Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2009 Nov 1.
Published in final edited form as: Arterioscler Thromb Vasc Biol. 2008 Aug 28;28(11):1928–1936. doi: 10.1161/ATVBAHA.108.162925

Chemokines as mediators of neovascularization

Ellen C Keeley a, Borna Mehrad b, Robert M Strieter b
PMCID: PMC2735456  NIHMSID: NIHMS125990  PMID: 18757292

Abstract

Chemokines are a superfamily of homologous heparin-binding proteins, first described for their role in recruiting leukocytes to sites of inflammation. Chemokines have since been recognized as key factors mediating both physiologic and pathologic neovascularization in such diverse clinical settings as malignancy, wound repair, chronic fibroproliferative disorders, myocardial ischemia and atherosclerosis. Members of the CXC chemokine family, structurally defined as containing the ELR aminoacid motif, are potent inducers of angiogenesis, whereas another subset of the CXC chemokines inhibit angiogenesis. In addition, CCL2, a CC chemokine ligand, has been implicated in arteriogenesis. In this manuscript, we review the current literature on the role of chemokines as mediators of neovascularization.

Introduction

Neovascularization is a general term that incorporates three forms of new blood vessel growth: vasculogenesis, angiogenesis, and arteriogenesis 1. Vasculogenesis is defined as the de novo formation of a capillary plexus by endothelial progenitor cells. During embryogenesis, vasculogenesis begins with formation of mesenchymal angioblasts into vascular structures 2. In the postnatal period, vasculogenesis may be mediated by bone marrow-derived endothelial progenitors 3, 4, but the contribution of vasculogenesis to neovascularization in the adult remains controversial 5. Angiogenesis is defined as the formation of new capillary networks from pre-existing capillaries and is the best understood form of neovascularization. Angiogenesis may proceed by “sprouting” or “intussusception” (the internal division of the preexisting capillary plexus 6, 7), resulting in the formation of new thin-walled endothelium-lined structures. Both forms of angiogenesis are triggered by tissue hypoxia 8, 9 and are followed by increased expression of hypoxia-inducible factor-1α (HIF-1α) protein, a nuclear transcription factor, which is the primary molecular event stimulating angiogenesis. Arteriogenesis is the formation of arteries, which are defined as blood vessels with 3 distinct wall layers and vasomotor properties. The mechanisms leading to arteriogenesis are incompletely defined, and may involve remodeling and enlargement of pre-existing vessels 10 or budding of new vessels from post-capillary venules 11. Both, however, are mediated by an increase in shear stress. The newly formed arteries, clinically described as collaterals, develop to bypass severe arterial stenoses to connect the proximal (high-pressure region) to the distal (low-pressure region) arterial system 12.

Although the initial triggers for vasculogenesis, angiogenesis and arteriogenesis differ, they are all affected by a variety of inflammatory cells and mediators that, depending on the particular physiologic or pathologic setting, modify the neovascularization process 1. This review will focus on the processes of angiogenesis and arteriogenesis, and will highlight the critical role that chemokines play in these important biologic processes.

Chemokines are a superfamily of homologous 8-10 kDa heparin-binding cytokine molecules that were described for their role in mediating leukocyte recruitment to sites of inflammation. Chemokine ligands and receptors are also recognized as critical mediators of neovascularization in diverse physiologic and pathologic settings and are involved in the pathogenesis of diverse disease settings including chronic inflammation, fibroproliferative disorders, malignancy, wound repair, and more recently, atherosclerosis 13, 14. The approximately 50 human chemokines are grouped into 4 families on the basis of conserved cysteine residues near their amino terminus, and are designated CC, CXC, C, and CX3C sub-families. In the CC chemokine family, the first two cysteine residues are adjacent, whereas the defining feature of the CXC chemokine family is that the first 2 cysteine residues are separated by a non-conserved amino acid, thus constituting the Cys-X-Cys or ‘CXC’ motif. The CXC chemokine ligands are further subdivided on the basis of the presence or absence of another 3 amino acid sequence, glutamic acid-leucine-arginine (the ‘ELR’ motif), immediately proximal to the CXC sequence. The ELR-positive CXC chemokines, which include IL-8/CXCL8, are potent neutrophil chemoattractants. Among the ELR-negative CXC chemokine ligands, CXCL9, CXCL10 and CXCL11 (previously designated MIG, IP-10 and I-TAC, respectively), are potently induced by both type 1 and type 2 interferons (IFN-α/β and IFN-γ) and attract mononuclear leukocytes, including activated Th-1 CD4 T cells, NK cells, and monocytes and myeloid dendritic cells to sites of inflammation. With regards to their role in neovascularization, the CC chemokine ligand CCL2 has also been shown to be a potent promoter of arteriogenesis. Among the CXC chemokines, the ELR-containing CXC chemokines are potent promoters of angiogenesis, whereas the interferon-inducible subset of the ELR-negative CXC chemokines are potent inhibitors of angiogenesis 15, 16 (Table).

Table.

Human chemokine ligands and receptors involved in angiogenesis and arteriogenesis. Modified from references 15 and 16.

Systematic nomenclature Old nomenclature Receptor
Angiogenic/arteriogenic
 CXCL1 Gro-α CXCR2
 CXCL2 Gro-β CXCR2
 CXCL3 Gro-γ CXCR2
 CXCL5 ENA-78 CXCR2
 CXCL6 GCP-2 CXCR2
 CXCL7 NAP-2 CXCR2
 CXCL8 IL-8 CXCR2
 CCL2 MCP-1 CCR2
 CCL11 Eotaxin CCR3
 CCL16 HCC-4/LEC CCR1
Angiostatic
 CXCL4, CXCL4L1 PF-4, PF-4var CXCR3B*
 CXCL9 Mig CXCR3B
 CXCL10 IP-10 CXCR3B
 CXCL11 I-TAC CXCR3B
 CXCL12 SDF-1 CXCR4
*

glycosaminoglycan binding may be involved, see text for details

Chemokines that promote neovascularization

All ELR-positive CXC chemokines are potent promoters of angiogenesis (Table). In the mouse, all ELR-positive CXC chemokine ligands signal via CXCR2, whereas the human ELR-positive CXC chemokines signal through 2 receptors, CXCR1 and CXCR2. Several observations support the notion that human CXCR2 is the primary receptor for chemokine-mediated angiogenesis: (1) all human ELR-positive CXC chemokines mediate angiogenesis and can bind CXCR2, whereas only two of these ligands, CXCL8 and CXCL6, have the ability to bind to CXCR1; (2) while both CXCR1 and CXCR2 are both expressed by endothelial cells, only the expression of CXCR2 is required for endothelial cell chemotaxis 17, 18; and (3) when the function of CXCR2 is blocked, the response of endothelial cells to CXCL8 is abrogated 19.

CXCR2 activation by the ELR-positive CXC chemokines can lead either to receptor internalization, receptor degradation, or recycling of the receptor to the cell membrane. CXCR2 internalization is critical to the generation of a chemotactic response: a mutation of this receptor which impairs internalization has been shown to markedly reduce chemotaxis 20. In this context, the local concentration of ligands dictates the fate of CXCR2: in the setting of low concentrations of ligand, internalized CXCR2 is targeted for recycling and returns to the cell surface whereas high concentrations of ELR-positive CXC chemokines result in targeting of internalized CXCR2 to endosomes for recycling or lysosomes for receptor proteolysis 20.

CXCR2 plays an integral role in mediating ELR-positive CXC chemokine angiogenesis in the cornea micropocket model in CXCR2+/+ and CXCR2 knockout mice 17: the ELR-positive CXC chemokine-mediated angiogenesis was inhibited in CXCR2 knockout mice and in the presence of CXCR2 neutralizing antibodies. In addition, recent reports have suggested that CXCR2 and ELR+ CXC chemokine ligands can mediate homing of circulating endothelial progenitor cells to sites of arterial injury 21, 22. These studies provide evidence that the CXCR2 receptor is important to ELR-positive CXC chemokine-mediated angiogenesis.

Another receptor shown to modulate the angiogenic effects of the ELR-positive CXC chemokines is the Duffy antigen receptor for chemokines (DARC). This receptor is a chemokine receptor that binds chemokines in the absence of detectable signal transduction events 23, 24. It acts as a decoy receptor that inhibits angiogenesis by sequestering ELR-positive CXC chemokines CXCL1, CXCL5 and CXCL8. Transgenic expression of DARC on mouse endothelial cells resulted in decreased angiogenic response of the animals to ELR-positive CXC chemokines in vivo 25. In a mouse model of prostate cancer, animals on a DARC-deficient background developed larger and more aggressive tumors with greater tumor-associated neovascularization and increased intra-tumor levels of angiogenic ELR-positive CXC chemokines 26. Similarly, in a human non-small cell lung cancer tumor cell line, over-expression of DARC resulted in binding of angiogenic ELR-positive CXC chemokines by the tumor cells and a marked decrease in tumor-mediated angiogenesis and metastases 24.

In addition to the CXC chemokine family, 3 members of the CC chemokine family, CCL2, CCL11 and CCL16, have also been implicated in neovascularization. CCL11 signals via the receptor CCR3 and mediates chemotaxis of human endothelial cells and promotes neovascularization in several models of angiogenesis, including chick chorioallantoic membrane neovascularization and Matrigel plug assays 27. CCL16 is primarily expressed in the liver, suggesting that it may play a role in the liver's vascular development and in angiogenesis associated with hepatic diseases 28. CCL16 has been shown to induce the migration of endothelial cells, promote endothelial differentiation into capillary-like structures, and to mediate angiogenesis in chick chorioallantoic membrane by activating CCR1 28.

CCL2 is the best-described CC chemokine mediator of neovascularization. Endothelial cells express CCR2, the receptor for CCL2, and display demonstrate chemotaxis and tube formation in response to CCL2 in vitro29-31. CCL2-mediated neovascularization has also been demonstrated in in vivo models including the Matrigel plug, chick chorioallantoic membrane, sponge and corneal implantation models 32-34. CCL2-mediated angiogenesis appears to be dependent on membrane type 1-matrix metalloproteinase (MT1-MMP): the absence or blockade of MT1-MMP activity resulted in decreased in vivo and in vitro angiogenesis induced by CCL2 30. CCL2 also significantly induced MT1-MMP surface expression, clustering, activity, and function in human endothelial cells. The angiogenic effect of CCL2 appears to be independent of its effects on leukocyte chemotaxis and is mediated via direct effects on the vascular endothelium 34 and may also mediate homing of endothelial progenitor cells to sites of vascular injury 35. Lastly, in vivo CCL2-induced angiogenesis has been associated with both the induction of vascular endothelial growth factor (VEGF)-A gene expression 36, and the transcription factor, MCP-1 induced protein 37.

CXC Chemokines that inhibit neovascularization

Several ELR-negative CXC chemokines have been implicated as inhibitors of angiogenesis (Table). Among these, human CXCL4, CXCL9, CXCL10, and CXCL11 all mediate angiostasis via the receptor CXCR3. The CXCR3 receptor, which was originally identified on murine endothelial cells 38, has been shown to mediate angiostatic activity 39. This receptor is unique in that it exists in three different splice variants, CXCR3A, CXCR3B, and CXCR3-alt, all generated by alternative splicing of mRNA of a single gene product. The expression of CXCR3A is strongly induced by interleukin (IL)-2, and it is primarily responsible for recruitment of leukocytes, most notably of Th1 lymphocytes 40-46. Conversely, CXCR3B is the main angiostatic variant of CXCR3 and is expressed on endothelial cells 39, 47, 48. The final variant, CXCR3-alt, is the most recently described and has been shown to have greater affinity for CXCL11 as compared to CXCL9 or CXCL10, but its role in angiogenesis is yet to be determined 49. CXCR3B is the main angiostatic receptor for CXCL4, CXCL9, CXCL10, and CXCL11 50, 51. The angiostatic chemokines, CXCL9, CXCL10, and CXCL11 are strongly induced by both type I and type II interferons (IFN-α/β and IFN-γ, respectively). These interferon-inducible ELR-negative CXC chemokines are potent inhibitors of angiogenesis in response to the angiogenic ELR-positive CXC chemokines, as well as to VEGF and basic fibroblast growth factor (bFGF). CXCL4, the first angiostatic CXC chemokine described 52, is a potent inhibitor of endothelial cell chemotaxis and proliferation, and has been shown to inhibit the angiogenic effect of VEGF and bFGF 53.

A unique feature of the CXCR3 ligands is that they mediate two distinct effects, namely inhibition of angiogenesis and promoting Th1-type cell mediated immunity via recruitment of CXCR3-expressing T and NK cells 41, 43, 45, 46, 54. The local production of IFN-γ at the site of inflammation induces a self-perpetuating cycle, promoting further expression of CXCL9, CXCL10, and CXCL11 and recruiting CXCR3-expressing cells that act as a further source of IFN-γ. These combined effects, which we have described as “immunoangiostasis”, can benefit the host in the context of anti-tumor immunity 55, 56. For example, in the context of renal cell carcinoma, the effectiveness of systemic IL-2 therapy was shown to be dependent on the receptor CXCR3; the therapy resulted in the up-regulation of CXCR3 on peripheral blood mononuclear cells, but the down-regulation of its ligands within the tumor. These anti-tumor effects of systemic IL-2 were substantially amplified when systemic administration of IL-2 was combined with over-expression of CXCL9 in the tumor, thereby augmenting the homing of IFN-γ producing leukocytes to the tumor microenvironment, inhibiting tumor-associated angiostasis, and enhancing immune responses against tumor antigens 55. Importantly, immunoangiostasis is operative not only in the context of IL-2-mediated effects in renal cell carcinoma, since a similar mechanism has been noted in IL-12-mediated regression of a mouse model of renal cell carcinoma 57. Similar findings have been shown in non-small cell lung carcinoma 58, 59; in addition to a reduction in angiogenesis, intratumoral injection of a recombinant CC chemokine, CCL21, induced tumor regression in immunocompetent mice, but not immunosuppressed mice suggesting that T cell immunity was required for the anti-tumor effect of CCL21. Moreover, this was associated with intra-tumor generation of IFN-γ, CXCL9 and CXCL10, and depletion studies demonstrated that CXCL9, CXCL10, and IFN-γ each attenuated the anti-tumor effects of CCL21 58.

In addition to binding CXCR3, both CXCL4 and CXCL10 ligands also bind to extra-cellular glycosaminoglycans. To determine whether the angiostatic properties of these ligands were mediated via this mechanism studies were performed using CXCL4 and CXCL10 variants with mutated binding sites for CXCR3 or glycosaminoglycans. The angiostatic activity CXCL4 was retained in cells that lacked surface heparin sulfate, and CXCL4 mutants that lacked heparin-affinity are capable of inhibiting angiogenesis 60-62, indicating that interaction with cell surface glycosaminoglycans is not essential for these effects. Similarly, when CXCL10 variants with mutated binding sites for CXCR3 or glycosaminoglycans were transfected into a human melanoma cell line, wild-type CXCL10 and CXCL10 mutants with partial or complete loss of glycosaminoglycans binding promoted significant reduction in tumor growth compared to control vector-transfected tumor cells, whereas tranfectants expressing mutants with loss of the CXCR3 binding domain did not inhibit tumor growth 63. While these studies demonstrated that the angiostatic effects of CXCR3 ligands are CXCR3-dependent, these ligands can also exert their angiostatic effects by CXCR3-independent mechanisms via interfering with the angiogenic effects of CXCL8, VEGF and bFGF 62, 64, 65. Heterodimerization with CXCL4 prevents the homodimerization of bFGF that is necessary for receptor binding 62, 66. Moreover, CXCL4 restricts VEGF165 binding to its receptors on endothelial cells by a similar mechanism 67. CXCL4 does not, however, appear to bind non-heparin-binding angiogenic peptides: for example, it does not bind to VEGF121 or its receptor 67-69.

The ligand-receptor relationship of CXCL4 is further complicated by the existence of its non-allelic variant, CXCL4L1 (previously designated PF-4var and SCYB4V1). CXCL4L1 differs from CXCL4 in 3 amino acids in the heparin-binding domain near the carboxy terminus 70. CXCL4L1 protein has been isolated from the α-granules of thrombin-activated human platelets 71. CXCL4L1 was >30-fold more potent than CXCL4 in inhibiting human microvascular endothelial cell chemotaxis induced by bFGF and CXCL8, and was also more potent in inhibiting in vitro wound-healing assay and bFGF- and CXCL8-induced angiogenesis in the rat corneal micropocket model 71, 72. CXCL4L1 was also more efficient than CXCL4 in inhibiting tumor-associated angiogenesis in B16 melanoma and A549 lung adenocarcinoma in immunocompromised mice 72.

The CXC chemokine ligand CXCL12 and its receptor, CXCR4, are critical to homing of progenitor cells in diverse biological settings 73. The precise role of this ligand-receptor pair in angiogenesis is not yet fully established. In the context of cancer biology, CXCR4 is expressed by many tumor lines and primary cancer cells, but its ligand, CXCL12 is not expressed within the cancer microenvironment 73-75. In contrast to studies of depletion of the ELR+ CXC chemokines and CXCR2, which show a parallel reduction in angiogenesis, tumor size and metastases, depletion of CXCL12 or CXCR4 does not affect tumor size or extent of primary tumor-associated angiogenesis 63, 76, 77. However, depletion of CXCL12 or CXCR4 was associated with decreased metastases in animal models of breast and lung cancer 74, 75, suggesting that the CXCL12-CXCR4 ligand-receptor pair regulates metastases independent of angiogenesis. On the other hand, CXCL12 is expressed in ischaemic tissues under the control of HIF-1α 78. In addition, in models of wound healing and several models of tissue hypoxia, CXCL12 mediate homing of endothelial progenitor cells to blood vessel walls in the ischaemic tissue 78-82. Interestingly, tissue expression of CXCL12 is associated with that of VEGF 83 and can also be induced by transgenic over-expression of VEGF, supporting the notion of cross-talk between chemokine- and cytokine-mediated angiogenesis 84.

Chemokines and neovascularization in human disease

Angiogenic and angiostatic chemokines have been implicated in the neovascularization process that occurs in diverse human diseases, including malignancy, chronic inflammatory and fibroproliferative disorders, wound repair, myocardial ischemia, heart failure and atherosclerosis.

Malignancy

Angiogenesis is essential for the development and progression of tumors; the increased metabolic demand brought on by rapid growth of neoplastic tissue requires commensurate increases in blood supply in order to maintain tissue integrity. When the increased metabolic demands outpace the pre-existing blood supply, a wide range of molecular signals within the microenvironment of the tumor result in the process of angiogenesis. Much of the tumor angiogenesis research has focused on the effects of the VEGF family 85. However, CXC chemokine-mediated angiogenesis has been shown to play a critical role in growth of many malignancies including lung, pancreatic, ovarian, prostate, melanoma, brain, and renal cell 86-92.

The ELR-positive CXC chemokines CXCL5 and CXCL8 have been shown to play an important role in human non-small cell lung cancer (NSCLC). In a SCID mouse model, human NSCLC tumor-derived CXCL8 levels were directly related to the extent of angiogenesis; when CXCL8 was depleted, however, there was a significant reduction in tumor size, tumor-induced angiogenesis, and metastases 76. It has also been shown that NSCLC cell lines that constitutively express higher levels of CXCL8 display greater virulence and angiogenic activity in mice 77, 93.

In addition to CXCL8, CXCL5 has been shown play an important role in angiogenesis, and is highly correlated with NSCLC-associated angiogenesis: there is a direct relationship between tissue levels of CXCL5 found in surgical specimens and the extent of capillary density consistent with tumor angiogenesis 94. Moreover, the expression of ELR-positive CXC chemokines in human NSCLC specimens correlate with clinical outcomes, including mortality 94, 95. Interestingly, while a significant correlation exists between CXCL5 and tumor-derived angiogenesis, tumor growth, and metastases, CXCL5 depletion does not completely inhibit tumor growth 96. This is thought to be due to functional redundancy between angiogenic ligands, and has been described in other disease settings 97.

In murine models of NSCLC, when the angiogenic CXC chemokines are neutralized angiogenic activity is decreased, and is followed by a reduction in tumor growth and metastases 23, 76, 96. In a syngeneic tumor model of lung cancer, CXCR2 knockout mice had reduced tumor growth, increased tumor-associated necrosis, and decreased tumor-associated angiogenesis and metastases compared to wildtype mice 23. In a different murine model, lung adenocarcinomas in mice with somatic activation of the oncogene KRAS were found to produce ELR-positive CXC chemokines, and again, neutralization of the CXCR2 receptor inhibited tumor development and apoptosis within the tumor 98.

The ELR-positive CXC chemokines have also been studied in human gastrointestinal cancers including pancreatic and colorectal malignancies. Human pancreatic cancer cell lines secrete the ELR-positive angiogenic CXC chemokines CXCL1 and CXCL8 99, but their expression differs across the different cell lines 97. When the different cancer cell lines were compared using the corneal micropocket model, tumor-induced angiogenesis was inhibited by blocking the receptor, CXCR2 in one cancer cell line, but not another; again supporting the concept of redundancy of angiogenic ligands, even within specific cancers. In colorectal cancer, in vivo tumor growth is also induced by increased expression of CXCL1 100.

Human ovarian and prostate cancers are highly dependent on successful angiogenesis for growth and metastatic potential. In one study of human ovarian cancer cell lines, in vitro expression of CXCL8 correlated with increased tumor neovascularization. Importantly, when the tumors were implanted into the peritoneum of immunocompromised mice, the mice had increased mortality rates 101. In this same study, the expression of VEGF correlated with ascites production, however, it was not associated with either the extent of angiogenesis or with mortality rates 101. Interestingly, in a separate study, the angiogenic potential of ascites fluid from patients with ovarian cancer was directly correlated with CXCL8 levels 102.

In human prostate cancer, tumorigenesis and metastases are dependent on angiogenesis 103, 104. Serum CXCL8 levels are markedly elevated in patients with prostate cancer, and these levels correlated with stage of the disease independent of prostate specific antigen levels 104, 105. In a SCID mouse model of human prostate cancer, different prostate cancer cell lines were found to use different ELR-positive CXC chemokine ligands: depletion of CXCL1 but not CXCL8 inhibited tumor-related angiogenesis in some cell lines, whereas the depletion of CXCL8 but not CXCL1 inhibited angiogenesis in other lines 106.

In patients with malignant melanoma, the angiogenic ELR-positive CXC chemokines, CXCL1, CXCL2 and CXCL3 are highly expressed 107. Sustained transgenic expression of CXCL1, CXCL2, and CXCL3 in immortalized murine melanocytes transformed their phenotype into one with the ability to form highly vascular tumors in immunocompetent mice 107, 108. Furthermore, in these same studies, depletion of CXCL1, CXCL2 or CXCL3 in vivo resulted in marked attenuation of tumor-associated angiogenesis and inhibition of tumor growth.

Glioblastoma multiforme tumors are also associated with marked angiogenesis 109, 110. While the mechanisms responsible with their increased growth and marked angiogenesis remain to be fully defined, a tumor suppressor gene appears to be important and is associated with the expression of angiogenic ELR-positive CXC chemokines in this disease. In one study, a candidate tumor suppressor gene was found to be down-regulated in human glioblastoma specimens compared with normal brain tissue. When implanted into immunocompromised mice, the specimens with the lowest expression of the tumor suppressor gene had the largest growth and degree of angiogenesis 110. The mechanism for this increased tumorigenicity was found to be CXCL8-dependent; inhibition of CXCL8 in vivo markedly reduced their tumor growth and tumor-associated angiogenesis.

Fibroproliferative disorders

The metabolic demands of proliferating tissue are higher than normal tissue and require an increased blood supply. Angiogenesis, therefore, plays a major role in the pathophysiology of such complex biologic processes as wound repair, chronic inflammatory and fibroproliferative diseases such as rheumatoid arthritis, psoriasis, idiopathic pulmonary fibrosis, bronchiolitis obliterans syndrome, acute respiratory distress syndrome, and atherosclerosis.

The ELR-positive CXC chemokines and their receptor, CXCR2 are important mediators of wound repair 111. CXCL1 and CXCR2 are expressed during wound healing by keratinocytes and endothelial cells in areas where epithelialization and neovascularization occur 112. In full-thickness excisional wounds, CXCR2 knockout mice demonstrated delayed healing; which was directly associated with impaired angiogenesis 113. Angiogenesis is critical to neovascularization of rheumatoid arthritic synovial tissue. In a model using whole human synovial tissue from patients with rheumatoid arthritis, the angiogenic ELR-positive CXC chemokines, CXCL8 and CXCL5 were potent mediators of the angiogenesis in the inflamed synovium compared to normal tissue 114. Psoriasis, a common inherited skin disease, is characterized by hyperproliferation of epidermal keratinocytes and excessive dermal angiogenesis. In a study of human psoriasis, media conditioned by keratinocytes from psoriatic patients induced a vigorous angiogenic response in the rat corneal micropocket model, and keratinocytes from psoriatic skin exhibited a 10- to 20-fold increase in CXCL8 production 115.

The role of chemokine mediated angiogenesis has been documented in several fibroproliferative lung diseases, including idiopathic pulmonary fibrosis (IPF), allograft bronchiolitis obliterans syndrome (BOS) and acute respiratory distress syndrome (ARDS). IPF is a chronic fibroproliferative lung disease characterized by progressive and disorganized tissue repair. Neovascularization was first recognized in the IPF lung in postmortem studies, and was described as extensive anastomoses between pulmonary and bronchial circulations 116, and was subsequently demonstrated in animal models of bleomycin-induced pulmonary fibrosis 117. Lung tissue and bronchoalveolar lavage fluid obtained from patients with IPF is strongly angiogenic secondary to over-expression of CXCL8, as compared to CXCL10 in the lung 118.

In a mouse model of bleomycin-induced pulmonary fibrosis, the expression and biological activity of chemokines have been studied 119, 120. In this model, the angiogenic ELR-positive CXC chemokine, CXCL2/3, was associated with increased pulmonary fibrosis and angiogenesis, while the angiostatic ELR negative CXC chemokine, CXCL10, had the opposite effect. Moreover, depletion of endogenous CXCL2/3, or administration of exogenous CXCL10, resulted in marked attenuation of lung fibrosis and a parallel reduction in angiogenesis. Finally, administration of exogenous CXCL11 in this model resulted in reduced lung fibrosis, as measured by lung collagen deposition, and this effect was abrogated with concomitant blockade of CXCR3 50. Bronchiolitis obliterans syndrome is a chronic fibroproliferative disorder of the lung and is the most common cause of death in lung transplant recipients 121. Human lung samples from patients with BOS demonstrate neovascularization, and lung tissue samples and bronchoalveolar lavage fluid have elevated levels of ELR-positive CXC chemokines. In the corneal micropocket model, there is an increase in angiogenesis, and it is inhibited by neutralizing the receptor, CXCR2 122. This CXCR2-dependent mechanism has been confirmed in a mouse model of heterotropic tracheal allograft transplantation 122. Acute respiratory distress syndrome (ARDS) is a severe manifestation of acute lung injury that quickly progresses to a fibroproliferative phase. Compared to ventilated patients without ARDS, ventilated patients with ARDS have elevated levels of angiogenic chemokines and reduced levels of angiostatic chemokines in bronchoalveolar lavage fluid samples 123.

Ischemic heart disease

Angiogenesis has also been implicated in the progression and instability of atherosclerotic plaques 124-128. The angiogenic ELR-positive CXC chemokine, CXCL8, has been shown to be over-expressed in human coronary artery plaque atherectomy specimens, as compared to control samples from internal mammary arteries without atherosclerosis 129. In this study, the CXCL8 co-localized with Factor VIII-related antigen expression on endothelial cells in the atherectomy specimens, and was the major mediator of plaque angiogenic activity in the rat cornea micropocket assay.

While a large number of chemokines are induced in the ischemic myocardium 130, 131, their specific contribution to angiogenesis has not been clearly established. The CXC chemokines CXCL8, CXCL10 and the CC chemokine, CCL2, are reproducibly upregulated in various animal models of myocardial ischemia. The ELR-negative CXC chemokine, CXCL10, is induced in canine 132 and murine 133 myocardial infarction models. It has been postulated that the upregulation of CXCL10 in the infarcted myocardium prevents early angiogenesis and granulation tissue formation until the wound is debrided and a fibrin-based matrix, necessary to support in-growth of tissue, is formed 134.

In a murine model of cardiomyopathy, interstitial fibrosis was preceded by a marked induction of the CC chemokine, CCL2 135. The wild type mice exhibited macrophage infiltration in the infarcted myocardium within days, and marked interstitial fibrosis within one week, which was accompanied by ventricular dysfunction. The CCL2 knockout mice, however, had markedly less interstitial fibrosis, less macrophage infiltration, and less ventricular dysfunction. In a different closed-chest model of reperfused murine myocardial infarction, CCL2 knockout mice had decreased and delayed macrophage infiltration into the infarcted tissue, and delayed replacement of myocytes with granulation tissue when compared to wild type mice 136. The absence of CCL2 resulted in attenuation of post-infarction left ventricular remodeling, a prolonged inflammatory phase, and delayed replacement of injured cardiomyocytes with granulation tissue. Interestingly, while CCL2 deficiency diminished myofibroblast accumulation, it did not significantly affect angiogenesis of the infarcted myocardium. Lastly, in a porcine microembolization model, CCL2 was upregulated and associated with a strong angiogenic response when compared to normal myocardium 137. CCL2 has also been implicated in ischemia-induced arteriogenesis in a murine hind-limb ischemia model 138. CCR2 knockout mice had reduced levels of arteriogenesis, and femoral artery occlusion lead to loss of structure and function. Moreover, local infusion of CCL2 into the proximal stump of the occluded femoral artery of rabbits has been shown to markedly increase the rate of arteriogenesis 139.

Conclusion

Chemokines were originally described for their role in recruitment of leukocytes, but have also been shown to play an integral biologic role in neovascularization in diverse human diseases, including cancers, fibroproliferative disorders, and ischemic heart disease. Chemokines may represent novel disease markers or therapeutic targets in these disorders.

Acknowledgments

Sources of funding: This work was supported by NIH grants HL73848 and an American Lung Association Career Investigator Award (Mehrad) and CA87879 and HL66027 (Strieter).

References

  • 1.Silvestre JS, Mallat Z, Tedgui A, Levy BI. Post-ischaemic neovascularization and inflammation. Cardiovasc Res. 2008;78:242–249. doi: 10.1093/cvr/cvn027. [DOI] [PubMed] [Google Scholar]
  • 2.Risau W, Flamme I. Vasculogenesis. Annu Rev Cell Dev Biol. 1995;11:73–91. doi: 10.1146/annurev.cb.11.110195.000445. [DOI] [PubMed] [Google Scholar]
  • 3.Asahara T, Masuda H, Takahashi T, Kalka C, Pastore C, Silver M, Kearne M, Magner M, Isner JM. Bone marrow origin of endothelial progenitor cells responsible for postnatal vasculogenesis in physiological and pathological neovascularization. Circ Res. 1999;85:221–228. doi: 10.1161/01.res.85.3.221. [DOI] [PubMed] [Google Scholar]
  • 4.Asahara T, Murohara T, Sullivan A, Silver M, van der Zee R, Li T, Witzenbichler B, Schatteman G, Isner JM. Isolation of putative progenitor endothelial cells for angiogenesis. Science. 1997;275:964–967. doi: 10.1126/science.275.5302.964. [DOI] [PubMed] [Google Scholar]
  • 5.Simons M, Ware JA. Therapeutic angiogenesis in cardiovascular disease. Nat Rev Drug Discov. 2003;2:863–871. doi: 10.1038/nrd1226. [DOI] [PubMed] [Google Scholar]
  • 6.Djonov V, Schmid M, Tschanz SA, Burri PH. Intussusceptive angiogenesis: its role in embryonic vascular network formation. Circ Res. 2000;86:286–292. doi: 10.1161/01.res.86.3.286. [DOI] [PubMed] [Google Scholar]
  • 7.Burri PH, Hlushchuk R, Djonov V. Intussusceptive angiogenesis: its emergence, its characteristics, and its significance. Dev Dyn. 2004;231:474–488. doi: 10.1002/dvdy.20184. [DOI] [PubMed] [Google Scholar]
  • 8.Madeddu P. Therapeutic angiogenesis and vasculogenesis for tissue regeneration. Exp Physiol. 2005;90:315–326. doi: 10.1113/expphysiol.2004.028571. [DOI] [PubMed] [Google Scholar]
  • 9.Schaper W, Scholz D. Factors regulating arteriogenesis. Arterioscler Thromb Vasc Biol. 2003;23:1143–1151. doi: 10.1161/01.ATV.0000069625.11230.96. [DOI] [PubMed] [Google Scholar]
  • 10.Helisch A, Schaper W. Arteriogenesis: the development and growth of collateral arteries. Microcirculation. 2003;10:83–97. doi: 10.1038/sj.mn.7800173. [DOI] [PubMed] [Google Scholar]
  • 11.Ware JA, Simons M. Angiogenesis in ischemic heart disease. Nat Med. 1997;3:158–164. doi: 10.1038/nm0297-158. [DOI] [PubMed] [Google Scholar]
  • 12.van Royen N, Piek JJ, Buschmann I, Hoefer I, Voskuil M, Schaper W. Stimulation of arteriogenesis; a new concept for the treatment of arterial occlusive disease. Cardiovasc Res. 2001;49:543–553. doi: 10.1016/s0008-6363(00)00206-6. [DOI] [PubMed] [Google Scholar]
  • 13.Belperio JA, Keane MP, Arenberg DA, Addison CL, Ehlert JE, Burdick MD, Strieter RM. CXC chemokines in angiogenesis. J Leukoc Biol. 2000;68:1–8. [PubMed] [Google Scholar]
  • 14.Strieter RM, Polverini PJ, Kunkel SL, Arenberg DA, Burdick MD, Kasper J, Dzuiba J, Van Damme J, Walz A, Marriott D, Chan S, Roczniak S, Shanafelt AB. The functional role of the ELR motif in CXC chemokine-mediated angiogenesis. J Biol Chem. 1995;270:27348–27357. doi: 10.1074/jbc.270.45.27348. [DOI] [PubMed] [Google Scholar]
  • 15.Chemokine/chemokine receptor nomenclature. Cytokine. 2003;21:48–49. doi: 10.1016/s1043-4666(02)00493-3. [DOI] [PubMed] [Google Scholar]
  • 16.Mehrad B, Keane MP, Strieter RM. Chemokines as mediators of angiogenesis. Thromb Haemost. 2007;97:755–762. [PMC free article] [PubMed] [Google Scholar]
  • 17.Addison CL, Daniel TO, Burdick MD, Liu H, Ehlert JE, Xue YY, Buechi L, Walz A, Richmond A, Strieter RM. The CXC chemokine receptor 2, CXCR2, is the putative receptor for ELR+ CXC chemokine-induced angiogenic activity. J Immunol. 2000;165:5269–5277. doi: 10.4049/jimmunol.165.9.5269. [DOI] [PubMed] [Google Scholar]
  • 18.Murdoch C, Monk PN, Finn A. Cxc chemokine receptor expression on human endothelial cells. Cytokine. 1999;11:704–712. doi: 10.1006/cyto.1998.0465. [DOI] [PubMed] [Google Scholar]
  • 19.Heidemann J, Ogawa H, Dwinell MB, Rafiee P, Maaser C, Gockel HR, Otterson MF, Ota DM, Lugering N, Domschke W, Binion DG. Angiogenic effects of interleukin 8 (CXCL8) in human intestinal microvascular endothelial cells are mediated by CXCR2. J Biol Chem. 2003;278:8508–8515. doi: 10.1074/jbc.M208231200. [DOI] [PubMed] [Google Scholar]
  • 20.Richmond A, Fan GH, Dhawan P, Yang J. How do chemokine/chemokine receptor activations affect tumorigenesis? 256Novartis Found Symp. 2004:74–89. discussion 89-91, 106-111, 266-109. [PubMed] [Google Scholar]
  • 21.Hristov M, Zernecke A, Bidzhekov K, Liehn EA, Shagdarsuren E, Ludwig A, Weber C. Importance of CXC chemokine receptor 2 in the homing of human peripheral blood endothelial progenitor cells to sites of arterial injury. Circ Res. 2007;100:590–597. doi: 10.1161/01.RES.0000259043.42571.68. [DOI] [PubMed] [Google Scholar]
  • 22.Kocher AA, Schuster MD, Bonaros N, Lietz K, Xiang G, Martens TP, Kurlansky PA, Sondermeijer H, Witkowski P, Boyle A, Homma S, Wang SF, Itescu S. Myocardial homing and neovascularization by human bone marrow angioblasts is regulated by IL-8/Gro CXC chemokines. J Mol Cell Cardiol. 2006;40:455–464. doi: 10.1016/j.yjmcc.2005.11.013. [DOI] [PubMed] [Google Scholar]
  • 23.Keane MP, Belperio JA, Xue YY, Burdick MD, Strieter RM. Depletion of CXCR2 inhibits tumor growth and angiogenesis in a murine model of lung cancer. J Immunol. 2004;172:2853–2860. doi: 10.4049/jimmunol.172.5.2853. [DOI] [PubMed] [Google Scholar]
  • 24.Addison CL, Belperio JA, Burdick MD, Strieter RM. Overexpression of the duffy antigen receptor for chemokines (DARC) by NSCLC tumor cells results in increased tumor necrosis. BMC Cancer. 2004;4:28. doi: 10.1186/1471-2407-4-28. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Du J, Luan J, Liu H, Daniel TO, Peiper S, Chen TS, Yu Y, Horton LW, Nanney LB, Strieter RM, Richmond A. Potential role for Duffy antigen chemokine-binding protein in angiogenesis and maintenance of homeostasis in response to stress. J Leukoc Biol. 2002;71:141–153. [PMC free article] [PubMed] [Google Scholar]
  • 26.Shen H, Schuster R, Stringer KF, Waltz SE, Lentsch AB. The Duffy antigen/receptor for chemokines (DARC) regulates prostate tumor growth. FASEB J. 2006;20:59–64. doi: 10.1096/fj.05-4764com. [DOI] [PubMed] [Google Scholar]
  • 27.Salcedo R, Young HA, Ponce ML, Ward JM, Kleinman HK, Murphy WJ, Oppenheim JJ. Eotaxin (CCL11) induces in vivo angiogenic responses by human CCR3+ endothelial cells. J Immunol. 2001;166:7571–7578. doi: 10.4049/jimmunol.166.12.7571. [DOI] [PubMed] [Google Scholar]
  • 28.Strasly M, Doronzo G, Capello P, Valdembri D, Arese M, Mitola S, Moore P, Alessandri G, Giovarelli M, Bussolino F. CCL16 activates an angiogenic program in vascular endothelial cells. Blood. 2004;103:40–49. doi: 10.1182/blood-2003-05-1387. [DOI] [PubMed] [Google Scholar]
  • 29.Stamatovic SM, Keep RF, Mostarica-Stojkovic M, Andjelkovic AV. CCL2 regulates angiogenesis via activation of Ets-1 transcription factor. J Immunol. 2006;177:2651–2661. doi: 10.4049/jimmunol.177.4.2651. [DOI] [PubMed] [Google Scholar]
  • 30.Galvez BG, Genis L, Matias-Roman S, Oblander SA, Tryggvason K, Apte SS, Arroyo AG. 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. 2005;280:1292–1298. doi: 10.1074/jbc.M408673200. [DOI] [PubMed] [Google Scholar]
  • 31.Weber KS, Nelson PJ, Grone HJ, Weber C. Expression of CCR2 by endothelial cells : implications for MCP-1 mediated wound injury repair and In vivo inflammatory activation of endothelium. Arterioscler Thromb Vasc Biol. 1999;19:2085–2093. doi: 10.1161/01.atv.19.9.2085. [DOI] [PubMed] [Google Scholar]
  • 32.Barcelos LS, Talvani A, Teixeira AS, Cassali GD, Andrade SP, Teixeira MM. Production and in vivo effects of chemokines CXCL1-3/KC and CCL2/JE in a model of inflammatory angiogenesis in mice. Inflamm Res. 2004;53:576–584. doi: 10.1007/s00011-004-1299-4. [DOI] [PubMed] [Google Scholar]
  • 33.Goede V, Brogelli L, Ziche M, Augustin HG. Induction of inflammatory angiogenesis by monocyte chemoattractant protein-1. Int J Cancer. 1999;82:765–770. doi: 10.1002/(sici)1097-0215(19990827)82:5<765::aid-ijc23>3.0.co;2-f. [DOI] [PubMed] [Google Scholar]
  • 34.Salcedo R, Ponce ML, Young HA, Wasserman K, Ward JM, Kleinman HK, Oppenheim JJ, Murphy WJ. Human endothelial cells express CCR2 and respond to MCP-1: direct role of MCP-1 in angiogenesis and tumor progression. Blood. 2000;96:34–40. [PubMed] [Google Scholar]
  • 35.Fujiyama S, Amano K, Uehira K, Yoshida M, Nishiwaki Y, Nozawa Y, Jin D, Takai S, Miyazaki M, Egashira K, Imada T, Iwasaka T, Matsubara H. Bone marrow monocyte lineage cells adhere on injured endothelium in a monocyte chemoattractant protein-1-dependent manner and accelerate reendothelialization as endothelial progenitor cells. Circ Res. 2003;93:980–989. doi: 10.1161/01.RES.0000099245.08637.CE. [DOI] [PubMed] [Google Scholar]
  • 36.Hong KH, Ryu J, Han KH. Monocyte chemoattractant protein-1-induced angiogenesis is mediated by vascular endothelial growth factor-A. Blood. 2005;105:1405–1407. doi: 10.1182/blood-2004-08-3178. [DOI] [PubMed] [Google Scholar]
  • 37.Niu J, Azfer A, Zhelyabovska O, Fatma S, Kolattukudy PE. Monocyte chemotactic protein (MCP)-1 promotes angiogenesis via a novel transcription factor, MCP-1-induced protein (MCPIP) J Biol Chem. 2008;283:14542–14551. doi: 10.1074/jbc.M802139200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Soto H, Wang W, Strieter RM, Copeland NG, Gilbert DJ, Jenkins NA, Hedrick J, Zlotnik A. The CC chemokine 6Ckine binds the CXC chemokine receptor CXCR3. Proc Natl Acad Sci U S A. 1998;95:8205–8210. doi: 10.1073/pnas.95.14.8205. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Romagnani P, Annunziato F, Lasagni L, Lazzeri E, Beltrame C, Francalanci M, Uguccioni M, Galli G, Cosmi L, Maurenzig L, Baggiolini M, Maggi E, Romagnani S, Serio M. Cell cycle-dependent expression of CXC chemokine receptor 3 by endothelial cells mediates angiostatic activity. J Clin Invest. 2001;107:53–63. doi: 10.1172/JCI9775. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Beider K, Nagler A, Wald O, Franitza S, Dagan-Berger M, Wald H, Giladi H, Brocke S, Hanna J, Mandelboim O, Darash-Yahana M, Galun E, Peled A. Involvement of CXCR4 and IL-2 in the homing and retention of human NK and NK T cells to the bone marrow and spleen of NOD/SCID mice. Blood. 2003;102:1951–1958. doi: 10.1182/blood-2002-10-3293. [DOI] [PubMed] [Google Scholar]
  • 41.Loetscher M, Gerber B, Loetscher P, Jones SA, Piali L, Clark-Lewis I, Baggiolini M, Moser B. Chemokine receptor specific for IP10 and mig: structure, function, and expression in activated T-lymphocytes. J Exp Med. 1996;184:963–969. doi: 10.1084/jem.184.3.963. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Loetscher M, Loetscher P, Brass N, Meese E, Moser B. Lymphocyte-specific chemokine receptor CXCR3: regulation, chemokine binding and gene localization. Eur J Immunol. 1998;28:3696–3705. doi: 10.1002/(SICI)1521-4141(199811)28:11<3696::AID-IMMU3696>3.0.CO;2-W. [DOI] [PubMed] [Google Scholar]
  • 43.Luster AD. Chemokines--chemotactic cytokines that mediate inflammation. N Engl J Med. 1998;338:436–445. doi: 10.1056/NEJM199802123380706. [DOI] [PubMed] [Google Scholar]
  • 44.Moser B, Loetscher P. Lymphocyte traffic control by chemokines. Nat Immunol. 2001;2:123–128. doi: 10.1038/84219. [DOI] [PubMed] [Google Scholar]
  • 45.Qin S, Rottman JB, Myers P, Kassam N, Weinblatt M, Loetscher M, Koch AE, Moser B, Mackay CR. The chemokine receptors CXCR3 and CCR5 mark subsets of T cells associated with certain inflammatory reactions. J Clin Invest. 1998;101:746–754. doi: 10.1172/JCI1422. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Rabin RL, Park MK, Liao F, Swofford R, Stephany D, Farber JM. Chemokine receptor responses on T cells are achieved through regulation of both receptor expression and signaling. J Immunol. 1999;162:3840–3850. [PubMed] [Google Scholar]
  • 47.Lasagni L, Francalanci M, Annunziato F, Lazzeri E, Giannini S, Cosmi L, Sagrinati C, Mazzinghi B, Orlando C, Maggi E, Marra F, Romagnani S, Serio M, Romagnani P. 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. 2003;197:1537–1549. doi: 10.1084/jem.20021897. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Salcedo R, Resau JH, Halverson D, Hudson EA, Dambach M, Powell D, Wasserman K, Oppenheim JJ. Differential expression and responsiveness of chemokine receptors (CXCR1-3) by human microvascular endothelial cells and umbilical vein endothelial cells. FASEB J. 2000;14:2055–2064. doi: 10.1096/fj.99-0963com. [DOI] [PubMed] [Google Scholar]
  • 49.Ehlert JE, Addison CA, Burdick MD, Kunkel SL, Strieter RM. Identification and partial characterization of a variant of human CXCR3 generated by posttranscriptional exon skipping. J Immunol. 2004;173:6234–6240. doi: 10.4049/jimmunol.173.10.6234. [DOI] [PubMed] [Google Scholar]
  • 50.Burdick MD, Murray LA, Keane MP, Xue YY, Zisman DA, Belperio JA, Strieter RM. CXCL11 attenuates bleomycin-induced pulmonary fibrosis via inhibition of vascular remodeling. Am J Respir Crit Care Med. 2005;171:261–268. doi: 10.1164/rccm.200409-1164OC. [DOI] [PubMed] [Google Scholar]
  • 51.Yang J, Richmond A. The angiostatic activity of interferon-inducible protein-10/CXCL10 in human melanoma depends on binding to CXCR3 but not to glycosaminoglycan. Mol Ther. 2004;9:846–855. doi: 10.1016/j.ymthe.2004.01.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Maione TE, Gray GS, Petro J, Hunt AJ, Donner AL, Bauer SI, Carson HF, Sharpe RJ. Inhibition of angiogenesis by recombinant human platelet factor-4 and related peptides. Science. 1990;247:77–79. doi: 10.1126/science.1688470. [DOI] [PubMed] [Google Scholar]
  • 53.Gupta SK, Singh JP. Inhibition of endothelial cell proliferation by platelet factor-4 involves a unique action on S phase progression. J Cell Biol. 1994;127:1121–1127. doi: 10.1083/jcb.127.4.1121. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Moser M. Regulation of Th1/Th2 development by antigen-presenting cells in vivo. Immunobiology. 2001;204:551–557. doi: 10.1078/0171-2985-00092. [DOI] [PubMed] [Google Scholar]
  • 55.Pan J, Burdick MD, Belperio JA, Xue YY, Gerard C, Sharma S, Dubinett SM, Strieter RM. CXCR3/CXCR3 ligand biological axis impairs RENCA tumor growth by a mechanism of immunoangiostasis. J Immunol. 2006;176:1456–1464. doi: 10.4049/jimmunol.176.3.1456. [DOI] [PubMed] [Google Scholar]
  • 56.Strieter RM, Belperio JA, Burdick MD, Sharma S, Dubinett SM, Keane MP. CXC chemokines: angiogenesis, immunoangiostasis, and metastases in lung cancer. Ann N Y Acad Sci. 2004;1028:351–360. doi: 10.1196/annals.1322.041. [DOI] [PubMed] [Google Scholar]
  • 57.Tannenbaum CS, Tubbs R, Armstrong D, Finke JH, Bukowski RM, Hamilton TA. The CXC chemokines IP-10 and Mig are necessary for IL-12-mediated regression of the mouse RENCA tumor. J Immunol. 1998;161:927–932. [PubMed] [Google Scholar]
  • 58.Sharma S, Stolina M, Luo J, Strieter RM, Burdick M, Zhu LX, Batra RK, Dubinett SM. Secondary lymphoid tissue chemokine mediates T cell-dependent antitumor responses in vivo. J Immunol. 2000;164:4558–4563. doi: 10.4049/jimmunol.164.9.4558. [DOI] [PubMed] [Google Scholar]
  • 59.Sharma S, Yang SC, Hillinger S, Zhu LX, Huang M, Batra RK, Lin JF, Burdick MD, Strieter RM, Dubinett SM. SLC/CCL21-mediated anti-tumor responses require IFNgamma, MIG/CXCL9 and IP-10/CXCL10. Mol Cancer. 2003;2:22. doi: 10.1186/1476-4598-2-22. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Bikfalvi A. Platelet factor 4: an inhibitor of angiogenesis. Semin Thromb Hemost. 2004;30:379–385. doi: 10.1055/s-2004-831051. [DOI] [PubMed] [Google Scholar]
  • 61.Bikfalvi A, Gimenez-Gallego G. The control of angiogenesis and tumor invasion by platelet factor-4 and platelet factor-4-derived molecules. Semin Thromb Hemost. 2004;30:137–144. doi: 10.1055/s-2004-822978. [DOI] [PubMed] [Google Scholar]
  • 62.Perollet C, Han ZC, Savona C, Caen JP, Bikfalvi A. Platelet factor 4 modulates fibroblast growth factor 2 (FGF-2) activity and inhibits FGF-2 dimerization. Blood. 1998;91:3289–3299. [PubMed] [Google Scholar]
  • 63.Arenberg DA, Kunkel SL, Polverini PJ, Morris SB, Burdick MD, Glass MC, Taub DT, Iannettoni MD, Whyte RI, Strieter RM. Interferon-gamma-inducible protein 10 (IP-10) is an angiostatic factor that inhibits human non-small cell lung cancer (NSCLC) tumorigenesis and spontaneous metastases. J Exp Med. 1996;184:981–992. doi: 10.1084/jem.184.3.981. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Dudek AZ, Nesmelova I, Mayo K, Verfaillie CM, Pitchford S, Slungaard A. Platelet factor 4 promotes adhesion of hematopoietic progenitor cells and binds IL-8: novel mechanisms for modulation of hematopoiesis. Blood. 2003;101:4687–4694. doi: 10.1182/blood-2002-08-2363. [DOI] [PubMed] [Google Scholar]
  • 65.Sulpice E, Bryckaert M, Lacour J, Contreres JO, Tobelem G. Platelet factor 4 inhibits FGF2-induced endothelial cell proliferation via the extracellular signal-regulated kinase pathway but not by the phosphatidylinositol 3-kinase pathway. Blood. 2002;100:3087–3094. doi: 10.1182/blood.V100.9.3087. [DOI] [PubMed] [Google Scholar]
  • 66.Jouan V, Canron X, Alemany M, Caen JP, Quentin G, Plouet J, Bikfalvi A. Inhibition of in vitro angiogenesis by platelet factor-4-derived peptides and mechanism of action. Blood. 1999;94:984–993. [PubMed] [Google Scholar]
  • 67.Gengrinovitch S, Greenberg SM, Cohen T, Gitay-Goren H, Rockwell P, Maione TE, Levi BZ, Neufeld G. Platelet factor-4 inhibits the mitogenic activity of VEGF121 and VEGF165 using several concurrent mechanisms. J Biol Chem. 1995;270:15059–15065. doi: 10.1074/jbc.270.25.15059. [DOI] [PubMed] [Google Scholar]
  • 68.Houck KA, Ferrara N, Winer J, Cachianes G, Li B, Leung DW. The vascular endothelial growth factor family: identification of a fourth molecular species and characterization of alternative splicing of RNA. Mol Endocrinol. 1991;5:1806–1814. doi: 10.1210/mend-5-12-1806. [DOI] [PubMed] [Google Scholar]
  • 69.Houck KA, Leung DW, Rowland AM, Winer J, Ferrara N. Dual regulation of vascular endothelial growth factor bioavailability by genetic and proteolytic mechanisms. J Biol Chem. 1992;267:26031–26037. [PubMed] [Google Scholar]
  • 70.Loscalzo J, Melnick B, Handin RI. The interaction of platelet factor four and glycosaminoglycans. Arch Biochem Biophys. 1985;240:446–455. doi: 10.1016/0003-9861(85)90049-9. [DOI] [PubMed] [Google Scholar]
  • 71.Struyf S, Burdick MD, Proost P, Van Damme J, Strieter RM. Platelets release CXCL4L1, a nonallelic variant of the chemokine platelet factor-4/CXCL4 and potent inhibitor of angiogenesis. Circ Res. 2004;95:855–857. doi: 10.1161/01.RES.0000146674.38319.07. [DOI] [PubMed] [Google Scholar]
  • 72.Struyf S, Burdick MD, Peeters E, Van den Broeck K, Dillen C, Proost P, Van Damme J, Strieter RM. Platelet factor-4 variant chemokine CXCL4L1 inhibits melanoma and lung carcinoma growth and metastasis by preventing angiogenesis. Cancer Res. 2007;67:5940–5948. doi: 10.1158/0008-5472.CAN-06-4682. [DOI] [PubMed] [Google Scholar]
  • 73.Salcedo R, Wasserman K, Young HA, Grimm MC, Howard OM, Anver MR, Kleinman HK, Murphy WJ, Oppenheim JJ. Vascular endothelial growth factor and basic fibroblast growth factor induce expression of CXCR4 on human endothelial cells: In vivo neovascularization induced by stromal-derived factor-1alpha. Am J Pathol. 1999;154:1125–1135. doi: 10.1016/s0002-9440(10)65365-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Muller A, Homey B, Soto H, Ge N, Catron D, Buchanan ME, McClanahan T, Murphy E, Yuan W, Wagner SN, Barrera JL, Mohar A, Verastegui E, Zlotnik A. Involvement of chemokine receptors in breast cancer metastasis. Nature. 2001;410:50–56. doi: 10.1038/35065016. [DOI] [PubMed] [Google Scholar]
  • 75.Phillips RJ, Burdick MD, Lutz M, Belperio JA, Keane MP, Strieter RM. The stromal derived factor-1/CXCL12-CXC chemokine receptor 4 biological axis in non-small cell lung cancer metastases. Am J Respir Crit Care Med. 2003;167:1676–1686. doi: 10.1164/rccm.200301-071OC. [DOI] [PubMed] [Google Scholar]
  • 76.Arenberg DA, Kunkel SL, Polverini PJ, Glass M, Burdick MD, Strieter RM. Inhibition of interleukin-8 reduces tumorigenesis of human non-small cell lung cancer in SCID mice. J Clin Invest. 1996;97:2792–2802. doi: 10.1172/JCI118734. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Smith DR, Polverini PJ, Kunkel SL, Orringer MB, Whyte RI, Burdick MD, Wilke CA, Strieter RM. Inhibition of interleukin 8 attenuates angiogenesis in bronchogenic carcinoma. J Exp Med. 1994;179:1409–1415. doi: 10.1084/jem.179.5.1409. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Ceradini DJ, Kulkarni AR, Callaghan MJ, Tepper OM, Bastidas N, Kleinman ME, Capla JM, Galiano RD, Levine JP, Gurtner GC. Progenitor cell trafficking is regulated by hypoxic gradients through HIF-1 induction of SDF-1. Nat Med. 2004;10:858–864. doi: 10.1038/nm1075. [DOI] [PubMed] [Google Scholar]
  • 79.Chen L, Tredget EE, Wu PY, Wu Y. Paracrine factors of mesenchymal stem cells recruit macrophages and endothelial lineage cells and enhance wound healing. PLoS ONE. 2008;3:e1886. doi: 10.1371/journal.pone.0001886. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.De Falco E, Porcelli D, Torella AR, Straino S, Iachininoto MG, Orlandi A, Truffa S, Biglioli P, Napolitano M, Capogrossi MC, Pesce M. SDF-1 involvement in endothelial phenotype and ischemia-induced recruitment of bone marrow progenitor cells. Blood. 2004;104:3472–3482. doi: 10.1182/blood-2003-12-4423. [DOI] [PubMed] [Google Scholar]
  • 81.Stellos K, Langer H, Daub K, Schoenberger T, Gauss A, Geisler T, Bigalke B, Mueller I, Schumm M, Schaefer I, Seizer P, Kraemer BF, Siegel-Axel D, May AE, Lindemann S, Gawaz M. Platelet-derived stromal cell-derived factor-1 regulates adhesion and promotes differentiation of human CD34+ cells to endothelial progenitor cells. Circulation. 2008;117:206–215. doi: 10.1161/CIRCULATIONAHA.107.714691. [DOI] [PubMed] [Google Scholar]
  • 82.Yamaguchi J, Kusano KF, Masuo O, Kawamoto A, Silver M, Murasawa S, Bosch-Marce M, Masuda H, Losordo DW, Isner JM, Asahara T. Stromal cell-derived factor-1 effects on ex vivo expanded endothelial progenitor cell recruitment for ischemic neovascularization. Circulation. 2003;107:1322–1328. doi: 10.1161/01.cir.0000055313.77510.22. [DOI] [PubMed] [Google Scholar]
  • 83.Guerin E, Sheridan C, Assheton D, Kent D, Wong D, Grant M, Hiscott P. SDF1-alpha is associated with VEGFR-2 in human choroidal neovascularisation. Microvasc Res. 2008;75:302–307. doi: 10.1016/j.mvr.2007.12.001. [DOI] [PubMed] [Google Scholar]
  • 84.Grunewald M, Avraham I, Dor Y, Bachar-Lustig E, Itin A, Jung S, Chimenti S, Landsman L, Abramovitch R, Keshet E. VEGF-induced adult neovascularization: recruitment, retention, and role of accessory cells. Cell. 2006;124:175–189. doi: 10.1016/j.cell.2005.10.036. [DOI] [PubMed] [Google Scholar]
  • 85.Kerbel RS. Tumor angiogenesis. N Engl J Med. 2008;358:2039–2049. doi: 10.1056/NEJMra0706596. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Chen Z, Malhotra PS, Thomas GR, Ondrey FG, Duffey DC, Smith CW, Enamorado I, Yeh NT, Kroog GS, Rudy S, McCullagh L, Mousa S, Quezado M, Herscher LL, Van Waes C. Expression of proinflammatory and proangiogenic cytokines in patients with head and neck cancer. Clin Cancer Res. 1999;5:1369–1379. In Process Citation. [PubMed] [Google Scholar]
  • 87.Cohen RF, Contrino J, Spiro JD, Mann EA, Chen LL, Kreutzer DL. Interleukin-8 expression by head and neck squamous cell carcinoma. Arch Otolaryngol Head Neck Surg. 1995;121:202–209. doi: 10.1001/archotol.1995.01890020064013. [DOI] [PubMed] [Google Scholar]
  • 88.Kitadai Y, Haruma K, Sumii K, Yamamoto S, Ue T, Yokozaki H, Yasui W, Ohmoto Y, Kajiyama G, Fidler IJ, Tahara E. Expression of interleukin-8 correlates with vascularity in human gastric carcinomas. Am J Pathol. 1998;152:93–100. [PMC free article] [PubMed] [Google Scholar]
  • 89.Mestas J, Burdick MD, Reckamp K, Pantuck A, Figlin RA, Strieter RM. The role of CXCR2/CXCR2 ligand biological axis in renal cell carcinoma. J Immunol. 2005;175:5351–5357. doi: 10.4049/jimmunol.175.8.5351. [DOI] [PubMed] [Google Scholar]
  • 90.Miller LJ, Kurtzman SH, Wang Y, Anderson KH, Lindquist RR, Kreutzer DL. Expression of interleukin-8 receptors on tumor cells and vascular endothelial cells in human breast cancer tissue. Anticancer Res. 1998;18:77–81. [PubMed] [Google Scholar]
  • 91.Richards BL, Eisma RJ, Spiro JD, Lindquist RL, Kreutzer DL. Coexpression of interleukin-8 receptors in head and neck squamous cell carcinoma. Am J Surg. 1997;174:507–512. doi: 10.1016/s0002-9610(97)00165-7. [DOI] [PubMed] [Google Scholar]
  • 92.Singh RK, Gutman M, Radinsky R, Bucana CD, Fidler IJ. Expression of interleukin 8 correlates with the metastatic potential of human melanoma cells in nude mice. Cancer Res. 1994;54:3242–3247. [PubMed] [Google Scholar]
  • 93.Yatsunami J, Tsuruta N, Ogata K, Wakamatsu K, Takayama K, Kawasaki M, Nakanishi Y, Hara N, Hayashi S. Interleukin-8 participates in angiogenesis in non-small cell, but not small cell carcinoma of the lung. Cancer Lett. 1997;120:101–108. doi: 10.1016/s0304-3835(97)00296-6. [DOI] [PubMed] [Google Scholar]
  • 94.White ES, Flaherty KR, Carskadon S, Brant A, Iannettoni MD, Yee J, Orringer MB, Arenberg DA. Macrophage migration inhibitory factor and CXC chemokine expression in non-small cell lung cancer: role in angiogenesis and prognosis. Clin Cancer Res. 2003;9:853–860. [PubMed] [Google Scholar]
  • 95.Chen JJ, Yao PL, Yuan A, Hong TM, Shun CT, Kuo ML, Lee YC, Yang PC. Up-regulation of tumor interleukin-8 expression by infiltrating macrophages: its correlation with tumor angiogenesis and patient survival in non-small cell lung cancer. Clin Cancer Res. 2003;9:729–737. [PubMed] [Google Scholar]
  • 96.Arenberg DA, Keane MP, DiGiovine B, Kunkel SL, Morris SB, Xue YY, Burdick MD, Glass MC, Iannettoni MD, Strieter RM. Epithelial-neutrophil activating peptide (ENA-78) is an important angiogenic factor in non-small cell lung cancer. J Clin Invest. 1998;102:465–472. doi: 10.1172/JCI3145. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Wente MN, Keane MP, Burdick MD, Friess H, Buchler MW, Ceyhan GO, Reber HA, Strieter RM, Hines OJ. Blockade of the chemokine receptor CXCR2 inhibits pancreatic cancer cell-induced angiogenesis. Cancer Lett. 2006;241:221–227. doi: 10.1016/j.canlet.2005.10.041. [DOI] [PubMed] [Google Scholar]
  • 98.Wislez M, Fujimoto N, Izzo JG, Hanna AE, Cody DD, Langley RR, Tang H, Burdick MD, Sato M, Minna JD, Mao L, Wistuba I, Strieter RM, Kurie JM. High expression of ligands for chemokine receptor CXCR2 in alveolar epithelial neoplasia induced by oncogenic kras. Cancer Res. 2006;66:4198–4207. doi: 10.1158/0008-5472.CAN-05-3842. [DOI] [PubMed] [Google Scholar]
  • 99.Takamori H, Oades ZG, Hoch OC, Burger M, Schraufstatter IU. Autocrine growth effect of IL-8 and GROalpha on a human pancreatic cancer cell line, Capan-1. Pancreas. 2000;21:52–56. doi: 10.1097/00006676-200007000-00051. [DOI] [PubMed] [Google Scholar]
  • 100.Wang D, Wang H, Brown J, Daikoku T, Ning W, Shi Q, Richmond A, Strieter R, Dey SK, DuBois RN. CXCL1 induced by prostaglandin E2 promotes angiogenesis in colorectal cancer. J Exp Med. 2006;203:941–951. doi: 10.1084/jem.20052124. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Yoneda J, Kuniyasu H, Crispens MA, Price JE, Bucana CD, Fidler IJ. Expression of angiogenesis-related genes and progression of human ovarian carcinomas in nude mice. J Natl Cancer Inst. 1998;90:447–454. doi: 10.1093/jnci/90.6.447. [DOI] [PubMed] [Google Scholar]
  • 102.Gawrychowski K, Skopinska-Rozewska E, Barcz E, Sommer E, Szaniawska B, Roszkowska-Purska K, Janik P, Zielinski J. Angiogenic activity and interleukin-8 content of human ovarian cancer ascites. Eur J Gynaecol Oncol. 1998;19:262–264. [PubMed] [Google Scholar]
  • 103.Bostwick DG, Iczkowski KA. Microvessel density in prostate cancer: prognostic and therapeutic utility. Semin Urol Oncol. 1998;16:118–123. [PubMed] [Google Scholar]
  • 104.Fregene TA, Khanuja PS, Noto AC, Gehani SK, Van Egmont EM, Luz DA, Pienta KJ. Tumor-associated angiogenesis in prostate cancer. Anticancer Res. 1993;13:2377–2381. [PubMed] [Google Scholar]
  • 105.Kim SJ, Uehara H, Karashima T, McCarty M, Shih N, Fidler IJ. Expression of interleukin-8 correlates with angiogenesis, tumorigenicity, and metastasis of human prostate cancer cells implanted orthotopically in nude mice. Neoplasia. 2001;3:33–42. doi: 10.1038/sj.neo.7900124. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Moore BB, Arenberg DA, Stoy K, Morgan T, Addison CL, Morris SB, Glass M, Wilke C, Xue YY, Sitterding S, Kunkel SL, Burdick MD, Strieter RM. Distinct CXC chemokines mediate tumorigenicity of prostate cancer cells. Am J Pathol. 1999;154:1503–1512. doi: 10.1016/S0002-9440(10)65404-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Luan J, Shattuck-Brandt R, Haghnegahdar H, Owen JD, Strieter R, Burdick M, Nirodi C, Beauchamp D, Johnson KN, Richmond A. Mechanism and biological significance of constitutive expression of MGSA/GRO chemokines in malignant melanoma tumor progression. J Leukoc Biol. 1997;62:588–597. doi: 10.1002/jlb.62.5.588. [DOI] [PubMed] [Google Scholar]
  • 108.Owen JD, Strieter R, Burdick M, Haghnegahdar H, Nanney L, Shattuck-Brandt R, Richmond A. Enhanced tumor-forming capacity for immortalized melanocytes expressing melanoma growth stimulatory activity/growth-regulated cytokine beta and gamma proteins. Int J Cancer. 1997;73:94–103. doi: 10.1002/(sici)1097-0215(19970926)73:1<94::aid-ijc15>3.0.co;2-5. [DOI] [PubMed] [Google Scholar]
  • 109.Charalambous C, Chen TC, Hofman FM. Characteristics of tumor-associated endothelial cells derived from glioblastoma multiforme. Neurosurg Focus. 2006;20:E22. doi: 10.3171/foc.2006.20.4.e22. [DOI] [PubMed] [Google Scholar]
  • 110.Garkavtsev I, Kozin SV, Chernova O, Xu L, Winkler F, Brown E, Barnett GH, Jain RK. The candidate tumour suppressor protein ING4 regulates brain tumour growth and angiogenesis. Nature. 2004;428:328–332. doi: 10.1038/nature02329. [DOI] [PubMed] [Google Scholar]
  • 111.Gillitzer R, Goebeler M. Chemokines in cutaneous wound healing. J Leukoc Biol. 2001;69:513–521. [PubMed] [Google Scholar]
  • 112.Nanney LB, Mueller SG, Bueno R, Peiper SC, Richmond A. Distributions of melanoma growth stimulatory activity of growth-regulated gene and the interleukin-8 receptor B in human wound repair. Am J Pathol. 1995;147:1248–1260. [PMC free article] [PubMed] [Google Scholar]
  • 113.Devalaraja RM, Nanney LB, Du J, Qian Q, Yu Y, Devalaraja MN, Richmond A. Delayed wound healing in CXCR2 knockout mice. J Invest Dermatol. 2000;115:234–244. doi: 10.1046/j.1523-1747.2000.00034.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Koch AE, Volin MV, Woods JM, Kunkel SL, Connors MA, Harlow LA, Woodruff DC, Burdick MD, Strieter RM. Regulation of angiogenesis by the C-X-C chemokines interleukin-8 and epithelial neutrophil activating peptide 78 in the rheumatoid joint. Arthritis Rheum. 2001;44:31–40. doi: 10.1002/1529-0131(200101)44:1<31::AID-ANR5>3.0.CO;2-4. [DOI] [PubMed] [Google Scholar]
  • 115.Nickoloff BJ, Mitra RS, Varani J, Dixit VM, Polverini PJ. Aberrant production of interleukin-8 and thrombospondin-1 by psoriatic keratinocytes mediates angiogenesis. Am J Pathol. 1994;144:820–828. [PMC free article] [PubMed] [Google Scholar]
  • 116.Turner-Warwick M. Precapillary Systemic-Pulmonary Anastomoses. Thorax. 1963;18:225–237. doi: 10.1136/thx.18.3.225. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Peao MN, Aguas AP, de Sa CM, Grande NR. Neoformation of blood vessels in association with rat lung fibrosis induced by bleomycin. Anat Rec. 1994;238:57–67. doi: 10.1002/ar.1092380108. [DOI] [PubMed] [Google Scholar]
  • 118.Keane MP, Arenberg DA, Lynch JP, 3rd, Whyte RI, Iannettoni MD, Burdick MD, Wilke CA, Morris SB, Glass MC, DiGiovine B, Kunkel SL, Strieter RM. The CXC chemokines, IL-8 and IP-10, regulate angiogenic activity in idiopathic pulmonary fibrosis. J Immunol. 1997;159:1437–1443. [PubMed] [Google Scholar]
  • 119.Keane MP, Belperio JA, Arenberg DA, Burdick MD, Xu ZJ, Xue YY, Strieter RM. IFN-gamma-inducible protein-10 attenuates bleomycin-induced pulmonary fibrosis via inhibition of angiogenesis. J Immunol. 1999;163:5686–5692. [PubMed] [Google Scholar]
  • 120.Keane MP, Belperio JA, Moore TA, Moore BB, Arenberg DA, Smith RE, Burdick MD, Kunkel SL, Strieter RM. Neutralization of the CXC chemokine, macrophage inflammatory protein-2, attenuates bleomycin-induced pulmonary fibrosis. J Immunol. 1999;162:5511–5518. [PubMed] [Google Scholar]
  • 121.Wilkes DS, Egan TM, Reynolds HY. Lung transplantation: opportunities for research and clinical advancement. Am J Respir Crit Care Med. 2005;172:944–955. doi: 10.1164/rccm.200501-098WS. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Belperio JA, Keane MP, Burdick MD, Gomperts B, Xue YY, Hong K, Mestas J, Ardehali A, Mehrad B, Saggar R, Lynch JP, Ross DJ, Strieter RM. Role of CXCR2/CXCR2 ligands in vascular remodeling during bronchiolitis obliterans syndrome. J Clin Invest. 2005;115:1150–1162. doi: 10.1172/JCI24233. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Keane MP, Donnelly SC, Belperio JA, Goodman RB, Dy M, Burdick MD, Fishbein MC, Strieter RM. Imbalance in the expression of CXC chemokines correlates with bronchoalveolar lavage fluid angiogenic activity and procollagen levels in acute respiratory distress syndrome. J Immunol. 2002;169:6515–6521. doi: 10.4049/jimmunol.169.11.6515. [DOI] [PubMed] [Google Scholar]
  • 124.Fowler S, Berberian PA, Shio H, Goldfischer S, Wolinsky H. Characterization of cell populations isolated from aortas of rhesus monkeys with experimental atherosclerosis. Circ Res. 1980;46:520–530. doi: 10.1161/01.res.46.4.520. [DOI] [PubMed] [Google Scholar]
  • 125.Joris I, Zand T, Nunnari JJ, Krolikowski FJ, Majno G. Studies on the pathogenesis of atherosclerosis. I. Adhesion and emigration of mononuclear cells in the aorta of hypercholesterolemic rats. Am J Pathol. 1983;113:341–358. [PMC free article] [PubMed] [Google Scholar]
  • 126.Munro JM, Cotran RS. The pathogenesis of atherosclerosis: atherogenesis and inflammation. Lab Invest. 1988;58:249–261. [PubMed] [Google Scholar]
  • 127.Sluimer JC, Gasc JM, van Wanroij JL, Kisters N, Groeneweg M, Sollewijn Gelpke MD, Cleutjens JP, van den Akker LH, Corvol P, Wouters BG, Daemen MJ, Bijnens AP. Hypoxia, hypoxia-inducible transcription factor, and macrophages in human atherosclerotic plaques are correlated with intraplaque angiogenesis. J Am Coll Cardiol. 2008;51:1258–1265. doi: 10.1016/j.jacc.2007.12.025. [DOI] [PubMed] [Google Scholar]
  • 128.Virmani R, Kolodgie FD, Burke AP, Finn AV, Gold HK, Tulenko TN, Wrenn SP, Narula J. Atherosclerotic plaque progression and vulnerability to rupture: angiogenesis as a source of intraplaque hemorrhage. Arterioscler Thromb Vasc Biol. 2005;25:2054–2061. doi: 10.1161/01.ATV.0000178991.71605.18. [DOI] [PubMed] [Google Scholar]
  • 129.Simonini A, Moscucci M, Muller DW, Bates ER, Pagani FD, Burdick MD, Strieter RM. IL-8 is an angiogenic factor in human coronary atherectomy tissue. Circulation. 2000;101:1519–1526. doi: 10.1161/01.cir.101.13.1519. [DOI] [PubMed] [Google Scholar]
  • 130.Damas JK, Eiken HG, Oie E, Bjerkeli V, Yndestad A, Ueland T, Tonnessen T, Geiran OR, Aass H, Simonsen S, Christensen G, Froland SS, Attramadal H, Gullestad L, Aukrust P. Myocardial expression of CC- and CXC-chemokines and their receptors in human end-stage heart failure. Cardiovasc Res. 2000;47:778–787. doi: 10.1016/s0008-6363(00)00142-5. [DOI] [PubMed] [Google Scholar]
  • 131.Lakshminarayanan V, Lewallen M, Frangogiannis NG, Evans AJ, Wedin KE, Michael LH, Entman ML. Reactive oxygen intermediates induce monocyte chemotactic protein-1 in vascular endothelium after brief ischemia. Am J Pathol. 2001;159:1301–1311. doi: 10.1016/S0002-9440(10)62517-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Frangogiannis NG, Mendoza LH, Lewallen M, Michael LH, Smith CW, Entman ML. Induction and suppression of interferon-inducible protein 10 in reperfused myocardial infarcts may regulate angiogenesis. FASEB J. 2001;15:1428–1430. doi: 10.1096/fj.00-0745fje. [DOI] [PubMed] [Google Scholar]
  • 133.Dewald O, Ren G, Duerr GD, Zoerlein M, Klemm C, Gersch C, Tincey S, Michael LH, Entman ML, Frangogiannis NG. Of mice and dogs: species-specific differences in the inflammatory response following myocardial infarction. Am J Pathol. 2004;164:665–677. doi: 10.1016/S0002-9440(10)63154-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Frangogiannis NG, Entman ML. Chemokines in myocardial ischemia. Trends Cardiovasc Med. 2005;15:163–169. doi: 10.1016/j.tcm.2005.06.005. [DOI] [PubMed] [Google Scholar]
  • 135.Frangogiannis NG, Dewald O, Xia Y, Ren G, Haudek S, Leucker T, Kraemer D, Taffet G, Rollins BJ, Entman ML. Critical role of monocyte chemoattractant protein-1/CC chemokine ligand 2 in the pathogenesis of ischemic cardiomyopathy. Circulation. 2007;115:584–592. doi: 10.1161/CIRCULATIONAHA.106.646091. [DOI] [PubMed] [Google Scholar]
  • 136.Dewald O, Zymek P, Winkelmann K, Koerting A, Ren G, Abou-Khamis T, Michael LH, Rollins BJ, Entman ML, Frangogiannis NG. CCL2/Monocyte Chemoattractant Protein-1 regulates inflammatory responses critical to healing myocardial infarcts. Circ Res. 2005;96:881–889. doi: 10.1161/01.RES.0000163017.13772.3a. [DOI] [PubMed] [Google Scholar]
  • 137.Zimmermann R, Arras M, Ullmann C, Strasser R, Sack S, Mollnau H, Schaper J, Schaper W. Time course of mitosis and collateral growth following coronary microembolization in the porcine heart. Cell Tissue Res. 1997;287:583–590. doi: 10.1007/s004410050781. [DOI] [PubMed] [Google Scholar]
  • 138.Heil M, Ziegelhoeffer T, Wagner S, Fernandez B, Helisch A, Martin S, Tribulova S, Kuziel WA, Bachmann G, Schaper W. Collateral artery growth (arteriogenesis) after experimental arterial occlusion is impaired in mice lacking CC-chemokine receptor-2. Circ Res. 2004;94:671–677. doi: 10.1161/01.RES.0000122041.73808.B5. [DOI] [PubMed] [Google Scholar]
  • 139.Ito WD, Arras M, Winkler B, Scholz D, Schaper J, Schaper W. Monocyte chemotactic protein-1 increases collateral and peripheral conductance after femoral artery occlusion. Circ Res. 1997;80:829–837. doi: 10.1161/01.res.80.6.829. [DOI] [PubMed] [Google Scholar]

RESOURCES