasts 
β-Thromboglobulin Fibroblasts 
NAP-2 Neutrophils, basophils 
non-ELR 
Platelet factor 4 Fibroblasts, endothelial cells 
IP-10 Activated T lymphocytes, TIL, ?endothelial cells, ?NK cells 
MIG Activated T lymphocytes, TIL 
SDF-1α T lymphocytes, CD34+ progenitors, ?B lymphocytes 
SDF-1β 
NameTarget Cells
ELR 
IL-8 Neutrophils, T lymphocytes, basophils, angiogenesis (?endothelial cells) 
GRO-α(MGSA) Neutrophils, melanoma cells, (?endothelial cells) 
GRO-β(MIP-2α) Neutrophils, (?endothelial cells) 
GRO-γ(MIP-2β) Neutrophils, (?endothelial cells) 
ENA-78 Neutrophils 
GCP-2 Neutrophils 
Platelet basic protein 
CTAP-III Fibroblasts 
β-Thromboglobulin Fibroblasts 
NAP-2 Neutrophils, basophils 
non-ELR 
Platelet factor 4 Fibroblasts, endothelial cells 
IP-10 Activated T lymphocytes, TIL, ?endothelial cells, ?NK cells 
MIG Activated T lymphocytes, TIL 
SDF-1α T lymphocytes, CD34+ progenitors, ?B lymphocytes 
SDF-1β 

See text for references.

IL-8 is produced by a variety of cell types including monocytes, T lymphocytes, neutrophils, fibroblasts, endothelial cells, and epithelial cells. Although subject to variable processing at the N-terminus, the most abundant form of naturally occurring IL-8 is 72 amino acids long.12,13 A 77–amino acid variant, occasionally called endothelial IL-8 because of its synthesis by these cells, is extended at the N-terminus. The longer protein is ≈10-fold less potent than the shorter protein in attracting and activating neutrophils in vitro, but has similar potency in vivo, perhaps due to proteolytic processing to the short form. The 77–amino acid form may be involved in neutrophil adherence to the endothelium as a prelude to diapedesis.14 (The long form has also been observed to inhibit neutrophil adhesion to activated endothelial cells, but this may be a vagary of the assay system.15 ) Other properties attributed to IL-8 include chemoattraction of T lymphocytes16 (but not in transendothelial assays17 ) and angiogenic activity (see below). The former is a direct effect because IL-8 receptors have been documented on CD8+, CD26 T cells,18 while the latter is controversial because of the reported absence of IL-8 receptors on endothelial cells.19 Finally, IL-8 stimulates histamine release from basophils.20,21 

GRO-α was so named because of its initial description as the product of a gene differentially expressed in transformed hamster cells that had suffered loss of growth control.22 Independently, the murine homolog had been cloned in a differential screening experiment as the platelet-derived growth factor (PDGF)-inducible KC gene, and the human protein was purified as MGSA, or melanoma growth stimulatory activity, because of its mitogenic effects on melanoma cell lines.23-25 However, GRO-α was also functionally identified as a neutrophil-specific chemoattractant secreted by activated mononuclear cells along with IL-8 and having similar potency.26,27 Thus, GRO-α finds a comfortable place within the chemokine family as a neutrophil chemoattractant, but it is instructive to note that its influences have always extended beyond leukocytes. GRO-β and GRO-γ are closely related proteins that are also potent neutrophil attractants.28 (The alternate designations, MIP-2α and MIP-2β, derive from their purification as neutrophil chemoattractants by the same investigators who isolated the CC chemokines, MIP-1α and MIP-1β.29,30 )

ENA78 is an ELR-containing CXC chemokine isolated from A549 cells, which are derived from type II pneumocytes that also secrete IL-8 and GRO-γ.31 Its sequence places it closer to the GRO proteins than IL-8, and like these chemokines it specifically attracts neutrophils. Similarly, GCP-2 was purified from the conditioned medium of MG63 osteosarcoma cells along with IL-8, GRO-α, GRO-γ, and IP-10.32 It is a neutrophil-specific chemoattractant and activator, but has a specific activity ≈5- to 10-fold lower than IL-8. Like IL-8, GCP-2 attracts neutrophils when injected into rabbit skin.

As mentioned above, PBP and its processed products CTAP-III and β-TG have ELR sequences but are poor neutrophil chemoattractants because of their extended N-terminal sequences. Removal of nine N-terminal amino acids from PBP33 produces CTAP-III which stimulates glycosaminoglycan production by connective tissue cells and is a very weak mitogen for fibroblasts (ED50 ≈ 100 nmol/L).34,35 Proteolytic removal of another four N-terminal amino acids produces β-TG, which is a chemoattractant for fibroblasts.36 Cleaving another 11 N-terminal residues results in NAP-2,37 whose potency as a neutrophil chemoattractant and activator has been estimated to be 2-fold11 to 100-fold38 less than that of IL-8.

Non-ELR CXC chemokines.The ELR-containing chemokines have an apparent uniformity of function which makes it easy to think of them as a family of neutrophil chemoattractants and activators (although this raises questions of redundancy). In contrast, descriptions of non-ELR CXC chemokines are a hodgepodge of disparate activities with as yet no clear underlying theme.

For example, long before the chemokine family was recognized or named, platelet factor 4 (PF4) was the first member to purified. It is found in platelet α-granules along with PBP and its processed products, but unlike those proteins it has no ELR motif and is an extremely weak attractant for neutrophils.39 It attracts fibroblasts in vitro but is ≈30-fold less potent than β-TG.36 However, one of its most interesting properties is its inhibition of angiogenesis, which appears to occur via direct inhibition of endothelial cell proliferation.40 Although PF4's ED50 for this effect is quite high, ≈250 nmol/L, a cleavage product is 30 to 50 times more potent (see below).41 

Another non-ELR CXC chemokine with antiangiogenic properties is IP-10, the product of an interferon-γ (IFN-γ)-inducible gene cloned from U937 cells.42 A variety of cell types express IP-10 in vitro including mononuclear cells, keratinocytes, fibroblasts, endothelial cells, and T lymphocytes.42 In the mouse, IFN-γ administration induces high levels of IP-10 expression in liver and kidney with lower levels in the spleen.43 Similar to other non-ELR chemokines, IP-10 is a poor neutrophil chemoattractant and activator.44 Some groups have ascribed T-lymphocyte chemoattractant properties to the protein both in vitro in Boyden chamber assays, and in vivo in severe combined immunodeficient (SCID) mice reconstituted with human peripheral blood lymphocytes (PBLs).45,46 However, others detect no such T-lymphocyte–directed activity in transendothelial models of T-cell migration.17 

MIG is another IFN-γ–inducible protein isolated from macrophages. It has chemoattractant activity in vitro for tumor-infiltrating lymphocytes (TIL), and for PBLs activated by syngeneic monocytes and phytohemogglutinin (PHA).47 IP-10 also attracts TIL, and MIG and IP-10 cross-desensitize in other measures of receptor activation, suggesting that on TIL they share the same receptor.47 This receptor, CXCR3, has recently been cloned and binds IP-10 and MIG selectively in vitro.48 

Finally, the genes encoding SDF-1α and SDF-1β were cloned from mouse bone marrow (BM) stromal cells (hence their designations as “stromal-derived factors”) using a signal sequence trap technique.49 Human SDF-1α is a low-potency, high-efficacy chemoattractant for T lymphocytes in vitro.50 However, perhaps relevant to its provenance, SDF-1 is also a potent chemoattractant for CD34+ hematopoietic progenitors.51 In vivo, targeted gene disruption of murine SDF-1α indicates that it is required for normal B-lymphocyte development and, surprisingly, for normal cardiac organogenesis because SDF-1α–deficient mice have nonfatal ventricular septal defects.52 Although it is a CXC chemokine, SDF-1's gene is exceptional in its location on human chromosome 10 rather than 4.53 Notably, a receptor for SDF-1α is CXCR4, and SDF-1α prevents infection of CD4/CXCR4-expressing cells by T-lymphocyte–tropic, syncytium-inducing strains of HIV-1.50,54 

CC chemokines are also functionally diverse. Like the CXC chemokines, their names more often reflect historical accidents of their cloning or isolation than their functions (Table 2).

Table 2.

CC Chemokines

NameTarget Cell
MCP-1 Monocytes, memory T lymphocytes, basophils, NK cells, hematopoietic progenitors, ?dendritic cells 
MCP-2 Monocytes, memory and naive T lymphocytes, eosinophils, basophils, NK cells 
MCP-3 Monocytes, Memory T lymphocytes, eosinophils, basophils, NK cells, dendritic cells 
MCP-4 Monocytes, T lymphocytes, eosinophils 
MCP-5 (mouse only) Monocytes, T lymphocytes, eosinophils 
MIP-1α Monocytes, T lymphocytes, NK cells, basophils, eosino-phils, dendritic cells, hematopoietic progenitors 
MIP-1β Monocytes, T lymphocytes, dendritic cells, NK cells, hematopoietic progenitors 
MIP-1γ (mouse only) Resting and activated T lymphocytes 
RANTES Memory T lymphocytes, eosinophils, basophils, NK cells, dendritic cells 
Eotaxin Eosinophils 
I309 Monocytes 
HCC-1 Monocytes, hematopoietic progenitors 
TARC T lymphocytes 
C10 (mouse only) 
CCF18 (mouse only) T lymphocytes, hematopoietic progenitors 
MIP-3α/LARC 
MIP-3β 
NameTarget Cell
MCP-1 Monocytes, memory T lymphocytes, basophils, NK cells, hematopoietic progenitors, ?dendritic cells 
MCP-2 Monocytes, memory and naive T lymphocytes, eosinophils, basophils, NK cells 
MCP-3 Monocytes, Memory T lymphocytes, eosinophils, basophils, NK cells, dendritic cells 
MCP-4 Monocytes, T lymphocytes, eosinophils 
MCP-5 (mouse only) Monocytes, T lymphocytes, eosinophils 
MIP-1α Monocytes, T lymphocytes, NK cells, basophils, eosino-phils, dendritic cells, hematopoietic progenitors 
MIP-1β Monocytes, T lymphocytes, dendritic cells, NK cells, hematopoietic progenitors 
MIP-1γ (mouse only) Resting and activated T lymphocytes 
RANTES Memory T lymphocytes, eosinophils, basophils, NK cells, dendritic cells 
Eotaxin Eosinophils 
I309 Monocytes 
HCC-1 Monocytes, hematopoietic progenitors 
TARC T lymphocytes 
C10 (mouse only) 
CCF18 (mouse only) T lymphocytes, hematopoietic progenitors 
MIP-3α/LARC 
MIP-3β 

See text for references.

MCP-1 (monocyte chemoattractant protein-1) was first purified from conditioned medium of baboon aortic smooth muscle cells in culture on the basis of its ability to attract monocytes, but not neutrophils, in vitro.55 The human version of MCP-1 (also called MCAF) was isolated soon thereafter from tumor cell lines.56,57 Its amino acid sequence turned out to be identical to the predicted amino acid sequence of the product of the human JE gene, a homolog of a PDGF-inducible gene cloned from murine 3T3 cells along with KC several years earlier.23,58,59 

MCP-1 is a potent chemoattractant for monocytes in vitro, with an ED50 similar to IL-8's for neutrophils (≈500 pmol/L). As part of its migration program in monocytes, MCP-1 induces the expression of integrins required for chemotaxis.60,61 In transendothelial migration assays in vitro, MCP-1 is a similarly potent attractant for activated CD4 and CD8 memory T lymphocytes (CD45RA/CD45RO+/CD29+/L-selectin and CD26+).62 In similar assays, MCP-1 attracts neither B lymphocytes nor natural killer (NK) cells.62 In contrast, in assays that do not involve endothelial cells, MCP-1 has been reported to attract NK cells63,64 as well as T lymphocytes.65 MCP-1 also induces granule release from NK cells and CD8+ T cells,63 and activates NK function in CD56+ cells.66 Finally, MCP-1 is a potent histamine-releasing factor for basophils,67-69 but does not attract or activate eosinophils.

MCP-2 and MCP-3 were isolated as novel monocyte chemoattractants from the conditioned medium of MG-63 osteosarcoma cells70 (although the cDNA encoding MCP-2 had been cloned earlier in a differential screen of IFN-γ–treated monocytes71 and murine MCP-3 had been cloned in another differential screen as FIC72). MCP-1 has generally been observed to be a slightly more potent and efficacious monocyte chemoattractant than MCP-2 or MCP-3.5,70,73,74 Like MCP-1, these chemokines attract CD4+ and CD8+ cells that are CD45RO+ in transendothelial assays, although MCP-2 appears to be unique in its ability to attract significant numbers of naive CD45RA+ cells.17 Again, there is some debate about MCP-2 and MCP-3 attracting NK cells with negative results in transendothelial assays,17 and positive results with acellular membrane assays.63,64,74 Like MCP-1, these chemokines activate basophils,75,76 but in contrast to MCP-1, MCP-2 and MCP-3 attract and activate eosinophils.75,77,78 MCP-3 also attracts dendritic cells.79 

The newest members of this family are MCP-480 and MCP-5.81 MCP-4 shares closest amino acid similarity with MCP-3 and eotaxin, and like those proteins it attracts monocytes, T lymphocytes, and eosinophils. On the latter cell type, it completely cross-desensitizes with eotaxin, suggesting that they share the same receptor (CCR3). In terms of relative chemoattractant potency, MCP-4 is equipotent with eotaxin in attracting eosinophils, but less potent than MCP-1 in attracting monocytes or T cells.80 (However, it should be noted that MCP-4's ED50 for all three cell types is the same, ≈10-fold higher than MCP-1's ED50 for monocytes or T cells). MCP-5 has so far only been identified in the mouse, but it attracts monocytes, eosinophils, and T lymphocytes, and activates human and murine CCR2.81,82 It is expressed abundantly in the lung during allergic inflammation and neutralizing antibodies reduce the number of infiltrating eosinophils in these models.82 

MIP-1α and MIP-1β were purified from lipopolysaccharide (LPS)-treated monocytic cell lines,83,84 hence their designations as “macrophage inflammatory proteins.” Some confusion has surrounded the activities of these proteins. Initial descriptions of neutrophil chemoattraction were based on using supraphysiological doses of chemokine.83 The current consensus is that MIP-1α attracts and activates monocytes more efficiently than MIP-1β, but less so than MCP-1.5,17,85 The MIP-1 proteins are also much less efficient than MCP-1 in promoting exocytosis by monocytes.5 MIP-1α may be a more important neutrophil attractant in mice.

In assays that do not involve endothelial cells, MIP-1α attracts predominantly CD8+ T cells while MIP-1β attracts CD4+ cells, although there is some overlap between subsets in response to both chemokines.86,87 In transendothelial assays, both MIP-1s attract CD4+ cells better than CD8+ cells, and MIP-1α is more potent and effective than MIP-1β for both subsets.17 MIP-1α also has the interesting property of being an inhibitor of hematopoietic stem cell proliferation (see below).88 

Other cellular targets for MIP-1α include dendritic cells,79 NK cells,63,89 basophils (weakly),75,90,91 and eosinophils.75,92 MIP-1β has no activity on dendritic cells,79 and minimal activity on NK cells.89 Both MIP-1α and MIP-1β activate NK function in CD56+ cells.66 

The cDNA encoding RANTES (regulated upon activation, normal Texpressed and secreted) was isolated in a T- versus B-lymphocyte differential screen, and found to be inducible by mitogens or antigen in a variety of T-cell lines and circulating lymphocytes.93 In vitro, RANTES is nearly as potent a chemoattractant for monocytes as MCP-1 but is much less effective in stimulating exocytosis.5,94 In endothelial cell–free assays, RANTES attracts CD4+, CD45R0+ T lymphocytes,94 but in transendothelial systems it attracts CD8+ cells as well as CD4+, and is the most potent CC chemokine for CD8+ chemoattraction.17 It also attracts and activates NK cells.66,89 RANTES is an important chemoattractant for eosinophils, which can also secrete it.92,95 Basophils are a RANTES target and, like MCP-1, RANTES induces histamine release.96 

Eotaxin was first isolated from bronchoalveolar lavage fluid in guinea pigs induced to respond to aerosol antigen challenge with an eosinophil-rich pulmonary infiltrate.97 The purified protein was shown to be unique among guinea pig chemokines in its ability to attract eosinophils specifically when injected intradermally. It has been suggested that eotaxin works in concert with IL-5 to elicit eosinophil infiltration: IL-5 stimulates eosinophil release from the BM, and eotaxin directs intravascular eosinophils to their local destination.98 Interestingly, in guinea pigs99 and mice,100,101 eotaxin is expressed at significant constitutive levels in several organs (including lung and heart) and its expression can be further induced upon antigen challenge. Human eotaxin attracts eosinophils when injected in primate skin and appears to use some of the same receptors on eosinophils as MCP-3 and RANTES.102,103 

A cDNA encoding I309 was isolated from a γδ T-cell line using subtractive hybridization from a B-lymphoblastoid line.104 Although not expressed in freshly isolated peripheral blood mononuclear cells (PBMCs), its expression is induced in αβ T-lymphocyte lines by PHA or anti-CD3. It is the likely human homolog of TCA3, the cDNA of which was identified by subtractive hybridization of murine activated T-cell clones versus B cells.105 Both chemokines are potent monocyte chemoattractants,106,107 although TCA3 has also been reported to attract neutrophils.107 Unlike many of the other CC chemokines that attract monocytes and T cells, I309 has no effects on T or NK cells63,106 and is not a histamine-releasing factor.108 

HCC-1 was isolated from the plasma of patients with chronic renal failure.109 Like SDF-1α, HCC-1 is expressed constitutively in a wide variety of tissues and in normal plasma. It does not attract any leukocyte subtype, but it does induce calcium flux and enzyme release specifically from monocytes. Its structure is distantly related to that of MIP-1α and it stimulates the proliferation of CD34+ cells, but less potently than MIP-1α (see below). (This chemokine is not to be confused with an antibody called HCC-1, which identifies a subset of CD34+ cells.110 )

Like SDF-1, TARC was isolated from PHA-stimulated PBMCs by a signal sequence trap strategy.111 The acronym stands for “thymus and activation-regulated chemokine. It is constitutively expressed in thymus and in vitro it binds and attracts T lymphocytes exclusively.

Several CC chemokines in addition to MCP-5 have so far been described only in the mouse. MIP-1γ was isolated from a murine macrophage cell line and found to have structural similarities to C10 (another CC chemokine of unknown function112 ), MIP-1α, and MIP-1β.113 It has been shown to be expressed by Langerhans cells and, perhaps relevantly, it attracts both resting and activated CD4 and CD8 cells in vitro, suggesting a possible role in immune cell recruitment.114 CCF18 is another murine CC chemokine closely related to C10 which is able to attract CD4 T-cell clones, and is expressed constitutively in several myeloid cell lines (especially P388D1 and 32D).115 It has been independently isolated as MIP-Related Protein-2 and shown to have myeloid stem cell inhibitory activity similar to MIP-1α.116 

Although some of these proteins have only been defined in the mouse, the question of whether their human homologs exist and have already been defined is not straightforward. This is a relevant concern because of the use of gene targeting in the mouse to infer chemokine function. Because chemokines are so closely related, simple assessments of homology by proportion of shared sequence may not be reliable. Similarly, because chemokine genes cluster, relatedness cannot be inferred by mapping. Instead, it may turn out that the issue of precise gene-to-gene interspecies homology may not be relevant. Rather, functional homology may be the major criterion for assigning a murine chemokine a role in physiology analogous to a human chemokine.

Because chemokines are so closely related in primary structure, there is great potential for using bioinformatics to isolate new chemokines. This approach, which has been pioneered in the chemokine field by the group at DNAX (Palo Alto, CA), uses computer-based homology searches to identify expressed sequence tags (ESTs) having significant homology to known chemokines. Novel ESTs are then used to search for other ESTs derived from the same cDNA, and when the process is complete, novel chemokine cDNA's may be identified. The CC chemokines MIP-3α and MIP-3β were recently identified this way.117 (MIP-3α was also cloned as LARC.117a ) Although their functions have not yet been described, MIP-3α is expressed in thymus, appendix, PBLs, and fetal liver, while MIP-3β is expressed only in thymus, lymph node, and appendix. Interestingly, the expression of both chemokines is suppressed by IL-10. The gene encoding MIP-3β maps to chromosome 9 and the gene encoding MIP-3α/LARC maps to chromosome 2, making them the first CC chemokine genes identified outside of the gene cluster at 17q11.2-12.

Genomic informatics was also used to identify the CX3C chemokine mentioned above.3,3a This is a membrane-bound protein with multiple structural domains: a chemokine portion at the N-terminus is followed by a long stretch of serine- and threonine-rich repeats that are heavily substituted with mucinlike polysaccharides; a transmembrane region then appears followed by a short cytoplasmic domain (see Fig 1). It appears that the protein may be cleaved physiologically near its membrane insertion, and that the free molecule attracts T lymphocytes and monocytes. Furthermore, the cell surface form of the CX3C chemokine enhances T-lymphocyte and monocyte adhesion to an expressing cell. The function of the mucin stalk is unclear but its structure is highly reminiscent of the C-terminal extension of murine MCP-1 which is also heavily substituted with sialylated carbohydrate.2 

Three-dimensional structure.Because chemokine amino acid sequences are so similar, it is not surprising that their secondary and tertiary structures are also similar (see Fig 1). Several chemokine structures have been solved and all share the same basic features (Fig 2A).118-121 A relatively disordered N-terminus is anchored to the rest of the molecule by disulfide bonds involving the two N-terminal domain cysteines. This is followed by an extended loop that leads into three antiparallel β-pleated sheets (a so-called Greek key) which provide a flat base over which the C-terminal α helix extends.

Fig. 2.

Chemokine structure. (A) Three-dimensional structures of IL-8 and MCP-1 monomers. The three β-pleated sheets of IL-8 are easier to appreciate than those of MCP-1 and are indicated. (B) Three-dimensional structures of IL-8 and MCP-1 dimers. The profound difference in dimer structure is indicated. For MCP-1 and other CC chemokines, dimerization near the N-terminus involves a short, so-called β0 sheet which is indicated.119,120 Reprinted with permission from Rollins BJ: Monocyte chemoattractant protein 1: A potential regulator of monocyte recruitment in inflammatory disease. Molecular Medicine Today, vol 2, p 198, 1996.266 

Fig. 2.

Chemokine structure. (A) Three-dimensional structures of IL-8 and MCP-1 monomers. The three β-pleated sheets of IL-8 are easier to appreciate than those of MCP-1 and are indicated. (B) Three-dimensional structures of IL-8 and MCP-1 dimers. The profound difference in dimer structure is indicated. For MCP-1 and other CC chemokines, dimerization near the N-terminus involves a short, so-called β0 sheet which is indicated.119,120 Reprinted with permission from Rollins BJ: Monocyte chemoattractant protein 1: A potential regulator of monocyte recruitment in inflammatory disease. Molecular Medicine Today, vol 2, p 198, 1996.266 

Close modal

So far, all structural analyses show that chemokines are multimers under conditions required for crystallization or NMR study. All are dimers except for PF4, which is a tetramer.122 However, the structures of the dimers differ profoundly depending on whether they are comprised of CXC or CC chemokines (Fig 2B). In the case of CXC chemokines, such as IL-8, the dimer interface occurs by solvent exclusion along the edge of the first β sheet.118 This creates an extended plane over which the two α helices from each subunit are arrayed in parallel, creating a structure reminiscent of the peptide binding groove of major histocompatibility complex (MHC) molecules. In the case of CC chemokines, such as MCP-1, the dimer interface forms primarily by interactions between short β sheets near the N-termini of the monomers.119-121 This creates an extended molecule that is more cylindrical than the compact CXC dimers.

What is the significance of the different shapes of chemokine dimers? To approach that question, one must first ask whether or not chemokine dimerization itself has any functional significance. Several observations militate against the importance of dimerization. First, physical measurements of monomer affinities by some investigators suggest that the dissociation constant (Kd) for dimer dissociation is several logs higher than the Kd for receptor binding or the ED50 for biological activity.123,124 This would indicate that chemokines are monomers at physiologically relevant concentrations. Furthermore, IL-8 variants that are incapable of forming dimers nonetheless have full biological activity on neutrophils.125 (Although this does not exclude the possibility that two nondimerizing monomers might bind to the same receptor to activate it.)

In contrast, others have used coprecipitation and crosslinking techniques to derive Kds for dimer dissociation that are in line with ED50's.126 In addition, N-terminal deletion variants of chemokines are potent inhibitors of the parent molecules from which they are derived.126-128 In some cases (eg, MCP-1) it has been shown that these variants do not bind efficiently to MCP-1 receptors, suggesting that they are not acting as competitive inhibitors. Instead, they form heterodimers with wild-type MCP-1, raising the possibility that they act as dominant negative inhibitors.126 Dimer-mediated activation of chemokine receptors would be consistent with this model, but not a necessary prerequisite.

Because most chemokines bind to cell-surface or connective-tissue components such as glycosaminoglycans, it may be that dimerization is favored when chemokines associate with these molecules.129 These interactions may make chemokine dimerization a critical process in vivo.

The role of dimerization remains controversial, but it is an important question. If it could be shown that dimer formation were an essential component of chemokine receptor activation, then the dimer interface would be a superb target for antichemokine drugs. This is an attractive notion because the dimerization surface may present a smaller target than the chemokine:receptor interface.

An interesting argument in favor of dimerization is the recent solution of the structure of macrophage migration inhibitory factor (MIF).130 This T-lymphocyte product was first isolated on the basis of its ability to inhibit macrophage migration, making it a cytokine with properties analogous, albeit inverse, to those of chemokines.131,132 The MIF monomer consists of four antiparallel β sheets over which two α helices are arrayed in parallel, creating a structure that looks very much like a CXC chemokine dimer. Because MIF has little amino acid sequence similarity with chemokines, this raises the intriguing possibility that cytokines related to cell migration have undergone convergent evolution toward a three-dimensional structure optimized for this type of interaction. This would support the notion that chemokines form dimers to recapitulate this structure.

Structure/activity relationships.Several lines of evidence point to the importance of the N-terminal domain in the function of CXC and CC chemokines. As mentioned earlier, the presence of the tripeptide ELR sequence near the N-terminus is essential for the activity of CXC chemokines that activate the receptors, CXCR1 or CXCR2.9,10 A further demonstration of the importance of this motif is that addition of ELR to the N-terminal domain of PF4 transforms it into a neutrophil chemoattractant that binds to IL-8 receptors.133 However, the context in which ELR is placed must be important because ELR-modified IP-10 remains inactive toward neutrophils. This suggests that other regions of CXC chemokines are also important for receptor activation. In fact, it has recently been shown that there is a hydrophobic pocket in IL-8 involving the loop beyond the N-terminal cysteines that is essential for binding to one of the IL-8 receptors (CXCR1).134-136 

Among CC chemokines, there is no amino acid sequence motif analogous to ELR that is associated with biologic effects on monocytes or T cells. Nonetheless, N-terminal regions of CC chemokines are clearly important for their activity as demonstrated by deletion analysis and single amino acid substitutions.127,137 Similar to CXC chemokines, there are also data implicating the loop structure between the N-terminal cysteines and the first β sheet.137 

The C-terminal α helices have also been shown to be important for imparting maximal biologic potency to chemokines. Part of this effect may be due to interactions between the α-helices and glycosaminoglycans. For example, IL-8's potency and efficacy are enhanced in the presence of heparan sulfate.138 

One of the confounding problems in understanding chemokine physiology is the fact that chemokine receptors bind several different chemokines, and chemokines bind several different receptors. This so-called promiscuity had been apparent through cross-desensitization experiments even before chemokine receptors were cloned. For example, treating a neutrophil with IL-8 produces a calcium flux and prevents a subsequent calcium flux in response to GRO-α, whereas treatment with GRO-α first does not prevent an IL-8–induced flux. This suggested the presence of two IL-8 receptors, one of which recognizes GRO-α as well as IL-8.27,139 The literature is replete with similar experiments involving a variety of chemokines and cell types. Although the information is critical for new receptor discovery, it can be confusing. The following discussion is therefore limited to data on interactions between chemokines and cloned receptors, with the understanding that there are more receptors waiting to be cloned.

So far, all chemokine receptors are members of the 7-transmembrane spanning (7-TMS), G-protein–coupled receptor family. Although similar to many other 7-TMS receptors, chemokine receptors have some unique structural signatures such as the amino acid sequence DRYLAIV in the second intracellular loop domain.140,141 These receptors are, for the most part, coupled to Gαi proteins, making cellular responses to chemokines inhibitable by pertussis toxin. As expected, receptor activation inhibits cyclic adenosine monophosphate (cAMP) production, but other signal transduction pathways are clearly involved as well. For example, chemoattractant responses to RANTES and MCP-1 can be inhibited by wortmannin, implicating PI3 kinase activation, and by inhibitors of MAPK activation.142-144 The precise mechanisms of coupling receptor activation to complex physiological responses such as chemotaxis are still being investigated.

Receptor-ligand specificities are summarized in Table 3. Four CXC chemokine receptors have been cloned and, as implied earlier, there are two receptors that bind IL-8 with high affinities (Kd < 5 nmol/L).145,146 All of the ELR-containing CXC chemokines bind to one receptor or the other. CXCR1 (IL-8RA) binds only IL-8, whereas CXCR2 (IL-8RB) binds IL-8, GRO-α, GRO-β, GRO-γ, NAP-2, and ENA-78.139,147-151 Much of the ligand specificity resides in the N-terminal domain of these two receptors.149,152 

Table 3.

Chemokine Receptors

ReceptorLigands
CXC Receptors 
CXCR1 IL-8 
CXCR2 IL-8, GRO-α, GRO-β, GRO-γ, NAP-2, ENA-78 
CXCR3 IP-10, MIG 
CXCR4 SDF-1α 
CC Receptors 
CCR1 MIP-1α, RANTES, MCP-3 
CCR2 MCP-1, MCP-3, MCP-5 
CCR3 Eoxtaxin, RANTES, MCP-2, MCP-3, ?MCP-4 
CCR4 MIP-1α, RANTES, MCP-1, TARC 
CCR5 MIP-1α, MIP-1β, RANTES 
CCR6 MIP-3α/LARC 
CCR7 MIP-3β/ELC 
ReceptorLigands
CXC Receptors 
CXCR1 IL-8 
CXCR2 IL-8, GRO-α, GRO-β, GRO-γ, NAP-2, ENA-78 
CXCR3 IP-10, MIG 
CXCR4 SDF-1α 
CC Receptors 
CCR1 MIP-1α, RANTES, MCP-3 
CCR2 MCP-1, MCP-3, MCP-5 
CCR3 Eoxtaxin, RANTES, MCP-2, MCP-3, ?MCP-4 
CCR4 MIP-1α, RANTES, MCP-1, TARC 
CCR5 MIP-1α, MIP-1β, RANTES 
CCR6 MIP-3α/LARC 
CCR7 MIP-3β/ELC 

Ligand assignments are based on in vitro binding assays showing Kd < 10 nmol/L. See text and “Note Added in Proof” for references.

So far, only two receptors have been cloned that bind non–ELR-containing CXC chemokines. IP-10's receptor had originally been identified as a heparan sulfate–containing proteoglycan,153 but it is now clear that it also has a specific heptahelical G-protein–coupled receptor called CXCR3 which also binds MIG.48 A receptor for SDF-1α was first cloned as LESTR, a so-called orphan chemokine receptor isolated by homology cloning based on motifs conserved in this receptor family.154 LESTR was independently cloned as fusin, the coreceptor with CD4 for laboratory-adapted, syncytium-inducing strains of HIV-1.155 Subsequent work identified SDF-1α as a ligand for LESTR/fusin, and the receptor was formally named CXCR4.50,54 

The CC chemokine receptors are much more confusing because they all display overlapping specificities. Most ligands also have overlapping specificities except (so far) for eotaxin, which binds only to CCR3, and MIP-1β, which binds only to CCR5. CCR1 binds MIP-1α, RANTES, and MCP-3 with high affinities.156-158 CCR2 occurs in two forms that are the results of alternative splicing.159 CCR2A and CCR2B differ only in their C-terminal intracellular tails, hence their ligand binding specificities defined to date (MCP-1, MCP-3, and MCP-5) are identical.81,158,160 Although the mRNAs for both receptors are expressed at nearly equivalent levels in monocytic cells, the C-terminal tail of CCR2B is rich in serines and threonines (expected phosphorylation sites in G-protein–coupled receptors) and is homologous to the C-terminal tail of CCR1.159 

CCR3 was cloned as a highly expressed receptor from eosinophils.102,103,161,162 Gratifyingly, this receptor binds the eosinophil chemoattractants, eotaxin, RANTES, and MCP-3, as well as MCP-2. (Eosinophils also express CCR1 and this may account for whatever effects MIP-1α has on eosinophil trafficking.162 ) Because MCP-4 cross-desensitizes eotaxin on eosinophils, CCR3 may also bind MCP-4.80 In an analogous manner, CCR4 was cloned from basophils, and its ligands include the major histamine-releasing CC chemokines, MIP-1α, RANTES, and MCP-1163 as well as TARC.163a Homology cloning resulted in the isolation of a cDNA encoding CCR5.164-166 Although it shares significant primary sequence similarity with CCR2, CCR5 binds MIP-1α, MIP-1β, and RANTES with high affinity.

In the discussion so far, it is apparent that chemokine/ligand promiscuity does not cross CC versus CXC boundaries. However, there is an exception to this rule in the Duffy RBC antigen, which is a heptahelical membrane protein that binds chemokines (as well as Plasmodium vivax).167,168 Because of this property, it is also known as DARC, the Duffy antigen receptor for chemokines. DARC binds several ELR-containing CXC and CC chemokines with high affinity.169 DARC is expressed on postcapillary endothelial cells even in Duffy-negative individuals whose RBCs selectively downregulate expression.170,171 Because DARC has not yet been shown to signal upon binding its ligands, it has been suggested that DARC may act as a chemokine sink, or as a way to present chemokines to circulating or diapedesing leukocytes. The significance of this mechanism is obscure, however, given the existence of an otherwise healthy individual carrying a truncated version of DARC.172 (Duffy is also a receptor for P vivax, and malarial invasion can be inhibited by some chemokine variants.173 )

There are several intriguing examples of DNA viruses that encode chemokine receptorlike molecules. Herpesvirus saimiri's ECRF3 predicts a heptahelical membrane protein with sequence similarity to CXCR1 and CXCR2. Expression of the recombinant protein in frog oocytes shows that it can bind and signal in response to IL-8, GRO-α, and NAP-2, making it a CXCR2-like molecule.174 Human cytomegalovirus has three open reading frames that predict heptahelical receptors. In particular, US28 is highly similar to CCR1, and when expressed in mammalian cells can bind a wide range of CC chemokines, but not CXC chemokines.156,175 These observations raise the interesting possibility that some viruses have selectively mimicked chemokine-related elements of host defense to inactivate them or to subvert them to virulent ends. Finally, the H saimiri–related virus isolated from Kaposi's sarcoma lesions also encodes a chemokine receptorlike molecule.176 Although it binds IL-8 with relatively high affinity, it is constitutively active when expressed in COS cells and enhances proliferation of transfected NRK cells. Thus, this viral product may be involved in transforming the cell of origin in Kaposi's sarcoma.

A recent insight into chemokine physiology comes from the demonstration that chemokine receptor expression can be regulated. For example, IL-2 strongly upregulates expression of CCR1 and CCR2 in circulating T cells.177 This provides another level of control over leukocyte migration because resting T lymphocytes may not be competent to migrate even in the presence of high concentrations of chemokines.

Because there are so many examples of a single chemokine receptor binding several chemokines, it might reasonably be asked whether any individual chemokine could possibly play an essential role in an inflammatory response in vivo. If the answer is no, then a therapy that targets a single chemokine would be doomed to failure because other chemokines with related activities could easily compensate for its inactivation.

In fact, there are two arguments for the essential importance of individual chemokines. The first is theoretical. By examining Table 3, it is apparent that each CC chemokine binds to a unique subset of receptors. If every combination of activated receptors produces a distinct biological response in vivo, then the seven cloned CC receptors could provide 127 different combinations of receptor activation and 127 potentially different physiological actions. (This could be viewed as a single cell responding differently depending on which of its receptors were activated, or as leukocyte subpopulations being defined by their patterns of chemokine receptor expression.) Considering the known CC chemokines, and probably more at 17q11.1-12 or elsewhere, these numbers are quite reasonable.

The second argument for the specificity of individual chemokines has to do with expression patterns. It may be that many chemokines elicit similar responses, but only one is expressed in vivo in a particular setting. For example, several CC chemokines attract monocytes in vitro, but perhaps only MCP-1 is expressed by arterial smooth muscle cells during atherogenesis. This possibility is currently being tested directly in animal models using passive immunization or gene targeting.

However, the first question that must be asked is whether the in vitro properties of chemokines accurately predict their in vivo properties. For the ELR-containing CXC chemokines, the data are relatively straightforward. For example, intradermal injection of IL-8 in rabbits produces neutrophil accumulation.16,178,179 For CC chemokines, injection results are more controversial. Some investigators see monocytic infiltrates after intradermal injection of MCP-1, -2, or -3 in rat or mouse skin, while others do not.70,180,181 

Regardless of the results, injection experiments are fraught with difficulty because of the possibilities of coinjecting contaminants and inducing unrelated tissue damage that could elicit leukocyte infiltration. Several groups have circumvented these problems by constructing transgenic mice that overexpress chemokines. Although these models have produced disparate results, in the aggregate they provide insight into how chemokines work in vivo.

Mice expressing IL-8 under the control of liver-specific promoter/enhancers did not develop neutrophil infiltrates in their livers.182 Instead, they had high serum levels of IL-8 that were associated with L-selectin shedding from circulating neutrophils and an inability to induce neutrophil extravasation in response to local stimuli. This is similar to the observation that intravenous administration of IL-8 in rabbits prevents local neutrophil accumulation (although L-selectin shedding was not observed).183,184 Along the same lines, mice overexpressing MCP-1 under the control of the MMTV-LTR had high serum levels of MCP-1 and no monocytic infiltrates in expressing organs.181 These mice were more susceptible than wild-type mice to intracellular pathogens such as Listeria monocytogenes and Mycobacterium tuberculosis, implicating MCP-1 in the host response to these organisms. This suggests that like IL-8 overexpression, MCP-1 overexpression rendered circulating monocytes incapable of responding to local physiological levels of MCP-1.

In contrast to these models, expression of chemokines at low levels in anatomically restricted areas can produce leukocytic infiltration consistent with their in vitro properties. For example, mice expressing murine GRO-α (KC) in the thymus or brain had neutrophil-rich infiltrates in these organs.185 There was no tissue damage associated with the acute infiltrate, although mice developed severe neurological disease long after GRO-α expression had declined and most of the neutrophils had departed.186 When murine MCP-1 was substituted for GRO-α, the infiltrates were monocytic, and, although quite mild, could be enhanced by systemic LPS treatment.187 In another model, mice expressing human MCP-1 under the control of a surfactant promoter had increased numbers of monocytes and lymphocytes in their bronchial lavage fluid.188 Similarly, mice expressing murine MCP-1 under the control of an insulin promoter had a substantial monocytic insulitis but no tissue damage or diabetes.188a This local effect of the insulin promoter–driven MCP-1 could be abrogated by mating these mice to the MMTV-LTR transgenic mice. Finally, MCP-1 expression controlled by a keratin promoter produced mice with increased numbers of Langerhans-like cells in their skin, and they responded to contact hypersensitivity challenges with exaggerated numbers of monocytes and T lymphocytes.189 

These results provide two insights into how chemokines work in vivo. First, chemokines exert their attractant effects only when they are expressed locally at low levels; systemically administered chemokines actually antagonize the local effects. Second, chemokines appear to attract leukocytes without necessarily activating them. This suggests that chemokine function in leukocyte trafficking is restricted to attraction, and that other signals are necessary for activation.

Lymphocyte trafficking.A critical component of systemic immunity is lymphocyte trafficking.190 This includes macroenvironmental movement, eg, migration of T lymphocytes from the BM and thymus to the spleen and lymph nodes, and microenvironmental movement, eg, migration of B cells from mantle zones to germinal centers. Because of the effects of chemokines on the migration of specific leukocyte subsets, it has been suggested that they are involved in controlling this migration.191 In one model, leukocytes engage in selectin-mediated rolling along the vascular endothelium until they encounter a high local concentration of chemokines (or other attractants), presumably presented in the context of cell-surface macromolecules192,193 or bound to the membrane like the CX3C chemokine.3,3a The chemokines are thought to upregulate integrin affinities, leading to strong adhesive interactions and diapedesis. Extravasated leukocytes then follow a gradient of chemokine concentration and come to rest at a site of high concentration. Leukocyte infiltration in inflammatory sites would then be a special case of this general model of leukocyte movement.

The specificity of chemokine targets makes this an attractive model. For example, RANTES and MCP-1 target only CD45RO+ T cells, suggesting a role in trafficking memory cells.17 However, as appealing as this mechanism may be, there has been only one direct demonstration of its existence in vivo, and ironically, the chemokine involved has yet to be identified. This example involves BRL1, a 7-TMS protein expressed on mature B and a subpopulation of T cells.194 BLR1 has the DRYLAIV sequence in its second intracellular domain, strongly suggesting that it may be a chemokine receptor (see above), but its ligand has not been identified. In BLR1−/− mice created by targeted gene disruption, B lymphocytes are found in T-cell–rich zones of primary splenic follicles rather than in the usual B-cell regions, and no germinal centers develop in secondary follicles.195 Furthermore, the animals have few or no Peyer's patches. Thus, the ligand for BLR1, which is probably a chemokine, may be responsible for attracting B cells that normally populate these areas. These results suggest that chemokines may be involved in basal leukocyte trafficking, and are consistent with the observation that pertussis toxin prevents the normal migration of B and T lymphocytes to their proper microenvironmental locations in the spleen.196 Precise roles for specific chemokines and their receptors in this process will become clearer as more mice are engineered with genetic alterations in this system.

Inflammatory diseases.It is a small step to extend the model of normal leukocyte trafficking to leukocyte infiltration in inflammatory disease. In this case, elaboration of chemokines by sentinel cells at an inflammatory focus may be responsible for inducing strong adhesive interactions between rolling leukocytes and the endothelium. Diapedesing cells are then attracted to the inflammatory site by the chemokine concentration gradient.

Most of the evidence for this model is indirect, coming from experiments that merely document chemokine overexpression at inflammatory foci. For example, MCP-1 expression can be detected in human atheromatous plaques and in the aortic walls of primates fed high-cholesterol diets, consistent with a model of atherogenesis in which MCP-1 in the vessel wall attracts monocytes that eventually become foam cells.197,198 Similarly, the presence of inflammatory cells in the joints of patients with rheumatoid arthritis has been explained by IL-8 and MCP-1 in synovial fluid.199,200 Chemokine expression can be documented in glomerulonephritis,201 asthma,202,203 inflammatory bowel disease,204 and allogeneic transplant rejection.205-207 

The critical question, of course, is whether chemokine expression is pathogenetically responsible for any of the manifestations of these diseases. This has been very difficult to address in humans because of the apparent absence of genetic abnormalities involving chemokines and their receptors (for an exception, see the section on HIV). In animals, chemokines can be shown to play important roles in inflammatory models that may or may not be relevant to disease. For example, passive immunization with anti–IL-8 or anti–MCP-1 antibodies can reduce the edema associated with delayed-type hypersensitivity reactions or the granulomas induced by Schistosoma mansoni eggs in sensitized mice.208,209 Anti–IL-8 can also significantly ameliorate tissue damage associated with reperfusion injury and chemical pneumonitis.210,211 These are striking results because the profoundly destructive inflammatory infiltrate associated with these models can be reversed by neutralizing IL-8 alone, without having to address any of numerous other proinflammatory byproducts of tissue damage.

Even if the singular role of IL-8 in these models is proved genetically, the relevance to human disease remains tenuous. In contrast, some animal models speak more directly to human illness. For example, experimental autoimmune encephalomyelitis closely mimics many of the manifestations of multiple sclerosis. Expression of MCP-1, IP-10, RANTES, MIP-1α, MIP-1β, MCP-3, and GRO-α occurs immediately before the appearance of infiltrating cells in the central nervous system (CNS).212-214 Interestingly, astrocytes appear to be the major source of MCP-1 and IP-10. A pathogenetic role for MIP-1α has been suggested by the observation that passive immunization with anti–MIP-1α antibodies ameliorates disease manifestations.215 

Another model that closely mimics human disease is pulmonary inflammation induced by aerosol challenge in ovalbumin sensitized mice. This produces an eosinophilic infiltrate and bronchospasm that, although imperfect, resembles human asthma. Soon after challenge, monocytes and macrophages appear in the lung in association with MCP-1 expression.216 This is followed by T-lymphocyte and eosinophil accumulation that are associated with RANTES and eotaxin expression.99,101,216 Passive immunization with anti-eotaxin antibodies reduces eosinophil accumulation by 50%, pointing to a role for this chemokine in asthma-like inflammation.216 

Finally, a genetic demonstration for the role of chemokines in viral disease has been shown. Mice engineered to be deficient in MIP-1α by targeted gene disruption show almost none of the presumably autoimmune myocarditis associated with Coxsackie virus injection.217 Furthermore, their pulmonary inflammatory responses to influenza virus are attenuated and clearance of virus is delayed. These results have two important implications. First, they show that MIP-1α plays a critical role in the inflammatory response to viruses. Second, they suggest that other chemokines with similar properties in vitro cannot substitute for MIP-1α in vivo.

HIV.Although it has been known for years that CD4 is an obligate receptor for HIV-1, expression of human CD4 in rodent cells is not sufficient to make them permissive for infection.218 This observation pointed to the existence of a coreceptor. Feng et al155 used an expression cloning strategy to identify the coreceptor for T-cell–tropic, syncytium-inducing HIV-1 strains, which they called “fusin.” We now know that fusin is CXCR4, and that one of its ligands, SDF-1, inhibits infection by these HIV-1 strains in cell expressing CD4 and CXCR4.50,54 

The first hint about a connection between chemokines and HIV-1 came from Cocchi et al,219 who showed that MIP-1α, MIP-1β, or RANTES could prevent infection by macrophage-tropic, nonsyncytium-inducing strains of HIV-1. This was reported soon after the first descriptions of the ligand specificity of CCR5, and led several groups to begin testing HIV-1's interactions with CCR5. It is now clear that CCR5 is the major coreceptor for these HIV-1 strains, but CCR3 and, to a lesser extent CCR2B, are also active.220-223 CCR5 coreceptor function has been proven genetically by the discovery that a significant proportion of individuals who have been exposed to HIV-1 over extended periods of time, but are not infected, are homozygous for an inactive variant of CCR5.224-226 

Evidence to date indicates that viral gp120 interacts directly with CD4 and with the chemokine receptor.227-229 Analyses of chimeric and mutated chemokine receptors point to the N-terminus and first extracellular loop as being critical specificity determinants for HIV-1 infection.230,231 Because some of these variants no longer signal in response to their cognate ligands, it has been suggested that receptor activation does not play a role in HIV-1 entry, but this remains to be proven rigorously.

These results identify chemokine receptors as new targets for anti-HIV therapy. Because individuals with inactive CCR5 seem to be otherwise well, it appears that full blockade of CCR5 is not detrimental. (However, because other CC chemokine receptors also act as coreceptors for some strains, the blockade might have to be more substantial and less specific.) These results also raise a series of interesting questions. First, why has the virus exploited chemokine receptors rather than other ubiquitous heptahelical receptors? If signaling is not part of HIV entry, this would suggest that the virus selected these receptors for structural reasons rather than for reasons of chemokine-like molecular mimicry. Second, the discovery of CCR5-negative individuals begs the question of what this receptor does. There are no obvious abnormalities in these individuals, and the inactive allele appears in certain populations with a frequency approaching 10% to 20%. The answers to these questions will provide as much insight into the physiology of chemokines as the pathophysiology of HIV.

Angiogenesis.CXC chemokines have been extensively investigated for their effects on angiogenesis in vivo and on endothelial cells in vitro. A variety of ELR-containing chemokines has been reported to be chemotactic for endothelial cells, including IL-8, ENA-78, GCP2, GRO-α, GRO-β, GRO-γ, and the processed products of PBP.232-234 All of these chemokines are also angiogenic in the rat cornea neovascularization assay.232,233 The non-ELR CXC chemokines, including PF4, IP-10, and MIG are not only inactive chemoattractants and nonangiogenic themselves, but they inhibit the angiogenic effects of ELR chemokines and basic FGF.40,232,235 Although these data appear to assign angiogenic properties to ELR chemokines and angiostatic properties to non-ELR CXC chemokines, there is a report that GRO-α and GRO-β inhibit endothelial cell proliferation in vitro and are potently anti-angiogenic in vivo.236 These disparate results may be related to differences in experimental systems.

PF4's angiostatic activity depends on the presence of its C-terminal heparin-binding domain and can be reversed by adding heparin. However, a PF4 variant that does not bind heparin (but does maintain the C-terminal α-helical structure) is an even more potent inhibitor.237 Another potent variant has a wild-type α-helix that binds heparin, but has undergone cleavage at Thr-16, suggesting that the determinants for angiostatic activity may be complex.41 

Because the non-ELR CXC chemokines only inhibit chemokine– or growth factor–induced angiogenesis, one model for their angiostatic properties would be interference with activation of growth factor receptors on endothelial cells. In fact, PF4 binds to full-length VEGF (VEGF165 ) and prevents its heparin-dependent interaction with its receptor.238 However, PF4 also inhibits the mitogenic effects of a truncated VEGF (VEGF121 ) whose receptor binding is not heparin-dependent and is not prevented by PF4. This is a strong argument for angiostatic chemokines exerting their effects through their own cognate receptors, and effecting events at a post–growth factor receptor activation stage.

The angiostatic effects of PF4 can be produced in vivo (eg, in corneal neovascularization models) and this has led to testing PF4 for its ability to inhibit tumor angiogenesis. Systemic PF4 administered to mice after inoculation with B16F10 melanoma cells reduces the number and size of lung metastases without preventing their homing to the lung.239,240 The non–heparin-binding variant is also potent in vivo.237 

In contrast, the observation that ELR chemokines are angiogenic suggests that they may play a role in promoting tumor angiogenesis. Several lung carcinomas have marginally increased levels of IL-8, and the angiogenic activity present in extracts from these tumors is almost entirely caused by IL-8.241 This has suggested a model in which the balance of ELR versus non-ELR chemokines produced by a tumor and its associated stroma dictate the degree of angiogenesis, and therefore aggressiveness, displayed by the tumor.233 Some evidence supporting this model comes from the beneficial effects of administering anti–IL-8 to SCID mice bearing human IL-8–expressing lung cancer xenografts.233 

Hematopoiesis.Some chemokines have dual effects on hematopoiesis depending on the maturity of the progenitors being treated. For example, MIP-1α, MIP-1β, GRO-β, and GRO-γ enhance the formation of colony-forming unit granulocyte-macrophage (CFU-GM) from unfractionated BM, but only in the presence of M- or GM-colony-stimulating factor (CSF).242-245 In contrast, several chemokines suppress the proliferation of more immature progenitors, eg, CFU-S, CFU-A (macrophage-rich colonies derived from stem cells), and CFU-GM (colonies that require stem cell factor [SCF] and GM-CSF), as well as CFU-GEMM and burst-forming unit erythroid (BFU-E) (colonies that require SCF and erythropoietin).88,116,243,246-248 This effect is a direct one on the progenitor cells because the suppression is more complete on CD34+-selected cells.244,247 The same effect can be demonstrated by in vivo administration of MIP-1α, and pretreatment of animals with MIP-1α can enhance myeloid recovery after treatment with S-phase–active chemotherapeutic agents.246,249-251 Furthermore, analysis of these in vivo experiments has led to the suggestion that MIP-1α's target is not the earliest long-term reconstituting stem cell, but a somewhat more mature cell246 which may be LinThy1+ in the mouse.242 Although the most direct explanation for these observations is that MIP-1α inhibits proliferation, it has also been suggested that MIP-1α may prevent factor-induced differentiation.248,252 

Broxmeyer's group has systematically studied chemokines in the methylcellulose culture system and they have identified the suppressive chemokines as MIP-1α, GRO-β, PF4, IL-8, MCP-1, IP-10, and CCF18.116,243,247 Mixtures of these chemokines are synergistic in their suppression. Chemokines that have no suppressive effects on hematopoietic progenitors include: MIP-1β, GRO-α, GRO-γ, NAP-2, and RANTES. In fact, pretreatment of progenitors with MIP-1β blocks the suppressive effects of MIP-1α, and pretreatment with GRO-α or GRO-γ blocks the effects of IL-8 and PF4.243 In general, chemokines exert their suppressive effects with an ED50 of ≈5 nmol/L. However, in the case of MIP-1α, which oligomerizes at high concentrations, the active form is the monomer, which is 1,000-fold more potent than the usual mixture of multimers.253 

Inspection of lists of suppressive and nonsuppressive chemokines suggests that this effect depends neither on CC versus CXC structure, nor on the presence or absence of the ELR motif. In fact, structure/activity analyses of IL-8 and PF4 show that neutrophil chemoattraction does not correlate with stem cell suppression.254 Some PF4/IL-8 chimeras were active at 10 pmol/L (similar to the MIP-1α monomers). Interestingly, mutations in IL-8 that involve the dimer interface destroy stem cell inhibitory activity, suggesting that dimerization may be important, although it remains possible that these amino acids are also required for interaction with the stem cell receptor for IL-8.

The fact that PF4 is as potent as IL-8 suggests that all suppressive effects do not occur via CXCR1 or CXCR2. Furthermore, the synergistic effects of some chemokines and the ability of chemokine pretreatment to block the effects of other chemokines implies the existence of multiple relevant receptors. At least one has been identified in mice. The gene encoding an IL-8 receptor homolog which binds murine MIP-2 (homologous to human GRO-β and GRO-γ) with high affinity has been disrupted in mice.255 Among other abnormalities, these mice have an expanded neutrophil compartment, especially when reared in a non–germ-free environment. BM progenitors from these mice are insensitive to the suppressive effects of IL-8 and murine MIP-2, but still sensitive to MIP-1α, PF4, IP-10, and MCP-1, arguing that multiple chemokine receptors are involved in hematopoiesis.256 

These observations are clearly relevant to hematopoiesis in vivo. In the IL-8 receptor homolog-deficient mice reared in a germ-free environment, the number of BM progenitor cells was similar to that in wild-type mice.256 However, although the number of progenitors increased in both mice when reared in specific pathogen-free conditions, the increase in number of progenitors in the receptor-deficient mice was much greater than in the wild type. This is consistent with a model in which endogenous ligands for this receptor are involved in an inhibitory pathway for myeloid progenitors. Interestingly, although MIP-1α was the first stem cell inhibitor to be identified, mice deficient for MIP-1α do not have an expanded progenitor pool.217 Either MIP-1α's effects are not relevant in vivo, or other chemokines can compensate for its absence.

Although the focus of this discussion has been on hematopoietic stem cells, MIP-1α also inhibits the proliferation of keratinocyte stem cells.257 Thus, chemokines may have a more general effect on suppressing stem cell proliferation.

Antitumor effects.Angiostasis is only one means by which chemokines limit tumor growth. For example, because some chemokines inhibit hematopoietic and epithelial stem cell proliferation, it might be imagined that they could have direct inhibitory effects on malignant cells as well. So far, however, there is only a single report in which IL-8 directly inhibited the growth of non–small cell lung cancer cell lines in vitro.258 

A more commonly investigated scenario exploits the leukocyte chemoattractant properties of chemokines to enhance a host antitumor response. Engineered expression of MCP-1, IP-10, TCA3, or lymphotactin in tumor cells has been shown to elicit antitumor effects when cells are injected in vivo.259-263 For example, MCP-1 expression in tumor cells prevents tumor formation after inoculation in athymic mice, suggesting that acute rejection is not T-cell–mediated.259 However, when syngeneic cells are irradiated and used as a tumor vaccine, engineered expression of MCP-1 in these cells prevents subsequent growth of wild-type tumor cells.264 This indicates that MCP-1 can enhance tumor specific immunity, presumably in a T-lymphocyte–dependent manner. Similar results have been reported for TCA3.262 Finally, lymphotactin itself has very little effect on tumor take or established tumors in syngeneic animals. However, when used in combination with IL-2, it has activity as a “therapeutic” tumor vaccine, reducing the size of established tumors.263 While the long-term immunity in immunocompetent animals in these models is T-lymphocyte–dependent, some effects may also reflect activation of NK cells by CC chemokines.66 

IP-10's effects are more complicated. On one hand, engineered IP-10 expression in tumor cells prevents their growth in syngeneic animals, but not in athymic animals, suggesting that this effect is T-cell–dependent.260 However, injection of IP-10 into established Burkitt's tumors in nu/nu mice leads to tumor necrosis in association with microvascular thrombosis.265 This may be related to IP-10's antiangiogenic activity.

It is striking to recall that a family of proteins as large and ubiquitously expressed as the chemokines and their receptors was unknown less than 10 years ago. Their recent history may be one reason why we still understand so little about the true roles played by chemokines in normal and diseased physiology. In the absence of clear-cut underlying principles, a review about chemokines will necessarily seem like a series of unconnected observations.

Nonetheless, the picture that is emerging from genetically modified mouse models is that most of the in vitro properties of chemokines are recapitulated in vivo. This suggests that many of the early hypotheses about the proinflammatory function of chemokines in disease may turn out to be accurate. If so, that will provide an enormous number of new drug targets for diseases ranging from atherosclerosis to acquired immunodeficiency syndrome. Now the focus can start to shift away from proof-of-concept experiments that document chemokine-mediated pathology to mechanistic studies that investigate ways to interrupt chemokine/receptor interactions or chemokine receptor signal transduction.

Two additional CC chemokine receptors have been characterized and their ligands identified. CCR6 is the receptor for MIP-3α/LARC and is expressed by CD4+ and CD8+ T lymphocytes and B lymphocytes, but not by NK cells, monocytes, or neutrophils.267 CCR7, formerly known as the orphan receptor EBI1, is the receptor for MIP-3β, also known as ELC (EBI1 ligand chemokine).268 CCR7 is expressed by activated T and B lymphocytes.

I apologize to any colleague whose work I may have unintentionally overlooked and not cited. I also want to thank Dr Craig Gerard for helpful advice and suggestions, and Laurie Geronimo for administrative assistance.

Supported by a grant from the National Institutes of Health. B.J.R. is a Scholar of The Leukemia Society.

Address reprint requests to Barrett J. Rollins, MD, PhD, Dana-Farber Cancer Institute, 44 Binney St, Boston, MA 02115.

1
Oppenheim
JJ
Zachariae
COC
Mukaida
N
Matsushima
K
Properties of the novel proinflammatory supergene “intercrine” family.
Annu Rev Immunol
9
1991
617
2
Ernst
CA
Zhang
YJ
Hancock
PR
Rutledge
BJ
Corless
CL
Rollins
BJ
Biochemical and biological characterization of murine MCP-1: Identification of two functional domains.
J Immunol
152
1994
3541
3
Bazan
JF
Bacon
KB
Hardiman
G
Wang
W
Soo
K
Rossi
D
Greaves
DR
Zlotnik
A
Schall
TJ
A new class of membrane-bound chemokine with a CX3C motif.
Nature
385
1997
640
3a
Pan Y, Lloyd C, Zhou H, Dolich S, Deeds J, Gonzalo JA, Vath J, Gosselin M, Ma J, Dussault B, Woolf E, Alperin G, Culpepper J, Gutierrez-Ramos JC, Gearing D: Neurotactin, a membrane-anchored chemokine upregulated in brain inflammation. Nature 387:611, 1997
4
Kelner
GS
Kennedy
J
Bacon
KB
Kleyensteuber
S
Largaespada
DA
Jenkins
NA
Copeland
NG
Bazan
JF
Moore
KW
Schall
TJ
Zlotnik
A
Lymphotactin: A cytokine that represents a new class of chemokine.
Science
266
1994
1395
5
Uguccioni
M
D'Apuzzo
M
Loetscher
M
Dewald
B
Baggiolini
M
Actions of the chemotactic cytokines MCP-1, MCP-2, MCP-3, RANTES, MIP-1α and MIP-1β on human monocytes.
Eur J Immunol
25
1995
64
6
Walz
A
Peveri
P
Aschauer
H
Baggiolini
M
Purification and amino acid sequencing of NAF, a novel neutrophil activating factor produced by monocytes.
Biochem Biophys Res Commun
149
1987
755
7
Yoshimura
T
Matsushima
K
Tanaka
S
Robinson
EA
Appella
E
Oppenheim
JJ
Leonard
EJ
Purification of a human monocyte-derived neutrophil chemotactic factor that has peptide sequence similarity to other host defense cytokines.
Proc Natl Acad Sci USA
84
1987
9233
8
Schroder
J-M
Mrowietz
U
Morita
E
Christophers
E
Purification and partial biochemical characterization of a human monocyte-derived, neutrophil-activating peptide that lacks interleukin 1 activity.
J Immunol
139
1987
3474
9
Clark-Lewis
I
Schumacher
S
Baggiolini
M
Moser
B
Structure-activity relationships of interleukin-8 determined using chemically synthesized analogs: Critical role of NH2-terminal residues and evidence for uncoupling of neutrophil chemotaxis, exocytosis, and receptor binding activities.
J Biol Chem
266
1991
23128
10
Hebert
CA
Viytangcol
RV
Baker
JB
Scanning mutagenesis of interleukin-8 identifies a cluster of residues required for receptor binding.
J Biol Chem
266
1991
18989
11
Walz
A
Dewald
B
von Tscharner
V
Baggiolini
M
Effects of the neutrophil-activating peptide NAP-2, platelet basic protein, connective tissue-activating peptide III, and platelet factor 4 on human neutrophils.
J Exp Med
170
1989
1745
12
Van Damme
J
Van Beeumen
J
Conings
R
Decock
B
Billiau
A
Purification of granulocyte chemotactic peptide/interleukin-8 reveals N-terminal sequence heterogeneity similar to that of β-thromboglobulin.
Eur J Biochem
181
1989
337
13
Yoshimura
T
Robinson
EA
Appella
E
Matsushima
K
Showalter
SD
Skeel
A
Leonard
EJ
Three forms of monocyte-derived neutrophil chemotactic factor (MDNCF) distinguished by different lengths of the amino-terminal sequence.
Mol Immunol
26
1989
87
14
Huber
AR
Kunkel
SL
Todd
RF
Weiss
SJ
Regulation of transendothelial neutrophil migration by endogenous interleukin-8.
Science
254
1991
99
15
Gimbrone
MAJ
Obin
MS
Brock
AF
Luis
EA
Hass
PE
Hebert
CA
Yip
YK
Leung
DW
Lowe
DG
Kohr
WJ
Darbonne
WC
Bechtol
KB
Baker
JB
Endothelial interleukin-8: A novel inhibitor of leukocyte-endothelial interactions.
Science
246
1989
1601
16
Larsen
CG
Anderson
AO
Appella
E
Oppenheim
JJ
Matsushima
K
Neutrophil activating protein (NAP-1) is also chemotactic for T lymphocytes.
Science
243
1989
1464
17
Roth
SJ
Carr
MW
Springer
TA
C-C chemokines, but not the C-X-C chemokines interleukin-8 and interferon-γ inducible protein-10, stimulate transendothelial chemotaxis of T lymphocytes.
Eur J Immunol
25
1995
3482
18
Oin
S
LaRosa
G
Campbell
JJ
Smith-Heath
H
Kassam
N
Shi
X
Zeng
L
Buthcher
EC
Mackay
CR
Expression of monocyte chemoattractant protein-1 and interleukin-8 receptors on subsets of T cells: Correlation with transendothelial chemotactic potential.
Eur J Immunol
26
1996
640
19
Petzelbauer
P
Watson
CA
Pfau
SE
Pober
JS
IL-8 and angiogenesis: Evidence that human endothelial cells lack receptors and do not respond to IL-8 in vitro.
Cytokine
7
1995
267
20
Dahinden
CA
Kurimoto
Y
De Weck
AL
Lindsey
I
Dewald
B
Baggiolini
M
The neutrophil activating peptide NAF/NAP-1 induces histamine and leukotriene release by interleukin 3-primed basophils.
J Exp Med
170
1989
1787
21
White
MV
Yoshimura
T
Hook
W
Kaliner
M
Leonard
EJ
Neutrophil attractant/activation protein (NAP-1) causes human basophil histamine release.
Immunol Lett
22
1989
151
22
Anisowicz
A
Bardwell
L
Sager
R
Constitutive overexpression of a growth-regulated gene in transformed chinese hamster and human cells.
Proc Natl Acad Sci USA
84
1987
7188
23
Cochran
BH
Reffel
AC
Stiles
CD
Molecular cloning of gene sequences regulated by platelet-derived growth factor.
Cell
33
1983
939
24
Oquendo
P
Alberta
J
Wen
D
Graycar
JL
Derynck
R
Stiles
CD
The platelet-derived growth factor-inducible KC gene encodes a secretory protein related to platelet α-granule proteins.
J Biol Chem
264
1989
4133
25
Richmond
A
Belentien
E
Thomas
HG
Flaggs
G
Barton
DE
Spiess
J
Bordoni
R
Francke
U
Derynck
R
Molecular characterization and chromosomal mapping of melanoma growth stimulatory activity, a growth factor structurally related to β-thromboglobulin.
EMBO J
7
1988
2025
26
Schroder
JM
Persoon
NLM
Christophers
E
Lipopolysaccharide-stimulated human monocytes secrete, apart from neutrophil-activating peptide/interleukin 8, a second neutrophil-activating protein: NH2-terminal amino acid sequence identity with melanoma growth stimulatory activity.
J Exp Med
171
1990
1091
27
Moser
B
Clark-Lewis
I
Zwahlen
R
Baggiolini
M
Neutrophil-activating properties of the melanoma growth-stimulating activity, MGSA.
J Exp Med
171
1990
1797
28
Haskill
S
Peace
A
Morris
J
Sporn
SA
Anisowicz
A
Lee
SW
Smith
T
Martin
G
Ralph
P
Sager
R
Identification of three related human GRO genes encoding cytokine functions.
Proc Natl Acad Sci USA
87
1990
7732
29
Wolpe
SD
Sherry
B
Juers
D
Davatelis
G
Yurt
RW
Cerami
A
Identification and characterization of macrophage inflammatory protein 2.
Proc Natl Acad Sci USA
86
1989
612
30
Tekamp-Olson
P
Gallegos
C
Bauer
D
McClain
J
Sherry
B
Fabre
M
van Deventer
DS
Cerami
A
Cloning and characterization of cDNAs for murine macrophage inflammatory protein 2 and its human homologues.
J Exp Med
172
1990
911
31
Walz
A
Burgener
R
Car
B
Baggiolini
M
Kunkel
SL
Strieter
RM
Structure and neutrophil-activating properties of a novel inflammatory peptide (ENA-78) with homology to interleukin 8.
J Exp Med
174
1991
1355
32
Proost
P
De Wolf-Peeters
C
Conings
R
Opdenakker
G
Billiau
A
Van Damme
J
Identification of a novel granulocyte chemotactic protein (GCP-2) from human tumor cells. In vitro and in vivo comparison with natural forms of GRO, IP-10, and IL-8.
J Immunol
150
1993
1000
33
Holt
JC
Harris
ME
Holt
AM
Lange
E
Henschen
A
Niewiarowski
S
Characterization of human platelet basic protein, a precursor form of low-affinity platelet factor 4 and β-thromboglobulin.
Biochemistry
25
1986
1988
34
Castor
CW
Miller
JW
Walz
DA
Structural and biological characteristics of connective tissue activating peptide (CTAP-III), a major human platelet-derived growth factor.
Proc Natl Acad Sci USA
80
1983
765
35
Mullenbach
GT
Tabrizi
A
Blacher
RW
Steimer
KS
Chemical synthesis and expression in yeast of a gene encoding connective tissue activating peptide-III.
J Biol Chem
261
1986
719
36
Senior
RM
Griffin
GL
Huang
JS
Walz
DA
Deuel
TF
Chemotactic activity of platelet alpha granule proteins for fibroblasts.
J Cell Biol
96
1983
382
37
Walz
A
Baggiolini
M
A novel cleavage product of β-thromboglobulin formed in cultures of stimulated mononuclear cells activates human neutrophils.
Biochem Biophys Res Commun
159
1989
969
38
Leonard
EJ
Yoshimura
T
Rot
A
Noer
K
Walz
A
Baggiolini
M
Walz
DA
Goetzl
EJ
Castor
CW
Chemotactic activity and receptor binding of neutrophil attractant/activation protein-1 (NAP-1) and structurally related host defense cytokines: Interaction of NAP-2 with the NAP-1 receptor.
J Leukoc Biol
49
1991
258
39
Deuel
TF
Senior
RM
Chang
D
Griffin
GL
Henrikson
RL
Kaiser
ET
Platelet factor 4 is chemotactic for neutrophils and monocytes.
Proc Natl Acad Sci USA
78
1981
4584
40
Maione
TE
Gray
GS
Petro
J
Hunt
AJ
Donner
AL
Bauer
SI
Carson
HF
Sharpe
RJ
Inhibition of angiogenesis by recombinant human platelet factor-4 and related peptides.
Science
247
1990
77
41
Gupta
SK
Hassel
T
Singh
JP
A potent inhibitor of endothelial cell proliferation is generated by proteolytic cleavage of the chemokine platelet factor 4.
Proc Natl Acad Sci USA
92
1995
7799
42
Luster
AD
Unkeless
JC
Ravetch
JV
γ-Interferon transcriptionally regulates an early-response gene containing homology to platetelt proteins.
Nature
315
1985
672
43
Narumi
S
Wyner
L
Stoler
MH
Tannenbaum
CS
Hamilton
TA
Tissue-specific expression of murine IP-10 mRNA following systemic treatment with interferon γ.
J Leukoc Biol
52
1992
27
44
Dewald
B
Moser
B
Barella
L
Schumacher
C
Baggiolini
M
Clark-Lewis
I
IP-10, a γ-interferon-inducible protein related to interleukin-8 lacks neutrophil actovating properties.
Immunol Lett
32
1992
81
45
Murphy
WJ
Tian
ZG
Asai
O
Funakoshi
S
Rotter
P
Henry
M
Strieter
RM
Kunkel
SL
Longo
DL
Taub
DD
Chemokines and T lymphocyte activation: II. Facilitation of human T cell trafficking in severe combined immunodeficiency mice.
J Immunol
156
1996
2104
46
Taub
DD
Longo
DL
Murphy
WJ
Human interferon-inducible protein-10 induces mononuclear cell infiltration in mice and promotes the migration of human T lymphocytes into the peripheral tissues and human peripheral blood lymphocytes-SCID mice.
Blood
87
1996
1423
47
Liao
F
Rabin
RL
Yannelli
JR
Koniaris
LG
Vanguri
P
Farber
JM
Human Mig chemokine: Biochemical and functional characterization.
J Exp Med
182
1995
1301
48
Loetscher
M
Gerber
B
Loetscher
P
Jones
SA
Piali
L
Clark-Lewis
I
Baggiolini
M
Moser
B
Chemokine receptor specific for IP10 and Mig: Structure, function, and expression in activated T-lymphocytes.
J Exp Med
184
1996
963
49
Tashiro
K
Tada
H
Heilker
R
Shirozu
M
Nakano
T
Honjo
T
Signal sequence trap: A cloning strategy for secreted proteins and type I membrane proteins.
Science
261
1993
600
50
Bleul
CC
Farzan
M
Choe
H
Parolin
C
Clark-Lewis
I
Sodroski
J
Springer
TA
The lymphocyte chemoattractant SDF-1 is a ligand for LESTR/fusin and blocks HIV-1 entry.
Nature
382
1996
829
51
Aiuti
A
Webb
IJ
Bleul
C
Springer
T
Guttierez-Ramos
JC
The chemokine SDF-1 is a chemoattractant for human CD34+ hematopoietic progenitor cells and provides a new mechanism to explain the mobilization of CD34+ progenitors to peripheral blood.
J Exp Med
185
1997
111
52
Nagasawa
T
Hirota
S
Tachibana
K
Takakura
N
Nishikawa
S
Kitamura
Y
Yoshida
N
Kikutani
H
Kishimoto
T
Defects of B-cell lymphopoiesis and bone-marrow myelopoiesis in mice lacking the CXC chemokine PBSF/SDF-1.
Nature
382
1996
635
53
Shirozu
M
Nakano
T
Inazawa
J
Tashiro
K
Tada
H
Shinohara
T
Honjo
T
Structure and chromosomal localization of the human stromal cell-derived factor 1 (SDF1) gene.
Genomics
28
1995
495
54
Oberlin
E
Amara
A
Bachelerie
F
Bessia
C
Virelizier
JL
Arenzana-Seisdedos
F
Schwartz
O
Heard
JM
Clark-Lewis
I
Legler
DF
Loetscher
M
Baggiolini
M
Moser
B
The CXC chemokine SDF-1 is the ligand for LESTR/fusin and prevents infection by T-cell-line-adapted HIV-1.
Nature
382
1996
833
55
Valente
AJ
Graves
DT
Vialle-Valentin
CE
Delgado
R
Schwartz
CJ
Purification of a monocyte chemotactic factor secreted by nonhuman primate vascular cells in culture.
Biochemistry
27
1988
4162
56
Matsushima
K
Larsen
CG
DuBois
GC
Oppenheim
JJ
Purification and characterization of a novel monocyte chemotactic and activating factor produced by a human myelomonocytic cell line.
J Exp Med
169
1989
1485
57
Yoshimura
T
Robinson
EA
Tanaka
S
Appella
E
Leonard
EJ
Purification and amino acid analysis of two human monocyte chemoattractants produced by phytohemagglutinin-stimulated human blood mononuclear leukocytes.
J Immunol
142
1989
1956
58
Rollins
BJ
Morrison
ED
Stiles
CD
Cloning and expression of JE, a gene inducible by platelet-derived growth factor and whose product has cytokine-like properties.
Proc Natl Acad Sci USA
85
1988
3738
59
Rollins
BJ
Stier
P
Ernst
TE
Wong
GG
The human homologue of the JE gene encodes a monocyte secretory protein.
Mol Cell Biol
9
1989
4687
60
Jiang
Y
Beller
DI
Frendl
G
Graves
DT
Monocyte chemoattractant protein-1 regulates adhesion molecule expression and cytokine production in human monocytes.
J Immunol
148
1992
2423
61
Vaddi
K
Newton
RC
Regulation of monocyte integrin expression by β-family chemokines.
J Immunol
153
1994
4721
62
Carr
MW
Roth
SJ
Luther
E
Rose
SS
Springer
TA
Monocyte chemoattractant protein 1 acts as a T-lymphocyte chemoattractant.
Proc Natl Acad Sci USA
91
1994
3652
63
Loetscher
P
Seitz
M
Clark-Lewis
I
Baggiolini
M
Moser
B
Activation of NK cells by CC chemokines. Chemotaxis, Ca2+ mobilization, and enzyme release.
J Immunol
156
1996
322
64
Allavena
P
Bianchi
G
Zhou
D
van Damme
J
Jilek
P
Sozzani
S
Mantovani
A
Induction of natural killer cell migration by monocyte chemotactic protein-1, -2 and -3.
Eur J Immunol
24
1994
3233
65
Loetscher
P
Seitz
M
Clark-Lewis
I
Baggiolini
M
Moser
B
Monocyte chemotactic proteins MCP-1, MCP-2, and MCP-3 are major attractants for human CD4+ and CD8+ T lymphocytes.
FASEB J
8
1994
1055
66
Maghazachi
AA
Al-Aoukaty
A
Schall
TJ
CC chemokines induce the generation of killer cells from CD56+ cells.
Eur J Immunol
26
1996
315
67
Bischoff
SC
Krieger
M
Brunner
T
Dahinden
CA
Monocyte chemotactic protein 1 is a potent activator of human basophils.
J Exp Med
125
1992
1271
68
Kuna
P
Reddigari
SR
Rucinski
D
Oppenheim
JJ
Kaplan
AP
Monocyte chemotactic and activating factor is a potent histamine-releasing factor for human basophils.
J Exp Med
175
1992
489
69
Alam
R
Lett-Brown
MA
Forsythe
PA
Anderson-Walters
DJ
Kenamore
C
Kormos
C
Grant
JA
Monocyte chemotactic and activating factor is a potent histamine-releasing factor for basophils.
J Clin Invest
89
1992
723
70
Van Damme
J
Proost
P
Lenaerts
JP
Opdenakker
G
Structural and functional identification of two human, tumor-derived monocyte chemotactic proteins (MCP-2 and MCP-3) belonging to the chemokine family.
J Exp Med
176
1992
59
71
Chang
HC
Hsu
F
Freeman
GJ
Griffin
JD
Reinherz
EL
Cloning and expression of a γ-interferon-inducible gene in monocytes: A new member of a cytokine gene family.
Int Immunol
1
1989
388
72
Heinrich
JN
Ryseck
RP
Macdonald-Bravo
H
Bravo
R
The product of a novel growth factor-activated gene, fic, is a biologically active “C-C”-type cytokine.
Mol Cell Biol
13
1993
2020
73
Sozzani
S
Locati
M
Zhou
D
Rieppi
M
Luini
W
Lamorte
G
Bianchi
G
Polentarutti
N
Allavena
P
Mantovani
A
Receptors, signal transduction, and spectrum of action of monocyte chemotactic protein-1 and related chemokines.
J Leukoc Biol
57
1995
788
74
Proost
P
Wuyts
A
Van Damme
J
Human monocyte chemotactic proteins-2 and -3: Structural and functional comparison with MCP-1.
J Leukoc Biol
59
1996
67
75
Dahinden
CA
Geiser
T
Brunner
T
von Tscharner
V
Caput
D
Ferrara
P
Minty
A
Baggiolini
M
Monocyte chemotactic protein 3 is a most effective basophil- and eosinophil-activating chemokine.
J Exp Med
179
1994
751
76
Alam
R
Forsythe
P
Stafford
S
Heinrich
J
Bravo
R
Proost
P
Van Damme
J
Monocyte chemotactic protein-2, monocyte chemotactic protein-3, and fibroblast-induced cytokine. Three new chemokines induce chemotaxis and activation of basophils.
J Immunol
153
1994
3155
77
Weber
M
Uguccioni
M
Ochensberger
B
Baggiolini
M
Clark-Lewis
I
Dahinden
CA
Monocyte chemotactic protein MCP-2 activates human basophil and eosinophil leukocytes similar to MCP-3.
J Immunol
154
1995
4166
78
Noso
N
Proost
P
Van Damme
J
Schroder
JM
Human monocyte chemotactic proteins-2 and 3 (MCP-2 and MCP-3) attract human eosinophils and desensitize the chemotactic responses towards RANTES.
Biochem Biophys Res Commun
200
1994
1470
79
Sozzani
S
Sallusto
F
Luini
W
Zhou
D
Piemonti
L
Allavena
P
Van Damme
J
Valitutti
S
Lanzavecchia
A
Mantovani
A
Migration of dendritic cells in reponse to formyl peptides, C5a, and a distinct set of chemokines.
J Immunol
155
1995
3292
80
Uguccioni
M
Loetscher
P
Forssmann
U
Dewald
B
Li
H
Lima
SH
Li
Y
Kreider
B
Garotta
G
Thelen
M
Baggiolini
M
Monocyte chemotactic protein 4 (MCP-4), a novel structural and functional analogue of MCP-3 and eotaxin.
J Exp Med
183
1996
2379
81
Sarafi
MN
Garcia-Zepeda
EA
MacLean
JA
Charo
IF
Luster
AD
Murine monocyte chemoattractant protein (MCP)-5: A novel CC chemokine that is a structural and functional homologue of human MCP-1.
J Exp Med
185
1997
99
82
Jia
GQ
Gonzalo
JA
Lloyd
C
Kremer
L
Lu
L
MartinezAC
Wershil BK
Guttierez-Ramos
JC
Distinct expression and function of the novel mouse chemokine monocyte chemotactic protein-5 in lung allergic inflammation.
J Exp Med
184
1996
1939
83
Wolpe
SD
Davatelis
G
Sherry
B
Beutler
B
Hesse
DG
Nguyen
HT
Moldawer
LL
Nathan
CF
Lowry
SF
Cerami
A
Macrophages secrete a novel heparin-binding protein with inflammatory and neutrophil chemotactic properties.
J Exp Med
167
1988
570
84
Sherry
B
Tekamp-Olson
P
Gallegos
C
Bauer
D
Davatelis
G
Wolpe
SD
Masiarz
F
Coit
D
Cerami
A
Resolution of the two components of macrophage inflammatory protein 1, and cloning and characterization of one of those components, macrophage inflammatory protein 1β.
J Exp Med
168
1988
2251
85
Fahey
TJI
Tracey
KJ
Tekamp-Olson
P
Cousens
LS
Jones
WG
Shires
GT
Cerami
A
Sherry
B
Macrophage inflammatory protein 1 modulates macrophage function.
J Immunol
148
1992
2764
86
Taub
DD
Conlon
K
Lloyd
AR
Oppenheim
JJ
Kelvin
DJ
Preferential migration of activated CD4+ and CD8+ T cells in response to MIP-1α and MIP-1β.
Science
260
1993
355
87
Schall
TJ
Bacon
K
Camp
RDR
Kaspari
JW
Goeddel
DV
Human macrophage inflammatory protein α (MIP-1α) and MIP-1β chemokines attract distinct populations of lymphocytes.
J Exp Med
177
1993
1821
88
Graham
GJ
Wright
EG
Hewick
R
Wolpe
SD
Wilkie
NM
Donaldson
D
Lorimore
S
Pragnell
IB
Identification and characterization of an inhibitor of haematopoietic stem cell proliferation.
Nature
344
1990
442
89
Taub
DD
Sayers
TJ
Carter
CR
Ortaldo
JR
α and β chemokines induce NK cell migration and enhance NK-mediated cytolysis.
J Immunol
155
1995
3877
90
Alam
R
Kumar
D
Anderson-Walters
D
Forsythe
PA
Macrophage inflammatory protein-1α and monocyte chemoattractant peptide-1 elicit immediate and late cutaneous reactions and activate murine mast cells in vivo.
J Immunol
152
1994
1298
91
Alam
R
Forsythe
PA
Stafford
S
Lett-Brown
MA
Grant
JA
Macrophage inflammatory protein-1α activates basophils and mast cells.
J Exp Med
176
1992
781
92
Rot
A
Krieger
M
Brunner
T
Bischoff
SC
Schall
TJ
Dahinden
CA
RANTES and macrophage inflammatory protein 1α induce the migration and activation of normal human eosinophil granulocytes.
J Exp Med
176
1992
1489
93
Schall
TJ
Jongstra
J
Dyer
BJ
Jorgensen
J
Clayberger
C
Davis
MM
Krensky
AM
A human T cell-specific molecule is a member of a new gene family.
J Immunol
141
1988
1018
94
Schall
TJ
Bacon
K
Toy
KJ
Goeddel
DV
Selective attraction of monocytes and T lymphocytes of the memory phenotype by cytokine RANTES.
Nature
347
1990
669
95
Lim
KG
Wan
HC
Bozza
PT
Resnick
MB
Wong
DT
Cruikshank
WW
Kornfeld
H
Center
DM
Weller
PF
Human eosinophils elaborate the lymphocyte chemoattractants. IL-16 (lymphocyte chemoattractant factor) and RANTES.
J Immunol
156
1996
2566
96
Kuna
P
Reddigari
SR
Schall
TJ
Rucinski
D
Viksman
MY
Kaplan
AP
RANTES, a monocyte and T lymphocyte chemotactic cytokine releases histamine from human basophils.
J Immunol
149
1992
636
97
Jose
PJ
Griffiths-Johnson
DS
Collins
PD
Walsh
DT
Moqbel
R
Totty
NF
Truong
O
Hsuan
JJ
Williams
TJ
Eotaxin: A potent eosinophil chemoattractant cytokine detected in a guinea pig model of allergic airways inflammation.
J Exp Med
179
1994
881
98
Collins
PD
Marleau
S
Griffiths-Johnson
DA
Jose
PJ
Williams
TJ
Cooperation between interleukin-5 and the chemokine eotaxin to induce eosinophil accumulation in vivo.
J Exp Med
182
1995
1169
99
Rothenberg
ME
Luster
AD
Lilly
CM
Drazen
JM
Leder
P
Constitutive and allergen-induced expression of eotaxin mRNA in the guinea pig lung.
J Exp Med
181
1995
1211
100
Rothenberg
ME
Luster
AD
Leder
P
Murine eotaxin: An eosinophil chemoattractant inducible in endothelial cells and in interleukin 4-induced tumor suppression.
Proc Natl Acad Sci USA
92
1995
8960
101
Gonzalo
JA
Jia
GQ
Aguirre
V
Friend
D
Coyle
AJ
Jenkins
NA
Lin
GS
Katz
H
Lichtman
A
Copeland
N
Kopf
M
Gutierrez-Ramos
JC
Mouse eotaxin expression parallels eosinophil accumulation during lung allergic inflammation but it is not restricted to a Th2-type response.
Immunity
4
1996
1
102
Kitaura
M
Nakajima
T
Imai
T
Harada
S
Combadiere
C
Tiffany
HL
Murphy
PM
Yoshie
O
Molecular cloning of human eotaxin, an eosinophil-selective CC chemokine, and identification of a specific eosinophil eotaxin receptor, CC chemokine receptor 3.
J Biol Chem
271
1996
7725
103
Ponath
PD
Qin
S
Ringler
DJ
Clark-Lewis
I
Wang
J
Kassam
N
Smith
H
Shi
X
Gonzalo
JA
Newman
W
Gutierrez-Ramos
JC
Mackay
CR
Cloning of the human eosinophil chemoattractant, eotaxin. Expression, receptor binding, and functional properties suggest a mechanism for the selective recruitment of eosinophils.
J Clin Invest
97
1996
604
104
Miller
MD
Hata
S
De WMR
Krangel MS
A novel polypeptide secreted by activated human T lymphocytes.
J Immunol
143
1989
2907
105
Burd
PR
Freeman
GJ
Wilson
SD
Berman
M
DeKruyff
R
Billings
PR
Dorf
ME
Cloning and characterization of a novel T cell activation gene.
J Immunol
139
1987
3126
106
Miller
MD
Krangel
MS
The human cytokine I-309 is a monocyte chemoattractant.
Proc Natl Acad Sci USA
89
1992
2950
107
Luo
Y
Laning
J
Devi
S
Mak
J
Schall
TJ
Dorf
ME
Biologic activities of the murine β-chemokine TCA3.
J Immunol
153
1994
4616
108
Kuna
P
Reddigari
SR
Schall
TJ
Rucinski
D
Sadick
M
Kaplan
AP
Characterization of the human basophil response to cytokines, growth factors, and histamine releasing factors of the intercrine/chemokine family.
J Immunol
150
1993
1932
109
Schulz-Knappe
P
Magert
HJ
Dewald
B
Meyer
M
Cetin
Y
Kubbies
M
Tomeczkowski
J
Kirchhoff
K
Raida
M
Adermann
K
Kist
A
Reinecke
M
Sillard
R
Pardigol
A
Uguccioni
M
Baggiolini
M
Forssmann
WG
HCC-1, a novel chemokine from human plasma.
J Exp Med
183
1996
295
110
Hill
B
Rozler
E
Travis
M
Chen
S
Zannetino
A
Simmons
P
Galy
A
Chen
B
Hoffman
R
High-level expression of a novel epitope of CD59 identifies a subset of CD34+ bone marrow cells highly enriched for pluripotent stem cells.
Exp Hematol
24
1996
936
111
Imai
T
Yoshida
T
Baba
M
Nishimura
M
Kakizaki
M
Yoshie
O
Molecular cloning of a novel T cell-directed CC chemokine expressed in thymus by signal sequence trap using Epstein-Barr virus vector.
J Biol Chem
271
1996
21514
112
Orlofsky
A
Berger
MS
Prystowsky
MB
Novel expression pattern of a new member of the MIP-1 family of cytokine-like genes.
Cell Regul
2
1991
403
113
Poltorak
AN
Bazzoni
F
Smirnova
II
Alejos
E
Thompson
P
Luheshi
G
Rothwell
N
Beutler
B
MIP-1γ: Molecular cloning, expression, and biological activities of a novel CC chemokine that is constitutively secreted in vivo.
J Inflammat
45
1995
207
114
Mohamadzadeh
M
Poltorak
AN
Bergstressor
PR
Beutler
B
Takashima
A
Dendritic cells produce macrophage inflammatory protein-1γ, a new member of the CC chemokine family.
J Immunol
156
1996
3102
115
Hara
T
Bacon
KB
Cho
LC
Yoshimura
A
Morikawa
Y
Copeland
NG
Gilbert
DJ
Jenkins
NA
Schall
TJ
Miyajima
A
Molecular cloning and functional characterization of a novel member of the C-C chemokine family.
J Immunol
155
1995
5352
116
Youn
BS
Jang
IK
Broxmeyer
HE
Cooper
S
Jenkins
NA
Gilbert
DJ
Copeland
NG
Elick
TA
Fraser
MJJ
Kwon
BS
A novel chemokine, macrophage inflammatory protein-related protein-2, inhibits colony formation of bone marrow myeloid progenitors.
J Immunol
155
1995
2661
117
Rossi
DL
Vicari
AP
Franz-Bacon
K
McClanahan
TK
Zlotnik
A
Identification through bioinformatics of two new macrophage proinflammatory human chemokines MIP-3α and MIP-3β.
J Immunol
158
1997
1033
117a
Hieshima K, Imai T, Opdenakker G, Van Damme J, Kusuda J, Tei H, Sakaki Y, Takatsuki K, Miura R, Yoshie O, Nomiyama H: Molecular cloning of a novel human CC chemokine liver and activation-regulated chemokine (LARC) expressed in liver. Chemotactic activity for lymphocytes and gene localization on chromosome 2. J Biol Chem 272:5846, 1997
118
Clore
GM
Appella
E
Yamada
M
Matsushima
K
Gronenborn
AM
Three-dimensional structure of interleukin 8 in solution.
Biochemistry
29
1990
1689
119
Lodi
PJ
Garrett
DS
Kuszewski
J
Tsang
ML
Weatherbee
JA
Leonard
WJ
Gronenborn
AM
Clore
GM
High-resolution solution structure of the β chemokine hMIP-1β by multidimensional NMR.
Science
263
1994
1762
120
Handel
TM
Domaille
PJ
Heteronuclear (1H, 13C, 15N) NMR assignments and solution structure of the monocyte chemoattractant protein-1 (MCP-1) dimer.
Biochemistry
35
1996
6569
121
Chung
CW
Cooke
RM
Proudfoot
AE
Wells
TN
The three-dimensional solution structure of RANTES.
Biochemistry
34
1995
9307
122
St Charles
R
Walz
D
Edwards
BFP
The three-dimensional structure of bovine platelet factor 4 at 3.0-A resolution.
J Biol Chem
264
1989
2092
123
Burrows
SD
Doyle
ML
Murphy
KP
Franklin
SG
White
JR
Brooks
I
McNulty
DE
Scott
MO
Knutson
JR
Porter
D
Young
PR
Hensley
P
Determination of monomer-dimer equilibrium of interleukin-8 reveals it is a monomer at physiological concentrations.
Biochemistry
33
1994
12741
124
Paolini
JF
Willard
D
Consler
T
Luther
M
Krangel
MS
The chemokines IL-8, monocyte chemoattractant protein-1 and I-309 are monomers at physiologically relevant concentrations.
J Immunol
153
1994
2704
125
Rajarathnam
K
Sykes
BD
Kay
CM
Dewald
B
Geiser
T
Baggiolini
M
Clark-Lewis
I
Neutrophil activation by monomeric interleukin-8.
Science
264
1994
90
126
Zhang
Y
Rollins
BJ
A dominant negative inhibitor indicates that monocyte chemoattractant protein 1 functions as a dimer.
Mol Cell Biol
15
1995
4851
127
Gong
JH
Clark-Lewis
I
Antagonists of monocyte chemoattractant protein 1 identified by modification of functionally critical NH2-terminal residues.
J Exp Med
181
1995
631
128
Moser
B
Dewald
B
Barella
L
Schumacher
C
Baggiolini
M
Clark-Lewis
I
Interleukin-8 antagonists generated by N-terminal modification.
J Biol Chem
268
1993
7125
129
Witt
DP
Lander
AD
Differential binding of chemokines to glycosaminoglycan subpopulations.
Current Biol
4
1994
394
130
Sun
HW
Bernhagen
J
Bucala
R
Lolis
E
Crystal structure at 2.6-A resolution of human macrophage migration inhibitory factor.
Proc Natl Acad Sci USA
93
1996
5191
131
David
JR
Delayed hypersensitivity in vitro: Its mediation by cell-free substances formed by lymphoid cell-antigen interaction.
Proc Natl Acad Sci USA
56
1966
72
132
Bloom
BR
Bennett
B
Mechanism of a reaction in vitro associated with delayed-type hypersensitivity.
Science
153
1966
80
133
Clark-Lewis
I
Dewald
B
Geiser
T
Moser
B
Baggiolini
M
Platelet factor 4 binds to interleukin 8 receptors and activates neutrophils when its N terminus is modified with Glu-Leu-Arg.
Proc Natl Acad Sci USA
90
1993
3574
134
Lowman
HB
Slagle
PH
DeForge
LE
Wirth
CM
Gillece-Castro
BL
Bourell
JH
Fairbrother
WJ
Exchanging interleukin-8 and melanoma growth-stimulating activity receptor binding specificities.
J Biol Chem
271
1996
14344
135
Hammond
ME
Shyamala
V
Siani
MA
Gallegos
CA
Feucht
PH
Abbott
J
Lapointe
GR
Moghadam
M
Khoja
H
Zakel
J
Tekamp-Olson
P
Receptor recognition and specificity of interleukin-8 is determined by residues that cluster near a surface-accessible hydrophobic pocket.
J Biol Chem
271
1996
8228
136
Williams
G
Borkakoti
N
Bottomley
GA
Cowan
I
Fallowfield
AG
Jones
PS
Kirtland
SJ
Price
GJ
Price
L
Mutagenesis studies of interleukin-8. Identification of a second epitope involved in receptor binding.
J Biol Chem
271
1996
9579
137
Zhang
YJ
Rutledge
BJ
Rollins
BJ
Structure/activity analysis of human monocyte chemoattractant protein-1 (MCP-1) by mutagenesis: Identification of a mutated protein that inhibits MCP-1-mediated monocyte chemotaxis.
J Biol Chem
269
1994
15918
138
Webb
LMC
Ehrengruber
MU
Clark-Lewis
I
Baggiolini
M
Rot
A
Binding to heparan sulfate or heparin enhances neutrophil responses to interleukin 8.
Proc Natl Acad Sci USA
90
1993
7158
139
Moser
B
Schumacher
C
von Tscharner
V
Clark-Lewis
I
Baggiolini
M
Neutrophil-activating peptide 2 and gro/melanoma growth-stimulatory activity interact with neutrophil-activating peptide 1/interleukin 8 receptors on human neutrophils.
J Biol Chem
266
1991
10666
140
Gerard
C
Gerard
NP
C5A anaphylatoxin and its seven transmembrane-segment receptor.
Annu Rev Immunol
12
1994
775
141
Murphy
PM
The molecular biology of leukocyte chemoattractant receptors.
Annu Rev Immunol
12
1994
593
142
Yen
H
Penfold
S
Zhang
Y
Rollins
BJ
MCP-1-mediated chemotaxis requires activation of non-overlapping signal transduction pathways.
J Leukoc Biol
61
1997
529
143
Dubois
PM
Palmer
D
Webb
ML
Ledbetter
JA
Shapiro
RA
Early signal transduction by the receptor to the chemokine monocyte chemoattractant protein-1 in a murine T cell hybrid.
J Immunol
156
1996
1356
144
Turner
L
Ward
SG
Westwick
J
RANTES-activated human T lymphocytes. A role for phosphoinositide 3-kinase.
J Immunol
155
1995
2437
145
Murphy
PM
Tiffany
HL
Cloning of complementary DNA encoding a functional human interleukin-8 receptor.
Science
253
1991
1280
146
Holmes
WE
Lee
J
Kuang
WJ
Rice
GC
Wood
WI
Structure and functional expression of a human interleukin-8 receptor.
Science
253
1991
1278
147
Geiser
T
Dewald
B
Ehrengruber
MU
Clark-Lewis
I
Baggiolini
M
The interleukin-8-related chemotactic cytokines GRO α, GRO β, and GRO γ activate human neutrophil and basophil leukocytes.
J Biol Chem
268
1993
15419
148
Lee
J
Horuk
R
Rice
GC
Bennett
GL
Camerato
T
Wood
WI
Characterization of two high affinity human interleukin-8 receptors.
J Biol Chem
267
1992
16283
149
LaRosa
GJ
Thomas
KM
Kaufmann
ME
Mark
R
White
M
Taylor
L
Gray
G
Witt
D
Navarro
J
Amino terminus of the interleukin-8 receptor is a major determinant of receptor subtype specificity.
J Biol Chem
267
1992
25402
150
Cerretti
DP
Kozlosky
CJ
Vanden Bos
T
Nelson
N
Gearing
DP
Beckmann
MP
Molecular characterization of receptors for human interleukin-8, GRO/melanoma growth-stimulatory activity and neutrophil activating peptide-2.
Mol Immunol
30
1993
359
151
Ahuja
SK
Murphy
PM
The CXC chemokines growth-regulated oncogene (GRO) alpha, GROβ, GROγ, neutrophil-activating peptide-2, and epithelial cell-derived neutrophil-activating peptide-78 are potent agonists for the type B, but not the type A, human interleukin-8 receptor.
J Biol Chem
271
1996
20545
152
Gayle RB3
Sleath
PR
Srinivason
S
Birks
CW
Weerawarna
KS
Cerretti
DP
Kozlosky
CJ
Nelson
N
Vanden Bos
T
Beckmann
MP
Importance of the amino terminus of the interleukin-8 receptor in ligand interactions.
J Biol Chem
268
1993
7283
153
Luster
AD
Greenberg
SM
Leder
P
The IP-10 chemokine binds to a specific cell surface heparan sulfate site shared with platelet factor 4 and inhibits endothelial cell proliferation.
J Exp Med
182
1995
219
154
Loetscher
M
Geiser
T
O'Reilly
T
Zwahlen
R
Baggiolini
M
Moser
B
Cloning of a human seven-transmembrane domain receptor, LESTR, that is highly expressed in leukocytes.
J Biol Chem
269
1994
232
155
Feng
Y
Broder
CC
Kennedy
PE
Berger
EA
HIV-1 entry cofactor: Functional cDNA cloning of a seven-transmembrane, G protein-coupled receptor.
Science
272
1996
872
156
Neote
K
DiGregorio
D
Mak
JY
Horuk
R
Schall
TJ
Molecular cloning, functional expression, and signaling characteristics of a C-C chemokine receptor.
Cell
72
1993
415
157
Gao
JL
Kuhns
DB
Tiffany
HL
McDermott
D
Li
X
Francke
U
Murphy
PM
Structure and functional expression of the human macrophage inflammatory protein 1 α/RANTES receptor.
J Exp Med
177
1993
1421
158
Combadiere
C
Ahuja
SK
Van Damme
J
Tiffany
HL
Gao
JL
Murphy
PM
Monocyte chemoattractant protein-3 is a functional ligand for CC chemokine receptors 1 and 2B.
J Biol Chem
270
1995
29671
159
Charo
IF
Myers
SJ
Herman
A
Franci
C
Connolly
AJ
Coughlin
SR
Molecular cloning and functional expression of two monocyte chemoattractant protein 1 receptors reveals alternative splicing of the carboxyl-terminal tails.
Proc Natl Acad Sci USA
91
1994
2752
160
Franci
C
Wong
LM
Van DJ
Proost P
Charo
IF
Monocyte chemoattractant protein-3, but not monocyte chemoattractant protein-2, is a functional ligand of the human monocyte chemoattractant protein-1 receptor.
J Immunol
154
1995
6511
161
Post
TW
Bozic
CR
Rothenberg
ME
Luster
AD
Gerard
N
Gerard
C
Molecular characterization of two murine eosinophil β chemokine receptors.
J Immunol
155
1995
5299
162
Daugherty
BL
Siciliano
SJ
DeMartino
JA
Malkowitz
L
Sirotina
A
Springer
MS
Cloning, expression, and characterization of the human eosinophil eotaxin receptor.
J Exp Med
183
1996
2349
163
Power
CA
Meyer
A
Nemeth
K
Bacon
KB
Hoogewerf
AJ
Proudfoot
AEI
Wells
TNC
Molecular cloning and functional expression of a novel CC chemokine receptor cDNA from a human basophilic cell line.
J Biol Chem
270
1995
19495
163a
Imai T, Baba M, Nishimura M, Kakizaki M, Takagi S, Yoshie O: The T cell-directed CC chemokine TARC is a highly specific biological ligand for CC chemokine receptor 4. J Biol Chem 272:15036, 1997
164
Samson
M
Labbe
O
Mollereau
C
Vassart
G
Parmentier
M
Molecular cloning and functional expression of a new human CC-chemokine receptor gene.
Biochemistry
35
1996
3362
165
Raport
CJ
Gosling
J
Schweickart
VL
Gray
PW
Charo
IF
Molecular cloning and functional characterization of a novel human CC chemokine receptor (CCR5) for RANTES, MIP-1β, and MIP-1α.
J Biol Chem
271
1996
17161
166
Combadiere
C
Ahuja
SK
Tiffany
HL
Murphy
PM
Cloning and functional expression of CC CKR5, a human monocyte CC chemokine receptor selective for MIP-1α, MIP-1β, and RANTES.
J Leukoc Biol
60
1996
147
167
Neote
K
Darbonne
W
Ogez
J
Horuk
R
Schall
TJ
Identification of a promiscuous inflammatory peptide receptor on the surface of red blood cells.
J Biol Chem
268
1993
12247
168
Horuk
R
Chitnis
CE
Darbonne
WC
Colby
TJ
Rybicki
A
Hadley
TJ
Miller
LH
A receptor for the malarial parasite Plasmodium vivax: the erythrocyte chemokine receptor.
Science
261
1993
1182
169
Szabo
MC
Soo
KS
Zlotnik
A
Schall
TJ
Chemokine class differences in binding to the Duffy antigen-erythrocyte chemokine receptor.
J Biol Chem
270
1995
25348
170
Hadley
TJ
Lu
ZH
Wasniowska
K
Martin
AW
Peiper
SC
Hesselgesser
J
Horuk
R
Postcapillary venule endothelial cells in kidney express a multispecific chemokine receptor that is structurally and functionally identical to the erythroid isoform, which is the Duffy blood group antigen.
J Clin Invest
94
1994
985
171
Peiper
SC
Wang
ZX
Neote
K
Martin
AW
Showell
HJ
Conklyn
MJ
Ogborne
K
Hadley
TJ
Lu
ZH
Hesselgesser
J
Horuk
R
The Duffy antigen/receptor for chemokines (DARC) is expressed in endothelial cells of Duffy negative individuals who lack the erythrocyte receptor.
J Exp Med
181
1995
1311
172
Mallinson
G
Soo
KS
Schall
TJ
Pisacka
M
Anstee
DJ
Mutations in the erythrocyte chemokine receptor (Duffy) gene: The molecular basis of the Fya/Fyb antigens and identification of a deletion in the Duffy gene of an apparently healthy individual with the Fy(a − b) phenotype.
Br J Haematol
90
1995
823
173
Hesselgesser
J
Chitnis
CE
Miller
LH
Yansura
DG
Simmons
LC
Fairbrother
WJ
Kotts
C
Wirth
C
Gillece-Castro
B
Horuk
R
A mutant of melanoma growth stimulating activity does not activate neutrophils but blocks erythrocyte invasion by malaria.
J Biol Chem
270
1995
11472
174
Ahuja
SK
Murphy
PM
Molecular piracy of mammalian interleukin-8 receptor type B by Herpesvirus saimiri.
J Biol Chem
268
1993
20691
175
Gao
JL
Murphy
PM
Human cytomegalovirus open reading frame US28 encodes a functional beta chemokine receptor.
J Biol Chem
269
1994
28539
176
Arvanitakis
L
Geras-Raaka
E
Varma
A
Gershengorn
MC
Cesarman
E
Human herpesvirus KSHV encodes a constitutively active G-protein-coupled receptor linked to cell proliferation.
Nature
385
1997
347
177
Loetscher
P
Seitz
M
Baggiolini
M
Moser
B
Interleukin-2 regulates CC chemokine receptor expression and chemotactic responsiveness in T lymphocytes.
J Exp Med
184
1996
569
178
Foster
SJ
Aked
DM
Schroder
JM
Christophers
E
Acute inflammatory effects of a monocyte-derived neutrophil-activating peptide in rabbit skin.
Immunology
67
1989
181
179
Colditz
I
Zwahlen
R
Dewald
B
Baggiolini
M
In vivo inflammatory activity of neutrophil-activating factor, a novel chemotactic peptide derived from human monocytes.
Am J Pathol
134
1989
755
180
Zachariae
COC
Anderson
AO
Thompson
HL
Appella
E
Mantovani
A
Oppenheim
JJ
Matsushima
K
Properties of monocyte chemotactic and activating factor (MCAF) purified from a human fibrosarcoma cell line.
J Exp Med
171
1990
2177
181
Rutledge
BJ
Rayburn
H
Rosenberg
R
North
RJ
Gladue
RP
Corless
CL
Rollins
BJ
High level monocyte chemoattractant protein-1 expression in transgenic mice increases their susceptibility to intracellular pathogens.
J Immunol
155
1995
4838
182
Simonet
WS
Hughes
TM
Nguyen
HQ
Trebasky
LD
Danilenko
DM
Medlock
ES
Long-term impairment of neutrophil migration in mice overexpressing human interleukin-8.
J Clin Invest
94
1994
1310
183
Ley
K
Baker
JB
Cybulsky
MI
Gimbrone
MA
Luscinskas
FW
Intravenous interleukin-8 inhibits granulocyte emigration from rabbit mesenteric vessels without altering L-selectin expression or leukocyte rolling.
J Immunol
151
1993
6347
184
Hechtman
DH
Cybulsky
MI
Fuchs
HJ
Baker
JB
Gimbrone
MA
Intravascular IL-8: Inhibitor of polymorphonuclear leukocyte accumulation at sites of acute inflammation.
J Immunol
147
1991
883
185
Lira
SA
Zalamea
P
Heinrich
JN
Funetes
ME
Carrasco
D
Lewin
AC
Barton
DS
Durham
S
Bravo
R
Expression of the chemokine N51/KC in the thymus and epidermis of transgenic mice results in a marked infiltration of a single class of inflammatory cells.
J Exp Med
180
1994
2039
186
Tani
M
Fuentes
ME
Peterson
JW
Trapp
BD
Durham
SK
Loy
JK
Bravo
R
Ransohoff
RM
Lira
SA
Neutrophil infiltration, glial reaction, and neurological disease in transgenic mice expressing the chemokine N51/KC in oligodendrocytes.
J Clin Invest
98
1996
529
187
Fuentes
ME
Durham
SK
Swerdel
MR
Lewin
AC
Barton
DS
Megill
JR
Bravo
R
Lira
SA
Controlled recruitment of monocytes/macrophages to specific organs through transgenic expression of MCP-1.
J Immunol
155
1995
5769
188
Gunn
MD
Nelken
NA
Liao
X
Willimas
LT
Monocyte chemoattractant protein-1 is sufficient for the chemotaxis of monocytes and lymphocytes in transgenic mice but requires an additional stimulus for inflammatory activation.
J Immunol
158
1997
376
188a
Grewal IS, Rutledge BJ, Fiorillo JA, Gu L, Gladue RP, Flavell RA, Rollins BJ: Transgenic monocyte chemoattractant protein-1 (MCP-1) in pancreatic islets produces monocyte-rich insulitis without diabetes: Abrogation by a second transgene expressing systemic MCP-1. J Immunol 159:401, 1997
189
Nakamura
K
Williams
IR
Kupper
TS
Keratinocyte-derived monocyte chemoattractant protein 1 (MCP-1): Analysis in a transgenic model demonstrates MCP-1 can recruit dendritic and Langerhans cells to skin.
J Invest Dermatol
105
1995
635
190
Butcher
EC
Picker
LJ
Lymphocyte homing and homeostasis.
Science
272
1996
60
191
McKay
CR
Chemokine receptors and T cell chemotaxis.
J Exp Med
184
1996
799
192
Butcher
EC
Leukocyte-endothelial cell recognition: Three (or more) steps to specificity and diversity.
Cell
67
1991
1033
193
Springer
T
Traffic signals for lymphocyte recirculation and leukocyte emigration: The multistep paradigm.
Cell
76
1994
301
194
Forster
R
Emrich
E
Kremmer
E
Lipp
M
Expression of the G protein-coupled receptor BLR1 defines mature, recirculating B cells and a subset of T-helper memory cells.
Blood
84
1994
830
195
Forster
R
Mattis
A
Kremmer
E
Wolf
E
Brem
G
Lipp
M
A putative chemokine receptor, BLR1, directs B cell migration to defined lymphoid organs and specific anatomic compartments of the spleen.
Cell
87
1996
1037
196
Cyster
JG
Goodnow
CC
Pertussis toxin inhibits migration of B and T lymphocytes into splenic white pulp cords.
J Exp Med
182
1995
581
197
Nelken
NA
Coughlin
SR
Gordon
D
Wilcox
JN
Monocyte chemoattractant protein-1 in human atheromatous plaques.
J Clin Invest
88
1991
1121
198
Yla-Herttuala
S
Lipton
BA
Rosenfeld
ME
Sarkioja
T
Yoshimura
T
Leonard
EJ
Witztum
JL
Steinberg
D
Expression of monocyte chemoattractant protein 1 in macrophage-rich areas of human and rabbit atherosclerotic lesions.
Proc Natl Acad Sci USA
88
1991
5252
199
Koch
AE
Kunkel
SL
Harlow
LA
Johnson
B
Evanoff
HL
Haines
GK
Burdick
MD
Pope
RM
Strieter
RM
Enhanced production of monocyte chemoattractant protein-1 in rheumatoid arthritis.
J Clin Invest
90
1992
772
200
Peichl
P
Ceska
M
Effenberger
F
Haberhauer
G
Broell
H
Lindley
IJD
Presence of NAP-1/IL-8 in synovial fluids indicates a possible pathogenetic role in rheumatoid arthritis.
Scand J Immunol
34
1991
333
201
Brown
Z
Robson
RL
Westwick
J
Regulation and expression of chemokines: Potential role in glomerulonephritis.
J Leukoc Biol
59
1996
75
202
Sousa
AR
Lane
SJ
Nakhosteen
JA
Yoshimura
T
Lee
TH
Poston
RN
Increased expression of the monocyte chemoattractant protein-1 in bronchial tissue from asthmatic subjects.
Am J Resp Cell Mol Biol
10
1994
142
203
Teran
LM
Noso
N
Carroll
M
Davies
DE
Holgate
S
Schroder
JM
Eosinophil recruitment following allergen challenge is associated with the release of the chemokine RANTES into asthmatic airways.
J Immunol
157
1996
1806
204
Grimm
MC
Elsbury
SK
Pavli
P
Doe
WF
Enhanced expression and production of monocyte chemoattractant protein-1 in inflammatory bowel disease mucosa.
J Leukoc Biol
59
1996
804
205
Pattison
J
Nelson
PJ
Huie
P
von Leuttichau
I
Farshid
G
Sibley
RK
Krensky
AM
RANTES chemokine expression in cell-mediated transplant rejection of the kidney.
Lancet
343
1994
209
206
Kondo
T
Novick
AC
Toma
H
Fairchild
RL
Induction of chemokine gene expression during allogeneic skin graft rejection.
Transplant
61
1996
1750
207
Adams
DH
Russell
ME
Hancock
WW
Sayegh
MH
Wyner
LR
Karnovsky
MJ
Chronic rejection in experimental cardiac transplantation: studies in the Lewis-F344 model.
Immunol Rev
134
1993
5
208
Chensue
SW
Warmington
KS
Lukacs
NW
Lincoln
PM
Burdick
MD
Strieter
RM
Kunkel
SL
Monocyte chemotactic protein expression during schistosome egg granuloma formation. Sequence of production, localization, contribution, and regulation.
Am J Pathol
146
1995
130
209
Larsen
CG
Thomsen
MK
Gesser
B
Thomsen
PD
Deleuran
BW
Nowak
J
Skodt
V
Thomsen
HK
Deleuran
M
Thestrup-Pedersen
K
Harada
A
Matsushima
K
Menné
T
The delayed-type hypersensitivity reaction is dependent on IL-8. Inhibition of a tuberculin skin reaction by an anti-IL-8 monoclonal antibody.
J Immunol
155
1995
2151
210
Folkesson
HG
Matthay
MA
Hebert
CA
Broaddus
VC
Acid aspiration-induced lung injury in rabbits is mediated by interleukin-8-dependent mechanisms.
J Clin Invest
96
1995
107
211
Sekido
N
Mukaida
N
Harada
A
Nakanishi
I
Watanabe
Y
Matsushima
K
Prevention of lung reperfusion injury in rabbits by a monoclonal antibody against interleukin-8.
Nature
365
1993
654
212
Ransohoff
RM
Hamilton
TA
Tani
M
Stoler
MH
Shick
HE
Major
JA
Estes
ML
Thomas
DM
Tuohy
VK
Astrocyte expression of mRNA encoding cytokines IP-10 and JE/MCP-1 in experimental autoimmune encephalomyelitis.
FASEB J
7
1993
592
213
Ransohoff
R
Lehman
P
Tuohy
V
Immunology of multiple sclerosis: New intricacies and new insights.
Curr Opin Neurol
7
1994
242
214
Godiska
R
Chantry
D
Dietsch
GN
Gray
PW
Chemokine expression in murine experimental allergic encephalomyelitis.
J Neuroimmunol
58
1995
167
215
Karpus
WJ
Lukacs
NW
McRae
BL
Strieter
RM
Kunkel
SL
Miller
SD
An important role for the chemokine macrophage inflammatory protein-1α in the pathogenesis of the T cell-mediated autoimmune disease, experimental autoimmune encephalomyelitis.
J Immunol
155
1995
5003
216
Gonzalo
JA
Lloyd
CM
Kremer
L
Finger
E
Martinez-A. C
Siegelman
MH
Cybulsky
M
Guttierez-Ramos
JC
Eosinophil recruitment to the lung in a murine model of allergic inflammation. The role of T cells, chemokines, and adhesion receptors.
J Clin Invest
98
1996
2332
217
Cook
DN
Beck
MA
Coffman
TM
Kirby
SL
Sheridan
JF
Pragnell
IB
Smithies
O
Requirement of MIP-1 alpha for an inflammatory response to viral infection.
Science
269
1995
1583
218
Maddon
PJ
Dalgleish
AG
McDougal
JS
Clapham
PR
Weiss
RA
Axel
R
The T4 gene encodes the AIDS virus receptor and is expressed in the immune system and the brain.
Cell
47
1986
333
219
Cocchi
F
DeVico
AL
Garzino
DA
Arya
SK
Gallo
RC
Lusso
P
Identification of RANTES, MIP-1α, and MIP-1β as the major HIV-suppressive factors produced by CD8+ T cells.
Science
270
1995
1811
220
Choe H, Farzan M, Sun Y, Sullivan N, Rollins B, Ponath PD, Wu L, Mackay CR, LaRosa G, Newman W, Gerard N, Gerard C, Sodroski J: The beta-chemokine receptors CCR3 and CCR5 facilitate infection by primary HIV-1 isolates. Cell 85:1135, 1196
221
Doranz
BJ
Rucker
J
Yi
Y
Smyth
RJ
Samson
M
Peiper
SC
Parmentier
M
Collman
RG
Doms
RW
A dual-tropic primary HIV-1 isolate that uses fusin and the beta-chemokine receptors CKR-5, CKR-3, and CKR-2b as fusion cofactors.
Cell
85
1996
1149
222
Deng
H
Liu
R
Ellmeier
W
Choe
S
Unutmaz
D
Burkhart
M
Di MP
Marmon S
Sutton
RE
Hill
CM
Davis
CB
Peiper
SC
Schall
TJ
Littman
DR
Landau
NR
Identification of a major co-receptor for primary isolates of HIV-1.
Nature
381
1996
661
223
Dragic
T
Litwin
V
Allaway
GP
Martin
SR
Huang
Y
Nagashima
KA
Cayanan
C
Maddon
PJ
Koup
RA
Moore
JP
Paxton
WA
HIV-1 entry into CD4+ cells is mediated by the chemokine receptor CC-CKR-5.
Nature
381
1996
667
224
Samson
M
Libert
F
Doranz
BJ
Rucker
J
Liesnard
C
Farber
CM
Saragosti
S
Lapouméroulie
C
Cognaux
J
Forceille
C
Muyldermans
G
Verhofstede
C
Burtonboy
G
Georges
M
Imai
T
Rana
S
Yi
Y
Smyth
RJ
Collman
RG
Doms
RW
Vassart
G
Parmentier
M
Resistance to HIV-1 infection in caucasian individuals bearing mutant alleles of the CCR-5 chemokine receptor gene.
Nature
382
1996
722
225
Liu
R
Paxton
WA
Choe
S
Ceradini
D
Martin
SR
Horuk
R
MacDonald
ME
Stuhlmann
H
Koup
RA
Landau
NR
Homozygous defect in HIV-1 coreceptor accounts for resistance of some multiply-exposed individuals to HIV-1 infection.
Cell
86
1996
367
226
Dean
M
Carrington
M
Winkler
C
Huttley
GA
Smith
MW
Allikmets
R
Goedert
JJ
Buchbinder
SP
Vittinghoff
E
Gomperts
E
Donfield
S
Vlahov
D
Kaslow
R
Saah
A
Rinaldo
C
Detels
R
Study
HGaD
Study
MAC
Study
MHC
Cohort
SFC
Study
A
O'Brien
SJ
Genetic restriction of HIV-1 infection and progression to AIDS by a deletion allele of the CKR5 structural gene.
Science
273
1996
1856
227
Wu
L
Gerard
NP
Wyatt
R
Choe
H
Parolin
C
Ruffing
N
Borsetti
A
Cardoso
AA
Desjardin
E
Newman
W
Gerard
C
Sodroski
J
CD4-induced interaction of primary HIV-1 gp120 glycoproteins with the chemokine receptor CCR-5.
Nature
384
1996
179
228
Trkola
A
Dragic
T
Arthos
J
Binley
JM
Olson
WC
Allaway
GP
Cheng-Mayer
C
Robinson
J
Maddon
PJ
Moore
JP
CD4-dependent, antibody-sensitive interactions between HIV-1 and its co-receptor CCR-5.
Nature
384
1996
184
229
Lapham
CK
Ouyang
J
Chandrasekhar
B
Nguyen
NY
Dimitrov
DS
Golding
H
Evidence for cell-surface association between fusin and the CD4-gp120 complex in human cell lines.
Science
274
1996
602
230
Atchison
RE
Gosling
J
Monteclaro
FS
Franci
C
Digilio
L
Charo
IF
Goldsmith
MA
Multiple extracellular elements of CCR5 and HIV-1 entry: Dissociation from response to chemokines.
Science
274
1996
1924
231
Farzan
M
Choe
H
Martin
KA
Sun
Y
Sidelkot
M
MacKay
CR
Gerard
NP
Sodrowski
J
Gerard
C
Macrophage inflammatory protein-1β-mediated signaling and HIV-1 entry are independent functions of the chemokine refeptor CCR5.
J Biol Chem
272
1997
6874
232
Strieter
RM
Polverini
PJ
Kunkel
SL
Arenberg
DA
Burdick
MD
Kasper
J
Dzuiba
J
Van DJ
Walz A
Marriott
D
Chan
SY
Roczniak
S
Shanafelt
AB
The functional role of the ELR motif in CXC chemokine-mediated angiogenesis.
J Biol Chem
270
1995
27348
233
Strieter
RM
Polverini
PJ
Arenberg
DA
Walz
A
Opdenakker
G
Van DJ
Kunkel SL
Role of C-X-C chemokines as regulators of angiogenesis in lung cancer.
J Leukoc Biol
57
1995
752
234
Koch
AE
Polverini
PJ
Kunkel
SL
Harlow
LA
DiPietro
LA
Elner
VM
Elner
SG
Strieter
RM
Interleukin-8 as a macrophage-derived mediator of angiogenesis.
Science
258
1992
1798
235
Angiolillo
AL
Sgadari
C
Taub
DD
Liao
F
Farber
JM
Maheshwari
S
Kleinman
HK
Reaman
GH
Tosato
G
Human interferon-inducible protein 10 is a potent inhibitor of angiogenesis in vivo.
J Exp Med
182
1995
155
236
Cao
Y
Chen
C
Weatherbee
JA
Tsang
M
Folkman
J
gro-β, a -C-X-C-chemokine, is an angiogenesis inhibitor that suppresses the growth of Lewis lung carcinoma in mice.
J Exp Med
182
1995
2069
237
Maione
TE
Gray
GS
Hunt
AJ
Sharpe
RJ
Inhibition of tumor growth in mice by an analogue of platelet factor 4 that lacks affinity for heparin and retains potent angiostatic activity.
Cancer Res
51
1991
2077
238
Gengrinovitch
S
Greenberg
SM
Cohen
T
Gitay
GH
Rockwell
P
Maione
TE
Levi
BZ
Neufeld
G
Platelet factor-4 inhibits the mitogenic activity of VEGF121 and VEGF165 using several concurrent mechanisms.
J Biol Chem
270
1995
15059
239
Kolber
DL
Knisely
TL
Maione
TE
Inhibition of development of murine melanoma lung metastases by systemic administration of recombinant platelet factor 4.
J Natl Cancer Inst
87
1995
304
240
Sharpe
RJ
Byers
HR
Scott
CF
Bauer
SI
Maione
TE
Growth inhibition of murine melanoma and human colon carcinoma by recombinant human platelet factor 4.
J Natl Cancer Inst
82
1990
848
241
Smith
DR
Polverini
PJ
Kunkel
SL
Orringer
MB
Whyte
RI
Burdick
MD
Wilke
CA
Strieter
RM
Inhibition of interleukin 8 attenuates angiogenesis in bronchogenic carcinoma.
J Exp Med
179
1994
1409
242
Keller
JR
Bartelmez
SH
Sitnicka
E
Ruscetti
FW
Ortiz
M
Gooya
JM
Jacobsen
SE
District and overlapping direct effects of macrophage inflammatory protein-1α and transforming growth factor β on hematopoietic progenitor/stem cell growth.
Blood
84
1994
2175
243
Broxmeyer
HE
Sherry
B
Cooper
S
Lu
L
Maze
R
Beckmann
MP
Cerami
A
Ralph
P
Comparative analysis of the human macrophage inflammatory protein family of cytokines (chemokines) on proliferation of human myeloid progenitor cells. Interacting effects involving suppression, synergistic suppression, and blocking of suppression.
J Immunol
150
1993
3448
244
Broxmeyer
HE
Sherry
B
Lu
L
Cooper
S
Oh
KO
Tekamp-Olson
P
Kwon
BS
Cerami
A
Enhancing and suppressing effects of recombinant murine macrophage inflammatory proteins on colony formation in vitro by bone marrow myeloid progenitor cells.
Blood
76
1990
1110
245
Broxmeyer
HE
Sherry
B
Lu
L
Cooper
S
Carow
C
Wolpe
SD
Cerami
A
Myelopoietic enhancing effects of murine macrophage inflammatory proteins 1 and 2 on colony formation in vitro by murine and human bone marrow granulocyte/macrophage progenitor cells.
J Exp Med
170
1989
1583
246
Quesniaux
VF
Graham
GJ
Pragnell
I
Donaldson
D
Wolpe
SD
Iscove
NN
Fagg
B
Use of 5-fluorouracil to analyze the effect of macrophage inflammatory protein-1α on long-term reconstituting stem cells in vivo.
Blood
81
1993
1497
247
Sarris
AH
Broxmeyer
HE
Wirthmueller
U
Karasavvas
N
Cooper
S
Lu
L
Krueger
J
Ravetch
JV
Human interferon-inducible protein 10: expression and purification of recombinant protein demonstrate inhibition of early human hematopoietic progenitors.
J Exp Med
178
1993
1127
248
Verfaillie
CM
Catanzarro
PM
Li
WN
Macrophage inflammatory protein 1α, interleukin 3 and diffusible marrow stromal factors maintain human hematopoietic stem cells for at least eight weeks in vitro.
J Exp Med
179
1994
643
249
Dunlop
DJ
Wright
EG
Lorimore
S
Graham
GJ
Holyoake
T
Kerr
DJ
Wolpe
SD
Pragnell
IB
Demonstration of stem cell inhibition and myeloprotective effects of SCI/rhMIP α in vivo.
Blood
79
1992
2221
250
Lord
BI
Dexter
TM
Clements
JM
Hunter
MA
Gearing
AJ
Macrophage-inflammatory protein protects multipotent hematopoietic cells from the cytotoxic effects of hydroxyurea in vivo.
Blood
79
1992
2605
251
Maze
R
Sherry
B
Kwon
BS
Cerami
A
Broxmeyer
HE
Myelosuppressive effects in vivo of purified recombinant murine macrophage inflammatory protein-1α.
J Immunol
149
1992
1004
252
Verfaillie
CM
Miller
JS
CD34+/CD33− cells reselected from macrophage inflammatory protein 1α+interleukin-3-supplemented “stroma-noncontact” cultures are highly enriched for long-term bone marrow culture initiating cells.
Blood
84
1994
1442
253
Mantel
C
Kim
YJ
Cooper
S
Kwon
B
Broxmeyer
HE
Polymerization of murine macrophage inflammatory protein 1α inactivates its myelosuppressive effects in vitro: The active form is a monomer.
Proc Natl Acad Sci USA
90
1993
2232
254
Daly
TJ
LaRosa
GJ
Dolich
S
Maione
TE
Cooper
S
Broxmeyer
HE
High activity suppression of myeloid progenitor proliferation by chimeric mutants of interleukin 8 and platelet factor 4.
J Biol Chem
270
1995
23282
255
Cacalano
G
Lee
J
Kikly
K
Ryan
AM
Pitts-Meek
S
Hultgren
B
Wood
WI
Moore
MW
Neutrophil and B cell expansion in mice that lack the murine IL-8 receptor homolog.
Science
265
1994
682
256
Broxmeyer
HE
Cooper
S
Cacalano
G
Hague
NL
Bailish
E
Moore
MW
Involvement of interleukin (IL) 8 receptor in negative regulation of myeloid progenitor cells in vivo: evidence from mice lacking the murine IL-8 receptor homologue.
J Exp Med
184
1996
1825
257
Parkinson
EK
Graham
GJ
Daubersies
P
Burns
JE
Heufler
C
Plumb
M
Schuler
G
Pragnell
IB
Hemopoietic stem cell inhibitor (SCI/MIP-1α) also inhibits clonogenic epidermal keratinocyte proliferation.
J Invest Dermatol
101
1993
113
258
Wang
J
Huang
M
Lee
P
Komanduri
K
Sharma
S
Chen
G
Dubinett
SM
Interleukin-8 inhibits non-small cell lung cancer proliferation: a possible role for regulation of tumor growth by autocrine and paracrine pathways.
J Interferon Cytokine Res
16
1996
53
259
Rollins
BJ
Sunday
ME
Suppression of tumor formation in vivo by expression of the JE gene in malignant cells.
Mol Cell Biol
11
1991
3125
260
Luster
AD
Leder
P
IP-10, a -C-X-C- chemokine, elicits a potent thymus-dependent antitumor response in vivo.
J Exp Med
178
1993
1057
261
Huang
S
Singh
RK
Xie
K
Gutman
M
Berry
KK
Bucana
CD
Fidler
IJ
Bar-Eli
M
Expression of the JE/MCP-1 gene suppresses metastatic potential in murine colon carcinoma cells.
Cancer Immunol Immunother
39
1994
231
262
Laning
J
Kawasaki
H
Tanaka
E
Luo
Y
Dorf
ME
Inhibition of in vivo tumor growth by the beta chemokine, TCA3.
J Immunol
153
1994
4625
263
Dilloo
D
Bacon
K
Holden
W
Zhong
W
Burdach
S
Zlotnik
A
Brenner
M
Combined chemokine and cytokine gene transfer enhances antitumor immunity.
Nat Med
2
1996
1090
264
Manome
Y
Wen
PY
Hershowitz
A
Tanaka
T
Rollins
BJ
Kufe
DW
Fine
HA
Monocyte chemoattractant protein-1 (MCP-1) gene transduction: an effective tumor vaccine strategy for non-intracranial tumors.
Cancer Immunol Immunother
41
1995
227
265
Sgadari
C
Angiolillo
AL
Cherney
BW
Pike
SE
Farber
JM
Koniaris
LG
Vanguri
P
Burd
PR
Sheikh
N
Gupta
G
Teruya-Feldstein
J
Toasato
G
Interferon-inducible protein-10 identified as a mediator of tumor necrosis in vivo.
Proc Natl Acad Sci USA
93
1996
13791
266
Rollins
BJ
Monocyte chemoattractant protein 1: A potential regulator of monocyte recruitment in inflammatory disease.
Mol Med Today
2
1996
198
267
Baba
M
Imai
T
Nishimura
M
Kakizaki
M
Takagi
S
Hieshima
K
Nomiyama
H
Yoshie
O
Identification of CCR6, the specific receptor for a novel lymphocyte-directed CC chemokine LARC.
J Biol Chem
23
1997
14893
268
Yoshida
R
Imau
T
Hieshima
K
Kusuda
J
Baba
M
Kitaura
M
Nishimura
M
Kakizaki
M
Nomiyama
H
Yoshie
O
Molecular cloning of a novel human CC chemokine EBI1-ligand chemokine that is a specific functional ligand for EBI1, CCR7.
J Biol Chem
272
1997
13803
Sign in via your Institution