Skip to main content
Access keys NCBI Homepage MyNCBI Homepage Main Content Main Navigation
Neoplasia. 2009 Jul; 11(7): 615–628.
PMCID: PMC2697348
PMID: 19568407

Endogenous Damage-Associated Molecular Pattern Molecules at the Crossroads of Inflammation and Cancer1

Abstract

Inflammatory mediators play important roles in the development and progression of cancer. Cellular stress, damage, inflammation, and necrotic cell death cause release of endogenous damage-associated molecular pattern (DAMP) molecules or alarmins, which alert the host of danger by triggering immune responses and activating repair mechanisms through their interaction with pattern recognition receptors. Recent studies show that abnormal persistence of these molecules in chronic inflammation and in tumor microenvironments underlies carcinogenesis and tumor progression, indicating that DAMP molecules and their receptors could provide novel targets for therapy. This review focuses on the role of DAMP molecules high-mobility group box 1 and S100 proteins in inflammation, tumor growth, and early metastatic events.

Introduction

A century and a half ago, German physician and pathologist Rudolph Virchow first reported that infectious diseases showed signs of a “tumor process” and that inflammatory cells were frequently present in tumor biopsies [1]. Noting similarities between wound healing and tumor stromal generation, Harold Dvorak referred to tumors as “wounds that do not heal” [2]. Although the association was largely ignored for many years, increasing evidence linking inflammation and tumorigenesis has triggered renewed interest in understanding the molecular and cellular mechanisms involved in cancer-related inflammation. Emerging information points to two different pathways linking the pathologies [3]: an extrinsic pathway mediated by chronic inflammation that increases the risk of tumorigenesis (inflammation-induced cancer) and an intrinsic pathway in which genetic alterations, in the absence of an underlying inflammation, initiate a tumor-driven host immune response leading to a tumor microenvironment composed of inflammatory cells (cancer-induced inflammation).

Prolonged subclinical inflammation and associated necrotic cell death also cause release of intracellular molecules that alert the immune system to danger, evoking responses leading to epithelial regeneration, angiogenesis, proliferation, and ultimately tumorigenesis. These events resemble normal injury-related stromal reactions and wound healing processes, which the tumor cells co-opt for growth, and further subvert to counteract normal regulatory immune responses. Recent studies show that these endogenous damage-associated molecular pattern (DAMP) molecules or alarmins, released from necrotic cells and activated leukocytes, play critical roles in both extrinsic and intrinsic pathways of cancer-related inflammation. These studies prompt a paradigm shift in understanding tumor development in adults as a pathological sequence initiated by cycles of inflammation and necrotic cell death [4,5]. In this review, we focus on the contribution of DAMP molecules high-mobility group box 1 (HMGB1) and proinflammatory S100 proteins to inflammation and cancer.

Inflammation and Cancer

Inflammation-Induced Cancer

It is estimated that infections and chronic inflammatory responses are involved in the pathogenesis of approximately 15% to 20% of human tumors, including gastric, colorectal, bladder, and liver cancers [6–10]. Other causes of chronic inflammation, including mechanical, physical, and chemical injury as well as dysregulated immune responses to injury, also predispose to cancers [10,11]. Agents modulating inflammation, including aspirin and nonsteroidal anti-inflammatory drugs, decrease the incidence of cancers [12–15], providing strong evidence for the extrinsic pathway of inflammation-based cancers. In addition, experimental animal models such as the 7,12-dimethyl benz[a]anthracene/12-O-tetradecanoylphorbol-13-acetate (DMBA/TPA)-induced papillomas and azoxymethane/dextran sulfate sodium (DSS)-induced colon cancers offer valuable insights into the initiation and progression of inflammation-based cancers [16–19].

Inflammation is an immune response to infection and tissue injury, characterized by the release of a complex regulatory network of mediators all aimed at combating the infectious or noxious agent, repairing damaged tissue, and restoring homeostasis [20]. In chronic inflammation, this response is exaggerated or sustained. Long-term inflammation is thought to lead to cancer because of the dysplastic degeneration of repaired epithelium by the continuous release of reactive oxygen and nitrogen species, which cause DNA damage resulting in genomic instability and providing a proliferative advantage for cells carrying mutations [21,22]. Recent evidence also indicate that chemokines and proinflammatory cytokines such as tumor necrosis factor α (TNFα), interleukin 6 (IL-6), and IL-23 play pleiotropic roles in tumor progression [23–28]. Transcription factors such as nuclear factor-κB (NF-κB) and signal transducer and activator of transcription 3 (STAT3) are activated in many tumors and serve as important molecular links between inflammation and cancer [29–31]. Nuclear factor-κB activation seems to play a complex role because NF-κB inhibition also induces certain tumors [32–34], and these contradictions have been incompletely understood [35,36].

Cancer-Induced Inflammation

The intrinsic pathway of cancer-related inflammation is orchestrated by the tumor itself without initiation by prior inflammation. Epithelial cells transformed by activation of oncogenes, by inactivation of tumor-suppressor genes, or by chromosomal rearrangement [37] secrete factors that recruit inflammatory cells to the tumor enabling the buildup of a microenvironment. The cancer cell coevolves with its associated stroma, and the cellular composition, size of the stromal component, and amplitude of response vary according to the tumor type and other factors. The stromal cells include macrophages and other myeloid cells, mast cells, endothelial cells, fibroblasts, dendritic cells, natural killer cells, and T and B cells [38]. Depending on the composition of the cells in the microenvironment and stage of disease, tumor stromal cells can stimulate or inhibit tumor growth. Tumor-associated macrophages (TAMs) secrete cytokines, chemokines, lipid mediators, growth and angiogenic factors, and matrix metalloproteinases that contribute to tissue remodeling and angiogenesis and regulate the adhesion, invasion, and motility of the tumor cells [6,39–43]. Lymphocytes may play a dual role in tumor progression [42]. Cytotoxic T cells could destroy tumor cells directly or through antibody-dependent killing. Conversely, although CD4+ and CD8+ T cells are major components of tumor microenvironments, they are largely ineffective in controlling tumor growth owing to the presence of Tregs and myeloid-derived suppressor cells (MDSCs) that are capable of suppressing T-cell proliferation. Myeloid-derived suppressor cells are a heterogeneous population of immature myeloid cells, characterized by the expression of Gr-1, CD11b, and IL4Rα [44–47]. They are normally present in low numbers in blood and lymphoid organs, but accumulate excessively in tumors, blood, and lymphoid organs of tumor-bearing mice, and are also found in various human cancers. Myeloid-derived suppressor cells suppress activation of CD4+ and CD8+ T cells, inducing their anergy or deletion, and promote the induction of Tregs, thus serving to blunt the body's immune responses to tumor antigens [48–50]. Immature dendritic cells (DCs) expressing HLA-DR accumulate in human tumors and provide a mechanism for immune evasion through inefficient antigen presentation [51,52]. Tumors also orchestrate other mechanisms to escape immune surveillance of the host [38].

As discussed below, DAMP molecules seem to play important roles in both inflammation-induced cancer and cancer-induced inflammation.

Damage-Associated Molecular Pattern Molecules

About two decades ago, the late Charles Janeway proposed that the immune system has evolved to protect the host, not against any innocuous foreign antigen but against infectious pathogens, and postulated that receptors on antigen-presenting cells of the innate immune system recognize pathogen-associated molecular pattern (PAMPs). It is now well established that cells of the innate immune system sense PAMPs through pattern recognition receptors (PRRs) such as Toll and Toll-like receptors (TLRs), the NOD1-like receptors, mannose receptor, and other scavenger receptors, and retinoic acid-inducible gene-1-like receptors, stimulation of which initiates a range of host defense mechanisms [53–59]. However, Janeway's model did not explain why strong immune responses are elicited against tissue transplants, ischemia-related injuries, tumors, and autoimmune diseases, none of which involve microbial components. In 1994, Matzinger [60] postulated that the immune system not only responds to pathogens but also senses and responds to intracellular alarm signals arising from nonphysiological cell death, damage, or stress. It is now known that necrosis of healthy cells in response to inflammation, ischemia, or hypoxia within tumors in fact releases endogenous molecules that alert the immune system of danger [61,62]. In live cells, the preexisting danger signals are hidden; apoptotic cells that are sequestered and cleared by phagocytes do not release their intracellular contents unless there is secondary necrosis while necrotic cells lose membrane integrity causing release of intracellular contents. These endogenous danger signals are called alarmins, and together with PAMPs that are microbial in origin, they are referred to as DAMPs [63,64]. The so-called signal 0 events initiated by DAMPs promote early innate and adaptive immune responses mediated through distinct receptors but interlinked to pathways orchestrated by cytokine, chemokine, and other inflammatory mediators. This early responsemediated byDAMPs is also called “sterile inflammation” because it is initiated in response to trauma, ischemia, and other tissue damage in the absence of pathogenic infection. The sequence of immune responses to injury is so robust and stereotypical that it is used by pathologists to date the time of tissue injury in autopsies. Recent studies suggest that radiotherapy and some chemotherapeutic agents may cause preapoptotic and apoptotic changes on cell surface of cancer cells with concomitant release of soluble mediators that trigger DC activation and antitumor immune responses [65,66]. These findings suggest that DAMP molecules could have both protumor and antitumor effects [67].

Apart from their release from necrotic cells, several DAMP molecules are also secreted from activated leukocytes in response to microbial components or cytokines [63]. The molecules lack secretion signals but they are actively secreted through by a nonclassical pathway. A recent study shows that this noncanonical secretion is mediated by activated caspase-1, suggesting regulation by inflammasomes [68]. Hence, there are different mechanisms by which DAMP molecules are released: 1) passive release from necrotic cells, 2) pulsatile release from apoptotic cells in response to radio or chemotherapy, and 3) induced release from activated immune cells by a noncanonical pathway. Because DAMP molecules promote the expression of cytokines, which in turn induce expression of DAMPs, signaling events mediated by DAMPs provide for a feed-forward cycle of inflammatory, tissue repair, and regeneration responses, which, when uncontrolled, may lead to carcinogenesis.

Endogenous DAMP molecules or alarmins include several intracellular proteins, DNA, RNA, and nucleotides (reviewed in Rock and Kono [61]). They are expressed in different cell types and function in normal cellular homeostasis. They are localized in the nucleus and cytoplasm (HMGB1), cytoplasm (S100 proteins), exosomes (heat shock proteins), and extracellular matrix (hyaluronic acid). On the basis of their origin and mechanism of action, the proinflammatory DAMP molecules can be classified as those that directly stimulate cells of the innate immune system and those that generate DAMPs from other extracellular molecules [61]. Here, we focus on two DAMP molecules HMGB1 and proinflammatory S100 proteins. They are released extracellularly by aforementioned processes, bind to PRRs, proteoglycans, and carboxylated glycans, trigger immune responses, promote tissue regeneration, and are implicated in inflammation and cancer.

High-Mobility Group Box 1

HMGB1 is a member of the nonhistone, chromatin-associated high-mobility group family of proteins [69]. It is a highly conserved gene expressed by all eukaryotic cells. During normal cellular homeostasis, HMGB1 is localized predominantly to the nucleus. It binds to the minor groove of DNA and facilitates the assembly of site-specific DNA-binding transcriptional complexes [69]. Nuclear functions of HMGB1 are essential to life because HMGB1 knockout mice die within 24 hours of birth from hypoglycemia [70].

Extracellular Release of HMGB1

HMGB1 is passively released by all cells upon necrotic cell death [71,72]. However, it can also be secreted by macrophages and DCs by a nonclassic pathway in response to lipopolysaccharide, interferon-γ, and TNFα [73,74]. Cytokine-mediated releases of HMGB1 from pituicytes and enterocytes have also been reported [75]. Lipopolysaccharide-mediated HMGB1 release is regulated by its hyperacetylation [76], whereas TNFα-induced secretion seems to be mediated through phosphorylation [77]. As mentioned earlier, this active secretion is through leaderless, non-Golgi-dependent pathways. Outside the cell, HMGB1 behaves as a cytokine, promoting inflammatory responses [74]. Exciting new reports show that HMGB1 secreted by apoptotic tumor cells after chemotherapy or radiation therapy promotes antitumor responses [65]. In addition, apoptotic cells activate macrophages that engulf them to secrete HMGB1 [78].

HMGB1 and Inflammation

HMGB1 released from necrotic cells induces inflammation [71]. It induces DC maturation, migration, and T-cell activation [79–81]. Monocytes, T cells, and endothelial cells release cytokines and inflammatory mediators in response to HMGB1, all of which augment the local inflammatory environment [79,80,82–86]. Administration of HMGB1 to mice significantly increases serum TNFα levels [83]. Purified recombinant HMGB1 also induces a delayed and biphasic release (3 and 8–10 hours after stimulation) of TNFα, IL-1α, IL-1β, IL-6, IL-8, and macrophage inflammatory protein 1α and β from human monocytes at concentrations within the pathological range observed in sepsis [83]. This effect is restricted to monocytes because it does not release cytokines from lymphocytes. HMGB1 also induces TNFα, IL-1β, IL-6, and nitric oxide production by murine macrophages through receptor for advanced glycation end products (RAGE)-dependent signaling pathways [82] and IL-6, monocyte chemoattractant protein 1, and thrombin-antithrombin complex levels in peritoneal lavage fluid and plasma of mice through TLR4 and RAGE-dependent mechanisms [87]. It has been suggested that the HMGB1 polypeptide itself has a weak proinflammatory activity and that binding to bacterial components including lipids may strengthen its effects [84]. HMGB1 also acquires enhanced proinflammatory activity through binding to cytokines such as IL-1β [88]. Furthermore, HMGB1 promotes the expression of intercellular adhesion molecule 1 and vascular cell adhesion molecule 1 on the surface of endothelial cells [85,86]. It mediates hemorrhagic shock-induced NAD(P)H oxidase activation and NF-κB-dependent gene expression in neutrophils [89,90]. HMGB1-DNA complexes promote maturation of immune cells and production of cytokines [91,92] while suppressing immune response in a few cell types [93].

Wang et al. [94] first described proinflammatory cytokine activity of HMGB1, when they showed that HMGB1 was a late mediator of endotoxin-mediated sepsis in mice. Since this seminal finding, HMGB1 has been implicated in the pathogenesis of a variety of sterile inflammatory conditions including rheumatoid arthritis [95–97], lupus erythematosus, and Sjögren syndrome [98,99], trauma and hemorrhagic shock [100–103], and ischemia-reperfusion injury of the liver, heart, kidney, and brain [104–107], providing evidence for its role as a danger signal. It was recently shown that HMGB1 released from late apoptotic cells remains bound to nucleosomes and that HMGB1-nucleosome complexes activate antigen-presenting cells and induce secretion of cytokines by macrophages and expression of costimulatory molecules in DCs [108]. Because autoantibodies against double-stranded DNA and nucleosomes are a characteristic of systemic lupus erythematosus, HMGB1 bound to nucleosomes could therefore contribute to the pathogenesis of systemic lupus erythematosus.

In addition to studies highlighting the proinflammatory effects of HMGB1 in models of multiple diseases in vivo, there is emerging evidence to suggest that HMGB1 participates in tissue repair and remodeling, a role that is increasingly recognized as a characteristic of damage-associated molecules. HMGB1 is proangiogenic [109,110]. It induces migration of mesangioblasts [111,112], endothelial progenitor cells [112], and myogenic cells [113] and chemotaxis and proliferation of smooth muscle cells [114,115]. It stimulates myogenesis [116] and promotes myoblast differentiation [117] and myocardial regeneration after infarction [118]. It also promotes enhanced arteriole density, granulation tissue deposition, and accelerated wound healing in diabetic skin [119].

HMGB1 and Cancer

HMGB1 is widely expressed inmany tumor cells and can be secreted by them or be released upon necrotic cell death [120,121]. Its expression is high in migrating growth cones and malignant cells [122]. It also binds tissue-type plasminogen activator and plasminogen, promoting plasmin production and hence tissue invasion [123,124]. Given its effects in tissue repair, wound healing, angiogenesis, and cell migration, HMGB1 could augment tumor growth and metastasis. In fact, studies suggest that up-regulation of HMGB1 is associated with a malignant phenotype of many cancers [125,126]. HMGB1 also mediates inflammation-based colon carcinogenesis [127]. HMGB1 is significantly elevated in serum and colonic tissue during acute inflammation induced by DSS and anti-HMGB1 treatment reduces severity and extent of inflammatory lesions [127]. HMGB1 released by necrotic colon cells seems to affect surrounding inflammatory cells such as macrophages inducing inflammatory cytokine production and tissue repair. Apc/Min+ mice, in which colon tumors were triggered by DSS-induced inflammation, also show a significant decrease in tumor numbers within the colon when treated with anti-HMGB1 [127]. Colon cancer-derived HMGB1 promotes growth inhibition and apoptosis of macrophages [128] suggesting a role for HMGB1 in the tumor microenvironment. Thus, targeting HMGB1 production or release, or blocking its interaction with its receptors and downstream signaling in macrophages, might have therapeutic applications in both inflammation and cancer [129]. However, as mentioned earlier, recent studies show that pulsed acute release of HMGB1 occurs after chemotherapy or radiotherapy [65], which promotes DC processing of apoptotic cells, DC maturation, clonal expansion of tumor-specific T cells, and antitumor immune response [66,130,131]. This suggests that HMGB1, depending on the pulsatile release from apoptotic tumor cells or chronic release from necrotic tumor cells, could have a paradoxical dual effect on tumors [67].

S100 Proteins

S100 proteins are a family of more than 20 homologous intracellular proteins characterized by calcium-binding EF hand motifs, low molecular weights, ability to form homodimers and heterodimers and oligomers, and tissue-specific expression [132–135]. Most of the S100 genes are clustered at the chromosomal region 1q21, a region frequently rearranged in epithelial tumors and tumors of soft tissues [132,135,136]. They have two distinct EF hand calcium-binding domains connected by a hinge region. The canonical C-terminal calcium binding EF hand is common to all EF hand proteins, whereas the N-terminal EF hand is characteristic of S100 proteins. Intracellular functions of S100 proteins have been extensively studied. These include calcium homeostasis, cell cycle regulation, cell growth and migration, cytoskeletal interactions, protein phosphorylation, and regulation of transcriptional factors among others [132–135].

Extracellular Functions of S100 Proteins

Extracellular functions have been reported for a few S100 proteins. The most well studied extracellular effects relate to the myeloid-specific S100 proteins, namely, S100A8, S100A9, and S100A12 [133,137,138]. They are expressed predominantly in cells of myeloid origin. S100A8 and S100A9 are present in neutrophils, monocytes, and myeloid progenitors and can be induced in keratinocytes during inflammation [136]. Expression is downregulated during macrophage and DC differentiation [139–141]. S100A12 expression is restricted to neutrophils and is not expressed in rodents [142,143]. Like other DAMP molecules, the proteins lack secretion signals required for classic Golgi-dependent transport and are released by an energy-dependent and tubulin-dependent process, which requires activation of protein kinase C [144]. When secreted into the extracellular medium in response to cell damage or activation, they become danger signals that activate other immune cells and endothelial cells [138].

Multimeric forms of S100 proteins seem to be necessary for the extracellular functions of S100 proteins [145,146]. Multimeric assemblies have been reported for S100A12, S100A4, and S100B. S100A8 and S100A9 function predominantly as S100A8/A9 heterodimers. Disruption of the S100A8 gene causes late embryonic lethality [147], whereas S100A9 null mice do not exhibit an obvious phenotype. However, targeted deletion of S100A9 leads to a complete lack of S100A8 and a functional S100A8/A9 complex in peripheral blood cells and cells of the bone marrow, despite normal mRNA levels of S100A8, suggesting that S100A9 expression is important for the stability of the S100A8 protein [148,149].

S100 Proteins and Inflammation

S100A8/A9 and S100A12 induce prothrombotic and proinflammatory responses in endothelial cells including induction of thrombospondin, chemokines, and adhesion molecules and stimulate proinflammatory cytokine production by macrophages [142,150–154]. Up-regulation of chemokines and adhesion molecules helps to promote further recruitment of leukocytes into inflamed tissues. S100A8/A9 and S100A12 are elevated early in tissues and serum in many pathological conditions associated with inflammation such as arthritis, inflammatory bowel disease, vasculitis, multiple sclerosis, psoriasis, and cystic fibrosis and are considered suitable biomarkers of inflammation [133,137,138].

S100A8/A9 and Cancer

It is becoming increasingly clear that S100A8/A9 proteins are involved in many aspects of tumor growth and metastasis. They are upregulated in many cancers including lung, gastric, colorectal, prostate, breast, and pancreatic cancers [136,155]. At low concentrations, S100A8/A9 promote tumor cell growth [19,156]. Elevated levels of S100A8/A9 in chronic inflammation and cancer suggest that the proteins play important roles in inflammation-mediated carcinogenesis.

Recent studies show that S100A8/A9 regulate the accumulation of MDSC in tumors [141,157]. S100A8 and S100A9 are downregulated during normal differentiation of myeloid precursors to DC and macrophages [139–141]. However, tumor-derived factors promote sustained up-regulation of S100A9 in myeloid precursors, which results in the inhibition of differentiation to DC and accumulation of MDSC [141]. These tumor-induced effects are not observed in cells from S100A9 null mice, which show less accumulation of MDSC, higher rate of tumor rejection, and lower tumor size than wild-type controls. This study also shows that activated STAT3 upregulates the expression of S100A8 and S100A9 in myeloid cells in vitro and in vivo. In a parallel study, we reported not only that S100A8/A9 are synthesized and secreted by MDSC but also that they have binding sites for S100A8/A9 [157]. Part of the binding is mediated by carboxylated glycans and by RAGE, leading to intracellular NF-κB signaling and MDSC migration. These findings strongly suggest that the S100A8/A9 proteins support an autocrine feedback loop that sustains accumulation of MDSC in tumors, alongside IL-6, IL-1β, prostaglandin E2, and complement components [158].

S100A8/A9 are also involved in early metastatic processes. Soluble factors such as vascular endothelial growth factor, transforming growth factor β, and TNFα expressed by primary tumors and/or TAMs induce expression of S100A8 and S100A9 in myeloid and endothelial cells of premetastatic lungs [159]. These changes in the local microenvironment termed “premetastatic niche” represent early events in tumor dissemination and dictate the pattern of metastasic spread [160]. Expression of S100A8/A9 in myeloid and endothelial cells in the lung promote homing of tumor cells to these premetastatic niches [159]. If the tumor cells encounter myeloid progenitors at the premetastatic sites, this could promote an angiogenic switch necessary for metastatic cell survival. These studies indicate that S100A8/A9 could be targeted to prevent tumor metastasis. Other S100 proteins have also been implicated in cancers, including S100B, S100A4, S100A7, S100A11, and S100P, and have been the subject of recent reviews [155,161].

Receptors That Detect Endogenous Danger Signals

It is becoming increasingly evident that intracellular mediators released upon necrotic cell death elicit inflammatory responses through recognition by signaling receptors, similar to the recognition of PAMPs. In fact, emerging evidence shows that some of the same PRRs that recognize PAMPs may also mediate responses to endogenous danger signals. By recognizing either pathogens or danger signals, PRRs seem to represent a common pathway to alert the host of danger and to promote tissue repair and regeneration.

Toll-like Receptors

Toll-like receptors or TLRs are a family of transmembrane receptors that recognize microbial molecular patterns or PAMPs and enable cells of the innate immune system to mount inflammatory responses against pathogens [59,162]. Different microbial moieties signal through different TLRs. Lipopolysaccharide from gram-negative bacteria is recognized by TLR4; double-stranded RNA activates TLR3; bacterial flagellin stimulates TLR5. TLR1, TLR2, and TLR6 recognize bacterial peptidoglycans, lipoproteins, lipoteichoic acids, lipoarabinomannan, and yeast zymosan; TLR7 recognizes single-stranded RNA; unmethylated CpG motifs in DNA are recognized by TLR9. Toll-like receptors that recognize viral nucleic acids such as TLR3, TLR7, and TLR9 are localized in endolysosomal compartments, whereas those that recognize bacterial protein and lipid ligands are expressed on the cell surface. All TLRs except TLR3 associate with myeloid differentiation factor 88 (MyD88), and this stimulates a kinase cascade resulting in the activation of mitogen-activated protein kinases (MAPKs), c-Jun N-terminal kinases, p38, and extracellular signal-regulated kinases, and NF-κB [163,164].

In addition to PAMPs, TLR2 and TLR4 also recognize endogenous danger signals [165,166]. HMGB1 binds to TLR2 and TLR4 [167,168]. The ability of HMGB1 to stimulate NF-κB activation and cytokine production in macrophages and to promote neutrophil recruitment in vivo in response to inflammation is dependent in part on TLR signaling pathways [103,169–171]. HMGB1 released by chemotherapy-induced cell death binds to TLR4 and induces antitumor T-cell immunity [65]. HMGB1 bound to nucleosomes from apoptotic cells induces anti-dsDNA and anti-histone immunoglobulin G responses in a TLR2-dependent manner [108]. HMGB1-RAGE interaction acts in a costimulatory manner for TLR9-mediated responses to DNA-containing immune complexes [91,92].

S100A8/A9 were recently shown to interact with TLR4, promoting endotoxin-induced shock [172]. The ability of S100A8 and S100A9 to promote premetastatic niches in lungs also requires TLR4-mediated signaling [173]. They induce serum amyloid A (SAA3) expression in premetastatic lungs, which attracts myeloid cells to the premetastatic niches. SAA3 stimulates TLR4 activity and promotes NF-κB activation [173]. S100A8/A9-SAA3-TLR4 paracrine cascade mediated through NF-κB could therefore be involved in early pulmonary metastasis.

Ligation of TLRs lead to several host defense events that protect the host from infection and damage induced injury [162,174]. TLRs promote tissue repair and regeneration through their angiogenic and antiapoptotic effects [64,174]. MyD88 is essential for the promotion of diethylnitrosamine-induced hepatocellular tumors, spontaneous and azoxymethane-induced intestinal tumorigenesis, and chemically induced skin tumors [175–177]. TLR4 signaling also promotes colitis-induced colon carcinogenesis [178]. Ligands involved in TLR-mediated tissue regeneration and carcinogenesis are unknown. It is likely that TLRs are activated by microbial entities in the gut and enterohepatic circulation or by endogenous ligands such as HMGB1, S100A8/A9, or extracellular matrix components released by necrotic cell death.

Receptor for Advanced Glycation End Products

RAGE, originally discovered as a receptor for advanced glycation end products (AGE), is a multiligand receptor of the immunoglobulin superfamily that plays a key role in immune and other signaling responses mediated by HMGB1 and many S100 proteins [161,179,180]. RAGE also binds other structurally unrelated ligands such as amyloid β peptide, transthyretin, and Mac-1 integrin. It is expressed on monocytes, macrophages, T cells, DCs, smooth muscle cells, immature myofibers, endothelial cells, embryonal neuronal, and tumor cells. Expression is high during embryonic development and low in healthy adult tissues, except in the lung where it is constitutively expressed at high levels [181]. RAGE is implicated in multiple pathologies including diabetes, inflammation, neuronal degeneration, and cancers, primarily as a receptor for DAMPs [161,182–184].

RAGE contains a single variable (V) domain containing two N-glycosylation sites, followed by two constant (C1 and C2) domains, a transmembrane segment and a short cytoplasmic tail necessary for ligand-induced signal transduction [185]. Most ligands bind to the V domain. Different RAGE splice variants exist and have recently been classified as RAGE, RAGE_v1 to RAGE_v19 [186]. The prevalent isoforms are full-length RAGE (RAGE), secreted RAGE that lacks the cytoplasmic and transmembrane domain (sRAGE, RAGE-v1), and N-terminal truncated RAGE (RAGE-v2). The relative expression of the isoforms is tissue-specific. RAGE-v1 or sRAGE is believed to regulate full-length RAGE activation through its ability to bind ligands extracellularly.

HMGB1 and RAGE

RAGE was the first identified receptor for HMGB1 [187]. RAGE-HMGB1 interactions mediate NF-κB-dependent production of cytokines and up-regulation of cell surface receptors [171]. HMGB1 stimulates endothelial progenitor cell migration to ischemic and tumor regions in a RAGE- and integrin-dependent manner [112] and RAGE mediates the proangiogenic effects of HMGB1 [109]. Effects of HMGB1 on mesangioblast homing, skeletal muscle regeneration, chemotaxis of smooth muscle cells, and myogenesis are partly mediated by RAGE [111,113,114,116]. RAGE-HMGB1 interactions also promote DC maturation, homing, and T-cell activation [79,80].

HMGB1-induced RAGE signaling also mediates embryonal neurite outgrowth [187]. This finding, combined with expression of HMGB1 at the leading edges of motile cells and ability to bind tissue-type plasminogen activator and plasminogen leading to the production of plasmin [122,123,188], suggested that HMGB1-RAGE interactions could promote tumor invasion and metastasis. In fact, overexpression of HMGB1, along with RAGE, has been associated with proliferation and metastasis of many tumors [120,124,128,189,190]. However, the tumorigenic effects may be tissue- and cell-dependent because the expressions of RAGE and HMGB1 and their interaction have also been shown to correlate negatively with tumor growth. For example, RAGE is constitutively expressed in the lung, and down-regulation of RAGE and HMGB1 is associated with increased aggressiveness of lung carcinomas [191]. RAGE-HMGB1 engagement also reduces tumor potential of rhabdomyosarcoma cells in vitro and in vivo, suggesting that reduced RAGE signaling may contribute to rhabdomyosarcomagenesis [117,192]. In patients with esophageal and oral squamous cell carcinomas, reduced expression of RAGE negatively correlates with tumor invasion and associated with better prognosis [193,194]. Soluble RAGE and HMGB1 are expressed in tumors of cartilage. Whereas RAGE expression correlates positively with tumor grade and survival, HMGB1 expression does not, suggesting distinct functions of the soluble form of RAGE and HMGB1 [195].

S100 Proteins and RAGE

S100B and S100A12 were the first of the S100 proteins shown to initiate intracellular signaling through interaction with RAGE [152,196]. Since then, a large number of S100 proteins have been shown to bind to RAGE, and some of these promote inflammation and cancer [145,146,155,161]. S100 proteins and RAGE are coexpressed in a variety of human tumors [197]. Like many S100 and DAMP proteins, S100B exhibits both intracellular and extracellular functions [198]. Intracellular S100B stimulates cell proliferation and migration and inhibits apoptosis and differentiation, whereas extracellular S100B exerts regulatory effects on a relatively larger number of cell types in an autocrine and paracrine manner through RAGE and possibly other receptors [198]. S100B is also implicated in diabetes and inflammation. S100B induces RAGE-dependent inflammatory gene expression and oxidative burst in monocytes, macrophages, microglia, and neurophils at high concentrations that could be relevant in local inflammatory environments in both acute and chronic inflammations [198]. S100B causes chemoattraction of RAGE-expressing encephalitogenic CD4+ TH1 T cells in a model of experimental autoimmune encephalomyelitis suggesting a role in the pathophysiology of multiple sclerosis [199]. S100B promotes RAGE-dependent activation of NF-κB in endothelial cells, inducing expression of vascular cell adhesion molecule 1, macrophage chemotactic protein 1, and RAGE [200,201]. It also triggers signaling pathways in smooth muscle cells in a RAGE-dependent manner, resulting in the up-regulation of macrophage chemotactic protein 1 and IL-6 [202]. Interaction of S100A7 or Psoriasin with RAGE mediates chemotaxis of leukocytes [203]. S100A8/A9 are upregulated in many inflammatory diseases, and RAGE and S100A8/A9 are coexpressed in tumors [156,204,205] and are linked to downstream signaling in tumor cells and endothelial cells [153,156,204]. Recent studies provide a more direct evidence of the interaction of S100A8/A9 to RAGE [19,156,206]. At low concentrations, S100A8/A9-induced NF-κB activation promote the growth of tumor cells. This effect is blocked by RAGE gene silencing or by treatment with anti-RAGE [156]. S100A8/A9 also promote LPS-induced cardiac myocyte dysfunction and RAGE coimmunoprecipitates with S100A8 and S100A9 suggesting a direct role for RAGE in S100A8/A9-mediated effects in cardiac myocytes [206]. More recently, we showed that S100A8/A9 binds to a subpopulation of RAGE modified by carboxylated glycans [19] suggesting a direct interaction between the proteins. S100A11 has been shown to modulate osteoarthritis through interaction with RAGE [207]. S100A12-mediated RAGE activation has been implicated in colon inflammation [152]. Functional interactions of RAGE and S100P in pancreatic and colon cancer cells have been demonstrated [208,209].

As mentioned above, most RAGE ligands, including AGEs, HMGB1, S100 proteins, and amyloid β peptide, are highly elevated in inflammatory foci, and RAGE-dependent inflammation promotes up-regulation of both ligands and receptor leading to a feed-forward signaling [182,210], amplifying the inflammatory environment that would promote tumorigenesis. In support of this, recent studies indicate a role for RAGE in inflammation-induced carcinogenesis. RAGE null mice are resistant to the onset of DMBA/TPA-induced skin carcinogenesis and azoxymethane/DSS-induced colon carcinogenesis [18,19]. In both these models, S100A8/A9 are strongly upregulated in stromal cells within the tumors. RAGE-/- mice show reduced levels of MDSC in the DMBA/TPA-induced skin carcinogenesis, implicating RAGE in S100A8/A9-induced MDSC recruitment [18].

Other Cell Surface Binding Sites for DAMP Molecules

DAMP molecules HMGB1, S100A8/A9, and S100A12 bind to a novel modification of N-glycans called carboxylated glycans, which are expressed on RAGE and other glycoproteins [211,212]. HMGB1 binds heparan sulfate proteoglycans, heparin, syndecan, and phosphocan [122,213]. S100A8/A9 also bind to heparan sulfate proteoglycans [214].

Signaling Pathways Activated by DAMP Ligation Converge on NF-κB

RAGE ligation by DAMPs leads to the activation of signaling pathways (Erk1/2 MAPKs, Cdc42/Rac SAP/JNK, and p38 MAPKs) implicated in cell proliferation and cell migration [124,152,215,216]. Toll-like receptors can signal through MyD88, IL-1 receptor-associated kinase, TNF receptor-associated factor, Akt, Cdc42/Rac, phosphatidyl inositol-3 kinase, and MAPKs [163,167,172,217]. Signaling pathways activated by DAMP ligation of the PRRs result in activation of NF-κB [218,219], which further promotes the expression of proinflammatory cytokines, chemokines, angiogenic factors, adhesion molecules, nitric oxide synthase, matrix metalloproteases, and antiapoptotic genes [29]. Chronic NF-κB activation and subsequent inflammation, angiogenesis, tissue repair, and regeneration could therefore lead to tumor development. In fact, specific inactivation of the classic NF-κB activation pathway in epithelial cells and macrophages reduces the formation of inflammation-associated colonic tumors in mice, suggesting that sustained NF-κB activation in either or both of these cells may provide a critical link between inflammation and cancer [17,29]. TNFα-induced NF-κB activation also promotes hepatitis-associated carcinoma in Mdr2 null mice [30].

Relative Importance of the Different Receptors and DAMP Molecules in Inflammation and Cancer, and Binding Specificities

PRRs represent a limited number of proteins by which cells recognize microbial entities and endogenous danger signals and orchestrate an immune response. However, the relative importance and contribution of the different DAMP molecules and of RAGE, TLRs, and other receptors in mediating inflammation and cancer are not completely understood and are likely to differ between cell types. If we learn more about the specificity of these interactions, we can also determine targets for inhibition. Binding specificity may be imparted by interactions of different domains on the ligands and receptors. The epitopes on DAMPs recognized by RAGE might be different from those recognized by TLRs. HMGB1 may interact with distinct receptors through its DNA-binding boxes or through the C-terminal domain. For example, amino acids 150 to 183 of HMGB1 interact with RAGE [220]. Similarly, distinct epitopes on RAGE and TLRs recognized by the ligands may also impart specificity. For example, it is becoming increasingly clear that different S100 proteins require different domains on RAGE for binding. S100A12 binds to V-C1 domains, S100B requires theV domain of RAGE and S100A6 interacts with V-C2 domain [221–224]. Posttranslational modifications such as glycosylation on the receptors, acetylation, or phosphorylation of ligands and formation of multimolecular assemblies as with S100 proteins could also play important roles in defining specificity of interactions and down stream signaling.

Our studies suggest that N-glycan modifications of RAGE may serve as unique ligand binding sites and may contribute to some of the pleiotropic binding ability of the receptor. RAGE has two N-glycosylation sites on the ligand binding V domain and both sites are occupied by complex and hybrid N-glycans. Our recent analysis of N-glycans on sRAGE suggests considerable heterogeneity of glycan structures on RAGE(Houliston et al., unpublished data). Several years ago, we identified a novel group of anionic N-glycans that contain an immunogenic carboxylate group unrelated to sialic or uronic acids [225]. These carboxylated glycans contain glutamic or aspartic acids [226]. Using a monoclonal antibody against the glycans (mAbGB3.1), we found that the glycans show restricted expression on mouse and human cells of myeloid lineage including monocytes, macrophages, and DCs and on endothelial cells [227,228]. They are absent or undetectable on normal epithelial cells. However, they are expressed on several tumor cells [19]. To identify glycan-binding proteins, we applied whole bovine lung homogenates through a column of carboxylated glycans and found that DAMP molecules HMGB1, S100A8/A9, and annexin I specifically bound to the column [211,212]. We found that a subpopulation of RAGE molecules is modified by carboxylated glycans [19,212] and that binding of HMGB1 to RAGE partially depends on carboxylated glycans [212]. The subpopulation of RAGE enriched for carboxylated glycans by mAbGB3.1 also showed 10- to 100-fold increase in binding potential (Bmax/Kd) for both S100A8/A9 and S100A12, suggesting that carboxylated glycans form critical binding sites for these ligands on RAGE. Conversely, based on our recent unpublished findings, S100A11, S100B, and S100A1 as well as AGE do not show enhanced binding to mAbGB3.1-enriched RAGE, suggesting that, although many S100 family members bind RAGE, they may bind to different structural domains or regions on the receptor, some involving carboxylated N-glycans on the V domain and others not, thus providing differential binding specificity. In support of this, we found that the mAbGB3.1-enriched population of RAGE forms higher-order complexes of S100A12 and that deglycosylation of RAGE reduced the ability to form multimeric complexes.

As further evidence of the importance of carboxylated glycans in mediating DAMP interactions, we found that inhibiting carboxylated glycan-dependent interactions of DAMP molecules using mAbGB3.1, blocked onset of T-cell-mediated colitis [228], colitis-dependent colon cancer [19], and recruitment of MDSC to secondary lymphoid organs and accumulation of MDSC in blood [157]. Proteomic analysis (Mud-PIT) of mAbGB3.1-immunoprecipitated proteins from macrophages revealed the presence of RAGE among other glycoproteins, but not TLR2 or TLR4, suggesting that this modification may not be present on all PRRs (Srikrishna et al., unpublished data). Our studies on colitis-associated carcinogenesis also show that S100A8/A9 and HMGB1 could participate in distinct events in disease progression through different receptors [19]: an acute inflammation phase involving TLR4 and a tumorigenesis and progression phase involving the glycans and RAGE expressed on tumor cells because mAbGB3.1 does not block early DSS-mediated colitis but blocks chronic inflammation and carcinogenesis. In addition, RAGE-/- mice are as susceptible to early DSS-induced injury as RAGE+/+ mice but are resistant to colitis-mediated cancer.

Current structural studies on RAGE are performed on the extracellular domains (VC1C2) that comprise soluble sRAGE but are often produced by expression systems (bacteria, insect cells, or yeast) that lack the complex glycosylation machinery of the mammalian systems. Structural and ligand-binding analysis is typically done on single or on tandem domains [221]. Nuclear magnetic resonance analysis of expressed VC1 domains show that they form an integrated structural unit that binds to Ca2+-S100B [223]. One of the critical residues in the interactive face (Thr27) is part of the unoccupied N-glycosylation sequon; the presence of a normal glycan chain on a native RAGE protein on cells would likely alter this interaction. S100A12 is shown to bind to the C1 domain of RAGE [221]; however, the protein used in this study has no N-glycans because it was expressed in Escherichia coli. On the basis of our findings, it is likely that both V and C1 domains are necessary for binding to S100A12. AGEs bind to the V domain [229], whereas S100A6 binds to the C1C2 domains [222]. Kd estimates for S100 ligands range from 5 to 500 nM depending on the analytical methods, specific ligands, and conditions. All studies concur that RAGE-ligand binding generates multimeric complexes of both RAGE (tetramer) and ligands (tetramer, hexamer, octomer) and that formation of these higher-order complexes may be essential for signal transduction [145, 224,230]. The efficiency of complex formation varies widely, sometimes requiring extremely high concentrations of protein (500 mM) and involving only a few percent of the molecules. It is unclear whether the in vitro conditions mimic the formation of signaling-competent RAGE-ligand multimers on a cell surface. Glycans are often regarded as impediments in high-resolution protein structural analysis, but their role in formation of signaling complexes is now well documented [231,232]. Studies on glycan-deficient RAGE, therefore, may not accurately reflect the appropriate in vitro or in vivo ligand binding, domain interactions, complex formation, or details of the signaling pathways. Detailed binding studies using fully glycosylated RAGE protein are therefore necessary.

Figures 1 and and22 provide a representation of the current findings on the role of DAMP molecules and PRRs in mediating inflammation and cancer.

An external file that holds a picture, illustration, etc.
Object name is neo1107_0615_fig001.jpg

Damage-associated molecular pattern molecules in inflammation. Inflammation can be initiated by microbial PAMPs or by intracellular danger signals such as HMGB1 and S100A8/A9 released from necrotic cells or secreted from monocytes. HMGB1 and S100A8/A9 bind to TLRs or RAGE and promote NF-κB signaling and expression of cytokines that act as growth factors for tissue repair and regeneration functioning through their respective receptors (generic representation for cytokine receptor(s) shown here). HMGB1 and S100A8/A9 promote expression of adhesion molecules and cytokine production by local vascular endothelium, which further attract neutrophils and monocytes. HMGB1 also promotes RAGE- and TLR-dependent smooth muscle and mesangioblast migration, angiogenesis, and tissue repair. Nuclear factor-κB-dependent proinflammatory cytokines in turn upregulate the expression of DAMPs and RAGE, leading to a pathological cycle of inflammation, necrosis, and tumorigenesis.

An external file that holds a picture, illustration, etc.
Object name is neo1107_0615_fig002.jpg

Damage-associated molecular pattern molecules in the tumor microenvironment. HMGB1 has a dual effect on tumors. Acute release of HMGB1 after antitumor treatments promotes maturation of DCs through interaction with TLR4 and clonal expansion of tumor antigen-specific T cells and antitumor responses. Conversely, persistent hypoxia in growing tumors leads to necrosis, causing chronic release of HMGB1, which activates protumor responses promoting angiogenesis and tumor growth through the recruitment of macrophages (TAM) and endothelial precursor cells (EPC) and activation of local endothelial cells through RAGE signaling. In the bone marrow, S100A8/A9 are downregulated during normal differentiation of myeloid precursors to DC and macrophages. However, tumor-derived factors promote sustained up-regulation of S100A9 in myeloid precursors through a STAT3 dependent process, which results in the inhibition ofDCdifferentiation and accumulation of MDSC. S100A8/A9 are synthesized and secreted by MDSC and bind carboxylated glycans on other MDSC. This promotes migration and accumulation of MDSC in blood and peripheral lymphoid organs, possibly through RAGE- and NF-κB-dependent pathways, thereby establishing an autocrine feedback loop that maintains MDSC levels and promoting immune suppression against tumors. S100A8/A9 promote tumor growth through RAGE- and carboxylated glycan-dependent pathways. Tumor-derived factors also induce expression of S100A8/A9 inmyeloid and endothelial cells in premetastatic niches within lungs, which promotes homing of tumor cells to lungs.

Conclusions and Future Perspectives

Recent studies show that chronic inflammation and necrotic cell death contribute to tumorigenesis. These studies provide novel insights into the functional role of danger signals such as HMGB1 and S100 proteins released during necrotic cell death and inflammation and the receptors that detect them such as TLRs and RAGE in mediating the pathology. Signaling responses mediated by DAMP molecules include production of cytokines and chemokines, recruitment of leukocytes such as MDSC and associated immune suppression, neoangiogenesis, stromagenesis, and epithelial proliferation. These represent homeostatic tissue repair and remodeling responses and have implications for carcinogenesis when chronic inflammatory states and necrotic cell death lead to uncontrolled responses. Studies suggest therapeutic strategies based on blocking the interactions of DAMP molecules and their receptors or downstream signaling pathways. These include administration of sRAGE, antibodies to RAGE, TLRs, HMGB1, S100 proteins, or carboxylated glycans, and inhibitors of MAPK pathways, NF-κB, and other signaling mediators. Recent studies suggest that pulsed release of HMGB1 after chemotherapy and radiotherapy in fact triggers antitumor immune responses. Further studies are therefore clearly needed to define which set of intracellular molecules constitute danger signals, understand specificity of interactions, whether early and late events are driven by different PRRs, what controlled stimuli could promote apoptosis and antitumor responses, and if inflammatory response to tissue injury can be selectively inhibited without affecting normal host defense mechanisms.

Acknowledgments

The authors thank colleagues and collaborators who have made invaluable contributions to their work through several years. The authors also thank the investigators whose excellent contributions to this field could not be cited owing to space constraints. The authors thank Lars Bode, Vandana Sharma, and Robert Houliston for their critical reading of this review.

Abbreviations

DAMPdamage-associated molecular pattern
DMBA/TPA7,12-dimethyl benz[a]anthracene/12-O-tetradecanoylphorbol-13-acetate
HMGB1high-mobility group box 1
ILinterleukin
MDSCmyeloid-derived suppressor cells
NF-κBnuclear factor-κB
PAMPpathogen-associated molecular pattern
PRRpattern recognition receptor
RAGEreceptor for advanced glycation end products
STAT3signal transducer and activator of transcription 3
TAMtumor-associated macrophage
TLRToll-like receptor
TNFαtumor necrosis factor α

Footnotes

1Work from our laboratory was supported by National Institutes of Health grants R01-CA92608 and R21-CA127780, the Broad Medical Research Program of the Eli and Edythe L. Broad Foundation, and the Crohn's and Colitis Foundation of America.

References

1. Balkwill F, Mantovani A. Inflammation and cancer: back to Virchow? Lancet. 2001;357:539–545. [PubMed] [Google Scholar]
2. Dvorak HF. Tumors: wounds that do not heal. Similarities between tumor stroma generation and wound healing. N Engl J Med. 1986;315:1650–1659. [PubMed] [Google Scholar]
3. Mantovani A, Allavena P, Sica A, Balkwill F. Cancer-related inflammation. Nature. 2008;454:436–444. [PubMed] [Google Scholar]
4. Vakkila J, Lotze MT. Inflammation and necrosis promote tumour growth. Nat Rev Immunol. 2004;4:641–648. [PubMed] [Google Scholar]
5. Coffelt SB, Scandurro AB. Tumors sound the alarmin(s) Cancer Res. 2008;68:6482–6485. [PMC free article] [PubMed] [Google Scholar]
6. Coussens LM, Werb Z. Inflammation and cancer. Nature. 2002;420:860–867. [PMC free article] [PubMed] [Google Scholar]
7. Allavena P, Garlanda C, Borrello MG, Sica A, Mantovani A. Pathways connecting inflammation and cancer. Curr Opin Genet Dev. 2008;18:3–10. [PubMed] [Google Scholar]
8. Fox JG, Wang TC. Inflammation, atrophy, and gastric cancer. J Clin Invest. 2007;117:60–69. [PMC free article] [PubMed] [Google Scholar]
9. Mostafa MH, Sheweita SA, O'Connor PJ. Relationship between schistosomiasis and bladder cancer. Clin Microbiol Rev. 1999;12:97–111. [PMC free article] [PubMed] [Google Scholar]
10. Xie J, Itzkowitz SH. Cancer in inflammatory bowel disease. World J Gastroenterol. 2008;14:378–389. [PMC free article] [PubMed] [Google Scholar]
11. Thun MJ, Henley SJ, Gansler T. Inflammation and cancer: an epidemiological perspective. Novartis Found Symp. 2004;256:6–21. discussion 22–28, 49–52, 266–269. [PubMed] [Google Scholar]
12. Gwyn K, Sinicrope FA. Chemoprevention of colorectal cancer. Am J Gastroenterol. 2002;97:13–21. [PubMed] [Google Scholar]
13. Thun MJ, Henley SJ, Patrono C. Nonsteroidal anti-inflammatory drugs as anticancer agents: mechanistic, pharmacologic, and clinical issues. J Natl Cancer Inst. 2002;94:252–266. [PubMed] [Google Scholar]
14. Backlund MG, Mann JR, Dubois RN. Mechanisms for the prevention of gastrointestinal cancer: the role of prostaglandin E2. Oncology. 2005;69(Suppl 1):28–32. [PubMed] [Google Scholar]
15. Jacobs EJ, Thun MJ, Bain EB, Rodriguez C, Henley SJ, Calle EE. A large cohort study of long-term daily use of adult-strength aspirin and cancer incidence. J Natl Cancer Inst. 2007;99:608–615. [PubMed] [Google Scholar]
16. Becker C, Fantini MC, Schramm C, Lehr HA, Wirtz S, Nikolaev A, Burg J, Strand S, Kiesslich R, Huber S, et al. TGF-beta suppresses tumor progression in colon cancer by inhibition of IL-6 trans-signaling. Immunity. 2004;21:491–501. [PubMed] [Google Scholar]
17. Greten FR, Eckmann L, Greten TF, Park JM, Li ZW, Egan LJ, Kagnoff MF, Karin M. IKKbeta links inflammation and tumorigenesis in a mouse model of colitis-associated cancer. Cell. 2004;118:285–296. [PubMed] [Google Scholar]
18. Gebhardt C, Riehl A, Durchdewald M, Nemeth J, Furstenberger G, Muller-Decker K, Enk A, Arnold B, Bierhaus A, Nawroth PP, et al. RAGE signaling sustains inflammation and promotes tumor development. J Exp Med. 2008;205:275–285. [PMC free article] [PubMed] [Google Scholar]
19. Turovskaya O, Foell D, Sinha P, Vogl T, Newlin R, Nayak J, Nguyen M, Olsson A, Nawroth PP, Bierhaus A, et al. RAGE, carboxylated glycans and S100A8/A9 play essential roles in colitis-associated carcinogenesis. Carcinogenesis. 2008;29:2035–2043. [PMC free article] [PubMed] [Google Scholar]
20. Medzhitov R. Origin and physiological roles of inflammation. Nature. 2008;454:428–435. [PubMed] [Google Scholar]
21. Jaiswal M, LaRusso NF, Burgart LJ, Gores GJ. Inflammatory cytokines induce DNA damage and inhibit DNA repair in cholangiocarcinoma cells by a nitric oxide-dependent mechanism. Cancer Res. 2000;60:184–190. [PubMed] [Google Scholar]
22. Marnett LJ. Oxyradicals and DNA damage. Carcinogenesis. 2000;21:361–370. [PubMed] [Google Scholar]
23. Ben-Baruch A. Inflammation-associated immune suppression in cancer: the roles played by cytokines, chemokines and additional mediators. Semin Cancer Biol. 2006;16:38–52. [PubMed] [Google Scholar]
24. Smyth MJ, Cretney E, Kershaw MH, Hayakawa Y. Cytokines in cancer immunity and immunotherapy. Immunol Rev. 2004;202:275–293. [PubMed] [Google Scholar]
25. Lin WW, Karin M. A cytokine-mediated link between innate immunity, inflammation, and cancer. J Clin Invest. 2007;117:1175–1183. [PMC free article] [PubMed] [Google Scholar]
26. Mumm JB, Oft M. Cytokine-based transformation of immune surveillance into tumor-promoting inflammation. Oncogene. 2008;27:5913–5919. [PubMed] [Google Scholar]
27. Fantini MC, Pallone F. Cytokines: from gut inflammation to colorectal cancer. Curr Drug Targets. 2008;9:375–380. [PubMed] [Google Scholar]
28. Naugler WE, Karin M. The wolf in sheep's clothing: the role of interleukin-6 in immunity, inflammation and cancer. Trends Mol Med. 2008;14:109–119. [PubMed] [Google Scholar]
29. Karin M, Greten FR. NF-kappaB: linking inflammation and immunity to cancer development and progression. Nat Rev Immunol. 2005;5:749–759. [PubMed] [Google Scholar]
30. Pikarsky E, Porat RM, Stein I, Abramovitch R, Amit S, Kasem S, Gutkovich-Pyest E, Urieli-Shoval S, Galun E, Ben-Neriah Y. NF-kappaB functions as a tumour promoter in inflammation-associated cancer. Nature. 2004;431:461–466. [PubMed] [Google Scholar]
31. Yu H, Kortylewski M, Pardoll D. Crosstalk between cancer and immune cells: role of STAT3 in the tumour microenvironment. Nat Rev Immunol. 2007;7:41–51. [PubMed] [Google Scholar]
32. van Hogerlinden M, Rozell BL, Ahrlund-Richter L, Toftgard R. Squamous cell carcinomas and increased apoptosis in skin with inhibited Rel/nuclear factor-kappaB signaling. Cancer Res. 1999;59:3299–3303. [PubMed] [Google Scholar]
33. Dajee M, Lazarov M, Zhang JY, Cai T, Green CL, Russell AJ, Marinkovich MP, Tao S, Lin Q, Kubo Y, et al. NF-kappaB blockade and oncogenic Ras trigger invasive human epidermal neoplasia. Nature. 2003;421:639–643. [PubMed] [Google Scholar]
34. Maeda S, Kamata H, Luo JL, Leffert H, Karin M. IKKbeta couples hepatocyte death to cytokine-driven compensatory proliferation that promotes chemical hepatocarcinogenesis. Cell. 2005;121:977–990. [PubMed] [Google Scholar]
35. Pikarsky E, Ben-Neriah Y. NF-kappaB inhibition: a double-edged sword in cancer? Eur J Cancer. 2006;42:779–784. [PubMed] [Google Scholar]
36. Vainer GW, Pikarsky E, Ben-Neriah Y. Contradictory functions of NF-kappaB in liver physiology and cancer. Cancer Lett. 2008;267:182–188. [PubMed] [Google Scholar]
37. Vogelstein B, Kinzler KW. Cancer genes and the pathways they control. Nat Med. 2004;10:789–799. [PubMed] [Google Scholar]
38. Whiteside TL. The tumor microenvironment and its role in promoting tumor growth. Oncogene. 2008;27:5904–5912. [PMC free article] [PubMed] [Google Scholar]
39. Egeblad M, Werb Z. New functions for the matrix metalloproteinases in cancer progression. Nat Rev Cancer. 2002;2:161–174. [PubMed] [Google Scholar]
40. Condeelis J, Pollard JW. Macrophages: obligate partners for tumor cell migration, invasion, and metastasis. Cell. 2006;124:263–266. [PubMed] [Google Scholar]
41. Sica A, Allavena P, Mantovani A. Cancer related inflammation: the macrophage connection. Cancer Lett. 2008;267:204–215. [PubMed] [Google Scholar]
42. Talmadge JE, Donkor M, Scholar E. Inflammatory cell infiltration of tumors: Jekyll or Hyde. Cancer Metastasis Rev. 2007;26:373–400. [PubMed] [Google Scholar]
43. Le Bitoux MA, Stamenkovic I. Tumor-host interactions: the role of inflammation. Histochem Cell Biol. 2008;130:1079–1090. [PubMed] [Google Scholar]
44. Sinha P, Clements VK, Miller S, Ostrand-Rosenberg S. Tumor immunity: a balancing act between T cell activation, macrophage activation and tumor-induced immune suppression. Cancer Immunol Immunother. 2005;54:1137–1142. [PMC free article] [PubMed] [Google Scholar]
45. Serafini P, Borrello I, Bronte V. Myeloid suppressor cells in cancer: recruitment, phenotype, properties, and mechanisms of immune suppression. Semin Cancer Biol. 2006;16:53–65. [PubMed] [Google Scholar]
46. Nagaraj S, Gabrilovich DI. Myeloid-derived suppressor cells. Adv Exp Med Biol. 2007;601:213–223. [PubMed] [Google Scholar]
47. Dolcetti L, Marigo I, Mantelli B, Peranzoni E, Zanovello P, Bronte V. Myeloid-derived suppressor cell role in tumor-related inflammation. Cancer Lett. 2008;267:216–225. [PubMed] [Google Scholar]
48. Nagaraj S, Gabrilovich DI. Tumor escape mechanism governed by myeloid-derived suppressor cells. Cancer Res. 2008;68:2561–2563. [PubMed] [Google Scholar]
49. Serafini P, De Santo C, Marigo I, Cingarlini S, Dolcetti L, Gallina G, Zanovello P, Bronte V. Derangement of immune responses by myeloid suppressor cells. Cancer Immunol Immunother. 2004;53:64–72. [PMC free article] [PubMed] [Google Scholar]
50. Wang RF. CD8+ regulatory T cells, their suppressive mechanisms, and regulation in cancer. Hum Immunol. 2008;69:811–814. [PubMed] [Google Scholar]
51. Pinzon-Charry A, Maxwell T, Prato S, Furnival C, Schmidt C, Lopez JA. HLA-DR+ immature cells exhibit reduced antigen-presenting cell function but respond to CD40 stimulation. Neoplasia. 2005;7:1123–1132. [PMC free article] [PubMed] [Google Scholar]
52. Pinzon-Charry A, Ho CS, Laherty R, Maxwell T, Walker D, Gardiner RA, O'Connor L, Pyke C, Schmidt C, Furnival C, et al. A population of HLA-DR+ immature cells accumulates in the blood dendritic cell compartment of patients with different types of cancer. Neoplasia. 2005;7:1112–1122. [PMC free article] [PubMed] [Google Scholar]
53. Kabelitz D, Medzhitov R. Innate immunity—cross-talk with adaptive immunity through pattern recognition receptors and cytokines. Curr Opin Immunol. 2007;19:1–3. [PubMed] [Google Scholar]
54. Palm NW, Medzhitov R. Pattern recognition receptors and control of adaptive immunity. Immunol Rev. 2009;227:221–233. [PubMed] [Google Scholar]
55. Park JH, Kim YG, Shaw M, Kanneganti TD, Fujimoto Y, Fukase K, Inohara N, Nunez G. Nod1/RICKandTLRsignaling regulate chemokine and antimicrobial innate immune responses in mesothelial cells. J Immunol. 2007;179:514–521. [PubMed] [Google Scholar]
56. Gazi U, Martinez-Pomares L. Influence of the mannose receptor in host immune responses. Immunobiology. in press. [PubMed] [Google Scholar]
57. McCartney SA, Colonna M. Viral sensors: diversity in pathogen recognition. Immunol Rev. 2009;227:87–94. [PMC free article] [PubMed] [Google Scholar]
58. Kumagai Y, Takeuchi O, Akira S. Pathogen recognition by innate receptors. J Infect Chemother. 2008;14:86–92. [PubMed] [Google Scholar]
59. Kawai T, Akira S. Pathogen recognition with Toll-like receptors. Curr Opin Immunol. 2005;17:338–344. [PubMed] [Google Scholar]
60. Matzinger P. Tolerance, danger, and the extended family. Annu Rev Immunol. 1994;12:991–1045. [PubMed] [Google Scholar]
61. Rock KL, Kono H. The inflammatory response to cell death. Annu Rev Pathol. 2008;3:99–126. [PMC free article] [PubMed] [Google Scholar]
62. Kono H, Rock KL. How dying cells alert the immune system to danger. Nat Rev Immunol. 2008;8:279–289. [PMC free article] [PubMed] [Google Scholar]
63. Bianchi ME. DAMPs, PAMPs and alarmins: all we need to know about danger. J Leukoc Biol. 2007;81:1–5. [PubMed] [Google Scholar]
64. Lotze MT, Zeh HJ, Rubartelli A, Sparvero LJ, Amoscato AA, Washburn NR, Devera ME, Liang X, Tor M, Billiar T. The grateful dead: damage-associated molecular pattern molecules and reduction/oxidation regulate immunity. Immunol Rev. 2007;220:60–81. [PubMed] [Google Scholar]
65. Apetoh L, Ghiringhelli F, Tesniere A, Obeid M, Ortiz C, Criollo A, Mignot G, Maiuri MC, Ullrich E, Saulnier P, et al. Toll-like receptor 4-dependent contribution of the immune system to anticancer chemotherapy and radiotherapy. Nat Med. 2007;13:1050–1059. [PubMed] [Google Scholar]
66. Tesniere A, Apetoh L, Ghiringhelli F, Joza N, Panaretakis T, Kepp O, Schlemmer F, Zitvogel L, Kroemer G. Immunogenic cancer cell death: a key-lock paradigm. Curr Opin Immunol. 2008;20:504–511. [PubMed] [Google Scholar]
67. Campana L, Bosurgi L, Rovere-Querini P. HMGB1: a two-headed signal regulating tumor progression and immunity. Curr Opin Immunol. 2008;20:518–523. [PubMed] [Google Scholar]
68. Keller M, Ruegg A, Werner S, Beer HD. Active caspase-1 is a regulator of unconventional protein secretion. Cell. 2008;132:818–831. [PubMed] [Google Scholar]
69. Thomas JO, Travers AA. HMG1 and 2, and related “architectural” DNA-binding proteins. Trends Biochem Sci. 2001;26:167–174. [PubMed] [Google Scholar]
70. Calogero S, Grassi F, Aguzzi A, Voigtlander T, Ferrier P, Ferrari S, Bianchi ME. The lack of chromosomal protein Hmg1 does not disrupt cell growth but causes lethal hypoglycaemia in newborn mice. Nat Genet. 1999;22:276–280. [PubMed] [Google Scholar]
71. Scaffidi P, Misteli T, Bianchi ME. Release of chromatin protein HMGB1 by necrotic cells triggers inflammation. Nature. 2002;418:191–195. [PubMed] [Google Scholar]
72. Raucci A, Palumbo R, Bianchi ME. HMGB1: a signal of necrosis. Autoimmunity. 2007;40:285–289. [PubMed] [Google Scholar]
73. Gardella S, Andrei C, Ferrera D, Lotti LV, Torrisi MR, Bianchi ME, Rubartelli A. The nuclear protein HMGB1 is secreted by monocytes via a non-classical, vesicle-mediated secretory pathway. EMBO Rep. 2002;3:995–1001. [PMC free article] [PubMed] [Google Scholar]
74. Yang H, Wang H, Czura CJ, Tracey KJ. HMGB1 as a cytokine and therapeutic target. J Endotoxin Res. 2002;8:469–472. [PubMed] [Google Scholar]
75. Wang H, Vishnubhakat JM, Bloom O, Zhang M, Ombrellino M, Sama A, Tracey KJ. Proinflammatory cytokines (tumor necrosis factor and interleukin 1) stimulate release of high mobility group protein-1 by pituicytes. Surgery. 1999;126:389–392. [PubMed] [Google Scholar]
76. Bonaldi T, Talamo F, Scaffidi P, Ferrera D, Porto A, Bachi A, Rubartelli A, Agresti A, Bianchi ME. Monocytic cells hyperacetylate chromatin protein HMGB1 to redirect it towards secretion. EMBO J. 2003;22:5551–5560. [PMC free article] [PubMed] [Google Scholar]
77. Youn JH, Shin JS. Nucleocytoplasmic shuttling of HMGB1 is regulated by phosphorylation that redirects it toward secretion. J Immunol. 2006;177:7889–7897. [PubMed] [Google Scholar]
78. Qin S, Wang H, Yuan R, Li H, Ochani M, Ochani K, Rosas-Ballina M, Czura CJ, Huston JM, Miller E, et al. Role of HMGB1 in apoptosis-mediated sepsis lethality. J Exp Med. 2006;203:1637–1642. [PMC free article] [PubMed] [Google Scholar]
79. Messmer D, Yang H, Telusma G, Knoll F, Li J, Messmer B, Tracey KJ, Chiorazzi N. High mobility group box protein 1: an endogenous signal for dendritic cell maturation and TH1 polarization. J Immunol. 2004;173:307–313. [PubMed] [Google Scholar]
80. Dumitriu IE, Baruah P, Valentinis B, Voll RE, Herrmann M, Nawroth PP, Arnold B, Bianchi ME, Manfredi AA, Rovere-Querini P. Release of high mobility group box 1 by dendritic cells controls T cell activation via the receptor for advanced glycation end products. J Immunol. 2005;174:7506–7515. [PubMed] [Google Scholar]
81. Yang D, Chen Q, Yang H, Tracey KJ, Bustin M, Oppenheim JJ. High mobility group box-1 protein induces the migration and activation of human dendritic cells and acts as an alarmin. J Leukoc Biol. 2007;81:59–66. [PubMed] [Google Scholar]
82. Kokkola R, Andersson A, Mullins G, Ostberg T, Treutiger CJ, Arnold B, Nawroth P, Andersson U, Harris RA, Harris HE. RAGE is the major receptor for the proinflammatory activity of HMGB1 in rodent macrophages. Scand J Immunol. 2005;61:1–9. [PubMed] [Google Scholar]
83. Andersson U, Wang H, Palmblad K, Aveberger AC, Bloom O, Erlandsson-Harris H, Janson A, Kokkola R, Zhang M, Yang H, et al. High mobility group 1 protein (HMG-1) stimulates proinflammatory cytokine synthesis in human monocytes. J Exp Med. 2000;192:565–570. [PMC free article] [PubMed] [Google Scholar]
84. Rouhiainen A, Tumova S, Valmu L, Kalkkinen N, Rauvala H. Pivotal advance: analysis of proinflammatory activity of highly purified eukaryotic recombinant HMGB1 (amphoterin) J Leukoc Biol. 2007;81:49–58. [PubMed] [Google Scholar]
85. Treutiger CJ, Mullins GE, Johansson AS, Rouhiainen A, Rauvala HM, Erlandsson-Harris H, Andersson U, Yang H, Tracey KJ, Andersson J, et al. High mobility group 1 B-box mediates activation of human endothelium. J Intern Med. 2003;254:375–385. [PubMed] [Google Scholar]
86. Fiuza C, Bustin M, Talwar S, Tropea M, Gerstenberger E, Shelhamer JH, Suffredini AF. Inflammation-promoting activity of HMGB1 on human microvascular endothelial cells. Blood. 2003;101:2652–2660. [PubMed] [Google Scholar]
87. van Zoelen MA, Yang H, Florquin S, Meijers JC, Akira S, Arnold B, Nawroth PP, Bierhaus A, Tracey KJ, van der Poll T. Role of Toll-like receptors 2 and 4, and the receptor for advanced glycation end products (RAGE) in Hmgb1 induced inflammation in vivo. Shock. 2009;31:280–284. [PMC free article] [PubMed] [Google Scholar]
88. Sha Y, Zmijewski J, Xu Z, Abraham E. HMGB1 develops enhanced proinflammatory activity by binding to cytokines. J Immunol. 2008;180:2531–2537. [PubMed] [Google Scholar]
89. Fan J, Li Y, Levy RM, Fan JJ, Hackam DJ, Vodovotz Y, Yang H, Tracey KJ, Billiar TR, Wilson MA. Hemorrhagic shock induces NAD(P)H oxidase activation in neutrophils: role of HMGB1-TLR4 signaling. J Immunol. 2007;178:6573–6580. [PubMed] [Google Scholar]
90. Park JS, Arcaroli J, Yum HK, Yang H, Wang H, Yang KY, Choe KH, Strassheim D, Pitts TM, Tracey KJ, et al. Activation of gene expression in human neutrophils by high mobility group box 1 protein. Am J Physiol Cell Physiol. 2003;284:C870–C879. [PubMed] [Google Scholar]
91. Tian J, Avalos AM, Mao SY, Chen B, Senthil K, Wu H, Parroche P, Drabic S, Golenbock D, Sirois C, et al. Toll-like receptor 9-dependent activation by DNA-containing immune complexes is mediated by HMGB1 and RAGE. Nat Immunol. 2007;8:487–496. [PubMed] [Google Scholar]
92. Ivanov S, Dragoi AM, Wang X, Dallacosta C, Louten J, Musco G, Sitia G, Yap GS, Wan Y, Biron CA, et al. A novel role for HMGB1 in TLR9-mediated inflammatory responses to CpG-DNA. Blood. 2007;110:1970–1981. [PMC free article] [PubMed] [Google Scholar]
93. Popovic PJ, DeMarco R, Lotze MT, Winikoff SE, Bartlett DL, Krieg AM, Guo ZS, Brown CK, Tracey KJ, Zeh HJ., III High mobility group B1 protein suppresses the human plasmacytoid dendritic cell response to TLR9 agonists. J Immunol. 2006;177:8701–8707. [PubMed] [Google Scholar]
94. Wang H, Bloom O, Zhang M, Vishnubhakat J, Ombrellino M, Che J, Frazier A, Yang H, Ivanova S, Borovikova L, et al. HMG-1 as a late mediator of endotoxin lethality in mice. Science. 1999;285:248–251. [PubMed] [Google Scholar]
95. Kokkola R, Sundberg E, Ulfgren AK, Palmblad K, Li J, Wang H, Ulloa L, Yang H, Yan XJ, Furie R, et al. High mobility group box chromosomal protein 1: a novel proinflammatory mediator in synovitis. Arthritis Rheum. 2002;46:2598–2603. [PubMed] [Google Scholar]
96. Taniguchi N, Kawahara K, Yone K, Hashiguchi T, Yamakuchi M, Goto M, Inoue K, Yamada S, Ijiri K, Matsunaga S, et al. High mobility group box chromosomal protein 1 plays a role in the pathogenesis of rheumatoid arthritis as a novel cytokine. Arthritis Rheum. 2003;48:971–981. [PubMed] [Google Scholar]
97. Pullerits R, Jonsson IM, Verdrengh M, Bokarewa M, Andersson U, Erlandsson-Harris H, Tarkowski A. High mobility group box chromosomal protein 1, a DNA binding cytokine, induces arthritis. Arthritis Rheum. 2003;48:1693–1700. [PubMed] [Google Scholar]
98. Ek M, Popovic K, Harris HE, Naucler CS, Wahren-Herlenius M. Increased extracellular levels of the novel proinflammatory cytokine high mobility group box chromosomal protein 1 in minor salivary glands of patients with Sjogren's syndrome. Arthritis Rheum. 2006;54:2289–2294. [PubMed] [Google Scholar]
99. Popovic K, Ek M, Espinosa A, Padyukov L, Harris HE, Wahren-Herlenius M, Nyberg F. Increased expression of the novel proinflammatory cytokine high mobility group box chromosomal protein 1 in skin lesions of patients with lupus erythematosus. Arthritis Rheum. 2005;52:3639–3645. [PubMed] [Google Scholar]
100. Barsness KA, Arcaroli J, Harken AH, Abraham E, Banerjee A, Reznikov L, McIntyre RC. Hemorrhage-induced acute lung injury is TLR-4 dependent. Am J Physiol Regul Integr Comp Physiol. 2004;287:R592–R599. [PubMed] [Google Scholar]
101. Kim JY, Park JS, Strassheim D, Douglas I, Diaz Del Valle F, Asehnoune K, Mitra S, Kwak SH, Yamada S, Maruyama I, et al. HMGB1 contributes to the development of acute lung injury after hemorrhage. Am J Physiol Lung Cell Mol Physiol. 2005;288:L958–L965. [PubMed] [Google Scholar]
102. Yang R, Harada T, Mollen KP, Prince JM, Levy RM, Englert JA, Gallowitsch-Puerta M, Yang L, Yang H, Tracey KJ, et al. Anti-HMGB1 neutralizing antibody ameliorates gut barrier dysfunction and improves survival after hemorrhagic shock. Mol Med. 2006;12:105–114. [PMC free article] [PubMed] [Google Scholar]
103. Levy RM, Mollen KP, Prince JM, Kaczorowski DJ, Vallabhaneni R, Liu S, Tracey KJ, Lotze MT, Hackam DJ, Fink MP, et al. Systemic inflammation and remote organ injury following trauma require HMGB1. Am J Physiol Regul Integr Comp Physiol. 2007;293:R1538–R1544. [PubMed] [Google Scholar]
104. Tsung A, Sahai R, Tanaka H, Nakao A, Fink MP, Lotze MT, Yang H, Li J, Tracey KJ, Geller DA, et al. The nuclear factor HMGB1 mediates hepatic injury after murine liver ischemia-reperfusion. J Exp Med. 2005;201:1135–1143. [PMC free article] [PubMed] [Google Scholar]
105. Goldstein RS, Gallowitsch-Puerta M, Yang L, Rosas-Ballina M, Huston JM, Czura CJ, Lee DC, Ward MF, Bruchfeld AN, Wang H, et al. Elevated high-mobility group box 1 levels in patients with cerebral and myocardial ischemia. Shock. 2006;25:571–574. [PubMed] [Google Scholar]
106. Wu H, Chen G, Wyburn KR, Yin J, Bertolino P, Eris JM, Alexander SI, Sharland AF, Chadban SJ. TLR4 activation mediates kidney ischemia/reperfusion injury. J Clin Invest. 2007;117:2847–2859. [PMC free article] [PubMed] [Google Scholar]
107. Faraco G, Fossati S, Bianchi ME, Patrone M, Pedrazzi M, Sparatore B, Moroni F, Chiarugi A. High mobility group box 1 protein is released by neural cells upon different stresses and worsens ischemic neurodegeneration in vitro and in vivo. J Neurochem. 2007;103:590–603. [PubMed] [Google Scholar]
108. Urbonaviciute V, Furnrohr BG, Meister S, Munoz L, Heyder P, De Marchis F, Bianchi ME, Kirschning C, Wagner H, Manfredi AA, et al. Induction of inflammatory and immune responses by HMGB1-nucleosome complexes: implications for the pathogenesis of SLE. J Exp Med. 2008;205:3007–3018. [PMC free article] [PubMed] [Google Scholar]
109. Mitola S, Belleri M, Urbinati C, Coltrini D, Sparatore B, Pedrazzi M, Melloni E, Presta M. Cutting edge: extracellular high mobility group box-1 protein is a proangiogenic cytokine. J Immunol. 2006;176:12–15. [PubMed] [Google Scholar]
110. Schlueter C, Weber H, Meyer B, Rogalla P, Roser K, Hauke S, Bullerdiek J. Angiogenetic signaling through hypoxia: HMGB1: an angiogenetic switch molecule. Am J Pathol. 2005;166:1259–1263. [PMC free article] [PubMed] [Google Scholar]
111. Palumbo R, Sampaolesi M, De Marchis F, Tonlorenzi R, Colombetti S, Mondino A, Cossu G, Bianchi ME. Extracellular HMGB1, a signal of tissue damage, induces mesoangioblast migration and proliferation. J Cell Biol. 2004;164:441–449. [PMC free article] [PubMed] [Google Scholar]
112. Chavakis E, Hain A, Vinci M, Carmona G, Bianchi ME, Vajkoczy P, Zeiher AM, Chavakis T, Dimmeler S. High-mobility group box 1 activates integrin-dependent homing of endothelial progenitor cells. Circ Res. 2007;100:204–212. [PubMed] [Google Scholar]
113. De Mori R, Straino S, Di Carlo A, Mangoni A, Pompilio G, Palumbo R, Bianchi ME, Capogrossi MC, Germani A. Multiple effects of high mobility group box protein 1 in skeletal muscle regeneration. Arterioscler Thromb Vasc Biol. 2007;27:2377–2383. [PubMed] [Google Scholar]
114. Degryse B, Bonaldi T, Scaffidi P, Muller S, Resnati M, Sanvito F, Arrigoni G, Bianchi ME. The high mobility group (HMG) boxes of the nuclear protein HMG1 induce chemotaxis and cytoskeleton reorganization in rat smooth muscle cells. J Cell Biol. 2001;152:1197–1206. [PMC free article] [PubMed] [Google Scholar]
115. Porto A, Palumbo R, Pieroni M, Aprigliano G, Chiesa R, Sanvito F, Maseri A, Bianchi ME. Smooth muscle cells in human atherosclerotic plaques secrete and proliferate in response to high mobility group box 1 protein. FASEB J. 2006;20:2565–2566. [PubMed] [Google Scholar]
116. Sorci G, Riuzzi F, Arcuri C, Giambanco I, Donato R. Amphoterin stimulates myogenesis and counteracts the antimyogenic factors basic fibroblast growth factor and S100B via RAGE binding. Mol Cell Biol. 2004;24:4880–4894. [PMC free article] [PubMed] [Google Scholar]
117. Riuzzi F, Sorci G, Donato R. The amphoterin (HMGB1)/receptor for advanced glycation end products (RAGE) pair modulates myoblast proliferation, apoptosis, adhesiveness, migration, and invasiveness. Functional inactivation of RAGE in L6 myoblasts results in tumor formation in vivo. J Biol Chem. 2006;281:8242–8253. [PubMed] [Google Scholar]
118. Limana F, Germani A, Zacheo A, Kajstura J, Di Carlo A, Borsellino G, Leoni O, Palumbo R, Battistini L, Rastaldo R, et al. Exogenous high-mobility group box 1 protein induces myocardial regeneration after infarction via enhanced cardiac C-kit+ cell proliferation and differentiation. Circ Res. 2005;97:e73–e83. [PubMed] [Google Scholar]
119. Straino S, Di Carlo A, Mangoni A, De Mori R, Guerra L, Maurelli R, Panacchia L, Di Giacomo F, Palumbo R, Di Campli C, et al. High-mobility group box 1 protein in human and murine skin: involvement in wound healing. J Invest Dermatol. 2008;128:1545–1553. [PubMed] [Google Scholar]
120. Ellerman JE, Brown CK, de Vera M, Zeh HJ, Billiar T, Rubartelli A, Lotze MT. Masquerader: high mobility group box-1 and cancer. Clin Cancer Res. 2007;13:2836–2848. [PubMed] [Google Scholar]
121. Klune JR, Dhupar R, Cardinal J, Billiar TR, Tsung A. HMGB1: endogenous danger signaling. Mol Med. 2008;14:476–484. [PMC free article] [PubMed] [Google Scholar]
122. Rauvala H, Huttunen HJ, Fages C, Kaksonen M, Kinnunen T, Imai S, Raulo E, Kilpelainen I. Heparin-binding proteins HB-GAM (pleiotrophin) and amphoterin in the regulation of cell motility. Matrix Biol. 2000;19:377–387. [PubMed] [Google Scholar]
123. Parkkinen J, Rauvala H. Interactions of plasminogen and tissue plasminogen activator (t-PA) with amphoterin. Enhancement of t-PA-catalyzed plasminogen activation by amphoterin. J Biol Chem. 1991;266:16730–16735. [PubMed] [Google Scholar]
124. Taguchi A, Blood DC, del Toro G, Canet A, Lee DC, Qu W, Tanji N, Lu Y, Lalla E, Fu C, et al. Blockade of RAGE-amphoterin signalling suppresses tumour growth and metastases. Nature. 2000;405:354–360. [PubMed] [Google Scholar]
125. Nestl A, Von Stein OD, Zatloukal K, Thies WG, Herrlich P, Hofmann M, Sleeman JP. Gene expression patterns associated with the metastatic phenotype in rodent and human tumors. Cancer Res. 2001;61:1569–1577. [PubMed] [Google Scholar]
126. Poser I, Golob M, Buettner R, Bosserhoff AK. Upregulation of HMG1 leads to melanoma inhibitory activity expression in malignant melanoma cells and contributes to their malignancy phenotype. Mol Cell Biol. 2003;23:2991–2998. [PMC free article] [PubMed] [Google Scholar]
127. Maeda S, Hikiba Y, Shibata W, Ohmae T, Yanai A, Ogura K, Yamada S, Omata M. Essential roles of high-mobility group box 1 in the development of murine colitis and colitis-associated cancer. Biochem Biophys Res Commun. 2007;360:394–400. [PubMed] [Google Scholar]
128. Kuniyasu H, Yano S, Sasaki T, Sasahira T, Sone S, Ohmori H. Colon cancer cell-derived high mobility group 1/amphoterin induces growth inhibition and apoptosis in macrophages. Am J Pathol. 2005;166:751–760. [PMC free article] [PubMed] [Google Scholar]
129. Lotze MT, Tracey KJ. High-mobility group box 1 protein (HMGB1): nuclear weapon in the immune arsenal. Nat Rev Immunol. 2005;5:331–342. [PubMed] [Google Scholar]
130. Tesniere A, Panaretakis T, Kepp O, Apetoh L, Ghiringhelli F, Zitvogel L, Kroemer G. Molecular characteristics of immunogenic cancer cell death. Cell Death Differ. 2008;15:3–12. [PubMed] [Google Scholar]
131. Dong Xda E, Ito N, Lotze MT, Demarco RA, Popovic P, Shand SH, Watkins S, Winikoff S, Brown CK, Bartlett DL, et al. High mobility group box I (HMGB1) release from tumor cells after treatment: implications for development of targeted chemoimmunotherapy. J Immunother. 2007;30:596–606. [PubMed] [Google Scholar]
132. Donato R. S100: a multigenic family of calcium-modulated proteins of the EF-hand type with intracellular and extracellular functional roles. Int J Biochem Cell Biol. 2001;33:637–668. [PubMed] [Google Scholar]
133. Roth J, Vogl T, Sorg C, Sunderkotter C. Phagocyte-specific S100 proteins: a novel group of proinflammatory molecules. Trends Immunol. 2003;24:155–158. [PubMed] [Google Scholar]
134. Marenholz I, Heizmann CW, Fritz G. S100 proteins in mouse and man: from evolution to function and pathology (including an update of the nomenclature) Biochem Biophys Res Commun. 2004;322:1111–1122. [PubMed] [Google Scholar]
135. Heizmann CW, Fritz G, Schafer BW. S100 proteins: structure, functions and pathology. Front Biosci. 2002;7:d1356–d1368. [PubMed] [Google Scholar]
136. Gebhardt C, Nemeth J, Angel P, Hess J. S100A8 and S100A9 in inflammation and cancer. Biochem Pharmacol. 2006;72:1622–1631. [PubMed] [Google Scholar]
137. Roth J, Goebeler M, Sorg C. S100A8 and S100A9 in inflammatory diseases. Lancet. 2001;357:1041. [PubMed] [Google Scholar]
138. Foell D, Wittkowski H, Vogl T, Roth J. S100 proteins expressed in phagocytes: a novel group of damage-associated molecular pattern molecules. J Leukoc Biol. 2007;81:28–37. [PubMed] [Google Scholar]
139. Lagasse E, Clerc RG. Cloning and expression of two human genes encoding calcium-binding proteins that are regulated during myeloid differentiation. Mol Cell Biol. 1998;8:2402–2410. [PMC free article] [PubMed] [Google Scholar]
140. Odink K, Cerletti N, Bruggen J, Clerc RG, Tarcsay L, Zwadlo G, Gerhards G, Schlegel R, Sorg C. Two calcium-binding proteins in infiltrate macrophages of rheumatoid arthritis. Nature. 1987;330:80–82. [PubMed] [Google Scholar]
141. Cheng P, Corzo CA, Luetteke N, Yu B, Nagaraj S, Bui MM, Ortiz M, Nacken W, Sorg C, Vogl T, et al. Inhibition of dendritic cell differentiation and accumulation of myeloid-derived suppressor cells in cancer is regulated by S100A9 protein. J Exp Med. 2008;205:2235–2249. [PMC free article] [PubMed] [Google Scholar]
142. Yang Z, Tao T, Raftery MJ, Youssef P, Di Girolamo N, Geczy CL. Proinflammatory properties of the human S100 protein S100A12. J Leukoc Biol. 2001;69:986–994. [PubMed] [Google Scholar]
143. Fuellen G, Foell D, Nacken W, Sorg C, Kerkhoff C. Absence of S100A12 in mouse: implications for RAGE-S100A12 interaction. Trends Immunol. 2003;24:622–624. [PubMed] [Google Scholar]
144. Rammes A, Roth J, Goebeler M, Klempt M, Hartmann M, Sorg C. Myeloid-related protein (MRP) 8 andMRP14, calcium-binding proteins of the S100 family, are secreted by activated monocytes via a novel, tubulin-dependent pathway. J Biol Chem. 1997;272:9496–9502. [PubMed] [Google Scholar]
145. Donato R. RAGE: a single receptor for several ligands and different cellular responses: the case of certain S100 proteins. Curr Mol Med. 2007;7:711–724. [PubMed] [Google Scholar]
146. Leclerc E, Fritz G, Vetter SW, Heizmann CW. Binding of S100 proteins to RAGE: an update. Biochim Biophys Acta. in press. [PubMed] [Google Scholar]
147. Passey RJ, Williams E, Lichanska AM, Wells C, Hu S, Geczy CL, Little MH, Hume DA. A null mutation in the inflammation-associated S100 protein S100A8 causes early resorption of themouse embryo. J Immunol. 1999;163:2209–2216. [PubMed] [Google Scholar]
148. Hobbs JA, May R, Tanousis K, McNeill E, Mathies M, Gebhardt C, Henderson R, Robinson MJ, Hogg N. Myeloid cell function in MRP-14 (S100A9) null mice. Mol Cell Biol. 2003;23:2564–2576. [PMC free article] [PubMed] [Google Scholar]
149. Manitz MP, Horst B, Seeliger S, Strey A, Skryabin BV, Gunzer M, Frings W, Schonlau F, Roth J, Sorg C, et al. Loss of S100A9 (MRP14) results in reduced interleukin-8-induced CD11b surface expression, a polarized microfilament system, and diminished responsiveness to chemoattractants in vitro. Mol Cell Biol. 2003;23:1034–1043. [PMC free article] [PubMed] [Google Scholar]
150. Pouliot P, Plante I, Raquil MA, Tessier PA, Olivier M. Myeloid-related proteins rapidly modulate macrophage nitric oxide production during innate immune response. J Immunol. 2008;181:3595–3601. [PubMed] [Google Scholar]
151. Sunahori K, Yamamura M, Yamana J, Takasugi K, Kawashima M, Yamamoto H, Chazin WJ, Nakatani Y, Yui S, Makino H. The S100A8/A9 heterodimer amplifies proinflammatory cytokine production by macrophages via activation of nuclear factor kappa B and p38 mitogen-activated protein kinase in rheumatoid arthritis. Arthritis Res Ther. 2006;8:R69. [PMC free article] [PubMed] [Google Scholar]
152. Hofmann MA, Drury S, Fu C, Qu W, Taguchi A, Lu Y, Avila C, Kambham N, Bierhaus A, Nawroth P, et al. RAGE mediates a novel proinflammatory axis: a central cell surface receptor for S100/calgranulin polypeptides. Cell. 1999;97:889–901. [PubMed] [Google Scholar]
153. Ehlermann P, Eggers K, Bierhaus A, Most P, Weichenhan D, Greten J, Nawroth PP, Katus HA, Remppis A. Increased proinflammatory endothelial response to S100A8/A9 after preactivation through advanced glycation end products. Cardiovasc Diabetol. 2006;5:6. [PMC free article] [PubMed] [Google Scholar]
154. Viemann D, Strey A, Janning A, Jurk K, Klimmek K, Vogl T, Hirono K, Ichida F, Foell D, Kehrel B, et al. Myeloid-related protein 8 and 14 induce a specific inflammatory response in human microvascular endothelial cells. Blood. 2004;105:2955–2962. [PubMed] [Google Scholar]
155. Salama I, Malone PS, Mihaimeed F, Jones JL. A review of the S100 proteins in cancer. Eur J Surg Oncol. 2008;34:357–364. [PubMed] [Google Scholar]
156. Ghavami S, Rashedi I, Dattilo BM, Eshraghi M, Chazin WJ, Hashemi M, Wesselborg S, Kerkhoff C, Los M. S100A8/A9 at low concentration promotes tumor cell growth via RAGE ligation and MAP kinase-dependent pathway. J Leukoc Biol. 2008;83:1484–1492. [PMC free article] [PubMed] [Google Scholar]
157. Sinha P, Okoro C, Foell D, Freeze HH, Ostrand-Rosenberg S, Srikrishna G. Proinflammatory s100 proteins regulate the accumulation of myeloid-derived suppressor cells. J Immunol. 2008;181:4666–4675. [PMC free article] [PubMed] [Google Scholar]
158. Ostrand-Rosenberg S. Cancer and complement. Nat Biotechnol. 2008;26:1348–1349. [PMC free article] [PubMed] [Google Scholar]
159. Hiratsuka S, Watanabe A, Aburatani H, Maru Y. Tumour-mediated upregulation of chemoattractants and recruitment of myeloid cells predetermines lung metastasis. Nat Cell Biol. 2006;8:1369–1375. [PubMed] [Google Scholar]
160. Kaplan RN, Rafii S, Lyden D. Preparing the “soil”: the premetastatic niche. Cancer Res. 2006;66:11089–11093. [PMC free article] [PubMed] [Google Scholar]
161. Logsdon CD, Fuentes MK, Huang EH, Arumugam T. RAGE and RAGE ligands in cancer. Curr Mol Med. 2007;7:777–789. [PubMed] [Google Scholar]
162. Rakoff-Nahoum S, Medzhitov R. Toll-like receptors and cancer. Nat Rev Cancer. 2009;9:57–63. [PubMed] [Google Scholar]
163. Akira S, Takeda K. Toll-like receptor signalling. Nat Rev Immunol. 2004;4:499–511. [PubMed] [Google Scholar]
164. Lee MS, Kim YJ. Signaling pathways downstream of pattern-recognition receptors and their cross talk. Annu Rev Biochem. 2007;76:447–480. [PubMed] [Google Scholar]
165. Miyake K. Innate immune sensing of pathogens and danger signals by cell surface Toll-like receptors. Semin Immunol. 2007;19:3–10. [PubMed] [Google Scholar]
166. Beutler B. Neo-ligands for innate immune receptors and the etiology of sterile inflammatory disease. Immunol Rev. 2007;220:113–128. [PubMed] [Google Scholar]
167. Park JS, Svetkauskaite D, He Q, Kim JY, Strassheim D, Ishizaka A, Abraham E. Involvement of Toll-like receptors 2 and 4 in cellular activation by high mobility group box 1 protein. J Biol Chem. 2004;279:7370–7377. [PubMed] [Google Scholar]
168. Yu M, Wang H, Ding A, Golenbock DT, Latz E, Czura CJ, Fenton MJ, Tracey KJ, Yang H. HMGB1 signals through toll-like receptor (TLR) 4 and TLR2. Shock. 2006;26:174–179. [PubMed] [Google Scholar]
169. van Zoelen MA, Yang H, Florquin S, Meijers JC, Akira S, Arnold B, Nawroth PP, Bierhaus A, Tracey KJ, van der Poll T. Role of Toll-like receptors 2 and 4, and the receptor for advanced glycation end products (Rage) in Hmgb1 induced inflammation in vivo. Shock. 2008;31:280–284. [PMC free article] [PubMed] [Google Scholar]
170. Zedler S, Faist E. The impact of endogenous triggers on trauma-associated inflammation. Curr Opin Crit Care. 2006;12:595–601. [PubMed] [Google Scholar]
171. Wang H, Yang H, Tracey KJ. Extracellular role of HMGB1 in inflammation and sepsis. J Intern Med. 2004;255:320–331. [PubMed] [Google Scholar]
172. Vogl T, Tenbrock K, Ludwig S, Leukert N, Ehrhardt C, van Zoelen MA, Nacken W, Foell D, van der Poll T, Sorg C, et al. Mrp8 and Mrp14 are endogenous activators of Toll-like receptor 4 promoting lethal, endotoxin-induced shock. Nat Med. 2007;13:1042–1049. [PubMed] [Google Scholar]
173. Hiratsuka S, Watanabe A, Sakurai Y, Akashi-Takamura S, Ishibashi S, Miyake K, Shibuya M, Akira S, Aburatani H, Maru Y. The S100A8-serum amyloid A3-TLR4 paracrine cascade establishes a pre-metastatic phase. Nat Cell Biol. 2008;10:1349–1355. [PubMed] [Google Scholar]
174. Kluwe J, Mencin A, Schwabe RF. Toll-like receptors, wound healing, and carcinogenesis. J Mol Med. 2009;87:125–138. [PMC free article] [PubMed] [Google Scholar]
175. Rakoff-Nahoum S, Medzhitov R. Regulation of spontaneous intestinal tumorigenesis through the adaptor protein MyD88. Science. 2007;317:124–127. [PubMed] [Google Scholar]
176. Swann JB, Vesely MD, Silva A, Sharkey J, Akira S, Schreiber RD, Smyth MJ. Demonstration of inflammation-induced cancer and cancer immunoediting during primary tumorigenesis. Proc Natl Acad Sci USA. 2008;105:652–656. [PMC free article] [PubMed] [Google Scholar]
177. Naugler WE, Sakurai T, Kim S, Maeda S, Kim K, Elsharkawy AM, Karin M. Gender disparity in liver cancer due to sex differences in MyD88-dependent IL-6 production. Science. 2007;317:121–124. [PubMed] [Google Scholar]
178. Fukata M, Chen A, Vamadevan AS, Cohen J, Breglio K, Krishnareddy S, Hsu D, Xu R, Harpaz N, Dannenberg AJ, et al. Toll-like receptor-4 promotes the development of colitis-associated colorectal tumors. Gastroenterology. 2007;133:1869–1881. [PMC free article] [PubMed] [Google Scholar]
179. Schmidt AM, Yan SD, Yan SF, Stern DM. The multiligand receptor RAGE as a progression factor amplifying immune and inflammatory responses. J Clin Invest. 2001;108:949–955. [PMC free article] [PubMed] [Google Scholar]
180. Bierhaus A, Humpert PM, Morcos M, Wendt T, Chavakis T, Arnold B, Stern DM, Nawroth PP. Understanding RAGE, the receptor for advanced glycation end products. J Mol Med. 2005;83:876–886. [PubMed] [Google Scholar]
181. Brett J, Schmidt AM, Yan SD, Zou YS, Weidman E, Pinsky D, Nowygrod R, Neeper M, Przysiecki C, Shaw A. Survey of the distribution of a newly characterized receptor for advanced glycation end products in tissues. Am J Pathol. 1993;143:1699–1712. [PMC free article] [PubMed] [Google Scholar]
182. Clynes R, Moser B, Yan SF, Ramasamy R, Herold K, Schmidt AM. Receptor for AGE (RAGE): weaving tangled webs within the inflammatory response. Curr Mol Med. 2007;7:743–751. [PubMed] [Google Scholar]
183. Yan SF, Ramasamy R, Schmidt AM. Receptor for AGE (RAGE) and its ligands-cast into leading roles in diabetes and the inflammatory response. J Mol Med. 2009;87:235–247. [PMC free article] [PubMed] [Google Scholar]
184. Chen X, Walker DG, Schmidt AM, Arancio O, Lue LF, Yan SD. RAGE: a potential target for Abeta-mediated cellular perturbation in Alzheimer's disease. Curr Mol Med. 2007;7:735–742. [PubMed] [Google Scholar]
185. Neeper M, Schmidt AM, Brett J, Yan SD, ang F, Pan YC, Elliston K, Stern D, Shaw A. Cloning and expression of a cell surface receptor for advanced glycosylation end products of proteins. J Biol Chem. 1992;267:14998–15004. [PubMed] [Google Scholar]
186. Hudson BI, Carter AM, Harja E, Kalea AZ, Arriero M, Yang H, Grant PJ, Schmidt AM. Identification, classification, and expression of RAGE gene splice variants. FASEB J. 2008;22:1572–1580. [PubMed] [Google Scholar]
187. Hori O, Brett J, Slattery T, Cao R, Zhang J, Chen JX, Nagashima M, Lundh ER, Vijay S, Nitecki D, et al. The receptor for advanced glycation end products (RAGE) is a cellular binding site for amphoterin. Mediation of neurite outgrowth and co-expression of RAGE and amphoterin in the developing nervous system. J Biol Chem. 1995;270:25752–25761. [PubMed] [Google Scholar]
188. Parkkinen J, Raulo E, Merenmies J, Nolo R, Kajander EO, Baumann M, Rauvala H. Amphoterin, the 30-kDa protein in a family of HMG1-type polypeptides. Enhanced expression in transformed cells, leading edge localization, and interactions with plasminogen activation. J Biol Chem. 1993;268:19726–19738. [PubMed] [Google Scholar]
189. Lotze MT, De Marco RA. Dealing with death: HMGB1 as a novel target for cancer therapy. Curr Opin Investig Drugs. 2003;4:1405–1409. [PubMed] [Google Scholar]
190. Sasahira T, Akama Y, Fujii K, Kuniyasu H. Expression of receptor for advanced glycation end products and MGB1/amphoterin in colorectal adenomas. Virchows Arch. 2005;446:411–415. [PubMed] [Google Scholar]
191. Bartling B, Hofmann HS, Weigle B, Silber RE, Simm A. Down-regulation of the receptor for advanced glycation end-products (RAGE) supports non-small cell lung carcinoma. Carcinogenesis. 2005;26:293–301. [PubMed] [Google Scholar]
192. Riuzzi F, Sorci G, Donato R. RAGE expression in rhabdomyosarcoma cells results in myogenic differentiation and reduced proliferation, migration, invasiveness, and tumor growth. Am J Pathol. 2007;171:947–961. [PMC free article] [PubMed] [Google Scholar]
193. Tateno T, Ueno S, Hiwatashi K, Matsumoto M, Okumura H, Setoyama T, Uchikado Y, Sakoda M, Kubo F, Ishigami S, et al. Expression of receptor for advanced glycation end products (RAGE) is related to prognosis in patients with esophageal squamous cell carcinoma. Ann Surg Oncol. 2009;16:440–446. [PubMed] [Google Scholar]
194. Landesberg R, Woo V, Huang L, Cozin M, Lu Y, Bailey C, Qu W, Pulse C, Schmidt AM. The expression of the receptor for glycation end-products (RAGE) in oral squamous cell carcinomas. Oral Surg Oral Med Oral Pathol Oral Radiol Endod. 2008;105:617–624. [PubMed] [Google Scholar]
195. Takeuchi A, Yamamoto Y, Tsuneyama K, Cheng C, Yonekura H, Watanabe T, Shimizu K, Tomita K, Yamamoto H, Tsuchiya H. Endogenous secretory receptor for advanced glycation endproducts as a novel prognostic marker in chondrosarcoma. Cancer. 2007;109:2532–2540. [PubMed] [Google Scholar]
196. Huttunen HJ, Kuja-Panula J, Sorci G, Agneletti AL, Donato R, Rauvala H. Coregulation of neurite outgrowth and cell survival by amphoterin and S100 proteins through receptor for advanced glycation end products (RAGE) activation. J Biol Chem. 2000;275:40096–40105. [PubMed] [Google Scholar]
197. Hsieh HL, Schafer BW, Sasaki N, Heizmann CW. Expression analysis of S100 proteins and RAGE in human tumors using tissue microarrays. Biochem Biophys Res Commun. 2003;307:375–381. [PubMed] [Google Scholar]
198. Donato R, Sorci G, Riuzzi F, Arcuri C, Bianchi R, Brozzi F, Tubaro C, Giambanco I. S100B's double life: intracellular regulator and extracellular signal. Biochim Biophys Acta. in press. [PubMed] [Google Scholar]
199. Yan SS, Wu ZY, Zhang HP, Furtado G, Chen X, Yan SF, Schmidt AM, Brown C, Stern A, LaFaille J, et al. Suppression of experimental autoimmune encephalomyelitis by selective blockade of encephalitogenic T-cell infiltration of the central nervous system. Nat Med. 2003;9:287–293. [PubMed] [Google Scholar]
200. Ding Y, Kantarci A, Hasturk H, Trackman PC, Malabanan A, Van Dyke TE. Activation of RAGE induces elevated O2- generation by mononuclear phagocytes in diabetes. J Leukoc Biol. 2007;81:520–527. [PMC free article] [PubMed] [Google Scholar]
201. Gao X, Zhang H, Schmidt AM, Zhang C. AGE/RAGE produces endothelial dysfunction in coronary arterioles in type 2 diabetic mice. Am J Physiol Heart Circ Physiol. 2008;295:H491–H498. [PMC free article] [PubMed] [Google Scholar]
202. Reddy MA, Li SL, Sahar S, Kim YS, Xu ZG, Lanting L, Natarajan R. Key role of Src kinase in S100B-induced activation of the receptor for advanced glycation end products in vascular smooth muscle cells. J Biol Chem. 2006;281:13685–13693. [PubMed] [Google Scholar]
203. Wolf R, Howard OM, Dong HF, Voscopoulos C, Boeshans K, Winston J, Divi R, Gunsior M, Goldsmith P, Ahvazi B, et al. Chemotactic activity of S100A7 (Psoriasin) is mediated by the receptor for advanced glycation end products and potentiates inflammation with highly homologous but functionally distinct S100A15. J Immunol. 2008;181:1499–1506. [PMC free article] [PubMed] [Google Scholar]
204. Hermani A, De Servi B, Medunjanin S, Tessier PA, Mayer D. S100A8 and S100A9 activate MAP kinase and NF-kappaB signaling pathways and trigger translocation of RAGE in human prostate cancer cells. Exp Cell Res. 2006;312:184–197. [PubMed] [Google Scholar]
205. Hermani A, Hess J, De Servi B, Medunjanin S, Grobholz R, Trojan L, Angel P, Mayer D. Calcium-binding proteins S100A8 and S100A9 as novel diagnostic markers in human prostate cancer. Clin Cancer Res. 2005;11:5146–5152. [PubMed] [Google Scholar]
206. Boyd JH, Kan B, Roberts H, Wang Y, Walley KR. S100A8 and S100A9 mediate endotoxin-induced cardiomyocyte dysfunction via the receptor for advanced glycation end products. Circ Res. 2008;102:1239–1246. [PubMed] [Google Scholar]
207. Cecil DL, Johnson K, Rediske J, Lotz M, Schmidt AM, Terkeltaub R. Inflammation-induced chondrocyte hypertrophy is driven by receptor for advanced glycation end products. J Immunol. 2005;175:8296–8302. [PubMed] [Google Scholar]
208. Fuentes MK, Nigavekar SS, Arumugam T, Logsdon CD, Schmidt AM, Park JC, Huang EH. RAGE activation by S100P in colon cancer stimulates growth, migration, and cell signaling pathways. Dis Colon Rectum. 2007;50:1230–1240. [PubMed] [Google Scholar]
209. Arumugam T, Simeone DM, Van Golen K, Logsdon CD. S100P promotes pancreatic cancer growth, survival, and invasion. Clin Cancer Res. 2005;11:5356–5364. [PubMed] [Google Scholar]
210. Bierhaus A, Schiekofer S, Schwaninger M, Andrassy M, Humpert PM, Chen J, Hong M, Luther T, Henle T, Kloting I, et al. Diabetes-associated sustained activation of the transcription factor nuclear factor-kappaB. Diabetes. 2001;50:2792–2808. [PubMed] [Google Scholar]
211. Srikrishna G, Panneerselvam K, Westphal V, Abraham V, Varki A, Freeze HH. Two proteins modulating transendothelial migration of leukocytes recognize novel carboxylated glycans on endothelial cells. J Immunol. 2001;166:4678–4688. [PubMed] [Google Scholar]
212. Srikrishna G, Huttunen HJ, Johansson L, Weigle B, Yamaguchi Y, Rauvala H, Freeze HH. N-glycans on the receptor for advanced glycation end products influence amphoterin binding and neurite outgrowth. J Neurochem. 2002;80:998–1008. [PubMed] [Google Scholar]
213. Salmivirta M, Rauvala H, Elenius K, Jalkanen M. Neurite growth-promoting protein (amphoterin, p30) binds syndecan. Exp Cell Res. 1992;200:444–451. [PubMed] [Google Scholar]
214. Robinson MJ, Tessier P, Poulsom R, Hogg N. The S100 family heterodimer, MRP-8/14, binds with high affinity to heparin and heparan sulfate glycosaminoglycans on endothelial cells. J Biol Chem. 2002;277:3658–3665. [PubMed] [Google Scholar]
215. Huttunen HJ, Fages C, Rauvala H. Receptor for advanced glycation end products (RAGE)-mediated neurite outgrowth and activation of NF-kappaB require the cytoplasmic domain of the receptor but different downstream signaling pathways. J Biol Chem. 1999;274:19919–19924. [PubMed] [Google Scholar]
216. Bianchi R, Giambanco I, Donato R. S100B/RAGE-dependent activation of microglia via NF-kappaB and AP-1 co-regulation of COX-2 expression by S100B, IL-1beta and TNF-alpha. Neurobiol Aging. in press. [PubMed] [Google Scholar]
217. Park JS, Gamboni-Robertson F, He Q, Svetkauskaite D, Kim JY, Strassheim D, Sohn JW, Yamada S, Maruyama I, Banerjee A, et al. High mobility group box 1 protein interacts with multiple Toll-like receptors. Am J Physiol Cell Physiol. 2006;290:C917–C924. [PubMed] [Google Scholar]
218. van Beijnum JR, Buurman WA, Griffioen AW. Convergence and amplification of toll-like receptor (TLR) and receptor for advanced glycation end products (RAGE) signaling pathways via high mobility group B1 (HMGB1) Angiogenesis. 2008;11:91–99. [PubMed] [Google Scholar]
219. Kawai T, Akira S. Signaling to NF-kappaB by Toll-like receptors. Trends Mol Med. 2007;13:460–469. [PubMed] [Google Scholar]
220. Huttunen HJ, Fages C, Kuja-Panula J, Ridley AJ, Rauvala H. Receptor for advanced glycation end products-binding COOH-terminal motif of amphoterin inhibits invasive migration and metastasis. Cancer Res. 2002;62:4805–4811. [PubMed] [Google Scholar]
221. Xie J, Burz DS, He W, Bronstein IB, Lednev I, Shekhtman A. Hexameric calgranulin C (S100A12) binds to the receptor for advanced glycated end products (RAGE) using symmetric hydrophobic target-binding patches. J Biol Chem. 2007;282:4218–4231. [PubMed] [Google Scholar]
222. Leclerc E, Fritz G, Weibel M, Heizmann CW, Galichet A. S100B and S100A6 differentially modulate cell survival by interacting with distinct RAGE (receptor for advanced glycation end products) immunoglobulin domains. J Biol Chem. 2007;282:31317–31331. [PubMed] [Google Scholar]
223. Dattilo BM, Fritz G, Leclerc E, Kooi CW, Heizmann CW, Chazin WJ. The extracellular region of the receptor for advanced glycation end products is composed of two independent structural units. Biochemistry. 2007;46:6957–6970. [PMC free article] [PubMed] [Google Scholar]
224. Ostendorp T, Leclerc E, Galichet A, Koch M, Demling N, Weigle B, Heizmann CW, Kroneck PM, Fritz G. Structural and functional insights into RAGE activation by multimeric S100B. EMBO J. 2007;26:3868–3878. [PMC free article] [PubMed] [Google Scholar]
225. Norgard-Sumnicht KE, Roux L, Toomre DK, Manzi A, Freeze HH, Varki A. Unusual anionic N-linked oligosaccharides from bovine lung. J Biol Chem. 1995;270:27634–27645. [PubMed] [Google Scholar]
226. Srikrishna G, Brive L, Freeze H. Novel carboxylated N-glycans contain oligosaccharide-linked glutamic acid. Biochem Biophys Res Commun. 2005;332:1020–1027. [PubMed] [Google Scholar]
227. Srikrishna G, Toomre D, Manzi A, Panneerselvam K, Freeze H, Varki A, Varki N. A novel anionic modification of N-glycans on mammalian endothelial cells is recognized by activated neutrophils and modulates acute inflammatory responses. J Immunol. 2001;166:624–632. [PubMed] [Google Scholar]
228. Srikrishna G, Turovskaya O, Shaikh R, Newlin R, Foell D, Murch S, Kronenberg M, Freeze HH. Carboxylated glycans mediate colitis through activation of NF-{kappa}B. J Immunol. 2005;175:5412–5422. [PubMed] [Google Scholar]
229. Allmen EU, Koch M, Fritz G, Legler DF. V domain of RAGE interacts with AGEs on prostate carcinoma cells. Prostate. 2008;68:748–758. [PubMed] [Google Scholar]
230. Moroz OV, Antson AA, Dodson EJ, Burrell HJ, Grist SJ, Lloyd RM, Maitland NJ, Dodson GG, Wilson KS, Lukanidin E, et al. The structure of S100A12 in a hexameric form and its proposed role in receptor signalling. Acta Crystallogr D Biol Crystallogr. 2002;58:407–413. [PubMed] [Google Scholar]
231. Sacchettini JC, Baum LG, Brewer CF. Multivalent proteincarbohydrate interactions. A new paradigm for supermolecular assembly and signal transduction. Biochemistry. 2001;40:3009–3015. [PubMed] [Google Scholar]
232. Demetriou M, Granovsky M, Quaggin S, Dennis JW. Negative regulation of T-cell activation and autoimmunity by Mgat5 N-glycosylation. Nature. 2001;409:733–739. [PubMed] [Google Scholar]

Articles from Neoplasia (New York, N.Y.) are provided here courtesy of Neoplasia Press

-