www.biochemj.org Biochem. J. (2008) 409, 635–649 (Printed in Great Britain)

635

doi:10.1042/BJ20071493

REVIEW ARTICLE

Chemokines and cancer: migration, intracellular signalling and intercellular communication in the microenvironment Morgan O’HAYRE1 , Catherina L. SALANGA1 , Tracy M. HANDEL1,2 and Samantha J. ALLEN1

Inappropriate chemokine/receptor expression or regulation is linked to many diseases, especially those characterized by an excessive cellular infiltrate, such as rheumatoid arthritis and other inflammatory disorders. There is now overwhelming evidence that chemokines are also involved in the progression of cancer, where they function in several capacities. First, specific chemokine– receptor pairs are involved in tumour metastasis. This is not surprising, in view of their role as chemoattractants in cell migration. Secondly, chemokines help to shape the tumour microenvironment, often in favour of tumour growth and metastasis, by recruitment of leucocytes and activation of pro-inflammatory mediators. Emerging evidence suggests that chemokine receptor signalling

also contributes to survival and proliferation, which may be particularly important for metastasized cells to adapt to foreign environments. However, there is considerable diversity and complexity in the chemokine network, both at the chemokine/receptor level and in the downstream signalling pathways they couple into, which may be key to a better understanding of how and why particular chemokines contribute to cancer growth and metastasis. Further investigation into these areas may identify targets that, if inhibited, could render cancer cells more susceptible to chemotherapy.

INTRODUCTION

CX3CL1 (fractalkine) and CXCL16, which have a chemokine domain tethered to a membrane-anchored mucin stalk [6,7]. Despite variable levels of sequence homology, they adopt a characteristic fold that consists of an N-terminal unstructured domain that is critical for signalling, a three-stranded β-sheet connected by loops and turns, and a C-terminal helix (Figure 1A). Although they bind their receptors as monomers in the context of cell migration, many chemokines dimerize or form higherorder oligomers that appear to be important for localization to glycosaminoglycans on cell surfaces and the ECM (extracellular matrix) [8] and possibly for signalling related to other processes separate from migration [9,10]. The chemokine receptors are seven-transmembrane GPCRs (G-protein-coupled receptors). As such, they have been best characterized with respect to signalling through heterotrimeric G-proteins, primarily involving Gi [11] (Figure 1B). However, there is ample evidence indicating that chemokine receptors also signal through other G-protein subtypes or even through non-G-protein-mediated pathways [12– 15]. Furthermore, although the α-subunit of G-proteins has traditionally been regarded as the major signalling subunit, the βγ -subunits are crucial for the activation of many chemokineinduced pathways. Two of the major pathways activated by Gβγ are PI3Kγ (phosphoinositide 3-kinase γ ) and PLC (phospholipase C), whereas Gαi proteins mainly inhibit adenylate cyclase and transduce signals through tyrosine kinases such as Src [11].

Chemokines comprise a superfamily of about 50 human ligands and 20 receptors that are pivotal regulators of cell migration [1,2]. Traditionally, chemokines and their receptors have been divided into four families on the basis of the pattern of cysteine residues in the ligands (CXC, CC, C and CX3C). They have also been functionally classified as being ‘homoeostatic’ or ‘inflammatory’. Homoeostatic chemokines are constitutively expressed and control leucocyte navigation during immune surveillance. Inflammatory chemokines, which constitute the vast majority, are inducible and control cell recruitment to sites of infection and inflammation [3]. Certain chemokines are also involved in developmental processes such as lymphopoiesis, cardiogenesis and development in the central nervous system [4]. For example, in lymphopoiesis, the transition from bone-marrowresident haematopoietic stem cells through development of T-cell precursors in the thymus, migration into secondary lymphoid organs for immune response initiation and maturation into circulating memory and effector T-cells involves sequentially co-ordinated changes in the profiles of chemokine receptor expression to guide the cells into the appropriate microenvironments [5]. Chemokines are 70–130 amino acid soluble proteins that contain one to three disulfide bonds, the exceptions being

Key words: cancer, chemokine receptors, chemokine, metastasis, signalling, tumour microenvironment.

Abbreviations and nomenclature used: Akt, PKB (protein kinase B); CAF, carcinoma-associated fibroblast; CLL, chronic lymphocytic leukaemia cell; DARC, Duffy antigen receptor for chemokines; DC, dendritic cell; EC, endothelial cell; ECM, extracellular matrix; EGF(R), epidermal growth factor (receptor); ELR, Glu-Leu-Arg; ERK, extracellular-signal-regulated kinase; FAK, focal-adhesion kinase; GPCR, G-protein-coupled receptor; GRK, G-protein receptor kinase; HIF-1, hypoxia-inducible factor 1; IL-2, interleukin-2; JNK, c-Jun N-terminal kinase; MAPK, mitogen-activated protein kinase; MCP, monocyte chemoattractant protein; MMP, matrix metalloprotease; NF-κB, nuclear factor κB; PI3Kγ, phosphoinositide 3-kinase γ; PKC, protein kinase C; PLC, phospholipase C; pVHL, von Hippel–Lindau tumour suppressor protein; RNAi, RNA interference; TAM, tumour-associated macrophage; VEGF, vascular endothelial growth factor. Chemokine nomenclature: chemokines are initially referred to as ‘new nomenclature (old nomenclature)’, for example ‘CX3CL1 (fractalkine)’ (where new=CCL#, CXCL#, CL# or CX3CL#; # represents a number and ‘L’ means ‘ligand’); subsequently new nomenclature alone (e.g. CX3CL1) is employed; the nomenclature for the receptors is CCR#, CXCR#, CR# or CX3CR#, where ‘R’ means ‘receptor’; chemoattractant receptor systems are referred to as ‘ligand:receptor’; for example, the chemoattractant CXCL12 and its receptor CXCR4 are referred to as ‘CXCL12:CXCR4’. 1 All authors contributed equally to this work. 2 To whom correspondence should be addressed (email [email protected]).  c The Authors Journal compilation  c 2008 Biochemical Society

Biochemical Journal

Skaggs School of Pharmacy and Pharmaceutical Science, University of California San Diego, La Jolla, CA 92093-0684, U.S.A.

636

Figure 1

M. O’Hayre and others

Molecular events in the classical activation of a chemokine GPCR involving G-proteins

(A) Structure of the IL-8 monomer (PDB ID 1IL8) [205]. The N-terminal signalling domain is highlighted; this region of the ligand is postulated to insert into the helical bundle of the receptor. It also contains the ELR motif in a subset of angiogenic CXC chemokines (discussed in the text), and in many chemokines is subject to proteolytic processing which modulates their activity. Additional receptor binding determinants are distributed along the rest of protein on the face shown, particularly the loop following the N-terminus. (B) Receptor activation. When a chemokine agonist binds to the extracellular side of its receptor, it stabilizes the receptor into a conformation that activates heterotrimeric G-proteins inside the cell by exposing important motifs such as the DRY box [206]. The G-proteins have three subunits: α, β and γ . The Gα subunit interacts directly with the GPCR C-terminal domain, intracellular loops two and three, and with the G-protein β subunit, which forms a tight complex with the γ subunit. In the inactive state, the Gα subunit binds GDP. Upon ligand binding and activation of the GPCR, GDP dissociates from Gα. GDP is then replaced by GTP, Gα-GTP dissociates from the receptor and from Gβγ , and both of these complexes subsequently activate a variety of downstream effectors that ultimately lead to the physiological response. Refraction to continued stimuli involves receptor desensitization and internalization by agonist dependent phosphorylation of the C-terminal tail of the GPCR by GRKs (G-protein receptor kinases) [207]. Receptor phosphorylation subsequently promotes binding of arrestins, which sterically block further interaction with G-proteins and mediate receptor internalization through clathrin-coated pits [208]. Endocytosis of a GPCR can lead to either lysosomal degradation or recycling back to the cell surface and re-sensitization. In addition to their involvement in internalization, β-arrestins can function as signal transducers by activating pathways such as Akt, PI3K, MAPK and NF-κB, which lead to a variety of cellular responses [209] (see Figure 2). An animated version of (B) can be accessed at http://www.BiochemJ.org/bj/409/0635/bj4090635add.htm.

Despite their structural homology and shared ability to induce chemotaxis, different chemokines can elicit other distinct cellular responses [16,17] and/or activate different pathways to elicit a particular response [18,19]. The signalling and physiological response downstream of receptor activation can also vary, depending on the chemokine/receptor combination, the cell type and the pathophysiological state [16,17]. Figure 2 summarizes the major signalling cascades and functional outcomes of chemokine/receptor activation. It is important to bear in mind that specific subsets or combinations of these pathways can be used to induce the functional response. In the last few years, the involvement of chemokines and their receptors in cancer, particularly metastasis, has been firmly established [20–22]. The association with metastasis is not unsurprising, since it is not a random process of cell migration. On the contrary, it has been known since the early 1900s that cancer cells have a propensity to metastasize to specific organs [23]. Furthermore, metastasis has many features in common with normal cell migration. However, key differences include abnormal chemokine receptor expression, regulation or utilization, often on cells that typically do not migrate. Chemokines then provide a physical address for the secondary destination of the tumour cells. The process by which tumours grow and metastasize is complex [24], with many steps required for primary tumour development and establishment of clinically significant secondary tumours (Figure 3). These steps include: (i) survival and growth of the primary tumour (ii) detachment of tumour cells from the primary lesion (iii) invasion into vascular or lymphatic vessels (iv) homing and adherence to the destination organs (v) survival, growth and ‘organogenesis’ of the metastasized cells in their new environment [21,25]  c The Authors Journal compilation  c 2008 Biochemical Society

Since alternative environments such as bone marrow and lymph node are not naturally compatible with cells from the breast, for example, cancer cells must both derive and provide signals to favourably shape the tumour microenvironment to become conducive to survival and growth [26–28]. The role of chemokines and their receptors in cancer can thus be divided into three broad categories which contribute to one or more of the above processes: (i) providing directional cues for migration/metastasis (ii) shaping the tumour microenvironment (iii) providing survival and/or growth signals In the present review we describe the role of chemokines and chemokine receptors in each of these processes. Table 1 summarizes the chemokine receptors and respective ligands involved in cancer and their general mechanism of tumorigenesis. Although their involvement in these three categories is fairly well established, the exact mechanisms of action are not well understood, and the underlying complexity of the chemokine network makes it difficult to characterize them definitively. Consideration of some of these complexities, discussed at the end of the present review, may be crucial to elucidating more precise mechanisms and thus enable the development of better cancer therapeutics. CHEMOKINES IN MIGRATION/METASTASIS OF CANCER CELLS

The leading cause of death in cancer patients is from metastasis, namely the formation of secondary tumours in organs distant from the original tumour. It is not a random process, but rather shows bias for particular tissues, and is ordered, specific and molecularly directed [20,24]. Breast cancer, for example, has a tendency to metastasize to the lymph nodes, bone marrow, lung and liver. When M¨uller et al. [20] highlighted the role

Chemokines and cancer

Figure 2

637

Chemokine receptor signalling in migration and survival/proliferation

One of the first events of cell migration involves cell polarization in response to a chemoattractant, whereby some receptors and signalling molecules localize toward the source of the chemoattractant, termed the leading edge, while other molecules distribute away at the trailing edge [210]. This process occurs via chemokine:receptor signalling through the class IB PI3Kγ , which activates Rac and subsequently PAK (p21-activated kinase). Protrusion of the leading edge to move in the direction of the chemoattractant is mediated by actin polymerization and focal adhesions activated as chemokines bind to their receptors. Gi -dependent signalling through PI3K and various protein tyrosine kinases induces the activation of Akt, Rac and Cdc42, which lead to downstream F-actin polymerization [31,32,211]. At the trailing edge, activation of ROCK (Rho-associated kinase) downstream of Rho is responsible for actomyosin contraction at the rear so the cell can progress forward [30,32]. Calcium release and PKC activation downstream of PLC can also play important roles in mediating adhesion events [146]. Activation of FAK, pyk2 (proline-rich tyrosine kinase 2 or FAK-related tyrosine kinase), and other tyrosine kinases are also important in this process. FAK activation is important in establishing focal adhesions and activating other molecules involved in cell movement, such as p130cas, crk and paxillin [37]. Integrin receptors that interact with the ECM to mediate cell adhesion, and secreted proteases such as MMPs that can aid in migration by degrading the ECM [24], can also be activated downstream of chemokine signalling. As described in more detail in the section on signalling, some chemokines, in normal function or in the context of cancer, also activate a variety of survival and proliferation pathways. Anti-apoptotic/survival signalling, transcription of growth and proliferation-related genes, and transcription of MMPs involved in migration and remodelling the microenvironment are all transduced downstream from Akt, ERK, PKC and tyrosine kinase (e.g. Src) activation. GRK phosphorylation of the C-terminus of chemokine receptors allows β-arrestin to bind, leading to receptor desensitization and internalization. However, β-arrestin binding also leads to the activation of several proteins including Src, MAPK (ERK, p38, JNK) and PI3K. Clearly, there is a large degree of overlap between the upstream signalling molecules underlying these various processes, as these pathways are able to elicit a broad spectrum of effects. Note that continuous lines indicate direct activation or inhibition of the downstream molecule, whereas broken lines indicate indirect activation.

for chemokines in directing organ-specific metastasis, it became clear that chemokine receptor expression patterns on cancer cells and the localization of the corresponding ligands could provide clues for understanding directional metastasis. It has now been established that several chemokines and their receptors play a role in the metastatic process by directing the migration of receptorbearing tumour cells to sites of metastases where the ligands are expressed. These findings make sense because of the parallels one can draw between lymphocyte trafficking and tumour migration. The general mechanisms involved in normal cell migration and metastasis are similar. Chemokines cause cell movement by inducing changes in cytoskeletal structure and dynamics. Actin polymerization leads to formation of protrusions (pseudopods) which, with the help of integrins, form focal adhesions with the ECM to help propel the cell forward [29]. Although multiple pathways contribute to chemokine-induced cell migration, PI3K, FAK (focal adhesion kinase) and the Rho family of GTPases (Rho, Rac, Cdc42) are particularly important [30–33]. ERK (extracellular-signal-regulated kinase) and PKC (protein kinase

C) signalling, independently or in conjunction with PI3K, may also be involved [34–37]. More detail regarding signalling involved in migration is shown in Figure 2. Altered chemokine receptor expression on cancer cells

What then distinguishes normal cells from metastatic cancer cells, enabling the cancer cells to migrate when they normally would not? Although there are many contributing factors, in numerous types of cancer the malignant cells exhibit increased or aberrant expression of particular chemokine receptors relative to their normal counterparts, notably CXCR4, CCR7 and CCR10 [22,25,38,39]. In a study of breast cancer, it was demonstrated that CXCR4 was strongly expressed on cancer cells compared with normal breast epithelial tissue, which does not express any CXCR4, and that antibodies against CXCR4 blocked metastasis in a mouse model of breast cancer [20]. Since that seminal publication, CXCR4 and its ligand CXCL12 [SDF-1α (stromalcell-derived factor-1α)] have been implicated in about 23 different  c The Authors Journal compilation  c 2008 Biochemical Society

638

M. O’Hayre and others

is often activated in cancer cells and can contribute to the transcription of chemokines [e.g. CXCL1 (Gro-α), CXCL8 (IL-8) and CXCL12] and chemokine receptors such as CXCR1, CXCR2 and CXCR4 [39,48]. At the post-transcriptional level, changes in receptor translation and desensitization by internalization and degradation provide other mechanisms for regulating chemokinereceptor expression. For example, enhanced CXCR4 translation in breast-cancer cells was shown to be associated with the oncogene HER2, which may help to protect CXCR4 from ligand-induced ubiquitination and degradation [42]. For more comprehensive reviews on mechanisms associated with chemokine receptor up-regulation in cancer, see [22,43,49]. THE TUMOUR MICROENVIRONMENT

Figure 3 Illustration of the various steps in cancer growth and metastasis where chemokines and receptors play a role In the primary lesion, tumour cells (dark blue) are supported by a network of cells in the microenvironment including fibroblasts (light blue), DCs (green) and TAMs (yellow). Chemokines produced by the tumour cells serve to recruit ECs, thereby promoting angiogenesis. They also recruit leucocytes, which produce other cytokines, growth factors and MMPs that enhance growth, proliferation and angiogenesis. Fibroblasts also produce angiogenic and survival/growth-promoting chemokines. Metastasis of cells is facilitated by up-regulation of particular chemokine receptors (such as CXCR4) on the tumour cells, which enables them to migrate to secondary tissues where the ligands are expressed. Similar to the primary site, paracrine and autocrine chemokine/cytokine signalling among cells within the microenvironment may be especially important for survival and growth of the metastasized cells.

types of cancer [38]. At least two other chemokine systems also play a direct role in metastatic homing of cancer cells: CCR10 in metastasis to skin, where CCL27 (CTACK) is expressed (e.g. melanoma), and CCR7 in lymph node metastasis, where CCL21 (SLC) is expressed [20,25,39,40]. Many reasons for altered chemokine/receptor expression have been identified. Chemokine receptor expression is regulated both at the transcriptional level and post-transcriptionally through RNA stability, translation and receptor desensitization and internalization [41–47]. The tumour microenvironment, and mutant proteins or altered signalling in the cancer cell itself, can also affect chemokine receptor levels. Conditions present within a tumour, such as hypoxia and a rich cytokine environment, including IL-2 (interleukin-2), can induce the transcription of certain chemokine receptors [41,44,45]. For example, hypoxia induces up-regulation of CXCR4 transcription via HIF-1 (hypoxia-inducible factor 1) and through transcript stabilization [42,44]. HIF-1 was also recently found to promote transcription of CCR7, CXCR1 and CXCR2 [46,47]. In renal carcinoma cells it was demonstrated that mutation of pVHL (von Hippel– Lindau tumour suppressor protein), normally responsible for targeting HIF-1 for cell degradation, results in constitutive activation of HIF-1 target genes, including CXCR4 [44]. NFκB (nuclear factor κB) is part of a key signalling pathway that  c The Authors Journal compilation  c 2008 Biochemical Society

A link between inflammation and cancer was observed over 150 years ago when Rudolf Virchow (cited in [50]) noted that cancers tend to occur at sites of chronic inflammation. Although the relationship between cancer and inflammation is complex, epidemiological studies indicate that inflammatory and infectious diseases are often associated with an increased risk of cancer [51]. In many ways, the microenvironment of tumours mimics that of tissues during the height of an inflammatory response to injury [51]. For example, they both contain a large number of cells from both the innate and adaptive immune system, recruited and activated by a complex profile of chemokines, cytokines, growth factors and proteases. However, unlike the organized morphology of normal tissue, and the ultimate resolution of the inflammation that occurs during healing, tumours exist in a state of chronic inflammation characterized by the presence of malignant cells, development of an aberrant vascular network and the persistence of inflammatory mediators. Within the tumour microenvironment, chemokines and their receptors play roles in modulating angiogenesis, cell recruitment, tumour survival and proliferation and, through these processes, help to define the progression of the cancer. Although the present review focuses primarily on the pro-tumorigenic roles of chemokines, it should be noted that many chemokines/receptors have anti-tumorigenic effects, which would be expected, given their physiological role in protecting the host [52]. Recruitment of leucocytes: tumour-associated macrophages and DCs (dendritic cells)

One important link between cancer and inflammation is the recruitment of cells, including neutrophils, macrophages, DCs, eosinophils, mast cells and lymphocytes. Of these, TAMs (tumour-associated macrophages) [53] represent an important component of solid tumours and may account for up to half of the tumour mass. They were first observed in tumours in the late 1970s and therefore represent one of the first specific links between the immune system and cancer. Macrophages are very heterogeneous and play many roles in the progression of cancer, depending upon the nature of their maturation/polarization. Although an oversimplification of their states, designating the two ends of the spectrum as M1 and M2 represents a convenient way to classify macrophage polarization [54]. M1 (classical) macrophage activation by IL-2/interferon-γ and IL-12 is characterized by high levels of antigen presentation, IL-12/IL23 production and development of a polarized type I response, leading to tumour cell cytotoxicity and necrosis. By contrast, M2 polarization is characterized by defective IL-12 production [55], leading to an IL-12low phenotype and is associated with the induction of angiogenic factors, cytokines and proteases.

Chemokines and cancer Table 1

639

Cancer-promoting properties of chemokines and their receptors

Only chemokines/receptors with pro-tumorigenic roles in cancer are listed. However, some of these chemokines/receptors are known to also mediate anti-tumorigenic effects depending on the context and these are indicated by the asterisk (∗ ). Some of the chemokine receptors are directly expressed on cancer cells (D), whereas others function indirectly (I) by recruiting TAMs, DCs or other non-malignant cells that can contribute to the tumour microenvironment. Abbreviations: ATLL, adult T-cell leukaemia/lymphoma; CLL, chronic lymphocytic leukaemia; CTCL, cutaneous T-cell lymphoma; HCC, hepatocellular carcinoma; MM, multiple myeloma. Chemokine receptor Ligand(s)

Direct (D)/indirect (I) effects

CXCR1/2

CXCL1, 2, 3, 5, 6, 7 and 8

CXCR3



CXCR4

CXCL12

CXCR5

Tumorigenic properties

Types of cancer

References

D and I

Angiogenesis (ELR+), invasion and metastasis, growth and proliferation, survival, MMP-2/9/MT1-MMP expression (CXCL8)

Colorectal, lung, melanoma, pancreatic, prostate, renal cell

[46,90,96,101,102,104,138,172,175–178]

D

Invasion and metastasis, survival, proliferation

Colorectal, melanoma, leukaemias

[21,34,179,180]

D and I

Angiogenesis, invasion and metastasis, growth and proliferation, survival, DC recruitment, MMP-9 expression

23 types

[38,103,142,162]

CXCL13

D

Invasion and metastasis, growth and proliferation

Carcinomas (pancreatic, colon, head and neck), CLL and lymphomas

[181–183]

CXCR7

CXCL12

D

Growth, survival

Breast, lung

CCR1



CCL3, 4, 5, 7, 16 and 23

D and I

TAM and DC recruitment, polarization, invasion and Breast, cervical, HCC, lung, MM, metastasis, angiogenesis, MMP-9/19 expression prostate, T-cell leukaemia (CCL5)

[22,66–68,103,106,184–189]

CCR2



CCL2, 7, 8 and 12

D and I

TAM and fibroblast recruitment, polarization, invasion and metastasis, angiogenesis, MMP-12/MT1-MMP expression (CCL2)

Breast, glioma, lung, melanoma, MM, prostate

[59–61,101,104,186,187,190–192]

CCR3



CCL5, 7, 11, 24 and 26

D and I

TAM and eosinophil recruitment, invasion and metastasis, angiogenesis, MMP-19 expression (CCL5)

Breast, cervical, CTCL, melanoma, renal cell

[66,68,106,184,185]

CCR4

CCL2, 3, 5, 17, 22

D and I

TAM and T-cell recruitment, invasion and metastasis ATLL, CTCL, Hodgkin’s lymphoma, ovarian

[72,193–196]

CCR5



D and I

TAM recruitment, polarization, invasion and metastasis, growth, MMP-19 expression (CCL5)

Breast, cervical, lung, MM, pancreatic, prostate

[66,68,106,184–188,197,198]

CCR6

CCL20

D and I

DC recruitment, invasion and metastasis proliferation,

Breast, colorectal, HCC

[71,199]

CCR7

CCL19, 21

D

Invasion and metastasis, survival

Breast, CLL, colorectal, gastric, head and neck, lung, melanoma

[20,21,200]

CCR9

CCL25

D

Invasion and metastasis, survival

Melanoma, prostate

[201,202]

CCR10

CCL27

D

Invasion and metastasis, growth, survival

ATLL, CTCL, melanoma

[40,193,201,203]

CX3CR1



D

Invasion and metastasis, survival

Prostate

[172,204]

CXCL9, 10 and 11

CCL3, 4, 5 and 8

CX3CL1

These factors serve to suppress the adaptive immune response and promote angiogenesis, matrix remodelling and tumour growth. A number of excellent reviews have presented evidence that TAMs function as M2-polarized macrophages [28,53,54] that promote a permissive environment for tumour growth. The in vivo significance of macrophage recruitment to tumour growth has been studied in a number of models and in clinical studies, and generally high numbers of TAMs correlate with enhanced vascularization and growth [51,56]. Early studies indicated a role for the pro-inflammatory MCP (monocyte chemoattractant protein) family of proteins, termed ‘tumour-derived chemotactic factors’ in macrophage recruitment to tumour sites [57]. Over the years, the role of CCL2 (MCP-1) in macrophage recruitment and activation via one of its receptors, CCR2, has been extensively studied [58]. CCL2 is produced by many types of tumours, including breast, pancreas, lung, cervix, ovary, melanomas, sarcomas and glial cell tumours, and by fibroblasts, ECs (endothelial cells) and macrophages at the tumour site [28]. CCL2 expression levels correlate with the extent of macrophage recruitment [59] and, for a number of cancers, expression levels of CCL2 are also related to prognosis [59,60]. The relationship between chemokines, macrophages and

[130,133]

prognosis is complex and dependent upon the cancer: this has been termed the ‘macrophage balance hypothesis’. For example, the effect of CCL2 upon tumour progression in melanoma cell lines is concentration-dependent. Whereas melanoma cell lines transfected with high levels of CCL2 promote tumour rejection, those transfected with lower levels support tumour growth [61]. A recent model for melanoma has directly correlated low levels of CCL2 with the presence of M2 macrophages and increased angiogenesis and tumour growth compared with macrophagedepleted melanomas or melanomas expressing no CCL2 at all [62]. Similarly, pancreatic cancer patients with higher circulating levels of CCL2 had higher survival rates than those with lower circulating levels [63]. However, higher levels of CCL2 are associated with increased malignancy in models of mammary adenocarcinomas [64,65]. CCL5 [RANTES (regulated upon activation, normal T-cell expressed and secreted)] also recruits macrophages to tumours, and its increased presence correlates with poor prognosis in breast- and cervical-cancer patients [66]. A direct link between macrophage recruitment by CCL5 and breast tumour growth was observed using a CCL5 variant that contains an extra methionine residue at its N-terminus and functions as a receptor antagonist.  c The Authors Journal compilation  c 2008 Biochemical Society

640

M. O’Hayre and others

Mice treated with the antagonist showed a decrease in tumour size and vascularization compared with control-treated mice [67]. The ability of CCL5 to induce the monocyte expression of CCL2 may also play an important role in monocyte recruitment and shaping the tumour microenvironment [68]. In total, about 20 chemokines have been detected in neoplastic tissues of various cancers (Table 1). TAMs themselves express chemokine receptors and selected chemokines [e.g. CCL2, CCL17 (TARC), CCL18 (PARC), CCL20 (MIP-3α) and CCL22 (MDC)], suggesting both autocrine and paracrine roles for these proteins in the microenvironment. Interestingly, on TAMs, the expression levels of chemokine receptors such as CCR2 is very low [69] and may be due to the transition of the cells from blood monocytes to tissue macrophages and/or may prevent these cells from migrating out of the tumour microenvironment once they arrive [70]. DCs are attracted to the tumour microenvironment by chemokines [71–74], where they are often found at low levels [75]. Like macrophages, DCs comprise a diverse population of cells and are likely to play multiple roles. Although a number of tumours contain mature plasmacytoid DCs capable of priming T-cells, those observed in ovarian cancers are thought to function as pro-angiogenic factors, acting through tumour necrosis factor-α and CXCL8 [72]. Thus DCs are likely to have both anti-tumorigenic (e.g. T-cell priming) and pro-tumorigenic (e.g. angiogenesis, immune tolerance) roles in cancer. A number of tumours, including primary cutaneous melanomas, ovarian tumours and breast carcinomas, also contain immature DCs [71,73,74]. This suggests that factors in the microenvironment may suppress DC maturation to dampen their anti-tumorigenic effects. Factors such as IL-6 and mononuclear phagocyte colony-stimulating factor present in the tumour environment may also prevent DC maturation by switching the differentiation from DC cells to macrophages [76]. Interestingly, differential localization of immature and mature DCs, most likely recruited by CCL20, has been observed in breast adenocarcinomas [71]. Immature cells incapable of T-cell priming are present within the tumour, whereas mature cells are located in the peritumoural areas, interacting in some cases with T-cells. Since the DC–T cell interaction usually only occurs in the lymph, it suggests that these cells may be mounting a tumour-specific response. Pro-tumorigenic effects of other stromal cells in the microenvironment

The complex interplay between cells in the microenvironment has previously been highlighted by the discovery of tumourpromoting fibroblasts in epithelial cancers [77,78]. Fibroblasts and myofibroblasts often represent a significant portion of stromal cells in carcinomas. Their direct effects upon carcinoma growth were highlighted in 1999, when it was demonstrated that stromal fibroblasts from human prostate carcinomas had tumour-stimulating properties [79]. In that study, fibroblasts from cancerous or non-cancerous tissues were isolated, mixed with immortal but non-tumorigenic prostate epithelial cells, and co-cultured or injected into immunodeficient mice. Only fibroblasts isolated from the carcinomas were able to stimulate tumour growth. These fibroblasts, termed CAFs (carcinoma-associated fibroblasts), include large populations of myofibroblasts that usually play crucial roles in wound repair, and this observation serves as another example of the association between inflammation and cancer. The pro-tumorigenic properties of CAFs in mammary carcinomas are partly mediated by CXCL12 [78]. CXCL12 is expressed at high levels by mammary carcinoma CAFs and plays  c The Authors Journal compilation  c 2008 Biochemical Society

two main roles. First, it directly stimulates tumour growth by binding to, and signalling through, CXCR4 on tumour cells. Secondly, it recruits ECs into tumours, facilitating angiogenesis (see below). However, CAFs exhibit complex and heterogeneous expression profiles, and it is not surprising that other CAF-derived chemokines (e.g. CCL2, CCL5 and CXCL8) also play roles in tumour growth [80–82].

Angiogenesis: recruitment of EC precursors

Angiogenesis is the process of forming new blood vessels and it requires the production of new ECs to form the vessel walls. It is necessary for tumour growth, and cells in the tumour microenvironment release a plethora of small molecules and proteins to promote this process [83–87]. Like leucocyte migration, blood vessel formation is directional and oriented towards the tumour to increase vascularization and enhance growth [88]. Chemoattractants contribute to this process both by recruiting precursor ECs and by inducing their proliferation. However, the role of chemokines in angiogenesis is quite variable: some are angiogenic, some have no effect on angiogenesis, and others are angiostatic (inhibit angiogenesis). This variability is especially evident within the CXC class of chemokines, as the presence (+) or absence (−) of the ELR (GluLeu-Arg) motif near their N-termini correlates with angiogenic or angiostatic characteristics [89–91]. Furthermore, introduction of the ELR motif into the ELR− chemokine CXCL9 [MIG (monokine-induced by interferon γ )] rendered it angiogenic, whereas mutating the ELR motif out of CXCL8 rendered it angiostatic [89]. CXCL12 is the one exception to the ELR correlation within the CXC chemokines, since it lacks the ELR motif, but is nevertherless angiogenic. However, it is believed to mediate its angiogenic effects in part through the induction of VEGF (vascular endothelial growth factor) [16,78]. Non-CXC chemokines can also have angiogenic and angiostatic properties that are not explained by the presence of an ELR motif. For example, CCL1 (I-309), CCL2, CCL11 (Eotaxin) and CX3CL1 can all induce angiogenesis, sometimes directly and sometimes through recruitment of other cells such as TAMs, which in turn release growth and angiogenic factors (e.g. VEGF and basic fibroblast growth factor) [92–94]. Other CC chemokines can inhibit angiogenesis, as has been demonstrated for CCL21 [95]. All of the ELR+ chemokines bind to CXCR2, which is expressed on ECs [96]. Of these, the TAM-derived CXCR2 ligand CXCL8 is thought to play a particularly important role in angiogenesis, where it is believed to induce the migration and proliferation of ECs [96]. The signalling pathways involved in angiogenesis are essentially the same pathways as those which induce cell migration, survival and proliferation of leucocytes (just specifically in ECs), as discussed elsewhere in the present review. Conversely, the angiostatic properties of many of the ELR− chemokines are thought to be due to activation of CXCR3B on ECs, which results in the inhibition of migration and proliferation [87,97]. CXCR3B is a splice variant of CXCR3 that is found on ECs, whereas the CXCR3A splice variant is expressed on mononuclear cells. Interestingly, CXCR3B contains 52 extra amino acids at its N-terminus relative to CXCR3A, which may modulate its chemokine binding and signalling properties. In contrast with CXCR3B, activation of CXCR3A results in activation of survival and proliferation pathways [97]. This celltype-specific expression of the CXCR3 splice variants provides an explanation for how ELR− chemokines can activate the standard chemotaxis and survival/proliferation pathways in mononuclear

Chemokines and cancer

cells while antagonizing these events in ECs in order to inhibit angiogenesis. Other contributors to the microenvironment: MMPs (matrix metalloproteases)

MMPs present in the tumour microenvironment have been classically associated with processes such as angiogenesis. However, it has become clear that the roles of MMPs are complex, and TAM-derived MMPs play both tumour-promoting and tumour-suppressing roles [98–100]. Chemokines and MMPs have an interesting interdependent relationship, since they control the function of one another. For example, CCL2 induces the expression of MMP-12 by monocytes/macrophages [101], CXCL8 induces MMP-2 and MMP-9 in ECs [102], CCL5 and CXCL12 induce MMP-9 expression in monocytes [103] and CCL2 and CXCL8 induce MT1-MMP [104]. In melanoma cells, the up-regulation of MMP-2 by CXCL8 is associated with increased tumour growth and metastasis [105], whereas the up-regulation of MMP-9 by CCL5 is thought to contribute to the progression of breast cancer [106]. Conversely, proteolysis of chemokines within their N-terminal signalling domains by MMPs can modulate their activity; examples include the MCPs (CCL2/7/8), CXCL11 (I-TAC), CXCL8 and CXCL12 [7,107]. In the case of the MCPs, the cleavage products have impaired activity and in fact can function as antagonists, whereas the cleavage of the six or seven N-terminal residues of CXCL8 by MMP-9 increases the activity of this chemokine. Truncated CXCL11 failed to attract lymphocytes and was much less potent in inhibiting microvascular EC migration, leading to the concept that the processed form(s) may lead to decreased numbers of tumourinfiltrating lymphocytes and a more angiogenic environment [107]. Overall, the ability of MMPs to dramatically alter the function of a given chemokine makes the full significance of their presence in the microenvironment difficult to define. Decoy receptors: chemokine receptors that dampen the pro-tumorigenic activities of chemokines?

Recently it has been suggested that the non-signalling ‘decoy chemokine receptors’ DARC (Duffy antigen receptor for chemokines) and D6 may play a role in the tumour microenvironment by binding and sequestering chemokines and thereby inhibiting tumour growth [108]. D6 has an altered DRYLAIV motif in the second intracellular loop (DKYLEIV instead of DRYLAIV) and DARC lacks this motif entirely. Consequently, these receptors fail to couple with G-proteins and cannot induce cell migration, despite their promiscuous chemokine binding profiles [109,110]. Instead, they efficiently internalize their ligands, thus dampening the immune response [110,111]. DARC binds to pro-inflammatory chemokines of both the CC and CXC families, including CCL2 and the angiogenic chemokines CXCL1 and CXCL8. Studies on in vivo models of mice transplanted with non-small cell lung cancer and breast cancer cell lines overexpressing DARC indicated that increased levels of DARC are associated with decreased rates of tumour growth, increased necrosis and decreased metastasis [112,113]. In breast-cancer cells, DARC was thought to be acting, at least in part, by decreasing CCL2 and MMP-9 expression levels, and low levels of DARC in human breast cancer samples were also correlated with a poor prognosis. Transgenic DARC-deficient mice in models of prostate cancer also indicated a role for DARC in reducing tumorigenesis and angiogenesis, by removing prostrate-derived angiogenic chemokines from the circulation.

641

The authors made the link between low levels of DARC expression on erythrocytes and increased mortality rates from prostate cancer [114]. D6 binds promiscuously, and with high affinity, to a number of CC chemokines [108]. It is thought to play a role in regulating and resolving inflammatory reactions by acting as a scavenger receptor [115]. D6-deficient mice show exaggerated inflammatory responses compared with wild-type mice in models of skin inflammation, with higher levels of inflammatory chemokines in the draining lymph nodes. With respect to cancer, D6 suppresses the development of chemically induced skin tumours, whereas D6-deficient mice have increased susceptibility to tumour development [116]. Interestingly, D6 binds major TAMrecruiting chemokines (CCL2/5/7/8), as well as chemokines postulated to skew inflammatory processes, both towards cellmediated (e.g. CCL3/4) and antibody-mediated (CCL2/11/17/22) responses. Although further studies are needed to explore the role of decoy receptors in anti-tumour strategies, the anti-tumorigenic properties of both D6 and DARC again highlight the contribution of chemokine-mediated inflammation to cancer. In summary, chemokines and their receptors play multiple roles in shaping the tumour microenvironment by their ability to attract leucocytes, particularly TAMs, which promote angiogenesis and activate other pro-tumorigenic enzymes and cytokines. However, in order for tumours to thrive, they must also circumvent apoptosis and find ways to promote survival and proliferation. It has recently become clear that tumours utilize chemokine-mediated signalling pathways in order to aid their survival, growth and proliferation, as discussed in the following section. SIGNALLING: SURVIVAL AND PROLIFERATION

Expression of chemokine receptors on cancer cells may provide the cells with more than a mechanism for migration from the primary tumour to a metastatic site. Receptor signalling may also provide a survival advantage. Chemokine signalling contributes to cancer cell survival in the primary tumour, but it is likely to be of particular importance to metastasized cells. What allows the metastasized cells to ‘feel at home’ in a foreign environment? How are they protected from the immune system? Although these issues involve a multitude of factors, chemokines are likely to be significant contributors. Furthermore, their role in survival signalling is not limited to metastatic cancers. Chemokines that serve as migratory signals to control the homing of lymphocytes to protective niches in the bone marrow and lymph nodes may also promote survival of leukaemia cells. The molecular strategies for survival and growth are often the result of utilizing, and sometimes reprogramming, existing physiological pathways [117]. In some cases the survival signals are likely to be related to the role of specific chemokine receptor pairs such as CXCL12:CXCR4 and CXCL12:CXCR7 in survival and growth during normal development [118–122]. In other cases, survival and proliferation signalling may result from redirecting existing migration pathways (Figure 2). Chemokines in survival

Although cancer cells generally have a strong propensity to survive and resist apoptotic stimuli, extracellular survival signals can aide, or even be necessary for, the survival of some cancer cells. For example, CLLs (chronic-lymphocytic-leukaemia cells) rapidly die when cultured in vitro unless they are co-cultured with stromal cells termed ‘Nurselike cells’ [123]. One of the factors secreted by Nurselike cells that contributes to the survival of  c The Authors Journal compilation  c 2008 Biochemical Society

642

M. O’Hayre and others

CLLs in vitro, and presumably in vivo, is CXCL12. CXCL12 has been shown to protect CLLs from spontaneous and drug-induced apoptosis [123], whereas small-molecule CXCR4 antagonists sensitize the cells to fludarabine-induced apoptosis [124]. In addition to leukaemia cells [123], addition of CXCL12 to in vitro cultures of numerous types of CXCR4-expressing cancer cells, including pancreatic adenocarcinoma [125], glioma [126] and breast cancer cells [127], results in their prolonged survival and protection from apoptosis when cultured under suboptimal conditions. CXCL12 also promotes in vivo survival of numerous cancer cells. Administration of antagonists of CXCR4 synergized with chemotherapy in cell killing and tumour regression of glioblastoma multiforme-derived tumour cells [128] and in a B16 murine model of melanoma [129]. Knockdown of CXCR4 expression through RNAi (RNA interference) or its pharmacological inhibition via AMD3100 (a bicyclam antagonist of the chemokine receptor CXCR4) in a murine 4T1 breast cancer model also decreased the formation of primary tumours and was found to delay and reduce the early growth and/or survival of the 4T1 cells in the lung [127]. These data clearly suggest a role for CXCL12:CXCR4 in survival and/or proliferation of both primary and metastasized cells. For many years, it was believed that CXCL12:CXCR4 functioned as an exclusive non-redundant pair, but CXCL12 is now known to bind to another chemokine receptor, namely CXCR7 (formerly the orphan GPCR RDC-1) [130,131]. Like CXCR4, CXCR7 plays a vital role in response to CXCL12 in various aspects of embryonic development [122,132]. Recent studies revealed that targeted deletion of CXCR7 results in postnatal death in > 95 % of the mice, and a different phenotype is observed compared with the CXCR4-knockout mouse, consisting of heart valve defects but normal haematopoietic and neural development [122]. Furthermore, overexpression of CXCR7 is observed in several types of cancers and has been shown to contribute to cell survival and tumour development independently of CXCR4 [130,133]. Small-molecule antagonists of CXCR7 interfere with tumour growth in mouse models of several different cancers. Overexpression of CXCR7 in the MDA-MB 435 breast-cancer cell line, which normally has undetectable levels of CXCR7, resulted in a growth advantage, owing to increased cell survival under suboptimal growth conditions [130]. Implantation of CXCR7-overexpressing MDA-MB 435 cells induced formation of larger tumours in SCID (severe combined immunodeficiency) mice than vector-control transfected cells, despite the absence of CXCR4. RNAi silencing of CXCR7 in the 4T1 breast-cancer cells also resulted in decreased tumour size compared with the wild-type or control RNAi cells. Finally, similar effects of CXCR7 on cell growth/survival were observed in lung-cancer cell lines [133]. Interestingly, in contrast with the activation of CXCR4 by CXCL12, CXCL12 does not induce calcium flux or migration upon engaging CXCR7, indicating that CXCR7 does not signal in the classic chemokine fashion [122,130]. Many theories have been proposed to explain this phenomenon, including CXCR7 signalling through different pathways, CXCR7 heterodimerization with other receptors or its function as a nonsignalling decoy receptor [122,134]. It should be noted that CXCR7 also binds CXCL11, but with lower affinity. Furthermore, CXCL11 is not necessary for development and is naturally absent in some mouse strains, including the C57BL/6 mouse strain which was used as the background CXCR7-knockout mouse mentioned above [133]. The possibility of CXCR7 functioning as a decoy receptor is not unprecedented, since D6 and DARC serve such a function. However, this hypothesis seems unlikely, owing to the importance of CXCR7 in development and the specificity  c The Authors Journal compilation  c 2008 Biochemical Society

of CXCR7 for CXCL12 and CXCL11 compared with D6 and DARC, which bind multiple CC and CXC chemokines [108,110]. Given the distinct roles of CXCR4 and CXCR7 in developmental processes and their non-identical CXCL12-binding domains [132], it will be interesting to determine the differences between these receptors in terms of their roles in cancer and whether they function independently and/or synergistically. Since CXCL12:CXCR4 and CXCL12:CXCR7 are important survival factors in development, it is understandable that they are also prominently involved in cancer cell survival. Yet several other chemokine-receptor pairs also contribute to cancer cell survival (Table 1). For example, the CCL27:CCR10 system is known to attract melanoma cells to the skin, but recent studies suggest that these proteins also promote tumour cell survival by helping to circumvent anti-tumour processes and by providing protection against apoptosis [40]. Chemokines in cell proliferation and tumour growth

Cancer cells frequently have growth and proliferative advantages over their normal cellular counterparts. As previously discussed, EC proliferation is important for the formation of new blood vessels in order to vascularize tumours and provide routes for metastasis. In addition to the recruitment of TAMs that secrete factors to promote cell proliferation and tumour growth, chemokines directly activate growth and proliferation pathways in the cancer cells themselves. Aberrant protein expression within the cancer cells, due to oncogenes or mutations in tumour suppressors, may then enhance the amplitude and/or duration of the chemokine-activated pathways. The ability of some chemokines to induce cell growth and proliferation in the context of cancer is well-established. CXCL1, CXCL2 and CXCL3 were originally named Gro/MGSA-α, -β and -γ respectively for ‘growth-related oncogene’ or ‘melanoma growth stimulatory activity/growth regulated protein’ [135]. These closely related chemokines were found to be expressed in approx. 70 % of melanomas and function as oncogenes [135,136]. All three chemokines bind to CXCR2 and cause activation of ERK, PI3K and tyrosine kinases to mediate cell proliferation and migration effects [90,91,136–139]. However, induction of cell proliferation is not limited to CXCR2 agonists; other chemokines, including CXCL12, also frequently signal growth and proliferation [27,26,127]. Downstream signalling in survival, growth and proliferation

There is a fair amount of overlap between the survival and proliferation signalling pathways, and this is understandable given that the two processes often work hand-in-hand. It has been demonstrated that stimulation of numerous cancer cells with CXCL12 and other chemokines activates the PI3K/Akt [protein kinase B (PKB)] pathway [11,18,27,46,123,140] which is well known to promote survival effects [137]. Although not all chemokines that promote Akt activation enhance the survival of cells (e.g. under low serum conditions), many do, and this pathway seems to be exploited by a variety of cancer cells [141–143]. Numerous downstream effectors and transcription factors of Akt, ERK1/2, and tyrosine kinase signalling can promote cell survival and proliferation (Figure 2). Chemokine signalling often activates NF-κB, which is commonly downstream of Akt, but can be activated through other pathways, such as the PKC one [144]. NF-κB dimerizes and translocates to the nucleus on activation, where it promotes transcription of various apoptosis inhibitors and cell-cycle-promoting genes [145]. Other downstream targets

Chemokines and cancer

of Akt include procaspase-9 and the pro-apoptotic Bcl-2 family member, BAD (Bcl-2/Bcl-XL -antagonist, causing cell death), both of which are inhibited upon phosphorylation. The FKHR (forkhead in rhabdomyosarcoma) family of transcription factors, which induce transcription of numerous apoptotic genes, are also inhibited by Akt [146]. Akt-induced activation of Mdm2/Hdm2 (murine double minute 2/human double minute 2), leading to p53 degradation and inhibition of GSK-3β (glycogen synthase kinase-3β), leading to stabilization of β-catenin, also results in downstream inhibition of negative regulators of cell cycle and activation of cell-cycle-promoting genes [147]. Furthermore, via inhibition of TSC2 (tuberous sclerosis complex 2), Akt leads to mTOR (mammalian target of rapamycin) activation, resulting in activation of p70S6K (p70 S6 kinase) and thus enhanced protein translation of numerous cell-growth regulators [137,148]. ERK1/2 signalling may also contribute to survival through some of these pathways, for example via phosphorylation and inhibition of procaspase-9 and BAD [149,150]. Furthermore, ERK1/2 [a MAPK (mitogen-activated protein kinase)] can itself localize to the nucleus and activate transcription factors involved in cell-cycle regulation and differentiation, thereby promoting cell proliferation [151]. Other MAPKs, including JNK (c-Jun N-terminal kinase), have also been implicated in chemokineinduced proliferation signalling [152]. Thus chemokine receptor signalling, resulting in activation of transcription factors involved in anti-apoptotic mechanisms, cell-cycle regulation, and growthfactor production, are yet other mechanisms whereby cancer cells exploit downstream chemokine signalling pathways. These protumorigenic pathways are likely to be particularly important for the ability of metastatic tumour cells to thrive in foreign environments. COMPLEXITY OF THE CHEMOKINE SYSTEM

As has been discussed above, it is clear that certain chemokine/ receptor pairs contribute to cancer progression and metastasis, whereas others are not implicated or may even mediate antitumorigenic effects [153]. Particular chemokine/receptor pairs such as CXCL12:CXCR4 have a dominant role in metastasis [39], likely related to their role in homoeostasis, development and the fact that the ligand is constitutively expressed. However, they cannot always be categorized as ‘pro-tumorigenic’ and ‘antitumorigenic’; although a particular chemokine may exhibit anti-tumorigenic properties in one context, it may still contribute to malignancy in other types of cancers. For example, one of the CCR7 ligands, CCL21, was shown to mediate anti-tumour effects by inhibiting angiogenesis, but it also provides important directional cues for metastasis of cancer cells to the lymph nodes [20,95]. Additionally, low levels of some chemokines, such as CCL2, may induce pro-tumorigenic properties, whereas higher levels inhibit tumorigenesis. Although it is desirable to compartmentalize the roles of chemokines in cancer in fairly defined ways, there are many complexities that should be appreciated. In this section we discuss some of the many possible mechanisms that can complicate the picture.

643

among ligands and receptors can result in quantitative and qualitative differences in the cellular response [154,155]. For example, it is clear that ligands of the same receptor can elicit different responses, even when their binding affinities are not too dissimilar. Ultimately this must be due to differences in the ligandinduced conformational states and dynamics of the receptors and how they couple into downstream pathways. In an elegant comparative study of CCL17 and CCL22, D’Ambrosio and co-workers showed quantitative differences in CCR4-mediated signalling [156]. CCL22 was much more effective than CCL17 in the induction of integrin-dependent T-cell adhesion, receptor desensitization and internalization. Furthermore, the authors showed that, although CCL22 is the higher-affinity ligand (but only by 2–3-fold), it dissociates more rapidly than CCL17, and they proposed the intriguing hypothesis that the frequency of association/dissociation may be a critical parameter in the activation of certain intracellular signalling pathways. Similarly, CCR7 binds both CCL19 and CCL21 with comparable affinities and demonstrates a similar efficacy in inducing chemotaxis and calcium mobilization. However, CCL19, but not CCL21, led to phosphorylation of CCR7 and subsequent β-arrestin-dependent desensitization in the H9 human T-lymphocyte cell line [157]. The potency and intensity of CCL19-mediated ERK1/2 activation was also higher than that of CCL21-mediated activation [157]. Distinct responses of a particular receptor to different ligands is also evident in the context of cancer, as the CXCR2 ligand CXCL1, but not CXCL8, was able to activate proliferation and tumour growth in prostate cancer cell lines [17]. Similarly, it is clear that the same chemokine binding to different receptors can also elicit different functional responses, as has already been discussed for CXCL12 binding to CXCR4 versus CXCR7. More quantitative and qualitative analysis of the responses of different chemokine/receptor pairs will be needed to appreciate fully how subtle shifts in binding affinity, kinetics and, ultimately, the induced receptor conformation, can lead to different outputs, which, in turn, will be influenced by cell type. In terms of cancer, this means that it may not always be straightforward to assign a role for a particular chemokine or receptor associated with a cancer. It is important to know what the relevant receptor is for an identified ligand and, vice versa, the relevant ligand for the receptor. Additionally, although the identification of chemokine/receptor mRNA transcripts is often used as evidence for the roles of particular proteins in cancer, the data may be misleading, as protein levels do not always track with mRNA levels. Furthermore, chemokines can be agonists of some receptors and antagonists of others, so what then is their true function? As an example of this concept, the ligands CXCL11, CXCL9 and CXCL10 (IP-10) are agonists of CXCR3, but antagonists of CCR3 [158], whereas CXCR3 may act as a decoy receptor of CCL11 [158]. N-terminal proteolytic processing may also activate or de-activate chemokines or change their specificity, and the question is: what is the predominant state in the tumour milieu? Other cell-dependent factors, discussed below, can also alter the response.

Diversity of chemokine/receptor responses

Complexities of intracellular signalling pathways

Owing to their structural homology and common chemoattractantrelated functions (e.g. migration), the diversity of chemokine signalling may be greatly underappreciated. The sheer number of chemokines and receptors, along with the fact that many chemokines bind the same receptor and many receptors engage multiple chemokines, offers the possibility of many outputs. Although such promiscuous partnering gives the appearance of redundancy, emerging evidence suggests that cross-reactivity

Signalling downstream of chemokine receptor activation is also complex, and many factors can influence the functional outcome. Although complexity contributes to fine-tuning of normal chemokine functions, it may also facilitate the ability of cancer cells to adapt various pathways for purposes not normally used in a particular cell type. As discussed above, there is significant overlap between the pathways that are operative in normal chemokine function and those that contribute to cancer. Migration,  c The Authors Journal compilation  c 2008 Biochemical Society

644

M. O’Hayre and others

for example, is critical both to classical chemokine function as well as tumour metastasis. PLC, Akt, ERK1/2 and tyrosine kinase pathways (independently and sometimes in conjunction) have all been implicated in migrational responses in normal cells [18,35,146,159,160]. However, these same signalling molecules also contribute to survival, growth and proliferation in cancer. Furthermore, they can be modulated in many ways in favour of the cancer cell. Paracrine signals from cells in the microenvironment, or autocrine signals, can influence receptor activation and/or regulation. This concept is a major reason why it is so difficult to recapitulate in vivo situations obtaining in in vitro systems. Other complicating factors include G-protein specificity and isoform availability, receptor dimerization, receptor cross-talk and altered signalling and regulation. Receptor cross-talk, in particular, has been demonstrated to modulate chemokine receptor signalling. For example, EGF (epidermal growth factor) and platelet-derived growth factor are established mediators in ovarian cancer growth and metastasis, and there is now clear evidence for a role of CXCL12:CXCR4 [161,162]. Stimulation with CXCL12 in several ovarian-cancer cell lines resulted in cell proliferation through CXCR4 and biphasic activation of ERK1/2 and Akt, which decreased upon addition of an EGFR (EGF receptor)-specific inhibitor, suggesting cross-talk between CXCR4 and EGFR. In addition, CXCL8 stimulation of CXCR1 and CXCR2 expressing ovarian-cancer cells activates ERK1/2 through interactions with EGFR and c-Src [163]. Dimer and higher-order oligomerization of GPCRs is also recognized as an important event in the activation and function of many GPCRs [164,165]. Although the relevance of chemokinereceptor homo- and hetero-dimer formation is still under study [166], it could affect ligand–receptor specificity, the activation of downstream signalling pathways and the duration of the signal in normal or malignant cells. To date, there is evidence suggesting homodimerization (CCR2, CCR5, CXCR1, CXCR2, CXCR4 and CXCR7) and heterodimerization (CCR2/CCR5, CCR2/CXCR4, CXCR1/CXCR2 and CXCR7/CXCR4) of several receptors. Since most of these receptors are implicated in cancer, dimerization could affect the cancer phenotype. Although cancer cells use much of the same machinery and signalling pathways as do normal cells, they do have altered characteristics, such as the expression of oncogenes or mutations in tumour suppressors that can change or exaggerate the response to chemokines. As has already been mentioned, the mutant pVHL tumour suppressor and the HER2 oncogene contribute to aberrant expression levels of chemokine receptors on cancer cells, thus altering how the cells would normally respond to chemokine signals. In addition, such oncogenes or mutant tumour suppressors could potentially have a dramatic effect on chemokine receptor signalling, leading to prolonged or enhanced pathway activation, or even activation of unique pathways. For example, mutation of PTEN (phosphatase and tensin homologue deleted on chromosome ten), a phosphatase that contributes to inactivation of Akt, could prolong chemokine-induced Akt activation, thus leading to aberrant activity [167]. Additionally, although inappropriate expression of particular chemokine receptors is certainly relevant to a variety of cancers, increased receptor levels do not always correlate with enhanced signalling [168]. For example, CXCR4 up-regulation was observed in both metastatic and non-metastatic breastcancer cell lines; however, only the metastatic lines expressed functional CXCR4 [168]. Thus an increase in receptor expression may not always translate into enhanced activity. Furthermore, there may be cancers in which receptor expression is unaltered, yet there are significant changes in the functional response  c The Authors Journal compilation  c 2008 Biochemical Society

and downstream signalling. Chemokine receptor mutations that cause constitutive activation or that impair desensitization could potentially contribute to tumorigenicity [169–171]. Although there are presently no known endogenously expressed chemokine receptors that exhibit these characteristics, the KSHV (Kaposi’s sarcoma herpes virus) GPCR is a CXCR2-like receptor that is constitutively active and contributes to the pathogenesis of Kaposi’s lesions [169]. Similarly, point mutations yielding constitutive activation of CXCR2 in mouse embryonic fibroblast NIH 3T3 cells resulted in cell transformation and induced proliferation [170]. Further adding to the complexity of the situation, many types of cancer cells express multiple chemokine receptors and/or chemokine ligands. CXCR1, CXCR2, CXCR3, CXCR4, CCR7 and CCR10 can all be expressed on melanoma cells and potentially contribute to malignancy [172]. Whether these different receptors contribute independently, redundantly and/or in a co-ordinated manner to the disease remains to be determined. In a tumour microenvironment there is a milieu of growth factors, cytokines and chemokines that most likely function in concert to shape the growth, survival and spread of cancer cells. Although it is critically important to gain a solid understanding of the individual contributions of each factor to the progression of cancer, they do not function in isolation, and it will also be necessary to consider the global picture in the context of complexities such as receptor cross-talk, altered signalling and interactions with other cell types. CONCLUSIONS

The well-established properties of chemokines in controlling cell migration have made them clear candidates for involvement in cancer-cell metastasis. However, the contribution of chemokines to other aspects of cancer, such as growth, proliferation, angiogenesis and cell survival, are also becoming areas of extensive investigation. Many important questions arise from such multifaceted effects of chemokines:

r which pathways are activated by chemokines to elicit different responses? r why are only select chemokines involved in cancer? r why do some chemokines have anti-tumorigenic functions while others clearly contribute to malignancy? r how does the same chemokine mediate different effects in normal cells and in different types of cancer cells? r are any of the receptors good drug targets for cancer? Deciphering signalling pathways activated by chemokines in various cancer cells will be critical to understanding how chemokines influence disease progression and may reveal potential downstream therapeutic targets and consequences of therapeutic intervention. An interesting paradigm that is emerging in the chemokine field, and may become more relevant in the cancer field in the future, is that of receptor cross-talk, both between other chemokine receptors (e.g. homo-/hetero-dimerization) and between different types of receptors at the cell surface. A number of chemokine receptors have been shown to form homodimers and/or heterodimers, and the complexes often show functional differences compared with their respective monomers [173]. In one recent study, the CXCR4 small-molecule antagonist AMD3100 was used to demonstrate heterodimerization of CCR2/CXCR4 and trans-inhibition of CCR2 [174]. In addition, TAK779, an antagonist of CCR2, CCR5 and CXCR3, was able to inhibit CXCL12 binding to CXCR4 in the context of the heterodimer. Antagonist trans-inhibition presents a unique example of

Chemokines and cancer

the functional consequences of heterodimer formation, with substantial implications in drug development. From the standpoint of drug development, it also illustrates the importance of understanding the complex network of interactions associated with chemokines and cancer. In doing so, one of the major challenges will be to find ways to recapitulate the influence of the microenvironment in experimental set-ups. This work was funded by a UARP (University of California AIDS Research Program) Fellowship to S. J. A. (TF06-SD-501), an NIH (National Institutes of Health) Training Grant in Cellular and Molecular Pharmacology (GM007752) to M. O., a Ruth L. Kirschstein NIGMS MARC (National Institute of General Medical Sciences Minority Access to Research Careers) Predoctoral Fellowship (F31) to C. L. S., and awards from the NIH (RO1-AI37113), the Department of Defense (BC060331), UARP (1D06-SD-206) and the Lymphoma Research Foundation to T. M. H.

REFERENCES 1 Rossi, D. and Zlotnik, A. (2000) The biology of chemokines and their receptors. Annu. Rev. Immunol. 18, 217–242 2 Loetscher, P., Moser, B. and Baggiolini, M. (2000) Chemokines and their receptors in lymphocyte traffic and HIV infection. Adv. Immunol. 74, 127–180 3 Moser, B. and Willimann, K. (2004) Chemokines: role in inflammation and immune surveillance. Ann. Rheum. Dis. 63, Suppl. 2, ii84–ii89 4 Nagasawa, T., Tachibana, K. and Kishimoto, T. (1998) A novel CXC chemokine PBSF/SDF-1 and its receptor CXCR4: their functions in development, hematopoiesis and HIV infection. Semin. Immunol. 10, 179–185 5 Moser, B. and Loetscher, P. (2001) Lymphocyte traffic control by chemokines. Nat. Immunol. 2, 123–128 6 Lau, E. K., Allen, S., Hsu, A. and Handel, T. M. (2004) Chemokine-receptor interactions: GPCRs, glycosaminoglycans and viral chemokine binding proteins. Adv. Protein Chem. 68, 351–391 7 Allen, S. J., Crown, S. E. and Handel, T. M. (2007) Chemokine: receptor structure, interactions, and antagonism. Annu. Rev. Immunol. 25, 787–820 8 Proudfoot, A. E. I., Handel, T. M., Johnson, Z., Lau, E. K., Liwang, P., Clark-Lewis, I., Borlat, F., Wells, T. N. C. and Kosco-Vilbois, M. (2003) Glycosaminoglycan binding and oligomerization are essential for the in vivo activity of certain chemokines. Proc. Natl. Acad. Sci. U.S.A. 100, 1885–1890 9 Appay, V., Brown, A., Cribbes, S., Randle, E. and Czaplewski, L. G. (1999) Aggregation of RANTES is responsible for its inflammatory properties. Characterization of nonaggregating, noninflammatory RANTES mutants. J. Biol. Chem. 274, 27505–27512 10 Murooka, T. T., Wong, M. M., Rahbar, R., Majchrzak-Kita, B., Proudfoot, A. E. and Fish, E. N. (2006) CCL5–CCR5-mediated apoptosis in T cells: requirement for glycosaminoglycan binding and CCL5 aggregation. J. Biol. Chem. 281, 25184–25194 11 Thelen, M. (2001) Dancing to the tune of chemokines. Nat. Immunol. 2, 129–134 12 Molon, B., Gri, G., Bettella, M., Gomez-Mouton, C., Lanzavecchia, A., Martinez, A. C., Manes, S. and Viola, A. (2005) T cell costimulation by chemokine receptors. Nat. Immunol. 6, 465–471 13 Bacon, K. B., Premack, B. A., Gardner, P. and Schall, T. J. (1995) Activation of dual T cell signalling pathways by the chemokine RANTES. Science 269, 1727–1730 14 Arai, H. and Charo, I. F. (1996) Differential regulation of G-protein-mediated signalling by chemokine receptors. J. Biol. Chem. 271, 21814–21819 15 Rodriguez-Frade, J. M., Mellado, M. and Martinez, A. C. (2001) Chemokine receptor dimerization: two are better than one. Trends Immunol. 22, 612–617 16 Kijowski, J., Baj-Krzyworzeka, M., Majka, M., Reca, R., Marquez, L. A., Christofidou-Solomidou, M., Janowska-Wieczorek, A. and Ratajczak, M. Z. (2001) The SDF-1–CXCR4 axis stimulates VEGF secretion and activates integrins but does not affect proliferation and survival in lymphohematopoietic cells. Stem Cells 19, 453–466 17 Moore, B. B., Arenberg, D. A., Stoy, K., Morgan, T., Addison, C. L., Morris, S. B., Glass, M., Wilke, C., Xue, Y. Y., Sitterding, S. et al. (1999) Distinct CXC chemokines mediate tumourigenicity of prostate cancer cells. Am. J. Pathol. 154, 1503–1512 18 Ganju, R. K., Brubaker, S. A., Meyer, J., Dutt, P., Yang, Y., Qin, S., Newman, W. and Groopman, J. E. (1998) The α-chemokine, stromal cell-derived factor-1α, binds to the transmembrane G-protein-coupled CXCR-4 receptor and activates multiple signal transduction pathways. J. Biol. Chem. 273, 23169–23175 19 Xue, X., Cai, Z., Seitz, G., Kanz, L., Weisel, K. C. and Mohle, R. (2007) Differential effects of G protein coupled receptors on hematopoietic progenitor cell growth depend on their signalling capacities. Ann. N.Y. Acad. Sci. 1106, 180–189 20 M¨uller, A., Homey, B., Soto, H., Ge, N., Catron, D., Buchanan, M. E., McClanahan, T., Murphy, E., Yuan, W., Wagner, S. N. et al. (2001) Involvement of chemokine receptors in breast cancer metastasis. Nature 410, 50–56

645

21 Zlotnik, A. (2006) Chemokines and cancer. Int. J. Cancer 119, 2026–2029 22 Balkwill, F. (2004) Cancer and the chemokine network. Nat. Rev. Cancer 4, 540–550 23 Murphy, P. M. (2001) Chemokines and the molecular basis of cancer metastasis. N. Engl. J. Med. 345, 833–835 24 Gupta, G. P. and Massague, J. (2006) Cancer metastasis: building a framework. Cell 127, 679–695 25 Murakami, T., Cardones, A. R. and Hwang, S. T. (2004) Chemokine receptors and melanoma metastasis. J. Dermatol. Sci. 36, 71–78 26 Ben-Baruch, A. (2003) Host microenvironment in breast cancer development: inflammatory cells, cytokines and chemokines in breast cancer progression: reciprocal tumour-microenvironment interactions. Breast Cancer Res. 5, 31–36 27 Kulbe, H., Levinson, N. R., Balkwill, F. and Wilson, J. L. (2004) The chemokine network in cancer – much more than directing cell movement. Int. J. Dev. Biol. 48, 489–496 28 Sica, A., Schioppa, T., Mantovani, A. and Allavena, P. (2006) Tumour-associated macrophages are a distinct M2 polarised population promoting tumour progression: potential targets of anti-cancer therapy. Eur. J. Cancer 42, 717–727 29 Friedl, P. and Wolf, K. (2003) Tumour-cell invasion and migration: diversity and escape mechanisms. Nat. Rev. Cancer 3, 362–374 30 Tan, W., Martin, D. and Gutkind, J. S. (2006) The Gα13-Rho signalling axis is required for SDF-1-induced migration through CXCR4. J. Biol. Chem. 281, 39542–39549 31 Tanaka, T., Bai, Z., Srinoulprasert, Y., Yang, B. G., Hayasaka, H. and Miyasaka, M. (2005) Chemokines in tumour progression and metastasis. Cancer Sci. 96, 317–322 32 Li, S., Guan, J. L. and Chien, S. (2005) Biochemistry and biomechanics of cell motility. Annu. Rev. Biomed. Eng. 7, 105–150 33 Zhao, M., Mueller, B. M., Discipio, R. G. and Schraufstatter, I. U. (2007) Akt plays an important role in breast cancer cell chemotaxis to CXCL12. Breast Cancer Res. Treat., doi:10.1007/s10549-007-9712-7 34 Zipin-Roitman, A., Meshel, T., Sagi-Assif, O., Shalmon, B., Avivi, C., Pfeffer, R. M., Witz, I. P. and Ben-Baruch, A. (2007) CXCL10 promotes invasion-related properties in human colorectal carcinoma cells. Cancer Res. 67, 3396–3405 35 Laudanna, C., Mochly-Rosen, D., Liron, T., Constantin, G. and Butcher, E. C. (1998) Evidence of ζ protein kinase C involvement in polymorphonuclear neutrophil integrin-dependent adhesion and chemotaxis. J. Biol. Chem. 273, 30306–30315 36 Scala, S., Giuliano, P., Ascierto, P. A., Ierano, C., Franco, R., Napolitano, M., Ottaiano, A., Lombardi, M. L., Luongo, M., Simeone, E. et al. (2006) Human melanoma metastases express functional CXCR4. Clin. Cancer Res. 12, 2427–2433 37 Webb, D. J., Donais, K., Whitmore, L. A., Thomas, S. M., Turner, C. E., Parsons, J. T. and Horwitz, A. F. (2004) FAK–Src signalling through paxillin, ERK and MLCK regulates adhesion disassembly. Nat. Cell. Biol. 6, 154–161 38 Balkwill, F. (2004) The significance of cancer cell expression of the chemokine receptor CXCR4. Semin. Cancer Biol. 14, 171–179 39 Kakinuma, T. and Hwang, S. T. (2006) Chemokines, chemokine receptors, and cancer metastasis. J. Leukocyte Biol. 79, 639–51. 40 Murakami, T., Cardones, A. R., Finkelstein, S. E., Restifo, N. P., Klaunberg, B. A., Nestle, F. O., Castillo, S. S., Dennis, P. A. and Hwang, S. T. (2003) Immune evasion by murine melanoma mediated through CC chemokine receptor-10. J. Exp. Med. 198, 1337–1347 41 Loetscher, P., Seitz, M., Baggiolini, M. and Moser, B. (1996) Interleukin-2 regulates CC chemokine receptor expression and chemotactic responsiveness in T lymphocytes. J. Exp. Med. 184, 569–577 42 Li, Y. M., Pan, Y., Wei, Y., Cheng, X., Zhou, B. P., Tan, M., Zhou, X., Xia, W., Hortobagyi, G. N., Yu, D. and Hung, M. C. (2004) Upregulation of CXCR4 is essential for HER2-mediated tumour metastasis. Cancer Cell 6, 459–469 43 Busillo, J. M. and Benovic, J. L. (2007) Regulation of CXCR4 signalling. Biochim. Biophys. Acta 1768, 952–963 44 Staller, P., Sulitkova, J., Lisztwan, J., Moch, H., Oakeley, E. J. and Krek, W. (2003) Chemokine receptor CXCR4 downregulated by von Hippel–Lindau tumour suppressor pVHL. Nature 425, 307–311 45 Schioppa, T., Uranchimeg, B., Saccani, A., Biswas, S. K., Doni, A., Rapisarda, A., Bernasconi, S., Saccani, S., Nebuloni, M., Vago, L. et al. (2003) Regulation of the chemokine receptor CXCR4 by hypoxia. J. Exp. Med. 198, 1391–1402 46 Maxwell, P. J., Gallagher, R., Seaton, A., Wilson, C., Scullin, P., Pettigrew, J., Stratford, I. J., Williams, K. J., Johnston, P. G. and Waugh, D. J. (2007) HIF-1 and NF-κB-mediated upregulation of CXCR1 and CXCR2 expression promotes cell survival in hypoxic prostate cancer cells. Oncogene 26, 7333–7345 47 Wilson, J. L., Burchell, J. and Grimshaw, M. J. (2006) Endothelins induce CCR7 expression by breast tumour cells via endothelin receptor A and hypoxia-inducible factor-1. Cancer Res. 66, 11802–11807 48 Helbig, G., Christopherson, K. W., Bhat-Naksharti, P., Kumar, S., Kishimoto, H., Miller, K. D., Broxmeyer, H. E. and Nakshatri, H. (2003) NF-κB promotes breast cancer cell migration and metastasis by inducing the expression of the chemokine receptor CXCR4. J. Biol. Chem. 278, 21631–21638  c The Authors Journal compilation  c 2008 Biochemical Society

646

M. O’Hayre and others

49 Luker, K. E. and Luker, G. D. (2006) Functions of CXCL12 and CXCR4 in breast cancer. Cancer Lett. 238, 30–41 50 Balkwill, F. R. and Mantovani, A. (2001) Inflammation and cancer: back to Virchow? Lancet 357, 539–545 51 Coussens, L. M. and Werb, Z. (2002) Inflammation and cancer. Nature 420, 860–867 52 Homey, B., Muller, A. and Zlotnik, A. (2002) Chemokines: agents for the immunotherapy of cancer? Nat. Rev. Immunol. 2, 175–184 53 Mantovani, A., Schioppa, T., Porta, C., Allavena, P. and Sica, A. (2006) Role of tumour-associated macrophages in tumour progression and invasion. Cancer Metastasis Rev. 25, 315–322 54 Mantovani, A., Sozzani, S., Locati, M., Allavena, P. and Sica, A. (2002) Macrophage polarization: tumour-associated macrophages as a paradigm for polarized M2 mononuclear phagocytes. Trends Immunol. 23, 549–555 55 Sica, A., Saccani, A., Bottazzi, B., Polentarutti, N., Vecchi, A., van Damme, J. and Mantovani, A. (2000) Autocrine production of IL-10 mediates defective IL-12 production and NF-κB activation in tumour-associated macrophages. J. Immunol. 164, 762–767 56 Lin, E. Y. and Pollard, J. W. (2007) Tumour-associated macrophages press the angiogenic switch in breast cancer. Cancer Res. 67, 5064–5066 57 Graves, D. T., Jiang, Y. L., Williamson, M. J. and Valente, A. J. (1989) Identification of monocyte chemotactic activity produced by malignant cells. Science 245, 1490–1493 58 Conti, I. and Rollins, B. J. (2004) CCL2 (monocyte chemoattractant protein-1) and cancer. Semin. Cancer Biol. 14, 149–154 59 Ueno, T., Toi, M., Saji, H., Muta, M., Bando, H., Kuroi, K., Koike, M., Inadera, H. and Matsushima, K. (2000) Significance of macrophage chemoattractant protein-1 in macrophage recruitment, angiogenesis, and survival in human breast cancer. Clin. Cancer Res. 6, 3282–3289 60 Saji, H., Koike, M., Yamori, T., Saji, S., Seiki, M., Matsushima, K. and Toi, M. (2001) Significant correlation of monocyte chemoattractant protein-1 expression with neovascularization and progression of breast carcinoma. Cancer 92, 1085–1091 61 Nesbit, M., Schaider, H., Miller, T. H. and Herlyn, M. (2001) Low-level monocyte chemoattractant protein-1 stimulation of monocytes leads to tumour formation in nontumourigenic melanoma cells. J. Immunol. 166, 6483–6490 62 Gazzaniga, S., Bravo, A. I., Guglielmotti, A., van Rooijen, N., Maschi, F., Vecchi, A., Mantovani, A., Mordoh, J. and Wainstok, R. (2007) Targeting tumour-associated macrophages and inhibition of MCP-1 reduce angiogenesis and tumour growth in a human melanoma xenograft. J. Invest. Dermatol. 127, 2031–2041 63 Monti, P., Leone, B. E., Marchesi, F., Balzano, G., Zerbi, A., Scaltrini, F., Pasquali, C., Calori, G., Pessi, F., Sperti, C. et al. (2003) The CC chemokine MCP-1/CCL2 in pancreatic cancer progression: regulation of expression and potential mechanisms of antimalignant activity. Cancer Res. 63, 7451–7461 64 Neumark, E., Anavi, R., Witz, I. P. and Ben-Baruch, A. (1999) MCP-1 expression as a potential contributor to the high malignancy phenotype of murine mammary adenocarcinoma cells. Immunol. Lett. 68, 141–146 65 Neumark, E., Cohn, M. A., Lukanidin, E., Witz, I. P. and Ben-Baruch, A. (2002) Possible co-regulation of genes associated with enhanced progression of mammary adenocarcinomas. Immunol. Lett. 82, 111–121 66 Niwa, Y., Akamatsu, H., Niwa, H., Sumi, H., Ozaki, Y. and Abe, A. (2001) Correlation of tissue and plasma RANTES levels with disease course in patients with breast or cervical cancer. Clin. Cancer Res. 7, 285–289 67 Robinson, S. C., Scott, K. A., Wilson, J. L., Thompson, R. G., Proudfoot, A. E. and Balkwill, F. (2003) A chemokine receptor antagonist inhibits experimental breast tumour growth. Cancer Res. 63, 8360–8365 68 Locati, M., Deuschle, U., Massardi, M. L., Martinez, F. O., Sironi, M., Sozzani, S., Bartfai, T. and Mantovani, A. (2002) Analysis of the gene expression profile activated by the CC chemokine ligand 5/RANTES and by lipopolysaccharide in human monocytes. J. Immunol. 168, 3557–3562 69 Sica, A., Saccani, A., Bottazzi, B., Bernasconi, S., Allavena, P., Gaetano, B., Fei, F., LaRosa, G., Scotton, C. J., Balkwill, F. R. and Mantovani, A. (2000) Defective expression of the monocyte chemotactic protein-1 receptor CCR2 in macrophages associated with human ovarian carcinoma. J. Immunol. 164, 733–738 70 Scotton, C. J., Milliken, D., Wilson, J. L., Raju, S. and Balkwill, F. R. (2001) Analysis of CC chemokine and chemokine receptor expression in solid ovarian tumours. Brit. J. Cancer 8, 891–897 71 Bell, D., Chomarat, P., Broyles, D., Netto, G., Harb, G. M., Lebecque, S., Valladeau, J., Davoust, J., Palucka, K. A. and Banchereau, J. (1999) In breast carcinoma tissue, immature dendritic cells reside within the tumour, whereas mature dendritic cells are located in peritumoural areas. J. Exp. Med. 190, 1417–1426 72 Curiel, T. J., Cheng, P., Mottram, P., Alvarez, X., Moons, L., Evdemon-Hogan, M., Wei, S., Zou, L., Kryczek, I., Hoyle, G. et al. (2004) Dendritic cell subsets differentially regulate angiogenesis in human ovarian cancer. Cancer Res. 64, 5535–5538  c The Authors Journal compilation  c 2008 Biochemical Society

73 Vermi, W., Bonecchi, R., Facchetti, F., Bianchi, D., Sozzani, S., Festa, S., Berenzi, A., Cella, M. and Colonna, M. (2003) Recruitment of immature plasmacytoid dendritic cells (plasmacytoid monocytes) and myeloid dendritic cells in primary cutaneous melanomas. J. Pathol. 200, 255–268 74 Zou, W., Machelon, V., Coulomb-L’Hermin, A., Borvak, J., Nome, F., Isaeva, T., Wei, S., Krzysiek, R., Durand-Gasselin, I., Gordon, A., Pustilnik, T. et al. (2001) Stromal-derived factor-1 in human tumours recruits and alters the function of plasmacytoid precursor dendritic cells. Nat. Med. 7, 1339–1346 75 Shurin, M. R., Shurin, G. V., Lokshin, A., Yurkovetsky, Z. R., Gutkin, D. W., Chatta, G., Zhong, H., Han, B. and Ferris, R. L. (2006) Intratumoural cytokines/chemokines/growth factors and tumour infiltrating dendritic cells: friends or enemies? Cancer Metastasis Rev. 25, 333–356 76 Chomarat, P., Banchereau, J., Davoust, J. and Palucka, A. K. (2000) IL-6 switches the differentiation of monocytes from dendritic cells to macrophages. Nat. Immunol. 1, 510–514 77 Orimo, A. and Weinberg, R. A. (2006) Stromal fibroblasts in cancer: a novel tumour-promoting cell type. Cell Cycle 5, 1597–1601 78 Orimo, A., Gupta, P. B., Sgroi, D. C., Arenzana-Seisdedos, F., Delaunay, T., Naeem, R., Carey, V. J., Richardson, A. L. and Weinberg, R. A. (2005) Stromal fibroblasts present in invasive human breast carcinomas promote tumour growth and angiogenesis through elevated SDF-1/CXCL12 secretion. Cell 121, 335–348 79 Olumi, A. F., Grossfeld, G. D., Hayward, S. W., Carroll, P. R., Tlsty, T. D. and Cunha, G. R. (1999) Carcinoma-associated fibroblasts direct tumour progression of initiated human prostatic epithelium. Cancer Res. 59, 5002–5011 80 Nazareth, M. R., Broderick, L., Simpson-Abelson, M. R., Kelleher, Jr, R. J., Yokota, S. J. and Bankert, R. B. (2007) Characterization of human lung tumour-associated fibroblasts and their ability to modulate the activation of tumour-associated T cells. J. Immunol. 178, 5552–5562 81 Mueller, L., Goumas, F. A., Affeldt, M., Sandtner, S., Gehling, U. M., Brilloff, S., Walter, J., Karnatz, N., Lamszus, K., Rogiers, X. and Broering, D. C. (2007) Stromal fibroblasts in colorectal liver metastases originate from resident fibroblasts and generate an inflammatory microenvironment. Am. J. Pathol 171, 1608–1618 82 Silzle, T., Kreutz, M., Dobler, M. A., Brockhoff, G., Knuechel, R. and Kunz-Schughart, L. A. (2003) Tumour-associated fibroblasts recruit blood monocytes into tumour tissue. Eur. J. Immunol. 33, 1311–1320 83 Goede, V., Brogelli, L., Ziche, M. and Augustin, H. G. (1999) Induction of inflammatory angiogenesis by monocyte chemoattractant protein-1. Int. J. Cancer 82, 765–770 84 Koch, A. E., Polverini, P. J., Kunkel, S. L., Harlow, L. A., DiPietro, L. A., Elner, V. M., Elner, S. G. and Strieter, R. M. (1992) Interleukin-8 as a macrophage-derived mediator of angiogenesis. Science 258, 1798–1801 85 Torisu, H., Ono, M., Kiryu, H., Furue, M., Ohmoto, Y., Nakayama, J., Nishioka, Y., Sone, S. and Kuwano, M. (2000) Macrophage infiltration correlates with tumour stage and angiogenesis in human malignant melanoma: possible involvement of TNFα and IL-1α. Int. J. Cancer 85, 182–188 86 Hu, D. E., Hori, Y. and Fan, T. P. (1993) Interleukin-8 stimulates angiogenesis in rats. Inflammation 17, 135–143 87 Strieter, R. M., Burdick, M. D., Mestas, J., Gomperts, B., Keane, M. P. and Belperio, J. A. (2006) Cancer CXC chemokine networks and tumour angiogenesis. Eur. J. Cancer 42, 768–778 88 Folkman, J. and Klagsbrun, M. (1987) Angiogenic factors. Science 235, 442–447 89 Strieter, R. M., Polverini, P. J., Kunkel, S. L., Arenberg, D. A., Burdick, M. D., Kasper, J., Dzuiba, J., Van Damme, J., Walz, A., Marriott, D. et al. (1995) The functional role of the ELR motif in CXC chemokine-mediated angiogenesis. J. Biol. Chem. 270, 27348–27357 90 Dhawan, P. and Richmond, A. (2002) Role of CXCL1 in tumourigenesis of melanoma. J. Leukocyte Biol. 72, 9–18 91 Schadendorf, D., Moller, A., Algermissen, B., Worm, M., Sticherling, M. and Czarnetzki, B. M. (1993) IL-8 produced by human malignant melanoma cells in vitro is an essential autocrine growth factor. J. Immunol. 151, 2667–2675 92 Salcedo, R., Young, H. A., Ponce, M. L., Ward, J. M., Kleinman, H. K., Murphy, W. J. and Oppenheim, J. J. (2001) Eotaxin (CCL11) induces in vivo angiogenic responses by human CCR3+ endothelial cells. J. Immunol. 166, 7571–7578 93 Lee, S. J., Namkoong, S., Kim, Y. M., Kim, C. K., Lee, H., Ha, K. S., Chung, H. T. and Kwon, Y. G. (2006) Fractalkine stimulates angiogenesis by activating the Raf-1/MEK/ERK- and PI3K/Akt/eNOS-dependent signal pathways. Am. J. Physiol. Heart Circ. Physiol. 291, H2836–H2846 94 Bernardini, G., Spinetti, G., Ribatti, D., Camarda, G., Morbidelli, L., Ziche, M., Santoni, A., Capogrossi, M. C. and Napolitano, M. (2000) I-309 binds to and activates endothelial cell functions and acts as an angiogenic molecule in vivo . Blood 96, 4039–4045 95 Vicari, A. P., Ait-Yahia, S., Chemin, K., Mueller, A., Zlotnik, A. and Caux, C. (2000) Antitumour effects of the mouse chemokine 6Ckine/SLC through angiostatic and immunological mechanisms. J. Immunol. 165, 1992–2000

Chemokines and cancer 96 Addison, C. L., Daniel, T. O., Burdick, M. D., Liu, H., Ehlert, J. E., Xue, Y. Y., Buechi, L., Walz, A., Richmond, A. and Strieter, R. M. (2000) The CXC chemokine receptor 2, CXCR2, is the putative receptor for ELR+ CXC chemokine-induced angiogenic activity. J. Immunol. 165, 5269–5277 97 Lasagni, L., Francalanci, M., Annunziato, F., Lazzeri, E., Giannini, S., Cosmi, L., Sagrinati, C., Mazzinghi, B., Orlando, C., Maggi, E. et al. (2003) An alternatively spliced variant of CXCR3 mediates the inhibition of endothelial cell growth induced by IP-10, Mig, and I-TAC, and acts as functional receptor for platelet factor 4. J. Exp. Med. 197, 1537–1549 98 Jodele, S., Blavier, L., Yoon, J. M. and DeClerck, Y. A. (2006) Modifying the soil to affect the seed: role of stromal-derived matrix metalloproteinases in cancer progression. Cancer Metastasis Rev. 25, 35–43 99 Giraudo, E., Inoue, M. and Hanahan, D. (2004) An amino-bisphosphonate targets MMP-9-expressing macrophages and angiogenesis to impair cervical carcinogenesis. J. Clin. Invest. 114, 623–633 100 Comerford, I. and Nibbs, R. J. (2005) Post-translational control of chemokines: a role for decoy receptors? Immunol. Lett. 96, 163–174 101 Wu, L., Fan, J., Matsumoto, S. and Watanabe, T. (2000) Induction and regulation of matrix metalloproteinase-12 by cytokines and CD40 signalling in monocyte/macrophages. Biochem. Biophys. Res. Commun. 269, 808–815 102 Li, A., Dubey, S., Varney, M. L., Dave, B. J. and Singh, R. K. (2003) IL-8 directly enhanced endothelial cell survival, proliferation, and matrix metalloproteinases production and regulated angiogenesis. J. Immunol. 170, 3369–3376 103 Klier, C. M., Nelson, E. L., Cohen, C. D., Horuk, R., Schlondorff, D. and Nelson, P. J. (2001) Chemokine-Induced secretion of gelatinase B in primary human monocytes. Biol. Chem. 382, 1405–1410 104 Galvez, B. G., Genis, L., Matias-Roman, S., Oblander, S. A., Tryggvason, K., Apte, S. S. and Arroyo, A. G. (2005) Membrane type 1-matrix metalloproteinase is regulated by chemokines monocyte-chemoattractant protein-1/CCL2 and interleukin-8/ CXCL8 in endothelial cells during angiogenesis. J. Biol. Chem. 280, 1292–1298 105 Luca, M., Huang, S., Gershenwald, J. E., Singh, R. K., Reich, R. and Bar-Eli, M. (1997) Expression of interleukin-8 by human melanoma cells up-regulates MMP-2 activity and increases tumour growth and metastasis. Am. J. Pathol. 151, 1105–1113 106 Azenshtein, E., Luboshits, G., Shina, S., Neumark, E., Shahbazian, D., Weil, M., Wigler, N., Keydar, I. and Ben-Baruch, A. (2002) The CC chemokine RANTES in breast carcinoma progression: regulation of expression and potential mechanisms of promalignant activity. Cancer Res. 62, 1093–1102 107 Proost, P., Mortier, A., Loos, T., Vandercappellen, J., Gouwy, M., Ronsse, I., Schutyser, E., Put, W., Parmentier, M., Struyf, S. and Van Damme, J. (2007) Proteolytic processing of CXCL11 by CD13/aminopeptidase N impairs CXCR3 and CXCR7 binding and signalling and reduces lymphocyte and endothelial cell migration. Blood 110, 37–44 108 Mantovani, A., Bonecchi, R. and Locati, M. (2006) Tuning inflammation and immunity by chemokine sequestration: decoys and more. Nat. Rev. Immunol. 6, 907–918 109 Graham, G. J. and McKimmie, C. S. (2006) Chemokine scavenging by D6: a movable feast? Trends Immunol. 27, 381–386 110 Locati, M., Torre, Y. M., Galliera, E., Bonecchi, R., Bodduluri, H., Vago, G., Vecchi, A. and Mantovani, A. (2005) Silent chemoattractant receptors: D6 as a decoy and scavenger receptor for inflammatory CC chemokines. Cytokine Growth Factor Rev. 16, 679–686 111 Rot, A. (2005) Contribution of Duffy antigen to chemokine function. Cytokine Growth Factor Rev. 16, 687–694 112 Wang, J., Ou, Z. L., Hou, Y. F., Luo, J. M., Shen, Z. Z., Ding, J. and Shao, Z. M. (2006) Enhanced expression of Duffy antigen receptor for chemokines by breast cancer cells attenuates growth and metastasis potential. Oncogene 25, 7201–7211 113 Addison, C. L., Belperio, J. A., Burdick, M. D. and Strieter, R. M. (2004) Overexpression of the duffy antigen receptor for chemokines (DARC) by NSCLC tumour cells results in increased tumour necrosis. BMC Cancer 4, 28 114 Shen, H., Schuster, R., Stringer, K. F., Waltz, S. E. and Lentsch, A. B. (2006) The Duffy antigen/receptor for chemokines (DARC) regulates prostate tumour growth. FASEB J. 20, 59–64 115 Martinez de la Torre, Y., Locati, M., Buracchi, C., Dupor, J., Cook, D. N., Bonecchi, R., Nebuloni, M., Rukavina, D., Vago, L., Vecchi, A., Lira, S. A. and Mantovani, A. (2005) Increased inflammation in mice deficient for the chemokine decoy receptor D6. Eur. J. Immunol. 35, 1342–1346 116 Nibbs, R. J., Gilchrist, D. S., King, V., Ferra, A., Forrow, S., Hunter, K. D. and Graham, G. J. (2007) The atypical chemokine receptor D6 suppresses the development of chemically induced skin tumours. J. Clin. Invest. 117, 1884–1892 117 Hendrix, M. J., Seftor, E. A., Seftor, R. E., Kasemeier-Kulesa, J., Kulesa, P. M. and Postovit, L. M. (2007) Reprogramming metastatic tumour cells with embryonic microenvironments. Nat. Rev. Cancer 7, 246–255

647

118 Stumm, R. K., Zhou, C., Ara, T., Lazarini, F., Dubois-Dalcq, M., Nagasawa, T., Hollt, V. and Schulz, S. (2003) CXCR4 regulates interneuron migration in the developing neocortex. J. Neurosci. 23, 5123–5130 119 Ma, Q., Jones, D., Borghesani, P. R., Segal, R. A., Nagasawa, T., Kishimoto, T., Bronson, R. T. and Springer, T. A. (1998) Impaired B-lymphopoiesis, myelopoiesis, and derailed cerebellar neuron migration in CXCR4- and SDF-1-deficient mice. Proc. Natl. Acad. Sci. U.S.A. 95, 9448–9453 120 Zou, Y. R., Kottmann, A. H., Kuroda, M., Taniuchi, I. and Littman, D. R. (1998) Function of the chemokine receptor CXCR4 in haematopoiesis and in cerebellar development. Nature 393, 595–599 121 Nagasawa, T., Hirota, S., Tachibana, K., Takakura, N., Nishikawa, S., Kitamura, Y., Yoshida, N., Kikutani, H. and Kishimoto, T. (1996) Defects of B-cell lymphopoiesis and bone-marrow myelopoiesis in mice lacking the CXC chemokine PBSF/SDF-1. Nature 382, 635–638 122 Sierro, F., Biben, C., Martinez-Munoz, L., Mellado, M., Ransohoff, R. M., Li, M., Woehl, B., Leung, H., Groom, J., Batten, M. et al. (2007) Disrupted cardiac development but normal hematopoiesis in mice deficient in the second CXCL12/SDF-1 receptor, CXCR7. Proc. Natl. Acad. Sci. U.S.A. 104, 14759–14764 123 Nishio, M., Endo, T., Tsukada, N., Ohata, J., Kitada, S., Reed, J. C., Zvaifler, N. J. and Kipps, T. J. (2005) Nurselike cells express BAFF and APRIL, which can promote survival of chronic lymphocytic leukemia cells via a paracrine pathway distinct from that of SDF-1α. Blood 106, 1012–1020 124 Burger, M., Hartmann, T., Krome, M., Rawluk, J., Tamamura, H., Fujii, N., Kipps, T. J. and Burger, J. A. (2005) Small peptide inhibitors of the CXCR4 chemokine receptor (CD184) antagonize the activation, migration, and antiapoptotic responses of CXCL12 in chronic lymphocytic leukemia B cells. Blood 106, 1824–1830 125 Marchesi, F., Monti, P., Leone, B. E., Zerbi, A., Vecchi, A., Piemonti, L., Mantovani, A. and Allavena, P. (2004) Increased survival, proliferation, and migration in metastatic human pancreatic tumour cells expressing functional CXCR4. Cancer Res. 64, 8420–8427 126 Zhou, Y., Larsen, P. H., Hao, C. and Yong, V. W. (2002) CXCR4 is a major chemokine receptor on glioma cells and mediates their survival. J. Biol. Chem. 277, 49481–49487 127 Smith, M. C., Luker, K. E., Garbow, J. R., Prior, J. L., Jackson, E., Piwnica-Worms, D. and Luker, G. D. (2004) CXCR4 regulates growth of both primary and metastatic breast cancer. Cancer Res. 64, 8604–8612 128 Redjal, N., Chan, J. A., Segal, R. A. and Kung, A. L. (2006) CXCR4 inhibition synergizes with cytotoxic chemotherapy in gliomas. Clin. Cancer Res. 12, 6765–6771 129 Lee, C. H., Kakinuma, T., Wang, J., Zhang, H., Palmer, D. C., Restifo, N. P. and Hwang, S. T. (2006) Sensitization of B16 tumour cells with a CXCR4 antagonist increases the efficacy of immunotherapy for established lung metastases. Mol. Cancer Ther. 5, 2592–2599 130 Burns, J. M., Summers, B. C., Wang, Y., Melikian, A., Berahovich, R., Miao, Z., Penfold, M. E., Sunshine, M. J., Littman, D. R., Kuo, C. J. et al. (2006) A novel chemokine receptor for SDF-1 and I-TAC involved in cell survival, cell adhesion, and tumour development. J. Exp. Med. 203, 2201–2213 131 Balabanian, K., Lagane, B., Infantino, S., Chow, K. Y., Harriague, J., Moepps, B., Arenzana-Seisdedos, F., Thelen, M. and Bachelerie, F. (2005) The chemokine SDF-1/CXCL12 binds to and signals through the orphan receptor RDC1 in T lymphocytes. J. Biol. Chem. 280, 35760–35766 132 Dambly-Chaudiere, C., Cubedo, N. and Ghysen, A. (2007) Control of cell migration in the development of the posterior lateral line: antagonistic interactions between the chemokine receptors CXCR4 and CXCR7/RDC1. BMC Dev. Biol. 7, 23 133 Miao, Z., Luker, K. E., Summers, B. C., Berahovich, R., Bhojani, M. S., Rehemtulla, A., Kleer, C. G., Essner, J. J., Nasevicius, A., Luker, G. D. et al. (2007) CXCR7 (RDC1) promotes breast and lung tumour growth in vivo and is expressed on tumour-associated vasculature. Proc. Natl. Acad. Sci. U.S.A. 104, 15735–15740 134 Perlin, J. R. and Talbot, W. S. (2007) Signals on the move: chemokine receptors and organogenesis in zebrafish. Science STKE 2007, PE45 135 Wang, D., Yang, W., Du, J., Devalaraja, M. N., Liang, P., Matsumoto, K., Tsubakimoto, K., Endo, T. and Richmond, A. (2000) MGSA/GRO-mediated melanocyte transformation involves induction of Ras expression. Oncogene 19, 4647–4659 136 Luan, J., Shattuck-Brandt, R., Haghnegahdar, H., Owen, J. D., Strieter, R., Burdick, M., Nirodi, C., Beauchamp, D., Johnson, K. N. and Richmond, A. (1997) Mechanism and biological significance of constitutive expression of MGSA/GRO chemokines in malignant melanoma tumour progression. J. Leukocyte Biol. 62, 588–597 137 Vivanco, I. and Sawyers, C. L. (2002) The phosphatidylinositol 3-Kinase AKT pathway in human cancer. Nat. Rev. Cancer 2, 489–501 138 Brew, R., Erikson, J. S., West, D. C., Kinsella, A. R., Slavin, J. and Christmas, S. E. (2000) Interleukin-8 as an autocrine growth factor for human colon carcinoma cells in vitro. Cytokine 12, 78–85 139 Miyamoto, M., Shimizu, Y., Okada, K., Kashii, Y., Higuchi, K. and Watanabe, A. (1998) Effect of interleukin-8 on production of tumour-associated substances and autocrine growth of human liver and pancreatic cancer cells. Cancer Immunol. Immunother. 47, 47–57  c The Authors Journal compilation  c 2008 Biochemical Society

648

M. O’Hayre and others

140 Curnock, A. P., Logan, M. K. and Ward, S. G. (2002) Chemokine signalling: pivoting around multiple phosphoinositide 3-kinases. Immunology 105, 125–136 141 Brand, S., Dambacher, J., Beigel, F., Olszak, T., Diebold, J., Otte, J.-M., Goeke, B. and Eichhorst, S. T. (2005) CXCR4 and CXCL12 are inversely expressed in colorectal cancer cells and modulate cancer cell migration, invasion and MMP-9 activation. Exp. Cell. Res. 310, 117–130 142 Burger, J. A. and Kipps, T. J. (2006) CXCR4: a key receptor in the crosstalk between tumour cells and their microenvironment. Blood 107, 1761–1767 143 Fujita, N. and Tsuruo, T. (2003) Survival-signalling pathway as a promising target for cancer chemotherapy. Cancer Chemother. Pharmacol. 52, Suppl. 1, S24–S28 144 Ye, R. D. (2001) Regulation of nuclear factor κB activation by G-protein-coupled receptors. J. Leukocyte Biol. 70, 839–848 145 Karin, M. (2006) Nuclear factor-κB in cancer development and progression. Nature 441, 431–436 146 Lee, B. C., Lee, T. H., Avraham, S. and Avraham, H. K. (2004) Involvement of the chemokine receptor CXCR4 and its ligand stromal cell-derived factor 1α in breast cancer cell migration through human brain microvascular endothelial cells. Mol. Cancer Res. 2, 327–338 147 Diehl, J. A., Cheng, M., Roussel, M. F. and Sherr, C. J. (1998) Glycogen synthase kinase-3β regulates cyclin D1 proteolysis and subcellular localization. Genes Dev. 12, 3499–3511 148 Shaw, R. J. and Cantley, L. C. (2006) Ras, PI(3)K and mTOR signalling controls tumour cell growth. Nature 441, 424–430 149 Allan, L. A., Morrice, N., Brady, S., Magee, G., Pathak, S. and Clarke, P. R. (2003) Inhibition of caspase-9 through phosphorylation at Thr 125 by ERK MAPK. Nat. Cell Biol. 5, 647–654 150 Bonni, A., Brunet, A., West, A. E., Datta, S. R., Takasu, M. A. and Greenberg, M. E. (1999) Cell survival promoted by the Ras–MAPK signalling pathway by transcription-dependent and -independent mechanisms. Science 286, 1358–1362 151 Kyriakis, J. M. (2000) MAP kinases and the regulation of nuclear receptors (2000) Science STKE 48, PE1 152 Sutton, A., Friand, V., Brule-Donneger, S., Chaigneau, T., Ziol, M., Sainte-Catherine, O., Poire, A., Saffar, L., Kraemer, M., Vassy, J. et al. (2007) Stromal cell-derived factor-1/chemokine (C-X-C motif) ligand 12 stimulates human hepatoma cell growth, migration, and invasion. Mol. Cancer Res. 5, 21–33 153 Rollins, B. J. (2006) Inflammatory chemokines in cancer growth and progression. Eur. J. Cancer 42, 760–767 154 Mantovani, A. (1999) The chemokine system: redundancy for robust outputs. Immunol. Today 20, 254–257 155 Devalaraja, M. N. and Richmond, A. (1999) Multiple chemotactic factors: fine control or redundancy? Trends Pharmacol. Sci. 20, 151–156 156 D’Ambrosio, D., Albanesi, C., Lang, R., Girolomoni, G., Sinigaglia, F. and Laudanna, C. (2002) Quantitative differences in chemokine receptor engagement generate diversity in integrin-dependent lymphocyte adhesion. J. Immunol. 169, 2303–2312 157 Kohout, T. A., Nicholas, S. L., Perry, S. J., Reinhart, G., Junger, S. and Struthers, R. S. (2004) Differential desensitization, receptor phosphorylation, β-arrestin recruitment, and ERK1/2 activation by the two endogenous ligands for the CC chemokine receptor 7. J. Biol. Chem. 279, 23214–23222 158 Loetscher, P., Pellegrino, A., Gong, J. H., Mattioli, I., Loetscher, M., Bardi, G., Baggiolini, M. and Clark-Lewis, I. (2001) The ligands of CXC chemokine receptor 3, I-TAC, Mig, and IP10, are natural antagonists for CCR3. J. Biol. Chem. 276, 2986–2991 159 Fernandis, A. Z., Cherla, R. P. and Ganju, R. K. (2003) Differential regulation of CXCR4-mediated T-cell chemotaxis and mitogen-activated protein kinase activation by the membrane tyrosine phosphatase, CD45. J. Biol. Chem. 278, 9536–9543 160 Weiss-Haljiti, C., Pasquali, C., Ji, H., Gillieron, C., Chabert, C., Curchod, M. L., Hirsch, E., Ridley, A. J., van Huijsduijnen, R. H., Camps, M. and Rommel, C. (2004) Involvement of phosphoinositide 3-kinase γ , Rac, and PAK signalling in chemokine-induced macrophage migration. J. Biol. Chem. 279, 43273–43284 161 Porcile, C., Bajetto, A., Barbieri, F., Barbero, S., Bonavia, R., Biglieri, M., Pirani, P., Florio, T. and Schettini, G. (2005) Stromal cell-derived factor-1α (SDF-1α/CXCL12) stimulates ovarian cancer cell growth through the EGF receptor transactivation. Exp. Cell Res. 308, 241–253 162 Scotton, C. J., Wilson, J. L., Scott, K., Stamp, G., Wilbanks, G. D., Fricker, S., Bridger, G. and Balkwill, F. R. (2002) Multiple actions of the chemokine CXCL12 on epithelial tumour cells in human ovarian cancer. Cancer Res. 62, 5930–5938 163 Venkatakrishnan, G., Salgia, R. and Groopman, J. E. (2000) Chemokine receptors CXCR-1/2 activate mitogen-activated protein kinase via the epidermal growth factor receptor in ovarian cancer cells. J. Biol. Chem. 275, 6868–6875 164 Terrillon, S. and Bouvier, M. (2004) Roles of G-protein-coupled receptor dimerization. EMBO Rep. 5, 30–34 165 Milligan, G. (2004) G protein-coupled receptor dimerization: function and ligand pharmacology. Mol. Pharmacol. 66, 1–7  c The Authors Journal compilation  c 2008 Biochemical Society

166 Springael, J. Y., Urizar, E. and Parmentier, M. (2005) Dimerization of chemokine receptors and its functional consequences. Cytokine Growth Factor Rev. 16, 611–623 167 Fox, J. A., Ung, K., Tanlimco, S. G. and Jirik, F. R. (2002) Disruption of a single Pten allele augments the chemotactic response of B lymphocytes to stromal cell-derived factor-1. J. Immunol. 169, 49–54 168 Holland, J. D., Kochetkova, M., Akekawatchai, C., Dottore, M., Lopez, A. and McColl, S. R. (2006) Differential functional activation of chemokine receptor CXCR4 is mediated by G proteins in breast cancer cells. Cancer Res. 66, 4117–4124 169 Arvanitakis, L., Geras-Raaka, E., Varma, A., Gershengorn, M. C. and Cesarman, E. (1997) Human herpesvirus KSHV encodes a constitutively active G-protein-coupled receptor linked to cell proliferation. Nature 385, 347–350 170 Burger, M., Burger, J. A., Hoch, R. C., Oades, Z., Takamori, H. and Schraufstatter, I. U. (1999) Point mutation causing constitutive signalling of CXCR2 leads to transforming activity similar to Kaposi’s sarcoma herpesvirus-G protein-coupled receptor. J. Immunol. 163, 2017–2022 171 Yang, T.-Y., Chen, S.-C., Leach, M. W., Manfra, D., Homey, B., Wiekowski, M., Sullivan, L., Jehn, C.-H., Narula, S. K., Chensue, S. W. and Lira, S. A. (2000) Transgenic expression of the chemokine receptor encoded by human herpesvirus 8 induces an angioproliferative disease resembling Kaposi’s sarcoma. J. Exp. Med. 191, 445–453 172 Slettenaar, V. I. and Wilson, J. L. (2006) The chemokine network: a target in cancer biology? Adv. Drug Deliv. Rev. 58, 962–974 173 Springael, J. Y., Le Minh, P. N., Urizar, E., Costagliola, S., Vassart, G. and Parmentier, M. (2006) Allosteric modulation of binding properties between units of chemokine receptor homo- and hetero-oligomers. Mol. Pharmacol. 69, 1652–1661 174 Sohy, D., Parmentier, M. and Springael, J. Y. (2007) Allosteric trans -inhibition by specific antagonists in CCR2/CXCR4 heterodimers. J. Biol. Chem. 282, 30062–30069 175 Yuan, A., Chen, J. J., Yao, P. L. and Yang, P. C. (2005) The role of interleukin-8 in cancer cells and microenvironment interaction. Front. Biosci. 10, 853–865 176 Zhu, Y. M., Bagstaff, S. M. and Woll, P. J. (2006) Production and upregulation of granulocyte chemotactic protein-2/CXCL6 by IL-1β and hypoxia in small cell lung cancer. Br. J. Cancer 94, 1936–1941 177 Varney, M. L., Li, A., Dave, B. J., Bucana, C. D., Johansson, S. L. and Singh, R. K. (2003) Expression of CXCR1 and CXCR2 receptors in malignant melanoma with different metastatic potential and their role in interleukin-8 (CXCL-8)-mediated modulation of metastatic phenotype. Clin. Exp. Metastasis 20, 723–731 178 Wente, M. N., Keane, M. P., Burdick, M. D., Friess, H., Buchler, M. W., Ceyhan, G. O., Reber, H. A., Strieter, R. M. and Hines, O. J. (2006) Blockade of the chemokine receptor CXCR2 inhibits pancreatic cancer cell-induced angiogenesis. Cancer Lett. 241, 221–227 179 Robledo, M. M., Bartolome, R. A., Longo, N., Rodriguez-Frade, J. M., Mellado, M., Longo, I., van Muijen, G. N., Sanchez-Mateos, P. and Teixido, J. (2001) Expression of functional chemokine receptors CXCR3 and CXCR4 on human melanoma cells. J. Biol. Chem. 276, 45098–45105 180 Berencsi, K., Meropol, N. J., Hoffman, J. P., Sigurdson, E., Giles, L., Rani, P., Somasundaram, R., Zhang, T., Kalabis, J., Caputo, L., Furth, E., Swoboda, R., Marincola, F. and Herlyn, D. (2007) Colon carcinoma cells induce CXCL11-dependent migration of CXCR3-expressing cytotoxic T lymphocytes in organotypic culture. Cancer Immunol. Immunother. 56, 359–370 181 Burkle, A., Niedermeier, M., Schmitt-Graff, A., Wierda, W. G., Keating, M. J. and Burger, J. A. (2007) Overexpression of the CXCR5 chemokine receptor, and its ligand, CXCL13 in B cell chronic lymphocytic leukemia. Blood 110, 3316–25. 182 Lopez-Giral, S., Quintana, N. E., Cabrerizo, M., Alfonso-Perez, M., Sala-Valdes, M., De Soria, V. G., Fernandez-Ranada, J. M., Fernandez-Ruiz, E. and Munoz, C. (2004) Chemokine receptors that mediate B cell homing to secondary lymphoid tissues are highly expressed in B cell chronic lymphocytic leukemia and non-Hodgkin lymphomas with widespread nodular dissemination. J. Leukoc. Biol. 76, 462–471 183 Meijer, J., Zeelenberg, I. S., Sipos, B. and Roos, E. (2006) The CXCR5 chemokine receptor is expressed by carcinoma cells and promotes growth of colon carcinoma in the liver. Cancer Res. 66, 9576–9582 184 Luboshits, G., Shina, S., Kaplan, O., Engelberg, S., Nass, D., Lifshitz-Mercer, B., Chaitchik, S., Keydar, I. and Ben-Baruch, A. (1999) Elevated expression of the CC chemokine regulated on activation, normal T cell expressed and secreted (RANTES) in advanced breast carcinoma. Cancer Res. 59, 4681–4687 185 Mrowietz, U., Schwenk, U., Maune, S., Bartels, J., Kupper, M., Fichtner, I., Schroder, J. M. and Schadendorf, D. (1999) The chemokine RANTES is secreted by human melanoma cells and is associated with enhanced tumour formation in nude mice. Br. J. Cancer 79, 1025–1031 186 Arenberg, D. A., Keane, M. P., DiGiovine, B., Kunkel, S. L., Strom, S. R., Burdick, M. D., Iannettoni, M. D. and Strieter, R. M. (2000) Macrophage infiltration in human non-small-cell lung cancer: the role of CC chemokines. Cancer Immunol. Immunother. 49, 63–70

Chemokines and cancer 187 Vaday, G. G., Peehl, D. M., Kadam, P. A. and Lawrence, D. M. (2006) Expression of CCL5 (RANTES) and CCR5 in prostate cancer. Prostate 66, 124–134 188 Menu, E., De Leenheer, E., De Raeve, H., Coulton, L., Imanishi, T., Miyashita, K., Van Valckenborgh, E., Van Riet, I., Van Camp, B., Horuk, R. et al. (2006) Role of CCR1 and CCR5 in homing and growth of multiple myeloma and in the development of osteolytic lesions: a study in the 5TMM model. Clin. Exp. Metastasis 23, 291–300 189 Wu, X., Fan, J., Wang, X., Zhou, J., Qiu, S., Yu, Y., Liu, Y. and Tang, Z. (2007) Downregulation of CCR1 inhibits human hepatocellular carcinoma cell invasion. Biochem. Biophys. Res. Commun. 355, 866–871 190 Liang, Y., Bollen, A. W. and Gupta, N. (2007) CC chemokine receptor-2A is frequently overexpressed in glioblastoma. J. Neurooncol. 86, 153–163 191 Lu, Y., Cai, Z., Xiao, G., Liu, Y., Keller, E. T., Yao, Z. and Zhang, J. (2007) CCR2 expression correlates with prostate cancer progression. J. Cell. Biochem. 101, 676–685 192 Vande Broek, I., Asosingh, K., Vanderkerken, K., Straetmans, N., Van Camp, B. and Van Riet, I. (2003) Chemokine receptor CCR2 is expressed by human multiple myeloma cells and mediates migration to bone marrow stromal cell-produced monocyte chemotactic proteins MCP-1, -2 and -3. Br. J. Cancer 88, 855–862 193 Harasawa, H., Yamada, Y., Hieshima, K., Jin, Z., Nakayama, T., Yoshie, O., Shimizu, K., Hasegawa, H., Hayashi, T., Imaizumi, Y. et al. (2006) Survey of chemokine receptor expression reveals frequent co-expression of skin-homing CCR4 and CCR10 in adult T-cell leukemia/lymphoma. Leuk. Lymphoma 47, 2163–2173 194 Ishida, T., Ishii, T., Inagaki, A., Yano, H., Kusumoto, S., Ri, M., Komatsu, H., Iida, S., Inagaki, H. and Ueda, R. (2006) The CCR4 as a novel-specific molecular target for immunotherapy in Hodgkin lymphoma. Leukemia 20, 2162–2168 195 Ishida, T., Utsunomiya, A., Iida, S., Inagaki, H., Takatsuka, Y., Kusumoto, S., Takeuchi, G., Shimizu, S., Ito, M., Komatsu, H. et al. (2003) Clinical significance of CCR4 expression in adult T-cell leukemia/lymphoma: its close association with skin involvement and unfavorable outcome. Clin. Cancer Res. 9, 3625–3634 196 Jones, D., O’Hara, C., Kraus, M. D., Perez-Atayde, A. R., Shahsafaei, A., Wu, L. and Dorfman, D. M. (2000) Expression pattern of T-cell-associated chemokine receptors and their chemokines correlates with specific subtypes of T-cell non-Hodgkin lymphoma. Blood 96, 685–690 197 Robinson, S. C., Scott, K. A., Wilson, J. L., Thompson, R. G., Proudfoot, A. E. and Balkwill, F. R. (2003) A chemokine receptor antagonist inhibits experimental breast tumour growth. Cancer Res. 63, 8360–8365

649

198 Manes, S., Mira, E., Colomer, R., Montero, S., Real, L. M., Gomez-Mouton, C., Jimenez-Baranda, S., Garzon, A., Lacalle, R. A., Harshman, K. et al. (2003) CCR5 expression influences the progression of human breast cancer in a p53-dependent manner. J. Exp. Med. 198, 1381–1389 199 Rubie, C., Frick, V. O., Wagner, M., Weber, C., Kruse, B., Kempf, K., Konig, J., Rau, B. and Schilling, M. (2006) Chemokine expression in hepatocellular carcinoma versus colorectal liver metastases. World J. Gastroenterol. 12, 6627–6633 200 Till, K. J., Lin, K., Zuzel, M. and Cawley, J. C. (2002) The chemokine receptor CCR7 and α4 integrin are important for migration of chronic lymphocytic leukemia cells into lymph nodes. Blood 99, 2977–2984 201 Hwang, S. T. (2004) Chemokine receptors in melanoma: CCR9 has a potential role in metastasis to the small bowel. J. Invest. Dermatol. 122, xiv–xv 202 Singh, S., Singh, U. P., Stiles, J. K., Grizzle, W. E. and Lillard, Jr, J. W. (2004) Expression and functional role of CCR9 in prostate cancer cell migration and invasion. Clin. Cancer Res. 10, 8743–8750 203 Notohamiprodjo, M., Segerer, S., Huss, R., Hildebrandt, B., Soler, D., Djafarzadeh, R., Buck, W., Nelson, P. J. and von Luettichau, I. (2005) CCR10 is expressed in cutaneous T-cell lymphoma. Int. J. Cancer 115, 641–647 204 Shulby, S. A., Dolloff, N. G., Stearns, M. E., Meucci, O. and Fatatis, A. (2004) CX3CR1-fractalkine expression regulates cellular mechanisms involved in adhesion, migration, and survival of human prostate cancer cells. Cancer Res. 64, 4693–4698 205 Clore, G. M., Appella, E., Yamada, M., Matsushima, K. and Gronenborn, A. M. (1990) Three-dimensional structure of interleukin 8 in solution. Biochemistry 29, 1689–1696 206 Schwartz, T. W., Frimurer, T. M., Holst, B., Rosenkilde, M. M. and Elling, C. E. (2006) Molecular mechanism of 7TM receptor activation – a global toggle switch model. Annu. Rev. Pharmacol. Toxicol. 46, 481–519 207 Reiter, E. and Lefkowitz, R. J. (2006) GRKs and β-arrestins: roles in receptor silencing, trafficking and signalling. Trends Endocrinol. Metab. 17, 159–165 208 Lefkowitz, R. J. and Whalen, E. J. (2004) β-Arrestins: traffic cops of cell signalling. Curr. Opin. Cell Biol. 16, 162–168 209 Lefkowitz, R. J. and Shenoy, S. K. (2005) Transduction of receptor signals by β-arrestins. Science 308, 512–517 210 Sanchez-Madrid, F. and del Pozo, M. A. (1999) Leukocyte polarization in cell migration and immune interactions. EMBO J. 18, 501–511 211 Cancelas, J. A., Jansen, M. and Williams, D. A. (2006) The role of chemokine activation of Rac GTPases in hematopoietic stem cell marrow homing, retention, and peripheral mobilization. Exp. Hematol. 34, 976–985

Received 31 October 2007/23 November 2007; accepted 3 December 2007 Published on the Internet 15 January 2008, doi:10.1042/BJ20071493

 c The Authors Journal compilation  c 2008 Biochemical Society

Biochemical Journal

domain that is critical for signalling, a three-stranded β-sheet connected by loops and turns, ... such, they have been best characterized with respect to signalling through heterotrimeric ..... in protecting the host [52]. Recruitment of leucocytes: ...

910KB Sizes 1 Downloads 195 Views

Recommend Documents

Biochemical Journal
kinases in various tiers phosphorylate and activate a large number .... subjected to Western blot analysis with anti-pERK (α pERK) and anti-total ERK (α gERK) antibodies ..... Taken together, these data indicate ..... in various stages of the study

Biochemical Journal - Beperk de Straling
has aroused concern about possible health problems that may be caused by the .... Subconfluent Rat1 of HeLa cells in 6-cm-diameter dishes or suspended ...

Biochemical Techniques.pdf
2. a) Give the principle of gel permeation chromatography. Explain how this technique. is used to establish the oligomeric nature of proteins. b) Describe the ...

Advances in Biochemical Engineering.pdf
There was a problem previewing this document. Retrying... Download. Connect more apps... Try one of the apps below to open or edit this item. Advances in ...

eBook Download Chemical, Biochemical, and Engineering ...
subject. By providing an applied and modern approach,. Stanley Sandler s Chemical,. Biochemical, and. Engineering. Thermodynamics, Fourth. Edition helps ...

Qualitative Computer Simulations for Biochemical ...
values, and the evaluation of a set of comparisons can be done by means of ... Qualitative computer models originate in frameworks such as qualitative physics ( ...