Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2011 Jul 1.
Published in final edited form as: Physiol Rev. 2011 Jan;91(1):221–264. doi: 10.1152/physrev.00052.2009

Hyaluronan as an Immune Regulator in Human Diseases

PAUL W NOBLE 1, JIURONG LIANG 1, DIANHUA JIANG 1
PMCID: PMC3051404  NIHMSID: NIHMS173532  PMID: 21248167

Abstract

Accumulation and turnover of extracellular matrix components are the hallmarks of tissue injury. Fragmented hyaluronan stimulates the expression of inflammatory genes by a variety of immune cells at the injury site. Hyaluronan binds to a number of cell surface proteins on a variety of cell types. Hyaluronan fragments signal through both Toll-like receptor (TLR) 4 and TLR2 as well as CD44 to stimulate inflammatory genes in inflammatory cells. Hyaluronan is also present on the cell surface of epithelial cells and provides protection against tissue damage by interacting with TLR2 and TLR4 on these parenchymal cells. Hyaluronan and hyaluronan-binding proteins regulate inflammation, tissue injury and repair through regulating inflammatory cell recruitment, release of inflammatory cytokines, and stem cell migration. This review focuses on the role of hyaluronan as an immune regulator in human diseases.

I. HYALURONAN

Extracellular matrix (ECM) plays an essential role in organogenesis, growth, function, and in many human diseases. Hyaluronan (or hyaluronic acid, HA), a major ECM component, is a non-sulfated glycosaminoglycan composed of repeating polymeric disaccharides D-glucuronic acid and N-acetyl-D-glucosamine linked by a glucuronidic β(1→3) bond (463, 464) (Figure 1. Hyaluronan structure). Hyaluronan forms specific stable tertiary structures in aqueous solution (374). During the last two decades, significant progress has been made in understanding the role of HA in both biological and pathological states, and the mechanisms of HA synthesis and degradation, and the mechanisms of HA regulation of these biological processes. There are several reviews on the roles of HA in different fields, such as in angiogenesis (388), reactive oxygen species (392), cancer (433, 434), chondrocytes (200), lung injury (303, 438), and in immune regulation (424). The current review will briefly summarize recent advances in understanding HA biology, and will focus on HA as an immune regulator in physiological and pathological conditions.

Figure 1. Hyaluronan structure.

Figure 1

Hyaluronan is composed of repeating polymeric disaccharides D-glucuronic acid (GlcA) and N-acetyl-D-glucosamine (GlcNAc) linked by a glucuronidic β(1→3) bond. Three disaccharide GlcA-GlcNAc are shown.

Hyaluronan is widely distributed from lower organisms such as simple bacteria (241, 243) to complex eukaryotes (333). In humans, HA is abundant in the vitreous of the eye (271), the umbilical cord (463), synovial fluid (127, 340), heart valves (435), skin (9, 182, 416), and skeletal tissues (9). HA can be released by many cell types (227, 432), although mesenchymal cells are believed to be predominant (433). Importantly, almost all the cell types are responsive to HA stimulation by changing their cell secretory and behavioral properties (177). HA-stem cell interactions have been particularly explored in hematopoietic stem cells (15, 193, 301), mesenchymal stem cells (43, 500, 504), and adult multipotent progenitor cells, pointing to the idea that HA and its binding proteins may regulate tissue injury and repair processes through their interaction with a variety of stem cells.

Hyaluronan has multiple functions, such as space filling, lubrication of joints, and provision of a matrix through which cells can migrate (433). HA is actively produced during tissue injury, tissue repair, and wound healing (182, 390). In addition to providing a framework for ingrowths of blood vessels (388, 390) and fibroblasts (234, 439), HA also regulates many aspects of molecular mechanisms of tissue repair, such as activation of inflammatory cells to mount an immunological response (149, 265, 422) and regulation of behavior of epithelial cells (22, 165, 175, 176, 505) and fibroblasts (17, 157). Elucidation of the role and mechanisms of HA is crucial in aiding the development of novel therapy for many diseases.

II. BIOSYNTHESIS OF HYALURONAN

Hyaluronan is synthesized by membrane-bound synthases on the inner surface of the plasma membrane (332), and the chains are extruded through pore-like structures into the extracellular space (333, 455). There are three mammalian hyaluronan synthases (HAS) (159, 260, 397, 460). The enzymes use UDP-α-N-acetyl-D-glucosamine and UDP-α-D-glucuronate as substrates (461) (Figure 2. HA synthase reaction). Based on their similarities and differences, the known HA synthase proteins have been divided into two categories, designated as Class I and Class II (71). The Streptococcal HA synthases and eukaryotic HA synthases are Class I family members, whereas the HA synthase from Pasteurella multocida is the only Class II member (71).

Figure 2. HA synthase reaction.

Figure 2

Hyaluronan synthase catalyzes the reaction by adding N-acetyl-D-glucosamine (GlcNAc) and D-glucuronic acid (GlcA) alternatively to expand HA chain.

Much of our knowledge of the mechanism of HAS has been based on bacterial HA synthases (334, 335). Using Pasteurella multocida as a model, Prehm proposed that a two-site mechanism in which the reducing end sugar of the growing HA chain would remain covalently bound to a terminal uridine diphosphate (UDP), and the next sugar to be added from the second site would be transferred as the UDP-sugar onto the reducing end sugar to displace its terminal UDP. The HA chain would then be in the second site (334, 335).

The major advance in this field was fueled by the cloning of HAS from prokaryotes and mammals (72, 73). The DeAngelis group reported the molecular cloning and characterization of the Group A Streptococcal gene encoding the protein HasA, as the S. pyogenes HAS was later proven to be responsible for HA synthesis (72, 73). Expression of these genes in either acapsular Streptococcus strains or Enteroccus faecalis conferred the organisms with the ability to synthesize HA and form a capsule, thus demonstrating that HasA is a bona fide HA synthase (72, 73). Subsequent cloning of mouse HAS (158, 394) and human HAS (383, 455) revealed high homology in their protein sequences among humans, mice, frogs, and even bacteria (395, 461). The amino acid sequence of human HAS1 shows significant homology to the hasA gene product of Streptococcus pyogenes, a glycosaminoglycan synthase from Xenopus laevis, and a murine HA synthase (455). Genomic location and genomic structure of these HA synthases have been determined (396). Since HAS1, HAS2, and HAS3 are located on different autosomes (396), suggesting that the HAS gene family may have arisen comparatively early in vertebrate evolution by sequential duplication of an ancestral HAS gene.

All the HAS isozymes are highly homologous in their amino acid sequences and have similar hydropathic features, suggesting that they are similarly organized within the membrane. The Weigel group proposed membrane topology for the HAS family proteins (461). Two types of membrane domains are present: transmembrane domains that span the membrane and membrane-associated domains that do not go all the way through the membrane. There are 6 – 8 transmembrane domains and two membrane-associated domains. Over 60% of the whole protein (including the amino and carboxyl termini) are inside the cell. Only about 5% of the protein is exposed to the outside of the cell (136).

In Xenopus, xhas1 produces HA with a molecular mass of around 40–200 kDa, while the product formed by xhas2 has a molecular mass above 1000 kDa (206). Expression of mammalian HA synthases led to HA biosynthesis in transfected mammalian cells (160, 395). The HAS1 protein alone is able to synthesize HA, and different amino acid residues on the cytoplasmic central loop domain are involved in transferring N-acetyl-D-glucosamine and D-glucuronic acid residues. HAS3 synthesizes HA with a smaller molecular mass than HA synthesized by HAS1 and HAS2 (160). Furthermore, comparisons of HA secreted into the culture media by stable HAS transfectants show that HAS1 and HAS3 generated HA with broad size distributions (molecular masses of 2 × 105 to approximately 2 × 106 Da), whereas HAS2 generated HA with a broad but extremely large size (average molecular mass of > 2 × 106 Da) (160). Subsequent studies suggested that all three HAS enzymes drive the biosynthesis and release of high molecular weight HA (1 × 106 Da) (398).

A. Hyaluronan synthase 1 (HAS1)

Human HAS1 gene is located at 19q13.3-q13.4, whereas mouse Has1 is located at Chromosome 17 (396). Human HAS1 protein shows significant homology with the hasA gene product of Streptococcus pyogenes, a glycosaminoglycan synthetase (DG42) from Xenopus laevis, and a murine HA synthase (395). HAS1 protein is a bona fide synthase since it alone is able to synthesize HA and different amino acid residues on the cytoplasmic central loop domain are involved in transferring N-acetyl-D-glucosamine and D-glucuronic acid residues, respectively (488).

HA accumulation and dysregulated expression of HA synthases have been demonstrated in many diseases in both animal studies and in humans. HA deposition is prominent in MRL-Fas(lpr) mice with renal disease and could be mediated by local synthesis through HAS1 and HAS2. The enhanced synthesis of HA could be promoted by proinflammatory cytokines in vivo (96). Has1 mRNA was expressed predominantly in bone marrow mesenchymal progenitor cells derived from multiple myeloma patients when compared to mesenchymal progenitor cells from normal individuals (43). Bone marrow mesenchymal progenitor cells from myeloma synthesize more HA than those from healthy donors, suggesting that myeloma mesenchymal progenitor cells could be an important component of the myeloma pathophysiology in vivo by their increased expression of extracellular matrix components relevant to plasma cell growth and survival (43). Accumulation of HA is a hallmark of rheumatoid arthritis. In human fibroblast-like synoviocytes, HAS1 expression can be enhanced by TGF-β transcription. The TGF-β-induced transcription is mediated by a p38 MEK dependent, not JNK, pathway. TGF-β treatment leads to an increase in synthase activity and in HA production (402).

B. Hyaluronan synthase 2 (HAS2)

Human hyaluronan synthase 2 is located on chromosome 8, whereas mouse Has2 is on chromosome 15 (396). In an in vitro transfection assay, Kimata and associates demonstrated that all three isoforms of HA synthase exhibited a de novo, massive formation of a HA matrix in nontransformed rat 3Y1 embryonic fibroblast cells (157). Hyaluronan and its matrix can modulate contact inhibition of cell growth and migration (157). Inhibition of a phosphatidylinositol 3-kinase pathway resulted in reacquisition of the normal phenotype of HAS2 transfectants, suggesting that the intracellular phosphatidylinositol 3-kinase signaling regulates diminution of contact inhibition induced by formation of the massive HA matrix (157).

Targeted deletion of HAS2 has been a major development in the field, providing in vivo evidence of the functions of HA (45). There are major abnormalities in heart and blood vessel development, resulting in an embryonic lethal phenotype (44, 45). HAS2-deficient embryos at embryonic day 9.5 completely lack endocardial cushions (44, 45), consistent with early observations that high content of HA is present in normal human heart valves (435). These defects resemble the features in naturally occurring versican knockouts (276). These studies suggest that HA-ECM interactions play a significant role in cardiac development. Elucidating the role for HAS2 in animal models has been difficult due to the embryonic lethal phenotype in the HAS2-deficient mice.

HAS2 also plays a role in limb development. Overexpression of Has2 in the mesoderm of the chick limb bud in vivo results in the formation of shortened and severely malformed limbs that lack one or more skeletal elements. Skeletal elements in limbs overexpressing Has2 are reduced in length, exhibit abnormal morphology, and are positioned inappropriately (236). In vitro, sustained HA production in micromass cultures of limb mesenchymal cells inhibits formation of precartilage condensations and subsequent chondrogenesis, indicating that proper regulation of HA is essential for formation of the precartilage condensations that trigger cartilage differentiation (236). Conditional inactivation of HAS2 showed a role for hyaluronan in skeletal growth, patterning, chondrocyte maturation and joint formation in the developing limb (258).

Recently, transcription factors Sp1 and Sp3 have been identified as principal mediators of HAS2 constitutive transcription. Sp1 and Sp3 bind to three sites immediately upstream of the HAS2 transcription initiation site and that mutation of the consensus recognition sequences within these sites ablated their transcriptional response. In contrast, NF-Y, CCAAT, and NF-κB binding proteins may not be involved in HAS2 transcription (278). A recent study showed that transgenic expression of Tbx2, a central intermediary of Bmp-Smad signaling, induced Has2 and TGF-β2 expression, facilitating endocardial cushions formation (382).

C. Hyaluronan synthase 3 (HAS3)

The human HAS3 gene is localized to chromosome 16q22.1 and the mouse Has3 to chromosome 8 (396). The HAS3 protein is transmembrane protein. Using green fluorescent protein to tag HA synthases in keratinocytes, HAS2 and HAS3 were found to travel through endoplasmic reticulum, Golgi, plasma membrane, and endocytic vesicles, and a distinct enrichment of plasma membrane HAS was observed in cell protrusions (351). The trafficking of HAS was correlated with its activity, since inhibition of HA synthesis by substrate UDP-glucuronic acid starvation using 4-methyl-umbelliferone prevented HAS access to the plasma membrane (351). Similarly, in cells transfected with green fluorescent protein-tagged Has3, the dorsal surface was decorated by up to 150 slender, 3 to 20-μm-long microvillus-like plasma membrane protrusions, which consisted of filamentous actin, CD44, and lipid raft microdomains, in addition to Has3 (212). Enzymatic activity of HAS was required for the growth of the microvilli. The microvilli induced by HAS3 gradually withered with the introduction of an inhibitor of HA synthesis and rapidly retracted by hyaluronidase digestion, whereas they were independent of HA receptors (212). Keratinocyte growth factor activates keratinocyte migration and stimulates wound healing. At same time, keratinocyte growth factor stimulates epidermal keratinocytes to accumulate intermediate-sized HA in the culture medium and within keratinocytes and leads to a rapid increase of Has2 and Has3 mRNA (188), suggesting that Has2 and Has3 are the targets of keratinocyte growth factor in keratinocytes. Enhanced HA synthesis acts an effector for the migratory response of keratinocytes in wound healing, whereas it may delay keratinocyte terminal differentiation (188).

D. Regulation of HAS expression and activity

In vitro, the expression of HAS isoforms can be regulated by growth factors and cytokines. For example, recombinant TNFα and interferons stimulate HA production by normal human lung fibroblasts (87). IL-1β and TNFα induce HAS-2 mRNA in fibroblasts (472). Epidermal growth factor induces HAS2 expression in rat epidermal keratinocytes (327). TGF-β activates HAS1, leading to an increase in HAS activity (402). Conversely, TGF-β suppresses HAS3 mRNA in human fibroblast-like synoviocytes (402). TGF-β reduces HAS2 mRNA slightly but does not significantly affect the expression of mRNAs for HAS1 and HAS3 in mesothelial cells (164). Stimulation of mesothelial cells with platelet-derived growth factor-BB induces HAS activity (137) and upregulation of HAS2 mRNA (164).

Dysregulation of HA synthases and their activities has been found during tissue injury (235, 415, 489), consistent with the findings that HA accumulates during a number of injuries. For example, HAS2 mRNAs are increased in rats after radiation-induced lung injury (235). Ventilation-induced low molecular weight HA production is dependent on de novo synthesis of HA through HAS3 by fibroblasts and plays a role in the inflammatory response of ventilator-induced lung injury (17). Similarly, epidermal HA is significantly increased after epidermal trauma in adult mice caused by tape stripping (415). The HA response is associated with a strong induction of HAS2 and HAS3 mRNA (415). Increased accumulation of HA and increased HAS expression have been noticed in autoimmune (96) and mechanical renal injury (489, 490).

III. HA DEGRADATION AND HYALURONIDASES

A. Hyaluronidases

Hyaluronidases (also called hyaluronoglucosaminidases) hydrolyze the hexosaminidic β(1–4) linkages between N-acetyl-D-glucosamine and D-glucuronic acid residues in HA and release HA fragments. Complete digestion of HA with hyaluronidase releases disaccharide D-glucuronic acid-N-acetyl-D-glucosamine. These enzymes also hydrolyze β(1–4) glycosidic linkages between N-acetyl-galactosamine or N-acetylgalactosamine sulfate and glucuronic acid in chondroitin, chondroitin 4-sulfate, chondroitin 6-sulfate, and dermatan. Some bacteria, such as Staphylococcus aureus, S. pyogenes, and Clostridium perfringens, produce hyaluronidases as a means for greater bacterial mobility through the host’s tissues and as an antigenic disguise that prevents recognition of bacteria by phagocytes (210). In humans, there are six hyaluronidases identified thus far: hyaluronidases 1 – 4, PH-20, and HYALP1 (400).

The human HYAL1 gene is located on 3p21.3-p21.2 (64) and the mouse gene is on chromosome 9 (62). The hyaluronidase 1 gene encodes a hyaluronidase found in the major parenchymal organs such as the liver, kidney, spleen, and heart. HYAL1 is also present in human serum (62) and urine (63). The enzyme intracellularly degrades HA, is active at an acidic pH, and is the major hyaluronidase in plasma (62). Mutations in the HYAL1 gene are associated with mucopolysaccharidosis type IX, or hyaluronidase deficiency (293).

HYAL2 is a glycosylphosphatidylinositol-anchored cell-surface receptor (341) in all mouse tissue types except brain. HYAL2 has very low hyaluronidase activity in comparison to serum hyaluronidase HYAL1 (341), and HYAL2 hyaluronidase activity has a pH optimum of below 4 (229). Furthermore, unlike HYAL1, the HYAL2 enzyme hydrolyzes only HA of high molecular mass, yielding intermediate-sized HA fragments of approximately 20 kDa, which are further hydrolyzed to small oligosaccharides by PH-20 (229). In addition, HYAL2 serves as a receptor for Jaagsiekte sheep retrovirus (341).

HYAL3 transcripts show strongest expression in testis and bone marrow but relatively weak expression in other organs (62). The role of HYAL3 in the degradation of HA is not clear. A recent study showed that Hyal3 overexpressed in cultured cells lacks intrinsic hyaluronidase activity and that Hyal3 may contribute to HA metabolism by augmenting the activity of Hyal1 (138). Hyal3 deficient mice showed a subtle change in the alveolar structure and extracellular matrix thickness in lung-tissues at 12–14 months-of-age, although there was no evidence of HA accumulation, suggesting that HYAL3 may not play a major role in constitutive HA degradation (14). The testicular enzyme PH-20 is encoded by the SPAM1 (sperm adhesion molecule 1) gene (220). This glycosylphosphatidylinositol-anchored enzyme is located on the human sperm surface and inner acrosomal membrane. PH-20 hyaluronidase is active at neutral pH. Mouse HYAL5, not HYALP1, degrades HA (350), involving in penetration of mouse sperm through cumulus mass (198).

Flavonoids, including silybin, apigenin, kaempferol, luteolin, and condensed tannin, were found to be inhibitors of hyaluronidase (216). Kinetic studies of these inhibitors showed that their mode of inhibition was competitive (216). They have been found effective in mice in vivo (215) and in inhibition of hyaluronidase activity (485). They are a useful tool in fertilization research (198). In addition, apigenin has significant anti-inflammatory activity that involves blocking nitric oxide-mediated cyclooxygenase-2 expression and monocyte adherence (226) or involves in vivo effectively blocking TNFα-induced ICAM-1 upregulation (317). Sodium cromoglycate and sodium aurothiomalate are potent inhibitors of hyaluronidases of Naja kaouthia and Calloselasma rhodostoma venoms and they can reduce the systemic and local tissue damage of Naja kaouthia and Calloselasma rhodostoma venoms when injected immediately at the sites of bites (485). Sulphated oligosaccharides such as verbascose, planteose and neomycin showed comparable inhibition on all hyaluronidases, thereby possessing much higher activity than that of the widely accepted hyaluronidase inhibitor apigenin (360).

B. Pathologic

Expression and activity of hyaluronidases have long been noticed in diseases such as rheumatoid arthritis (180, 349) and periodontal disease (143). Patients with advanced scleroderma have decreased serum HYAL-1 activity and elevated circulating levels of HA (297). In one study, six patients with bone or connective tissue abnormalities had lower levels of HYAL-1 activity than did healthy donors (100). Reactive oxygen species accumulate at sites of tissue injury and may provide a mechanism for generating HA fragments in vivo. Reactive oxygen species degrade ECM components such as collagen, laminin, and HA in vitro (21). This may further exaggerate the inflammation state at the sites of tissue injury because the fragmented HA in turn augments inflammatory responses. However, HA has the capacity to absorb reactive oxygen species, playing an active role in protecting articular tissues by scavenging reactive oxygen species (364).

Animal studies showed hyaluronidase significantly reduced mortality and reduced “enzymatic” infarct size in gerbils with experimental cerebrovascular accidents (483). With small numbers of patients, Maroko and associates found that hyaluronidase intravenously accelerated the reduction of myocardial ischemic injury in patients with acute myocardial infarction (250) and reduced the frequency of electrocardiographic signs of myocardial necrosis (251). However, the effectiveness of hyaluronidase for patients with acute myocardial infarction was questionable. In a primate model that has a minimal collateral blood supply, hyaluronidase did not significantly reduce the ultimate infarct size (336). A multicentered, randomised, blind study from 1978 to 1988 did not find a beneficial effect of hyaluronidase on mortality or infarct size (352).

IV. HYALURONAN AS A SIGNALING MOLECULE

A. Hyaluronan Fragments

During the last 15 years, we and others have established the concept that HA plays different roles depending on its molecular weight (149, 264, 265, 306, 425, 465). In its native state, such as in normal synovial fluid, HA generally exists as a high molecular weight polymer, usually in excess of 1000 kDa. However, under certain conditions such as tissue injury and inflammation, HA is more polydisperse, with a preponderance of lower-molecular-weight forms (357).

Fragmentation of HA occurs during many pathological conditions, a possible consequence of dysregulated expression of HA synthases and HA degradation enzymes during tissue injury and inflammation (55). Altered expression of HAS2 and HYAL2 is involved in the turnover of HA during the early phase of lung injury in irradiated rats (235). Hyaluronan fragmentation can also be a result of the release of reactive oxygen species during tissue injury (3, 51, 281, 316, 392). We recently demonstrated that extracellular superoxide dismutase inhibits inflammation in response to lung injury by preventing oxidative fragmentation of HA (106). Extracellular superoxide dismutase directly binds to HA and significantly inhibits oxidant-induced degradation of HA (106).

We have found that lower-molecular-weight forms of HA (<500 kDa, many preparations in 100 – 250 kDa), but not the native form (>1000 kDa), induce inflammatory responses in inflammatory but not resident macrophages (144, 148, 149, 151, 152, 264, 265, 306). HA inhibits human neutrophil elastase-induced airway responses and, within the range of 150–300 kDa, dose rather than molecular weight may be the most important determinant of pretreatment time resulting in a protective effect (375). HA fragments augment steady-state mRNA, protein, and inhibitory activity of PAI-1 as well as diminish the baseline levels of uPA mRNA and inhibit uPA activity in an alveolar macrophage cell line and in freshly isolated inflammatory alveolar macrophages from bleomycin-treated rats, suggesting that the regulation of PAI-1 and uPA by HA fragments may provide a mechanism for regulating fibrinolytic activity during lung inflammation (150). HA fragments induced macrophage metalloelastase mRNA, protein expression as well as enzyme activity in mouse macrophages and in inflammatory alveolar macrophages from bleomycin-injured rat lungs. HA fragments may be an important mechanism for the expression of macrophage metalloelastase by macrophages in inflammatory lung disorders (152).

West and colleagues showed that oligosaccharides derived from high molecular weight HA have distinct functions when compared with the larger molecular forms of HA (465). Whereas HA oligomers of 8 – 16 disaccharides stimulated angiogenesis in vivo and endothelial proliferation in vitro, native, high molecular weight HA had no effect (465). Furthermore, this group confirmed that the high molecular mass HA (about 1000 kDa) is anti-angiogenic, whereas HA fragments of 3–25 disaccharide units induce angiogenesis (74, 354).

Fragmented HA containing 6 – 40-mers enhanced CD44 cleavage by tumor cells, whereas large polymer HA failed to enhance CD44 cleavage (405). A 6.9-kDa HA (36-mers) also promoted tumor cell motility in a CD44-dependent manner (404, 405). Enhanced expression of matrix metalloprotease (MMP)-9 and MMP-13 was induced in lung carcinoma cells by only small HA fragments containing 6-mers to 40-mers, but not by HA preparations with molecular weight greater than 600 kDa (97). Similarly, Termeer and colleagues demonstrated that HA oligomers of tetra- and hexasaccharide size, but not higher-molecular-weight HA species, induced immunophenotypic maturation of human monocyte-derived dendritic cells (425). Others found that HA hexamers or smaller were not active. Fragmented HA (80 – 800 kDa), but neither purified high molecular weight HA nor HA hexamers, markedly increased MCP-1 mRNA and protein expression (22), and ICAM-1 and VCAM-1 steady-state mRNA and cell-surface expression by murine kidney tubular epithelial cells (310).

B. HA-inducible genes

Many studies have shown that HA is biological active and is able to stimulate expression of a variety of genes. Below is the summary of these genes.

V. HYALURONAN BINDING PROTEINS

Much of HA exists in ECM in soluble form. HA also covalently binds to a variety of proteins to influence the functions of these proteins (432). HA binding proteins include receptors such as CD44, RHAMM, TNFIP6, Brevican, SHAP, LYVE1, and many other proteins. Some HA binding proteins are associated with cell membranes, whereas others are found in the extracellular matrix. Structurally, the link module (204) and the B(X7)B motif (where B is arginine or lysine and X is any nonacidic amino acid) (481) are thought to constitute the HA-binding region. Goetinck and associates identified that the sites for interaction with HA are in the tandemly repeated sequences of link protein and that there are four potential sites available for that interaction (113).

Day and colleague suggested that HA as a free polymer in solution exists in a stiffened highly dynamic ensemble of chaotically interchanging semiordered states (70). Under the organizing influence of proteins, these states may be coaxed towards a number of families of ordered and variously shaped structures of similar energy. When present as a simple polymer, HA is likely to be a perfect dynamic colonizer of the extracellular space (70). The specificity of protein–HA interactions, combined with the repeating nature of the polysaccharide, may drive the formation of periodic filamentous complexes that are likely to have significantly different quaternary organizations depending on the conformation of the bound sugar and the nature of the stabilizing protein–protein associations (70). These states may be interchangeable depending on the force. Furthermore, Day and associates proposed that that HA cross-linking is part of a protective mechanism, promoting adhesion of leukocytes to the HA complexes rather than enabling contact with inflammation-promoting receptors on the underlying tissues (69). Thus, leukocytes are maintained in a non-activated state by appropriate receptor clustering or receptor co-engagement. Hyaluronan networks serve as scaffolds to prevent the loss of extracellular matrix components during inflammation and to sequester proinflammatory mediators (69).

A. CD44

CD44 is the major cell-surface HA binding protein (10). CD44 is a widely researched molecule and it has been shown to have multiple roles in a variety of biomedical fields. CD44 is a polymorphic type I transmembrane glycoprotein whose diversity is determined by differential splicing of at least 10 variable exons encoding a segment of the extracellular domain and by cell-type-specific glycosylation (232). It is widely expressed in almost every cell type in the human and in the mouse. It is also an abundant protein on many cell types. It is believed to be one of the most abundant cell surface proteins on macrophages. Although glycosaminoglycan side chains associated with some CD44 isoforms can bind a subset of heparin-binding growth factors, cytokines, and ECM proteins such as fibronectin, most of the functions ascribed to CD44 thus far can be attributed to its ability to bind and internalize HA (381). N-glycosylation regulates CD44 structure by altering both the affinity and avidity of CD44-HA binding (422). Most cells express the standard isoform, which is an 85-kDa protein that undergoes posttranslational modification (232). Most cells including stromal cells such as fibroblasts and smooth muscle cells, epithelial cells, and immune cells all express CD44 (381).

HA-CD44 interactions play an important role in development, inflammation, T cell recruitment and activation, and tumor growth and metastasis (232). CD44 cytoplasmic domain can be phosphorylated when ligand binds to transduce signaling. CD44 is also required for BMP-7 signaling. The cytoplasmic domain of CD44 is required for BMP-7 induced Smad1 nuclear translocation (325). Smad1 was found to interact with the cytoplasmic domain of CD44 demonstrated with coimmunoprecipitation. The integrity of extracellular HA-cell interactions is required for BMP-7-mediated Smad1 phosphorylation, nuclear translocation of Smad1 or Smad4, and SBE4-luciferase reporter activation, suggesting a functional link between the BMP signaling cascade and CD44 (325).

When CD44 phosphorylation mutants were transfected into a murine fibroblast line expressing low levels of endogenous CD44, CD44 phosphorylation mutants were as efficient as wild type CD44 in mediating cell adhesion but were unable to support HA-dependent fibroblast migration, demonstrating a control mechanism specific for CD44-mediated cell motility (323). Work with transformed cells indicates that these cells use the cell surface matrix receptor CD44 for migration and invasion (139). Lung tissue from patients who died from acute alveolar fibrosis after lung injury reveals CD44-expressing mesenchymal cells throughout newly formed fibrotic tissue. CD44 was found uniformly over the cell surface and was found densely labeling filopodia and lamellipodia. By blocking the function of CD44 with monoclonal antibodies, fibroblast invasion into a fibrin matrix was inhibited (408). Fibroblast CD44 functions as an adhesion receptor for provisional matrix proteins and is capable of mediating fibroblast migration and invasion of the wound provisional matrix resulting in the formation of fibrotic tissue (408). Hyaluronan inhibits PDGF-BB-induced activation of PDGF receptor and cell motility, while blockage of the binding of HA to CD44 restored PDGF-receptor activation and motility, indicating that CD44 mediates the inhibiting effect on PDGF receptor (234). PDGF receptor and CD44 form a complex and the inhibitory effect of HA is neutralized by inhibition of tyrosine phosphatases, suggesting that HA-activated CD44 modulate PDGF receptor signaling by recruiting tyrosine phosphatase to the receptor (234).

The role of CD44 was investigated by genetic targeting in vivo. Transgenic mice expressing an antisense CD44 cDNA under the control of keratin-5 promoter had defective keratinocyte proliferation indicating the role of CD44 in the regulation of keratinocyte proliferation in response to extracellular stimuli (192). Given the importance of CD44 as demonstrated in many biological and pathological conditions, it was very surprising that the genetic disruption of CD44 in the mouse did not show much abnormality (371). Our laboratory has demonstrated that CD44 is required for the resolution of pulmonary inflammation (421). CD44-deficient mice succumb to unremitting inflammation following bleomycin-induced noninfectious lung injury, characterized by impaired clearance of apoptotic neutrophils, persistent accumulation of HA fragments at the site of tissue injury, and impaired activation of TGF-β1 (421). The phenotype was partially reversed by reconstitution with CD44+ cells, thus demonstrating a critical role for this receptor on leukocytes in resolving lung inflammation. CD44 deficiency results in enhanced inflammation in E. coli but not S. pneumoniae-induced pneumonia, suggesting a role for CD44 in limiting the inflammatory response to E. coli (454). We and others demonstrated that inflammation is more aggravated in CD44-knockout mice than in wild type mice (237, 294). We further found that the induction of the negative regulators of TLR signaling IL-1R-associated kinase-M (202), Toll-interacting protein, and A20 by intratracheal lipopolysaccharide in vivo and in macrophages in vitro was significantly reduced in CD44−/− mice, suggesting that CD44 plays a role in preventing exaggerated inflammatory responses to LPS by promoting the expression of negative regulators of TLR4 signaling (237).

In addition, recent studies suggested that CD44 expression was strongly induced in the infarcted myocardium and was localized on infiltrating leukocytes, wound myofibroblasts, and vascular cells. Although CD44 deficient mice showed enhanced inflammation following myocardial infarction, CD44 deficient infarcts showed decreased fibroblast infiltration, reduced collagen deposition, and diminished proliferative activity (156). Isolated primary CD44 deficient cardiac fibroblasts had reduced proliferation upon stimulation with serum and decreased collagen synthesis in response to TGF-β in comparison to wild type fibroblasts, suggesting that CD44-mediated interactions are critically involved in infarct healing (156). After tissue injury, fibroblast migration from the peri-wound collagenous stroma into the fibrin-laden wound is critical for granulation tissue formation and subsequent healing. Fibroblast transmigration or invasive migration from a collagen matrix into a fibrin matrix required the presence of fibronectin, several integrins, CD44 (59), and syndecan-4 (238). A recent study suggested that CD44 expression is not a general requirement for cell migration and gradient sensing; rather, it elicits a ligand-specific response (441). Expression of CD44 alone is not sufficient to drive chemotaxis towards HA, as NIH-3T3 fibroblasts were unable to respond to a HA gradient even when transfected with high levels of human CD44 (441). For NIH-3T3 cells to bind exogenous HA, it was necessary to both increase the level of receptor expression and remove a HA pericellular matrix (441). CD44 is required for myoblast migration and differentiation, since primary myoblasts from CD44-deficient mice displayed attenuated differentiation and subsequent myotube formation at early times in a differentiation-inducing in vitro environment (287). Chemotaxis of CD44 deficient myoblasts toward hepatocyte growth factor and basic fibroblast growth factor was totally abrogated (287). Very recently, it was demonstrated that CD44-deficient fibroblasts have fewer stress fibers, and focal adhesion complexes (1). Migration of CD44-deficient fibroblasts was increased in velocity, but was directionless (1).

B. RHAMM

RHAMM (for receptor for hyaluronan-mediated motility expressed protein, also called CD168) binds to biotinylated HA (129, 482). RHAMM is a functional receptor in many cell types including endothelial cells (240, 365). It is believed to be an HA receptor involved in tumor cell locomotion (129). Transfection experiments in fibroblasts suggest that RHAMM plays a role in Ras-dependent oncogenesis (122), but this role has been challenged by others (145), reflecting the complexity of the interactions between HA and HABPs in biological and pathological conditions. Nevertheless, RHAMM-HA interactions do play an important role in tissue injury and repair (491). Upregulation of RHAMM expression in bovine aortic smooth muscle cells was readily detected in an in vitro injury model (366). While HA stimulated the random locomotion of bovine aortic smooth muscle cells, an antibody to RHAMM that blocks HA binding with this receptor abolished smooth muscle cell migration following injury (366), suggesting that RHAMM is necessary for the migration of smooth muscle cells during wound repair.

Increased expression of RHAMM in macrophages after bleomycin injury in rats has been reported with a function blocking anti-RHAMM antibody (491). HA-stimulated macrophage chemotaxis was also inhibited by anti-RHAMM antibody (491). Daily administration of anti-RHAMM antibody to injured animals resulted in a 40% decrease in macrophage accumulation and lung architecture were improved with anti-RHAMM antibody treatment (491).

Inflammation is more aggravated in CD44-knockout mice than in wild type mice (294). Compensation for the loss of the CD44 gene occurs not because of enhanced expression of the redundant RHAMM gene, but rather because the loss of CD44 allows increased accumulation of the HA substrate, with which both CD44 and RHAMM engage, thus enabling augmented signaling through RHAMM (294). RHAMM knockout mice are viable and show no significant defect, suggesting that RHAMM is not an absolute requirement for normal development (431). RHAMM is expressed at high levels in aggressive fibromatosis. When crossed with mice that harbor a targeted mutation in the tumor suppressor APC gene predisposing animals to tumors including aggressive fibromatosis tumor, RHAMM deficiency significantly decreased the number of aggressive fibromatosis tumors formed (431). RHAMM promotes fibroblast proliferation under conditions of low density (431). RHAMM-deficient fibroblasts fail to repair wounds in an in vitro wound healing assay (430). ERK1,2 activation and fibroblast migration/differentiation is also defective during the repair of RHAMM-deficient excisional skin wounds and results in aberrant granulation tissue in vivo. These results identify RHAMM as an essential regulator of CD44-ERK1,2 fibroblast motogenic signaling required for wound repair (430).

C. TSG-6/TNFIP6

Tumor necrosis factor-stimulated gene-6 (TSG-6), also called TNFα-induced protein 6 (TNFIP6) contains Link module and binds HA (204, 228, 322). The solution structure of the Link module from human TNFIP6 was determined and found to consist of two alpha helices and two antiparallel beta sheets arranged around a large hydrophobic core. This defines the consensus fold for the Link module family, which includes CD44, cartilage link protein, and aggrecan (204). The interaction of TNFIP6 with HA is pH-dependent (32). TNFIP6 expression is upregulated in many cell types in response to a variety of proinflammatory mediators and growth factors. TNFIP6 is induced by IL-1 and TNFα (228, 474), and LPS (474). This protein is detected in several inflammatory disease states such as rheumatoid arthritis and in the context of inflammation-like processes is often associated with extracellular matrix remodeling. TNFIP6 is a potent inhibitor of neutrophil migration in an in vivo model of acute inflammation (473).

The recombinant Link module from human TNFIP6 has an inhibitory effect on neutrophil influx into zymosan A-stimulated murine air pouches, equivalent to that of full-length protein, resulting in a significant reduction in the concentrations of various inflammatory mediators in air pouch exudates (110). TNFIP6 is expressed by cultured vascular smooth muscle cells in response to serum and growth factor IL-1 but not to TNF or IL-6 (484). TNFIP6 is upregulated in proliferating vascular smooth muscle in the rat neointima after injury (484). TNFIP6-overexpressing cells grew faster than control vascular smooth muscle cells (484). TNFIP6 forms both covalent and non-covalent complexes with inter-α-inhibitor and potentiates its anti-plasmin activity (273).

Knockout of Tnfip6 in mice with a BALB/c background did not change the onset of proteoglycan-induced arthritis, but progression and severity were significantly greater in Tnfip6-deficient mice when compared to wild type BALB/c mice (409). An early and more extensive infiltration of the synovium with neutrophil leukocytes was the most prominent histopathologic feature of proteoglycan-induced arthritis in Tnfip6-deficient mice (409). Thus, TNFIP6 is a multifunctional anti-inflammatory protein that is produced at the site of inflammation and can be retained by the HA-rich extracellular matrix. A major effect of TNFIP6 is the inhibition of the extravasation of PMN cells, predominantly neutrophils, into the site of inflammation, most likely via a CD44/HA/TNFIP6-mediated blocking mechanism (409). TNFIP6 is a down-regulated gene during osteoblastic differentiation. TNFIP6 inhibits osteoblastic differentiation of human mesenchymal stem cells induced by osteogenic differentiation medium and BMP-2 (437). TNFIP6 is specifically expressed by expanding cumulus cell-oocyte complexes (103). Cumulus cell-oocyte complexes fail to expand in Tnfip6-deficient female mice because of the inability of the cumulus cells to assemble their HA-rich extracellular matrix, demonstrating that TNFIP6 is a key catalyst in the formation of the cumulus extracellular matrix and indispensable for female fertility (104).

D. Brevican

Brevican (also call BEHAB, for brain enriched hyaluronan binding) is a brain-specific proteoglycan in perineuronal nets. Brevican occurs as secreted and cell-surface, glycosylphosphatidylinositol-anchored, isoforms (166). Brevican binds not only to HA (167, 378) but also to chondroitin sulfate (480). Brevican increases the invasiveness of glioma cells in vivo and has been suggested to play a role in central nervous system fiber tract development (108, 167). Brevican deficient mice are viable and fertile and have a normal life span (41). Extracellular matrix perineuronal nets formed but appeared to be less prominent in mutant than in wild type mice. Brevican-deficient mice showed significant deficits in the maintenance of hippocampal long-term potentiation, but no significant deficits in learning and memory (41). Expression of brevican was upregulated in response to a stab wound to the adult rat brain (166).

E. Neurocan

Neurocan is a nervous tissue-specific chondroitin sulfate proteoglycan of the aggrecan family which has been shown to interact with neural cell adhesion molecules (348) and tenascin (347). Neurocan binds to HA and chondroitin sulfate in the brain (75). Through its interactions with neural cell adhesion and extracellular matrix molecules, neurocan is a potent inhibitor of neuronal and glial adhesion and neurite outgrowth (249). Neurocan-deficient mice are viable and fertile and have no obvious deficits in reproduction and general performance (499). Brain anatomy, morphology, and ultrastructure are similar to those of wild type mice, indicating that neurocan has a redundant function in the development of the brain (499). The distribution and function of neurocan and HA are associated with the remodeling process of brain (346). Hyaluronan is organized into fiber-like structures along migratory pathways in the developing mouse cerebellum. Thus, HA-rich fibers are concentrated at sites where specific neural precursor cell types migrate, and the anisotropic orientation of these fibers suggests that they may support guided neural migration during brain development (18). Tissue injuries such as brain injury (12, 266) and spinal cord injury (181) induced neurocan expression. Proteoglycan decorin suppresses neurocan and brevican expression and inflammation, and promotes axon growth across adult rat spinal cord injuries (68).

F. HABP1/C1QBP

HABP1 (for hyaluronan-binding protein 1, also called C1QBP, for complement component 1, q subcomponent binding protein) was first purified from the liver (67) and the brain (66). cDNA cloning confirmed that HABP1 is identical to complement component 1, q subcomponent binding protein (C1QBP) (111), and to pre-mRNA splicing factor SF2 (147, 208). It is about 64 to 68 kDa and binds to HA (66, 67) as well as mannosylated albumin (213). In addition to its mitochondrial location, HABP1 was also localized in the Golgi and completely dispersed throughout the cell during mitosis. This distinctive distribution pattern of HABP1 during mitosis resembles its ligand HA (377).

HABP1 exists as a highly acidic, noncovalently associated trimer in equilibrium with a small fraction of a covalently linked dimer of trimers (172). Structural studies demonstrated that HABP1 exhibits structural plasticity, which is influenced by the ionic environment under in vitro conditions near physiological pH (173). HA interacts only with compact HABP1, while C1q and mannosylated albumin can bind to loosely held oligomeric HABP1 as well. Thus, structural changes in HABP1 mediated by changes in the ionic environment are responsible for recognizing different ligands (173). The more stable conformation of HABP1 was attained at higher ionic strength or at acidic pH showed maximum affinity toward HA (171). Therefore, cellular HA-HABP1 interaction can be regulated by pH and ionic strength. Endogenous HABP1 can be phosphorylated, and was coimmunoprecipitated with activated extracellular signal-regulated kinases (246). HABP1 is a substrate for extracellular signal-regulated kinases and an integral part of the mitogen-activated protein kinase cascade (246). In addition to HA, HABP1/C1QBP also binds to the globular heads of complement subcomponent C1q molecules and inhibit C1 activation (111).

HABP1 overexpressing cells showed extensive vacuolation and reduced growth rate, underwent apoptosis in normal rat skin fibroblasts (187, 267). Overexpression of HABP1 was seen during Leishmania donovani infection. HABP1 binds with 2 proteins of promastigotes as well as amastigotes of L. donovani, suggesting a possible role for HABP in adhesion during the interaction of promastigotes and macrophages (344). It is unknown whether HA is required in this interaction. HABP1 has been demonstrated to play a role in spermatogenic differentiation. Stage-specific expression of HABP1 precursor during spermatogenesis in the rat (24). The sperm surface HABP1 level was correlated with the degree of sperm motility, an important criterion for fertilization (112). The appearance of HABP1 proprotein in the pachytene spermatocytes and the round spermatids during the initial stages of postnatal testis development suggests that this expression may be crucial for spermatogenesis (427).

G. HARE

HARE (for hyaluronan receptor for endocytosis, also called stabilin-2 and FEEL-2) was identified in the abundant expression of 175- and approximately 300-kDa HARE species from sinusoidal endothelial cells of the liver, lymph node, and spleen (497). The expression of HA receptors Stab2 and LYVE1 depended on the tissue and developmental stage of liver endothelial cells in mice (307). The human gene that encodes HARE is on chromosome 12 (495). The HARE protein features multiple B(X7)B HA-binding motifs, several fasciclin-like adhesion domains, and 18–20 epidermal-growth-factor domains (329). HARE protein binds to HA (130, 262), dermatan sulfate, and chondroitin sulfates A, C, D, and E, but not chondroitin, heparin, heparan sulfate, or keratan sulfate (131, 459).

While HARE/Stab-2 mRNA is confined to the liver and spleen, HARE/Stab-2 protein expression can be detected in the endothelial sinuses of the liver, lymph nodes, spleen, and bone marrow, and in specialized structures of the eye, heart, brain, and kidney in mice (93). Cloning and functional expression of the rat 175-kDa HARE demonstrated that the recombinant HARE is an authentic endocytic receptor for HA and was expressed on the cell surface (498). HARE was substantially colocalized with clathrin, but not with internalized HA that was delivered to lysosomes (498). The liver contains two distinct endothelial cell types: vascular and sinusoidal. The anti-HARE monoclonal antibodies showed diffuse strong staining of nonneoplastic liver sinusoidal endothelium. No staining of nonsinusoidal endothelium or the endothelial lining of the hemangiomas was seen with anti-HARE (82). Antibodies to HARE are not only markers for endothelial cells and the sinusoidal cells of rat liver, spleen, and lymph nodes, but also proved a useful tool in studying the functions of HARE and the physiological significance of HA clearance (458, 497). HARE mediates systemic clearance of glycosaminoglycans from the circulatory and lymphatic systems via coated pit-mediated uptake. A majority of each HARE isoform was intracellular, within the endocytic system, suggesting transient surface residency typical of an active endocytic recycling receptor (130).

H. LYVE1

LYVE1 (for lymphatic vessel endothelial hyaluronan receptor 1, also called CRSBP-1 for cell surface retention sequence binding protein-1) was cloned as a lymph-specific HA receptor on the lymph vessel wall in human (19) and in mice (337). It is a type I integral membrane glycoprotein, homologue of CD44, contains a link module, and binds both soluble and immobilized HA (19). Expression of mouse LYVE1 remains restricted to the lymphatics in homozygous knockout mice lacking a functional gene for CD44 (337). However, LYVE1 is also present in normal hepatic blood sinusoidal endothelial cells in mice and humans (282). LYVE1 expression was also found on the endothelial cells of the lymphatic sinus and in reticular cells in the lymph nodes. (476). LYVE1 is expressed in DC-SIGN+ macrophages within the chorionic villi in the rapidly differentiating placenta (33). TNFa promotes LYVE1 internalization from cell surface to lysosomes leading to its degradation. However, the internalization of LYVE1 is independent of HA (179). LYVE1 has being widely used as a lymphatic vessel specific marker (4, 163, 280, 479). It also offers a prognostic parameter for head and neck squamous cell carcinomas, since intratumoral LYVE1+ lymphatic vessels were clearly associated with a higher risk for local relapse as well as with poor disease-specific prognosis (259). LYVE1 is expressed in tissue macrophages. In murine tumor models and excisional wound healing, LYVE1 expression occurred in a subset of CD11b+, F4/80+ tissue macrophages (370). DC-SIGN+CD163+ macrophages have been shown to express LYVE-1 (33). LYVE1 is also implicated in the trafficking of cells within lymphatic vessels and lymph nodes (162).

LYVE1 plays a role in the transport of HA from tissue to lymph by uptaking HA via lymphatic endothelial cells (163, 337). The occurrence of LYVE1-expressing lymphatic compartments and the alteration of chemokine CCL21 expression in the lymphatics may be involved in defective thymocyte differentiation and migration, and play a significant role in insulitic and diabetic processes (174). Despite the implied importance of this molecule, LYVE1-deficient mice are grossly normal (105, 155). One study did not observe obvious alterations in lymphatic vessel ultrastructure or function or any apparent changes in secondary lymphoid tissue structure or cellularity in LYVE1-deficient mice (105). HA homeostasis is unperturbed in LYVE-deficient mice (105). This suggests that LYVE1 is not obligatory for normal lymphatic development and function and that either compensatory receptors exist or LYVE1 has a more specific role than previously envisioned (105). However, a second study identified morphological and functional alterations of lymphatic capillary vessels in certain tissues, marked by constitutively increased interstitial-lymphatic flow and lack of typical irregularly shaped lumens in LYVE1-deficient mice (155). LYVE1 is not required for either entry or migration of dendritic cells through the afferent lymphatics, and leukocyte populations are unaltered in LYVE1-deficient mice (105), in line with the finding that lymphatic adhesion/transmigration is largely mediated by ICAM-1 and VCAM-1 rather than LYVE1 (178).

I. SHAP

SHAP (for serum-derived hyaluronan-associated protein) was originally identified as a serum-derived HA-associated protein since it covalently binds to HA (154, 487). It was later found that SHAP is identical to inter-α (globulin) inhibitor heavy chain 2 (ITIH2), which belongs to the inter-α (globulin) inhibitor (ITI) family of structurally related plasma serine protease inhibitors involved in extracellular matrix stabilization. The ITI family consists of multiple proteins made up of a given combination of polypeptide chains after complex posttranslational maturation. There are 4 heavy chains and 1 light chain of ITI in the human, produced in the liver and circulating in peripheral blood. Mass spectrometric analyses showed that the C-terminal aspartic acid of each ITI heavy chain was esterified to the C6-hydroxyl group of an internal N-acetylglucosamine of HA chain (494). Biochemical analysis demonstrated the multivalent feature of the SHAP-HA complex, where an HA chain with a molecular weight of 2000 kDa has as many as five covalently bound SHAP (486). Under both static and flowing conditions, CD44-positive cells adhered preferentially to the immobilized SHAP-HA complex than to HA. The enhanced adhesion is exclusively mediated by the CD44-HA, but not by ITI-HA interaction (501). The ITI complex also includes bikunin, a 30-kD subunit (35). The bikunin precursor, α-1-microglobulin, is a single chain plasma glycoprotein involved in regulation of the inflammatory process (268). Alpha-1-microglobulin is processed to mature α-microglobulin and bikunin (447). Bikunin circulates in peripheral blood as a free peptide and as a complex with IgA (268).

SHAP-HA complex was isolated from pathological synovial fluid from human arthritis patients, and the amount of complex correlates positively with the degree of inflammation (494). Furthermore, SHAP-HA complex levels in rheumatoid arthritis sera were extremely high compared to control levels, but in osteoarthritis sera no marked increase was observed compared to controls (196), suggesting that the SHAP-HA complex plays different roles in rheumatoid arthritis and in osteroarthritis. Both serum levels of the SHAP-HA complex and HA in the patients with of chronic hepatitis, liver cirrhosis, and hepatocellular carcinoma were significantly higher than those of normal individuals, and these levels correlate with of the stages of liver fibrosis. Thus SHAP-HA complex level is an indicator for the progression of the stages of liver fibrosis (379). Targeting the bikunin gene in mice abolished the formation of the SHAP-HA complex, leading to the absence of matrix in ovulated oocytes and severe female infertility (503). Restoration of the formation of the SHAP-HA complex by administration of ITI fully rescued the defects (503). The chondroitin sulfate moiety of bikunin is essential for presenting SHAP to HA, which is indispensable for ovulation and fertilization in mammals (502).

VI. ROLE OF CD44 IN HA SIGNALING

Although numerous studies suggest that HA fragments may signal through a CD44-dependent tyrosine kinase pathway (37, 193, 389) and that anti-CD44 antibodies partially inhibit HA fragment signaling in macrophages (144), other studies suggest that anti-CD44 antibodies have no effect on macrophage metalloelastase expression (152). With the aid of CD44-deficient mice, studies have demonstrated that HA is able to stimulate chemokine production in absence of CD44, suggesting that the presence of CD44 is not required to mediate HA signaling (97, 175, 193). The cumulative interpretation of these data is that the role of CD44 in regulating HA interactions depends on the cell type.

VII. HYALURONAN SIGNALING THROUGH TOLL-LIKE RECEPTORS

HA is a component of the cell coat of groups A and C of Streptococcus (241, 243) and P. multocida (50). The repeating disaccharide structure of HA has features of pathogen-associated molecular patterns. Many pathogen-associated molecular patterns on pathogens utilize Toll-like receptors to initiate innate immune responses (2, 413). The innate immune system uses TLRs to recognize microbes and initiate host defense (2, 413). TLR4 has been unequivocally identified as the transmembrane signaling protein required for LPS signal transduction in macrophages (414). TLR2 is the mediator of macrophage recognition of mycobacteria (442) and gram-positive organisms (414). Stimulation of chemokine gene expression by HA fragments was abolished in the MyD88-deficient macrophages (175). Chemokine MIP-2 expression was reduced but remained present in both TLR2- and TLR4-deficient macrophages. HA fragment-induced chemokine and cytokine expression was completely abolished in TLR2 −/− TLR4 −/− peritoneal macrophages (175). The human HA degradation products purified from the serum of patients with acute lung injury were of similar molecular mass (peak at 200 kDa) to the in vitro-generated HA degradation products and stimulated chemokine production in wild type macrophages but not in either TLR2 −/− TLR4 −/− or MyD88 −/− peritoneal macrophages (175).

HA oligosaccharides induce maturation of dendritic cells via TLR4 (423). HA oligosaccharide treatment results in distinct phosphorylation of MAP kinases, nuclear translocation of NF-κB, and TNFα production (423). Priming of alloimmunity by HA-activated dendritic cells is dependent on signaling via TIRAP, a TLR adaptor downstream of TLR2 and TLR4 (426). However, this effect is independent of alternate TLR adaptors, MyD88, or TRIF (426). We further demonstrated that HA accumulates during skin transplant rejection and suggested that fragments of HA can act as innate immune agonists that activate alloimmunity (426). Injury-induced, Toll-like receptor-triggered signaling pathways, such as HA fragments (426), high-mobility group box 1 protein (345), involved in establishing innate alloimmunity utilize adaptor proteins and transcription factors that play a crucial role in the host’s defense against pathogens (218). Horton and colleagues reported similar findings, using a commercial source of HA fragments, although they suggested a primary role for TLR2 (368).

Hyaluronan upregulates IL-1R-associated kinase-M, a negative regulator of TLR signaling, to deactivate human monocytes, mediated through the engagement with CD44 and TLR4 (79). Small HA fragments stimulate expression of MMP2 and IL-8, leading to NF-κB activation, enhanced the motility of melanoma cells, in part in a TLR4-dependent manner (451). HA inhibits osteoclast differentiation through TLR4 by interfering with M-CSF signaling, suggesting a role of HA-TLR interaction in the regulation of bone metabolism (56).

VIII. HYALURONAN AS AN IMMUNE REGULATOR

A. T cells

CD44 is expressed on T cells (221). The interaction of cell-surface HA and CD44 on T cells is manifested by polarization, spreading, and colocalization of cell-surface CD44 with a rearranged actin cytoskeleton. Thus, cytokines and chemokines present in the vicinities of blood vessel walls or present intravascularly in tissues where immune reactions take place can rapidly activate the CD44 molecules expressed on T cells (8). Naive splenic T cells did not bind fluoresceinated HA constitutively. HA binding requires the activation of splenic T cells by a CD44-specific monoclonal antibody (231). Lesley and colleagues suggested that CD44 functions associated with HA binding involve a regulated process (230, 231).

T cell activation is associated with increased surface levels of CD44. CD44 expression on T cells but not dendritic cells plays a critical role in antigen-specific T cell responsiveness (81). Activation of CD44 and ability to engage in rolling occurs directly through polyclonal as well as antigen-specific T cell receptor-initiated signaling, suggesting potential roles for the CD44-HA interaction (78). CD44 activation does not appear to be the result of overt changes in glycosylation (78).

Leukocytes extravasate from the blood in response to physiologic or pathologic demands by means of complementary ligand interactions between leukocytes and endothelial cells, via a multistep process. Binding of CD44 on activated T lymphocytes to endothelial HA mediates a primary adhesive interaction under shear stress, permitting extravasation at sites of inflammation (76, 77). The integrin α4 (VLA-4), not integrin αL, is used in secondary adhesion after CD44-mediated primary adhesion of human and mouse T cells in vitro, and by mouse T cells in an in vivo model. Thus, extravasation of activated T cells initiated by CD44 binding to HA depends upon integrin α 4-mediated firm adhesion, which may explain the frequent association of these adhesion receptors with diverse chronic inflammatory processes (385). CD44 and integrin α 4 are physically associated on the cell surface and that the cytoplasmic portion of CD44 is necessary for this association. Disruption of the association through deletion of the CD44 cytoplasmic tail concordantly prevents firm adhesion of integrin α 4 on its ligand VCAM-1 under shear stress, while leaving rolling interactions between CD44 and HA intact, demonstrating that co-anchoring within CD44 and VLA-4 bimolecular complex between a primary and secondary adhesion molecule regulates the ability of a cell to firmly adhere (292).

Skin γδT cells play specialized roles in keratinocyte proliferation during wound repair. γδT cells are required for HA deposition in the extracellular matrix and subsequent macrophage infiltration into wound sites (165). γδT cell-derived keratinocyte growth factors induce epithelial cell production of HA. In turn, HA recruits macrophages to the site of damage. These results demonstrate a novel function for skin γδT cells in inflammation and provide a new perspective on T cell regulation of ECM molecules (165).

While lymphocytes from CD44 −/− mice preferentially homed to lymph nodes, their entry into the inflamed synovial joints was delayed when compared to wild type cells (401). CD44-deficient lymphocytes from animals with chronic arthritis expressed markedly reduced levels of the lymph node homing receptor, L-selectin. Down-modulation of L-selectin from CD44 −/− cells in arthritic condition might be a counter-regulatory response, which, by extending lymphocyte transit time in the circulation at the expense of lymph node homing, allows CD44-deficient cells to gain entry to the site of chronic inflammation via secondary adhesion mechanisms (401).

CD4+CD25+ regulatory T cells (Treg) are fundamental to the maintenance of peripheral tolerance. Siegelman and colleague found that the level of functionally active, HA-binding form of CD44act is strikingly correlated with superior suppressor activity, suggesting that CD44 is more than a cell surface marker and plays a role in regulating Treg cell functions (98). The expression of other surface markers and Foxp3 are similar irrespective of HA binding and associated degree of suppressor potency (98). Furthermore, high molecule weight HA enhances human CD4+CD25+ regulatory T cell functional suppression of responder cell proliferation, whereas low molecule weight HA does not (34). In addition, high molecule weight HA also up-regulates the transcription factor FOXP3 on CD4+CD25+ regulatory T cells. These effects are only seen with activated CD4+CD25+ regulatory T cells and are associated with the expression of CD44 isomers that more highly bind high molecule weight HA (34).

B. PMN

Both polarization and directed migration of neutrophils are dependent on the expression of CD44 and its interaction with HA, which could modulate neutrophil migration into inflamed tissues (7). TNFIP6 inhibits polymorphonuclear cell efflux and neutrophil invasion into the inflammatory site (409). A recent study suggested that CD44 does not appear to affect subsequent migration within inflamed tissues, although CD44 mediates neutrophil adhesion and emigration (195). CD44 deficiency leads to enhanced neutrophil migration and lung injury in Escherichia coli pneumonia in mice CD44 deficiency results in enhanced inflammation in E. coli but not S. pneumoniae-induced pneumonia, suggesting a role for CD44 in limiting the inflammatory response to E. coli (454). This study is consistent with our recent observation that inflammation is more aggravated in CD44-knockout mice than in wild type mice (237). Endothelial CD44 rather than neutrophil CD44 mediates neutrophil migration (356). Ligation of CD44 on neutrophils with anti-CD44 and HA induced IL-6 gene transcription and IL-6 protein secretion. Interferon γ-induced IL-6 production is dependent on CD44 cross-linking (373). The cross-linking of specific epitopes of the CD44 molecule can rapidly induce neutrophil apoptosis in vitro and inhibit neutrophil-dependent renal injury in vivo (412). Thus physiological ligands of the CD44 molecule may play an important role in eliminating neutrophils from sites of inflammation, including inflammatory kidney disease (412).

C. Macrophages

Ligation of human macrophage surface CD44 by bivalent monoclonal antibodies rapidly and profoundly augments the capacity of macrophages to phagocytose apoptotic neutrophils in vitro (450). CD44-deficient mice succumb to unremitting inflammation following noninfectious lung injury, characterized by impaired clearance of apoptotic neutrophils, persistent accumulation of HA fragments at the site of tissue injury, and impaired activation of transforming growth factor-β1 (421). This phenotype was partially reversed by reconstitution with CD44+ cells, thus demonstrating a critical role for this receptor in resolving lung inflammation (421). CD44 regulates phagocytosis of apoptotic neutrophil granulocytes, but not apoptotic lymphocytes, by human macrophages (132).

There are several reports demonstrating that HA fragments can influence dendritic cell maturation. Goldstein and colleagues showed that 135-kDa fragments of HA induce dendritic cell maturation and initiate alloimmunity (426). Activation of dendritic cells with HA enhances their ability to stimulate allogeneic and antigen-specific T cells markedly (81).

Langerhans cells are skin-specific members of the dendritic cell family and crucial for the initiation of cutaneous immune responses. Systemic, local, or topical administration of HA blocking peptide prevented hapten-induced Langerhans cell migration from the epidermis and inhibited the hapten-induced maturation of Langerhans cells in vivo (283). Small HA fragments of tetra- and hexasaccharide size increase dendritic cell production of the cytokines IL-1β, TNFα, and IL-12 as well as their allostimulatory capacity (425). These small HA fragments induce dendritic cell maturation independently of CD44 or RHAMM and are dependent upon TLR4 (423).

D. Mast cells

Mast cell granules are a rich source of HA, and this may account for the striking concurrence of HA accumulation with a mastocytotic condition in many tissues undergoing pathologic changes (84). Cultured human mast cells adhere to HA-coated surfaces (102). Human mast cells expressing standard form of CD44, but not the v5, v6, v7, and v8 variants, appears to mediate the attachment of these cells to HA (102). However, CD44 was not found to be consistently elevated in serum obtained from patients with mastocytosis or individuals experiencing anaphylaxis (102).

The number of mast cells was higher in patients with sarcoidosis than in controls (86). Sarcoidosis patients had significantly increased bronchoalveolar lavage (BAL) fluid concentrations of HA, fibronectin, and type III procollagen peptide, as well as albumin and lymphocytes, compared to controls (86). Mast cell activation and stress play a role in the pathogenesis of interstitial cystitis (363). Intravesical sodium hyaluronate has been used to treat interstitial cystitis due to its possible replenishment of bladder glycosaminoglycans (186). HA inhibited mast cell activation and the secretion of proinflammatory mediators, providing a mechanism of HA as a therapeutic option for interstitial cystitis (36).

E. Eosinophils

The infiltration, accumulation and degranulation of eosinophils in the lung is a hallmark of active asthma (42). HA increased GM-CSF, TGF-β and ICAM-1 expression by eosinophils, leading to eosinophil survival. HA-eosinophil interaction may contribute to the regulation of airway inflammation and airway remodeling (313). HA stimulates the growth of CD34+ progenitor cells into specifically differentiated, mature eosinophils (126). HA contributes to long-term eosinophil survival in vivo by enhancing GM-CSF production in asthma conditions (90). HA triggers granulocyte GM-CSF mRNA stabilization in eosinophils yet engages differential intracellular pathways and mRNA binding proteins (90). Recently, Pin1, a cis-trans isomerase, was identified as an essential component of the ribonucleoprotein complex responsible for GM-CSF mRNA stabilization, cytokine secretion and the survival of HA-activated eosinophils (380).

The expression of CD44 on alveolar eosinophils and the concentration of soluble CD44 were increased in BAL of patients with eosinophilic pneumonia patients, and the increase was correlated with Th2 cytokine IL-5 in BAL (190). In addition, a high concentration of HA was observed in BAL of eosinophilic pneumonia patients (190). Intraperitoneal administration of anti-CD44 monoclonal antibody prevented both lymphocyte and eosinophil accumulation in the lung, blocked antigen-induced elevation of Th2 cytokines, chemokines (CCL11, CCL17), and inhibited the increased levels of HA and leukotriene concentrations in BAL, demonstrating a critical role for HA-CD44 interaction in development of allergic respiratory inflammation (189).

IX. HA AND NON-LEUKOCYTES

A. Epithelial cells

Hyaluronan induces MCP-1 expression in renal tubular epithelial cells (22). HA is synthesized in high molecular weight form at the apical pole of human airway epithelial cells, covering the luminal surface (279). High molecular form HA is broken down by reactive oxygen species to form low molecular weight fragments that signal via RHAMM to stimulate ciliary beat frequency (248). Low molecular weight HA fragments stimulated ciliary beat frequency, an effect blocked by anti-RHAMM antibody and genistein (248), suggesting a role of HA-RHAMM interaction in airway mucosal host defense.

The transforming effects of hepatocyte growth factor and β-catenin are dependent on HA-cell interactions. Increased expression of HA is sufficient to induce epithelial-mesenchymal transition and acquisition of transformed properties in phenotypically normal epithelial cells in vitro (505). HA plays a central role in the transition of epithelia to mesenchyme in the embryo and in the acquisition of transformed properties in carcinoma cells (434). Overexpression of HAS2 on airway epithelial cells under control of the Clara cell specific CC10 promoter leading to an increased production of high molecular weight HA was protective against mortality, lung injury, and epithelial cell apoptosis (175).

B. Endothelial cells

Oligosaccharides of HA enhance the production of collagens by endothelial cells (354). Early-response gene signaling is induced by angiogenic oligosaccharides of HA in endothelial cells, while is inhibited by high molecular weight HA (74). HA fragments, but not high molecular weight HA, induce iNOS in liver, having the greatest effects on endothelial and Kupffer cells. Thus, HA fragments may be an important stimulus for NO production in various forms of liver disease, particularly as a cofactor with inflammatory cytokines (353).

Local cytokine production within inflamed vascular beds may enhance surface HA expression on endothelial cells, thereby creating local sites receptive to the CD44/HA interaction and thus extravasation of inflammatory cells (277). IL 15 induces endothelial HA expression in vitro and promotes activated T cell extravasation through a CD44-dependent pathway in vivo (91). Endogenous components of the extracellular matrix HA can stimulate endothelia to trigger recognition of injury in the initial stages of the wound defense and repair response (417). HA-induced proliferation of endothelial cells is CD44 receptor mediated and accompanied by early-response-gene activation (389). HARE/stabilin-2 is found on sinusoidal endothelial cells of the liver, lymph node, and spleen (496). HARE mediates the systemic clearance of HA from the circulatory and lymphatic systems, through a clathrin-dependent, coated pit-mediated uptake (497).

Hyaluronan fragments stimulate endothelial cell differentiation. Hyaluronan dodecasaccharides induce capillary endothelial cell sprouting by the binding to CD44 and by the induction of CXCL1/GRO1 (411). HA regulates vascular endothelial cell barrier function through CD44v10 isoform interaction with S1P receptors, S1P receptor transactivation, and RhoA/Rac1 signaling to the EC cytoskeleton (386). CD44 is an important regulator of HGF/c-Met-mediated in vitro and in vivo barrier enhancement, a process with essential involvement of Tiam1, Rac1, dynamin 2, and cortactin (387).

C. Fibroblasts

Fibroblasts are the main cell type that release HA (164, 242, 397, 429). Fibroblasts synthasize all three isoforms of HA synthases and release HA upon tissue injury (17, 255) and the stimulation of inflammatory factors such as IL-1β and TNFα (472). On the other hand, HA fragments stimulate fibroblasts to release cytokines and to regulate inflammatory responses (242, 255), and facilitate TGF-β-mediated fibroblast proliferation (270). Furthermore, upon myofibroblastic differentiation, more HA appeared in the conditioned medium and became associated with the cells (170). Moreover, HA modulates TGF-β dependent myofibroblast differentiation (456, 457).

Hyaluronan-binding proteins regulate fibroblast functions through their interactions with HA. For example, acute lung injury fibroblast migration and invasion of a fibrin matrix is mediated by CD44 (408). HA modulates contact inhibition of cell growth and migration (157). RHAMM deficient fibroblasts are defective in CD44-mediated ERK1/2 motogenic signaling, leading to defective skin wound repair, demonstrating a role of RHAMM in CD44 signaling and in fibroblast migration (430). A recent study suggested that HA-activated CD44 modulates PDGF receptor signaling by recruiting tyrosine phosphatases to the receptor and influences cell motility (234).

Consistent with the role of CD44 in mediating fibroblast invasion and subsequent tissue fibrosis, immunohistochemical analysis of lung tissue from patients who died from acute alveolar fibrosis after lung injury reveals CD44-expressing mesenchymal cells throughout newly formed fibrotic tissue (408). Anti-CD44 antibody blocked fibroblast migration on the provisional matrix proteins fibronectin, fibrinogen, and HA (408). Therefore, fibroblast CD44 functions as an adhesion receptor for provisional matrix proteins and is capable of mediating fibroblast migration and invasion of the wound provisional matrix resulting in the formation of fibrotic tissue (408). A recent study reported that high molecular weight HA increased Snail2 leading to fibroblast invasion (60).

D. Smooth muscle cells

HA is able to promote smooth muscle cell migration in various conditions (448, 471). In addition, growth factors and cytokines regulate HA production in smooth muscle cells. For example, PDGF stimulates HA production in vascular smooth muscle cells (318, 319), while IL-15 inhibits HA production and smooth muscle cell migration (161). Furthermore, HA binding proteins regulate smooth muscle cell proliferation and migration during tissue injury. For example, RHAMM is necessary for the migration of smooth muscle cells and the expression and distribution of this receptor is tightly regulated following wounding of BASMC monolayers (366). HA-induced migration depends exclusively on RHAMM-mediated PI3K-dependent Rac activation (117).

X. HYALURONAN AND STEM CELLS

A. HA and embryonic stem cells

CD44 is expressed in discrete embryonic structures and the differentiation process of embryonic stem cells is accompanied by an induction of CD44 mRNA and protein (469). HA synthase 2 and RHAMM are differentially expressed during all stages of preimplantation human embryos and human embryonic stem cells (58). RHAMM expression is significantly downregulated during differentiation of human embryonic stem cells, in contrast to HAS2, which is significantly upregulated. RHAMM knockdown results in downregulation of several pluripotency markers in human embryonic stem cells, induction of early extraembryonic lineages, loss of cell viability, and changes in the human embryonic stem cell cycle, highlighting an important role for RHAMM in maintenance of human embryonic stem cell pluripotency, viability, and cell cycle control (58). HAS2 knockdown results in suppression of human embryonic stem cell differentiation without affecting human embryonic stem cells pluripotency, suggesting an intrinsic role for HAS2 in the human embryonic stem cells differentiation process (58). Addition of exogenous HA to the differentiation medium enhances human embryonic stem cell differentiation to mesodermal and cardiac lineages (58). During stem cell differentiation, HA (and other glycosaminoglycans) synthesis was enhanced by 13- and 24-fold, most likely due to increased expression of HA synthase-2 (289).

B. HA and hematopoietic stem cells

The fate of hematopoietic stem cells (HSC) is determined by microenvironmental niches. Hyaluronan is part of the extracellular environment in bone marrow. HA not only provides a physical scaffold or support within the marrow to facilitate localization and retention of HSCs to the stem cell niche but moreover, through ligation with its counter-receptors is able to directly affect the cellular functions of HSCs (133).

HA is required for in vitro hematopoiesis, since the removal of HA with hyaluronidase in long-term bone marrow cultures results in reduced production of both progenitor and mature cells (193). HA enhanced hematopoietic activity, committed progenitors, and the total number of mature bone marrow cells (256). HA synthesized by primitive hemopoietic cells participates in their lodgment at the endosteum following transplantation (301). Therefore, HA is required for the regulation of the hematopoiesis-supportive function of bone marrow accessory cells and participates in hematopoietic niche assembly (256). In vivo exposure to cigarette smoke and in vitro treatment of long-term bone marrow cultures with nicotine, a major constituent of cigarette smoke, result in inhibition of hematopoiesis (193, 194, 256).

CXCL12/SDF-1 is a major chemokine in promoting HSC homing, migration, and proliferation (219, 269, 372). HA oligosaccharides enhance CXCL12-dependent chemotactic effect on peripheral blood hematopoietic CD34+ cells (367). HA-CD44 interaction is essential for homing into the bone marrow and spleen of non-obese diabetic/severe combined immunodeficient mice and engraftment by human HSCs (15). Hematopoietic progenitor cells migrating on HA toward a gradient of CXCL12 acquired spread and polarized morphology with CD44 concentrating at the pseudopodia at the leading edge. These morphologic alterations were not observed when the progenitors were first exposed to monoclonal anti-CD44 antibodies, demonstrating a crosstalk between CD44 and CXCR4 signaling (15).

Although RHAMM and CD44 are expressed by the mobilized blood hematopoietic progenitor cells, function blocking monoclonal antibodies identified RHAMM as a major HA binding receptor, with a less consistent participation by CD44. The G-CSF-associated alterations in RHAMM distribution and the RHAMM-dependent motility of hematopoietic progenitor cells suggest a potential role for HA and RHAMM in the trafficking of hematopoietic progenitor cells and the possible use of HA as a mobilizing agent in vivo (328).

C. HA and mesenchymal stem cells

CD44 has long been used as a marker for mesenchymal stem cells (MSC) (504). Later studies demonstrated that HA and HA-binding proteins play an active role in maintaining and differentiation of mesenchymal stem cells. HA improves human mesenchymal stem cell culture growth and promotes chondrogenic differentiation of human mesenchymal stem cells (135, 191). CD44 and HA help mesenchymal stem cells move to injury area to promote tissue regeneration (331). HA promotes a CD44-dependent migration of the mesenchymal stem cells (142). CD44+, not CD44−/−, stem cells injected into mice with acute renal failure migrated to the injured kidney where HA expression was increased, to accelerate morphological or functional recovery (142). Thus, CD44-HA interactions recruit exogenous MSC to injured renal tissue and enhance renal regeneration (142). MSCs actively migrated to cardiac allografts and contributed to graft fibrosis and, to a lesser extent, to myocardial regeneration (477). MSC, in response to PDGF stimulation, express high levels of CD44 standard isoform, which facilitates cell migration through interaction with extracellular HA. Such a migratory mechanism could be critical for the recruitment of MSCs into wound sites for the proposition of tissue regeneration, as well as for the migration of fibroblast progenitors to allografts in the development of graft fibrosis (500).

An HA-based scaffold has been developed for tissue regeneration. Cartilage regeneration using mesenchymal stem cells and a hybrid poly-(lactic-co-glycolic acid)-gelatin/chondroitin/hyaluronate hybrid scaffold had better chondrocyte morphology, integration, continuous subchondral bone, and much thicker newly formed cartilage (94). HA mixed esters of butyric and retinoic Acid drive cardiac and endothelial fate in term placenta human mesenchymal stem cells and enhance cardiac repair in infarcted rat hearts (445). MSCs survive well in the HA-based prototype ligament scaffold, as assessed after 2 days from seeding, and express CD44, a receptor important for scaffold interaction, and proteins responsible for the functional characteristics of the ligaments (61).

XI. HA-TLR INTERACTION IN NONINFECTIOUS TISSUE INJURY

It is becoming clear that Toll-like receptors not only play a role in the recognition of pathogens and in the initiation of immune responses but also have a fundamental role in noninfectious disease pathogenesis. Recent studies from Medzhitov and colleagues suggest that TLRs may have homeostatic functions in the gut epithelium (342). These investigators demonstrated that the stimulation of TLR by commensal intestinal flora is critical for protecting against intestinal epithelial injury (342). This elegant work emphasizes the importance of cross talk between factors involved in innate immunity and cytoprotection in the maintenance of intestinal epithelial homeostasis. We demonstrated that TLR2−/−TLR4−/− mice developed a reduced inflammatory response to lung injury, with a decrease in transepithelial neutrophil migration and reduced expression of MIP-2. However, these mice had a higher mortality rate when compared with wild type mice (175). Inhibition of HA binding with the peptide in vivo recapitulated the phenotype observed in the TLR2−/−TLR4−/− mice after lung injury, enhanced lung injury and increased epithelial cell apoptosis, and decreased transmigration of neutrophils (175). After the administration of a high dose of bleomycin, increasing production of high molecular weight HA by overexpression of HAS2 under direction of the lung-specific CC10 promoter was protective against mortality, lung injury, and epithelial cell apoptosis (175). To investigate the hypothesis that HA and TLR interactions are important in lung injury and repair processes, we asked if HA on the epithelial cell surface plays a role in lung injury. Isolated lung alveolar epithelial cells have increased rates of apoptosis at baseline and exhibit greater apoptosis in response to bleomycin. The exogenous addition of high molecular weight HA is protective against bleomycin-induced apoptosis (175). We found that bleomycin induces both NF-κB activation and apoptosis in primary lung epithelial cells (175). Thus, HA regulates basal NF-κB activation in epithelial cells. NF-κB regulates apoptosis (224) and HA fragments can activate NF-κB in macrophages (306). Primary epithelial cells from TLR2−/−TLR4−/− mice have a significant increase in spontaneous apoptosis relative to wild type (175). Furthermore, we made the remarkable observation that cell-surface HA is severely abrogated in TLR2−/−TLR4−/− epithelial cells (175), although the cause of the loss of HA in these mice is unclear. Epithelial cell-surface HA promotes basal NF-κB activation in a TLR-dependent manner and this activation has a protective effect against injury (309).

TLR4 plays a protective role in oxidant-mediated lung injury by maintaining appropriate levels of antiapoptotic responses in the face of oxidant stress (493). TLR4 also maintains constitutive lung integrity by modulating oxidant generation, preventing the development of emphysema (492). On the other hand, TLR4 (and TLR2) have a deteriorative role during non-infectious tissue injury. Hemorrhage-induced lung TNFα production, neutrophil accumulation, and protein permeability, but not NF-κB activation, is dependent on a functional TLR4 (20). TLR4-TLR2 cross talk activates a positive-feedback signal leading to alveolar macrophage priming and exaggerated lung inflammation in response to invading pathogens during hemorrhage-induced acute lung injury (95). Recently, it was shown that TLR4 signaling through the MyD88-dependent pathway was required for the full development of kidney ischemia/reperfusion injury, as both TLR4- and MyD88-deficient mice were protected against kidney dysfunction, tubular damage, neutrophil and macrophage accumulation, and expression of proinflammatory cytokines and chemokines (478). Upregulation of the endogenous ligands HA, high-mobility group box 1, and biglycan, providing circumstantial evidence that one or more of these ligands may be the source of TLR4 activation (478). The discrepancy on the roles of TLR in tissue injury may be due to the injury models utilized, the degree of the injury, and the timing of endpoints examined.

Although both HA and LPS activate TLR, they provoke different sets of gene expression (418). For example, cultured cells exposed to HA showed a pattern of gene induction that mimics the response seen in mouse skin after sterile injury with an increase in molecules such as TGF-β2 and matrix metalloproteinase-13. These factors were not induced by LPS despite the mutual dependence of both HA and LPS on TLR4, suggesting that TLRs may play different roles in infectious inflammation and in sterile inflammation (418). Gallo and associates made an important observation recently that a unique complex of TLR4, MD-2, and CD44 recognizes HA in noninfectious inflammation (418), different from the TLR4, MD-2 and CD14 complex that recognizes LPS during infection. However, MD-2 is usually associated lipid recognition and binding (199, 314, 321). As HA is a pure glycosaminoglycan and contains no lipid, it is against the involvement of MD-2 in HA-TLR interaction. How MD-2 fits into HA-TLR interaction is to be determined. Furthermore, Gallo and colleagues reported that LPS-TLR4-CD44 signaling is mediated (in part) by A20 (286).

A TLR4 polymorphism (Asp299Gly) attenuates receptor signaling and diminishes the inflammatory response to gram-negative pathogens (315). A recent study found that the Asp299Gly TLR4 polymorphism is associated with a decreased risk of atherosclerosis. This finding pointed to the notion that the innate immunity plays a role in atherogenesis (197). Similarly, inactivation of the MyD88 pathway led to a reduction in atherosclerosis through a decrease in macrophage recruitment to the artery wall that was associated with reduced chemokine levels (29). How HA plays a role in this process, it is to be seen.

XII. LUNG DISEASES

A. Pulmonary Fibrosis

Idiopathic Pulmonary Fibrosis (IPF) is a fatal disease of unknown origin with an average survival of 2–3 years from the time of diagnosis for which no effective medical therapies currently exist (304). Lung transplantation is available for IPF patients under the age of 60, but most patients are older than this when the diagnosis is made (253). Although there is a familial form of pulmonary fibrosis, it is rare and the vast majority of cases are sporadic (25). Patients suffocate from unremitting deposition of collagen in the gas exchanging portions of the lung. The molecular mechanisms of lung injury, inflammation and fibrosis are largely unknown. HA levels were higher in BAL fluid in patients with IPF than in healthy controls (27). The HA levels were also correlated with the severity of the disease. Patients with deteriorated conditions had higher lavage fluid concentrations of HA than the patients those whose disease was stable (27). Fibroblast clones derived from primary fibroblast cultures from the lung tissue of patients with pulmonary fibrosis secreted much greater amounts of HA and proteoglycan decorin than the clones from normal individuals (468). HA levels in BAL fluid in sarcoidosis patients were significantly higher than those of controls (26). The increases were significantly higher in clinically active than in inactive sarcoidosis (26, 31). The concentrations of HA and fibronectin were higher in patients with sarcoidosis compared to the healthy nonsmoking controls (86). We recently reported that serum inter-alpha-trypsin inhibitor and matrix hyaluronan promote angiogenesis in fibrotic lung injury (107).

Elevated HA levels have been reported in experimental pulmonary fibrosis. There was a transient histological accumulation of HA in the alveolar interstitium, corresponding to increases in HA levels in BAL fluid and lung tissue extracts in bleomycin-induced alveolitis in rats (295, 296). Increased extracellular matrix components, fibronectin and HA, as well as polymorphonuclear cells, were detected in BAL fluid in quartz exposed rats (85). While HA in lung tissue and BAL fluid peaked on days 3–7 and then gradually declined towards normal values on days 21–30, lung tissue collagen contents increased between days 7 and 30 (141). The mechanism involved in bleomycin-induced increased HA production in rat lung was associated with growth factors such as PDGF-BB (420), and a decreased HA binding capacity of alveolar macrophages may account for the impairment of internalization and thereby degradation of excessive HA during the early phase of fibrotic lung injury (419).

Although many studies have suggested that HA plays a causative role in the pathogenesis of pulmonary fibrosis, it has been difficult to demonstrate this directly in vivo. One of the challenges is that the disruption of HAS2 allele leads to embryonic leathality and leaves no surviving mice to be subject to injury models. Conditional knockout HAS2 in specific cell types will provide excellent models to dissect the role of HA in biology and pathology.

B. Asthma

HA appears in low concentrations in BAL fluid from healthy individuals, while increased amounts have been reported in lavage fluid from patients with allergic asthma (358, 391). Furthermore, HA levels in BAL fluids were significantly increased in patients with persistent asthma, in comparison to patients with intermittent asthma (449), were correlated with the pulmonary function of the patients (39).

One of the reasons for increased HA levels in asthma is an increase in TGF-β levels which promote airway smooth muscle cells to secret glycosaminoglycans such as HA (30). IL-1β and TNFα were the most potent stimulators of HA synthesis and when combined, caused synergistic increases in HA accumulation in human lung fibroblasts. Fluticasone inhibited IL-1β and TNFα induced HA synthesis, and attenuated IL-1β and TNFα stimulated HA synthase-2 mRNA (472). On the other hand, low molecular weight HA increased GM-CSF, TGF-β and intercellular adhesion molecule-1 expression by eosinophils, leading to eosinophil survival. HA-eosinophil interaction may contribute to the regulation of airway inflammation and airway remodeling (313).

Fibroblasts from subjects with the most hyperresponsive airways in asthma produced much more total proteoglycans such as HA, perlecan, and versican than cells from subjects with less hyperresponsive or normal responsive airways (467). Hyaluronan, secreted from submucosal gland cells, plays a role in mucosal host defense by retaining lactoperoxidase and possibly other substances important for first line host defense at the apical surface ‘ready for use’ and protected from ciliary clearance (359).

Aerosol HA administration significantly reduces the bronchial hyper-reactivity to muscular exercise in asthmatics (326). Such an effect could be attributed to the correction of the pathological remodeling, one of the main features of asthma: a correction which could be attributed to the unique physicochemical properties of this major component of the loose connective amorphous matrix of the airways, which is undoubtedly involved in the remodeling process (326). Others found that inhaled HA does not significantly protect against exercise-induced bronchoconstriction in a randomized double-blinded placebo-controlled crossover study, suggesting that HA at the testing dose is not effective as a prophylaxis for exercise-induced bronchoconstriction in patients with asthma (214).

C. COPD/Pulmonary Emphysema

Young COPD patients with higher concentrations of HA and fibronectin in BAL had higher inflammatory cells in BAL, and lower in pulmonary function measurements, than the COPD patients with lower HA levels (393). HA levels were significantly higher in the sputum from patients with COPD than in those from controls. The COPD population appeared to consist of two subpopulations with either high or moderate HA levels. The subgroup of patients with high HA levels had lower FEV1 than the moderate HA group (393). In addition, neutrophil influx and levels of IL-8, and the soluble TNF receptors were significantly higher in patients with high HA levels than in those with moderate HA levels and controls, indicating a relationship between HA levels, local inflammation and severity of disease, and suggesting enhanced degradation of HA in the lungs of patients with COPD (80). Hyaluronan, chondroitin 4-sulfate, and chondroitin 6-sulfate levels decreased significantly in animals exposed to ozone for 20 months when compared with those in control animals (338). Cigarette smoke exposure leads to enhanced deposition of mostly low molecular weight HA in alveolar and bronchial walls by altering the expression of HA modulating enzymes (40). However it is uncertain if the changes of these glycosaminoglycans contribute the pathogenesis of emphysema.

Aerosolized low molecular weight HA following endotoxin administration significantly increased lung inflammation, whereas pretreatment with HA had the opposite effect (288). Aerosolized HA may be an effective means of preventing pulmonary emphysema and perhaps other lung diseases that involve elastic fiber injury (48). Although clinical trials involving nebulized HA are not expected to yield a measurable treatment effect for at least several years, it is proposed that the special ability of this polysaccharide to retain water may increase the elasticity of lung elastic fibers, producing a relatively rapid improvement in pulmonary mechanics (49). Compared to untreated/smoked controls, aerosolized HA-treated animals showed statistically significant reductions in mean linear intercept and elastic fiber breakdown products in BAL fluid. The aerosolized HA showed preferential binding to elastic fibers, suggesting that it may protect them from injury (47). The effect of HA on inflammation appears to be related to its molecular size, with larger polysaccharide chains having anti-inflammatory activity and smaller ones having proinflammatory properties. The breakdown of inhaled HA into smaller fragments could possibly induce an inflammatory reaction in the lung that counteracts any beneficial effect. Consequently, the proposed therapeutic use of HA will require development of treatment strategies aimed at minimizing its proinflammatory activity (46).

D. Respiratory distress syndrome

The HA concentrations in BAL fluid and in serum from patients with adult respiratory distress syndrome was much higher than that seen in control patients (125, 209). HA staining can be seen in lung sections of patients with severe established adult respiratory distress syndrome (252). The observed accumulation of HA in the small airways in adult respiratory distress syndrome may be expected to immobilize water and thereby contribute to the interstitial and alveolar edema. The inverse correlation was seen between BAL fluid HA and pulmonary oxygenation index (125). In addition, HA concentrations in lung extracts increased with progressively severe respiratory distress syndrome in premature monkeys (184). Bleomycin and hyperoxia cause an increase in lung HA (185, 421). We demonstrated that CD44 plays a role in removing accumulated HA fragments at the site of tissue injury to resolve lung inflammation (421).

XIII. Kidney diseases

A. Nephritis

Crescent formation is a major feature of rapidly progressive glomerulonephritis and is generally associated with a poor prognosis. Crescents are formed by accumulation of monocyte/macrophages and plasma proteins in Bowman’s space, by proliferation of parietal epithelial cells and fibroblasts, and by deposition of the extracellular matrix. Marked accumulation of hyaluronate was demonstrated in developing and sclerosing crescents, in association with local infiltration of T lymphocytes and monocyte/macrophages, cells known to express CD44 (302). CD44 is constitutively expressed in the normal kidney and is dramatically up-regulated in rat crescentic anti-glomerular basement membrane disease, suggesting possible roles for the CD44-HA interaction in leukocyte recruitment, renal fibrosis and tubular cell-matrix and cell-cell interactions during the induction and progression of crescentic glomerulonephritis (183). The expression of CD44, HA, and osteopontin was up-regulated at the early stage of the crescent formation in patients with crescentic glomerulonephritis, suggesting that cell-matrix interactions mediated by the CD44-osteopontin and CD44-HA may play important roles in the formation and progression of the crescents (290). Furthermore, the expression of CD44 in the interstitium correlated with the severity of chronic glomerular lesions (362). The glomerular and interstitial CD44 and HA expression correlated with proteinuria, and the interstitial CD44 and HA expression correlated with creatinine clearance rate. It is believed that CD44 participates in the progression of IgA nephropathy by binding HA and osteopontin (362). CD44 is expressed de novo by tubular epithelial cells in areas of tubular injury in kidneys of kd/kd mice which develop a spontaneous and chronic tubulointerstitial renal disease, but not in normal control kidneys (384). CD44 positive lymphocytes and macrophages also infiltrate the kidney to kd/kd mice. HA also accumulates in kd/kd kidneys in the interstitial space, particularly in cortical areas of tubular injury. The expression of osteopontin is enhanced in kd/kd kidneys, predominantly in areas of tubular injury (384).

HA plays a significant role in thromboxane-mediated immune events in the kidney, where HA stimulates cyclooxygenase-2 expression and subsequent thromboxane A2 production (406). In experimental anti-Thy-1 nephritis, there was an early glomerular influx of CD44+ macrophages and de novo CD44 expression by mesangial cells, suggesting that cell-matrix interactions mediated by the CD44-HA receptor is involved in mesangial cell proliferation in rat anti-Thy-1 nephritis (300). Leukocyte infiltration into tissues in inflammation is a multistep process involving the sequential engagement of adhesion molecules such as selectins. Synthesized sulphated HA showed strong inhibitory effects on the binding of P- and L-selectin in vitro and is minimally antigenic. Thus, sulphated HA is considered to be a candidate selectin-blocking agent for clinical use. Sulfated HA inhibits intraglomerular infiltration of macrophages and prevents progression of experimental crescentic glomerulonephritis (257, 312). Sulfated HA reduced proteinuria, macrophage infiltration, and crescent formation in a dose-dependent manner and reduced urinary protein excretion. Sulfated polysaccharides might be beneficial for the treatment of crescentic glomerulonephritis (257, 312).

B. Lupus Nephritis

Increased serum concentration of HA was noted in dermatomyositis patients (211). HA expression was increased in the mesangium, and in the periglomerular and tubular distribution in kidney biopsies of patients with lupus nephritis (490). Lupus nephritis patients showed increased levels of circulating HA, especially during active disease, which correlated with anti-DNA antibody titers (490). Fibroblasts derived from active lesions of nephrogenic fibrosing dermopathy synthesize elevated levels of HA when compared with normal controls (83). Given that HA plays a pivotal role during inflammatory responses, influences cellular behavior and assists in the recruitment of lymphocytes to sites of injury, it is likely that HA contributes to the pathogenesis of lupus nephritis (490).

C. Renal failure

Serum HA levels were significantly increased in patients with renal insufficiency and with end-stage renal failure when compared with the levels measured in healthy controls. Significant correlations were found between serum HA and degree of impaired renal function (123, 146). Hyaluronan levels were significantly greater in the subgroup with lower glomerular filtration rates, were associated with an inflammatory state, suggesting impaired renal elimination of proinflammatory cytokines, increased generation of cytokines in uremia, or an adverse effect of inflammation on renal function (324). Similarly, Terney and associates found that patients with deteriorated clinical condition often had greater HA levels, suggesting that HA may be a biochemical marker of patients whose condition deteriorates despite renal replacement therapy (440). Serum HA is mainly in a high molecular weight form (440). In addition, serum HA is raised in active vasculitis (466).

Serum HA concentrations predict survival in patients with chronic renal failure on maintenance haemodialysis (475). Serum HA is an accurate predictor of mortality and morbidity over an 18-month period in patients treated by continuous ambulatory peritoneal dialysis. Large quantities of HA are excreted in peritoneal dialysate, which in part represents local HA production (239). Markedly elevated serum HA levels are found in predialysis patients with malnutrition, inflammation, and atherosclerotic cardiovascular disease and that serum HA is a risk predictor of poor survival in dialysis (399).

HA turnover is cleared rapidly in the circulation by both the liver and the kidney. Evidence suggests that high molecular size HA chains, which are anti-inflammatory, antiangiogenic, and immuno-suppressive are cleared by the liver (298). By contrast, intermediate-sized fragments, which are highly angiogenic, inflammatory, and a stimulus for fibrous deposition, are cleared by the kidney. The accumulation of HA fragments in renal failure can account for HA deposition in the dermis and may be a mechanism for the nephrogenic fibrosing dermopathy that can accompany these lesions (298).

XIV. ARTHRITIS

Hyaluronan was first isolated from synovial fluid from patients with rheumatoid arthritis more than 50 years ago (101, 340). The serum HA concentration in patients with rheumatoid arthritis was significantly greater than that in healthy controls (89, 247). The intrinsic viscosity of HA in synovial fluid decreases significantly in mild and severe arthritis compared to that in normal individuals (203). However, while some studies found that the serum levels of HA may not correlate with the severity of the disease (233), others found that serum levels of HA correlate with the severity of the disease and with an objective functional capacity score and with an articular index based on the total amount of cartilage in involved joints (114). Amount of serum HA may be a useful measure of disease activity in patients with rheumatoid arthritis and is a better correlate of clinical disease activity in patients with rheumatoid arthritis than erythrocyte sedimentation rate or C-reactive protein (88). In an experimental condition, serum HA increased as the arthritic lesions developed, correlating with the severity of the disease (28, 115).

HA is susceptible to degradation by excessive reactive oxygen species in rheumatoid arthritis patients and HA markedly decreased the O2, H2O2, and OH in protecting articular tissues from oxidative damage (364). Degradation of hyaluronate in arthritic synovial tissue may be inhibited by radical scavengers (369). The hyaluronate-derived low-molecular-mass oligosaccharide species and formate (HCOO) are suggested as novel markers of reactive oxygen radical activity in the inflamed rheumatoid joint during exercise-induced hypoxic/reperfusion injury (118).

Sodium HA has been used in an intra-articular treatment of arthritis in race-horses (11) as well as in patients with arthritis (205). Sodium hyaluronate and glucocorticoid treatments had a significant positive effect according to the patients’ subjective evaluation (205). In a clinical trial in Japan, significant improvement in pain symptoms and inflammation was observed after the 5 injections of sodium hyaluronate intra-articularly into the knees of 25 patients with rheumatoid arthritis (116). Local therapeutic effects of HA in antigen-induced arthritis in rats are clearly biphasic, with inhibition of inflammation and cartilage damage in the early chronic phase but with promotion of joint swelling, inflammation and cartilage damage in the late chronic phase (355).

It has been suggested that HA-binding proteins play a role in arthritis. CD44 is up-regulated on many synovial cell types in patients with rheumatoid arthritis, and the level of CD44 present in synovial tissue is correlated with the degree of synovial inflammation (134). An anti-CD44 antibody abrogates tissue swelling and leukocyte infiltration (272). On the other hand, treatment with recombinant TNFIP6 protein had a potent ameliorative effect, manifested by decreases in the disease incidence, arthritis index, and footpad swelling, significantly reduced levels of IgG1, IgG2a, and IgG2b antibodies against bovine and murine type II collagen (274). In comparison with wild type mice, the progression and severity of proteoglycan-induced arthritis were significantly greater in Tnfip6-deficient mice with more extensive infiltration of the synovium with neutrophil leukocytes, and elevated serum levels of IL-6 and amyloid A (409).

XV. BRAIN INJURY

Significantly elevated HA levels were found in cerebrospinal fluid patients with spinal stenosis, head injury and cerebral infarction (223). In addition, HA-binding proteins were also upregulated during brain injury. CD44 expression was strongly activated in the area surrounding the injury within 2 days and then persisted for over 2 months (403). BEHAB/brevican is upregulated in response to central nerve system injury (166). Versican was upregulated in central nerve system injury (13). The concomitant induction of CD44 and HAS-2 mRNA expression was detected in the microglia, macrophages, and microvessels of the ischemic brain tissue (452). HA accumulates after injury and activates microglia and macrophages. Although HA-mediated cytokine release and MAPK signaling in microglia was lower than from peritoneal macrophages, resident microglia does respond to extracellular mediators after brain ischemia (453).

In a rat middle cerebral artery occlusion model HA accumulation was seen in stroke-affected areas and hyaluronidase-1 and 2, and CD44 were upregulated after stroke (5). Increased oligosaccharide HA production soon after stroke may be detrimental through enhancement of the inflammatory response, while activation of HA-induced cellular signaling pathways in neurons and microvessels may impact on the remodeling process by stimulating angiogenesis and revascularization, as well as the survival of susceptible neurons (6). HA binding to the cultured astrocytes stimulated Rac1 signaling and cytoskeleton-mediated migration. HA binding to astrocytes stimulated Rac1-dependent protein kinase N-γ kinase activity which, in turn, up-regulated the phosphorylation of the cytoskeletal protein, cortactin, and attenuated the ability of cortactin to cross-link F-actin. HA-CD44-induced astrocyte function may provide important insights into novel therapeutic treatments for tissue repair following central nervous system injury (38). HA accumulates in demyelinated lesions from individuals with multiple sclerosis and in mice with experimental autoimmune encephalomyelitis (16). The addition of high molecular weight HA to oligodendrocyte progenitor cultures reversibly inhibits progenitor-cell maturation, whereas degrading HA in astrocyte-oligodendrocyte progenitor cocultures promotes oligodendrocyte maturation. High molecular weight hyaluronan may therefore contribute substantially to remyelination failure by preventing the maturation of oligodendrocyte progenitors that are recruited to demyelinating lesions (16).

An HA-based scaffold has been developed for tissue regeneration (94). When a HA-poly-D-lysine copolymer hydrogel with an open porous structure was implanted in brain tissue, macrophages and multinucleated foreign body giant cells were found at the site of implantation of the hydrogel, and astrocytes between the hydrogel and the surrounding tissue, demonstrating the promise of the HA-poly-D-lysine hydrogel as a scaffold material for the repair of defects in the brain (428). Similarly, HA hydrogels modified with laminin created a scaffold supported cell infiltration and angiogenesis, and simultaneously inhibit the formation of glial scar (153).

XVI. HEART DISEASES

HA has been implied in the development and the progression of atherogenesis. HA is expressed in all aortic layers. The highest concentration of the human aorta HA was found in the tunica media, exhibiting a negative concentration gradient from the tunica media to the atheromatic plaque (320). HA acts as a negative regulator on the PDGF-induced vascular smooth muscle cell proliferation and as a positive regulator on the PDGF-induced vascular smooth muscle cell migration (320). HA deposits and cyclooxygenase-2 expression are colocalized in the human internal carotid artery (407). In atherosclerosis, HA associates with leukocytes and vascular smooth muscle cells, and is involved in vascular remodeling. HAS1 and HAS2 are upregulated in response to prostaglandins via Gs-coupled prostaglandin receptors in human vascular smooth muscle cells (99).

Accumulation of HA in myocardial interstitial tissue parallels development of transplantation edema in heart allografts in rats (124). Low molecular weight HA increases the uptake of oxidized low-density lipoprotein into monocytes (410). HA retains low density lipoprotein by forming a HA-low-density lipoprotein complex through the macrophage scavenger receptor CD204 (376). CD44 is upregulated in atherosclerotic lesions of apoE-deficient mice. Low molecular weight forms of HA stimulate VCAM-1 expression and proliferation of cultured primary aortic smooth muscle cells, whereas high molecular weight forms of HA inhibit smooth muscle cell proliferation. CD44 plays a critical role in the progression of atherosclerosis through multiple mechanisms (65). Organization of HA- and versican-rich pericellular matrices may facilitate migration and mitosis by diminishing cell surface adhesivity and affecting cell shape through steric exclusion and the viscous properties of HA proteoglycan gels (92). Versican interacts with HA to create expanded viscoelastic pericellular matrices that are required for arterial smooth muscle cell proliferation and migration (470). Versican is prominent in advanced lesions of atherosclerosis, at the borders of lipid-filled necrotic cores as well as at the plaque-thrombus interface, suggesting roles in lipid accumulation, inflammation, and thrombosis (470).

Overproduction of HA in the aorta resulted in thinning of the elastic lamellae in HAS2 transgenic mice, leading to increased mechanical stiffness and strength (53). Overproduction of HA in the genetic background of the ApoE-deficient mouse strain promoted atherosclerosis development in the aorta, suggesting that accumulation of HA accelerates the progression of atherosclerosis (53).

Infarct healing is dependent on an inflammatory reaction that results in leukocyte infiltration and clearance of the wound of dead cells and matrix debris. CD44 expression was markedly induced in the infarcted myocardium and was localized on infiltrating leukocytes, wound myofibroblasts, and vascular cells (156). CD44−/− mice showed enhanced inflammation, decreased fibroblast infiltration, reduced collagen deposition, and diminished proliferative activity (156). Isolated CD44−/− cardiac fibroblasts had reduced proliferation upon stimulation with serum and decreased collagen synthesis in response to TGF-β in comparison to wild type fibroblasts. Thus, CD44-mediated interactions are critically involved in infarct healing in resolution of the postinfarction inflammatory reaction and regulates fibroblast function (156).

CD44 is expressed abundantly in the embryonic myocardium. The differentiation process is accompanied by an induction of CD44 mRNA and protein (469). Synthesized mixed esters of HA with butyric and retinoic acid primed the expression of cardiogenic genes and elicited a remarkable increase in cardiomyocyte yield in mouse embryonic stem cells (444), demonstrating the potential for chemically modifying the gene program of cardiac differentiation without the aid of gene transfer (444). The mixed esters of HA with butyric and retinoic acid enhanced the expression of VEGF, VEGF receptor KDR, and HGF, primed stem cell differentiation into endothelial cells, and increased the transcription of the cardiac lineage-promoting genes GATA-4 and Nkx-2.5 (445, 446).

The Asp299Gly TLR4 polymorphism is associated with a decreased risk of atherosclerosis (197). Similarly, inactivation of the MyD88 pathway led to a reduction in atherosclerosis through a decrease in macrophage recruitment to the artery wall that was associated with reduced chemokine levels (29). These findings pointed to the notion that the innate immunity plays a role in atherogenesis. Whether (or how) HA plays a role in this process, it is to be seen.

XVII. DIABETES

Inflammatory destruction of insulin-producing β cells in the pancreatic islets is the hallmark of insulin-dependent diabetes mellitus. Increased HA levels were seen in diabetic patients (299) and in the glomeruli from diabetic rats (245). Insulin treatment promoted the proliferative response of aorta to injury and this was associated mainly with increased HA production (54). Increased HA, hyaluronidase production and HA degradation were observed in injured aorta of insulin-resistant rats (55). However, others found a decrease of HA in skin in diabetic patients (23) or rats with chronic diabetes mellitus (52). This discrepancy may be due to the measurement of HA during different stages of the disease, inflammatory status and treatment.

Cardiovascular disease contributes to mortality in type 1 diabetes mellitus. Accumulation of HA around smooth muscle cells in lesions of atherosclerosis in diabetic patients. Serum HA levels correlate with poor blood glucose control and diabetic angiopathy and that it could be used as a marker of diabetic angiopathy (275). HA and hyaluronidase were significantly increased in type 1 diabetes when compared to controls. Plasma HA and hyaluronidase correlated in type 1 diabetes. Type 1 diabetes patients show structural changes of the arterial wall associated with increased HA metabolism. These data may lend further support to altered glycosaminoglycan metabolism in type 1 diabetes as a potential mechanism involved in accelerated atherogenesis (299). In a porcine model of atheroscerosis, diabetes was associated with multiple extracellular matrix changes including an increase in HA staining that have been associated with increased lesion instability, greater atherogenic lipoprotein retention and accelerated atherogenesis (261).

In addition, HA binding proteins were also increased in patients with diabetes or in experimental diabetic animal models. The occurrence of LYVE1-expressing lymphatic compartments and the alteration of CCL21 expression in the lymphatics may be involved in defective thymocyte differentiation and migration, and play a significant role in insulitic and diabetic processes (174). Injection of anti-CD44 monoclonal antibody 1 hour before cell transfer of diabetogenic splenocytes and subsequently on alternate days for 4 weeks induced considerable resistance to diabetes in recipient mice, reflected by reduced insulitis (462). A similar antidiabetic effect was observed even when the anti-CD44 monoclonal antibody administration was initiated at the time of disease onset (462). Administration of the enzyme hyaluronidase also induced appreciable resistance to insulin-dependent diabetes mellitus, suggesting that the CD44-HA interaction is involved in the development of the disease.

HA plays an active role in the development of diabetes. HA recruits monocytes to the injury area (128). HA production in response to a raised glucose environment in diabetes can contribute to mesangial hypercellularity (244). HA increases vascular smooth muscle cell expression of PAI-1, a phenomenon that may alter the balance between proteolysis and its inhibition in vessels of patients with type 2 diabetes, thereby contributing to the acceleration of macroangiopathy (254). HA facilitates corneal epithelial wound healing in diabetic rats, which suggests that one possible mechanism of its stimulatory effect lies in its binding to a provisional fibronectin matrix, in both diabetic and non-diabetic rats (291). A regimen consisting of moist wound healing using HA-containing dressings may be a useful adjunct to appropriate diabetic foot ulcer care (443).

Recently, HA-insulin complex was developed to test whether it could be used to treat diabetic patients by oral administration. Glucose-lowering activity was demonstrated after oral administration of the HA-insulin complex to diabetic rats (168). The HA-insulin complex was active after oral administration and the complexed insulin significantly decreased blood glucose concentrations within 1 hour after oral administration in rats (169).

XVIII. LIVER DISEASES

Serum HA levels have long been used as a marker for liver fibrosis. Serum concentrations of HA were higher in patients with active hepatitis compared to that of normal individuals (119, 120, 343). HA is believed to be a better marker than other markers such as amino terminal propeptide of type III procollagen with regard to prediction of development of cirrhosis as well as prediction of symptoms (284, 308). Furthermore, HA levels have a negative correlation with time of survival and HA is a sensitive marker for liver damage in primary biliary cirrhosis (308) and in patients with compensated hepatitis C virus cirrhosis (121). Serum HA levels in children with chronic hepatitis B is a better fibrosis marker than laminin for diagnosing children with advanced liver fibrosis (225). A modification of the Child-Pugh classification of liver cirrhosis by inclusion of HA significantly improves the predictive power of the Child-Pugh classification in alcoholic cirrhosis (207), and in patients with compensated hepatitis C virus cirrhosis (121). Serum HA can also be used as a marker for liver fibrosis in patients with asymptomatic chronic viral hepatitis B with portal inflammation, in assessing and monitoring time trends in liver disease, substituting for repeated biopsies (330). Moreover, HA levels can be used to monitor the antifibrotic effect of lamivudine in children with chronic hepatitis B long-term lamivudine treatment (225). The serum HA level is regarded as a useful predictor for hepatic regeneration after hepatectomy (311). In addition, hyaluronidase levels are elevated after liver injury and measures of circulating hyaluronidase activity may be used to assess liver damage (109).

Increased ascitic levels of HA in liver cirrhosis is found in patients with cirrhosis, suggesting that simultaneous increased synthesis of HA by the peritoneal cells and a reduction of degradation by liver endothelial cells occur in patients with cirrhosis with ascites. This event of increased HA synthesis may be contributory to remodeling and regeneration of the peritoneal lining (217). Hyaluronan analysis indicated that a certain glycosaminoglycan level is required in ascites before its appearance in plasma. The simultaneous increased HA levels in ascitic fluid do not seem to be derived from the systemic circulation. Correlation analyses for TGF-β and IL-6 indicated a strong dependence of the production of HA on cytokine levels and, to a lesser extent, on IL-1β levels, in ascitic fluid of cirrhotic patients (361).

Hyaluronan is excreted in human from the liver and urine (222). The increased level of circulating endogenous HA found in patients with alcoholic cirrhosis is caused by a combination of increased supply to and decreased extraction from plasma (140). HA turnover occurs systemically from the lymph and serum as well as locally by the same cells responsible for its synthesis. Local turnover involves receptor-mediated uptake and delivery to lysosomes. Some fraction of the HA bound to CD44 becomes internalized and delivered to lysosomes regulated possibly by alternatively spliced isoforms of CD44, changes in CD44 phosphorylation, changes in cytoskeletal binding proteins or the activity of extracellular proteolytic activity (201).

In an IL-2-induced vascular leak syndrome model, CD44−/− mice showed a markedly reduced vascular leak syndrome in the lungs and liver, suggesting that CD44 plays a key role in endothelial cell injury by cytotoxic lymphocytes (339). Treatment with HA enhanced the IL-2-induced edema and lymphocytic infiltration in these organs and caused marked increase in IL-2-induced lymphokine-activated killer cell activity, whereas administration of anti-CD44 monoclonal antibodies caused a significant decrease in edema and lymphokine-activated killer cell activity but similar levels of lymphocytic infiltration (285). On the other hand, in a Con A induced liver injury model, CD44−/− mice showed a markedly increased hepatitis, increased production of cytokines such as TNFα, IL-2 and IFN-γ, T cells from CD44−/− mice were more resistant to activation-induced cell death when compared with the wild type mice. Activated T cells use CD44 to undergo apoptosis, and dysregulation in this pathway could lead to increased pathogenesis in a number of diseases, including hepatitis (57). Administration of staphylococcal enterotoxin B to CD44−/− mice caused significantly enhanced liver damage which correlated with elevated numbers of T cells, NK cells, NKT cells, and macrophages in the liver and increased production of TNFα and interferon-γ compared to wild type mice (263). These studies demonstrated a protective role for CD44 during hepatic injury.

XIX. SUMMARY

Hyaluronan as an important part of extracellular matrix accumulates during inflammation and tissue injury. Hyaluronan is degradated to smaller species by reactive oxygen species and possible by hyaluronidases at the injury site. Hyaluronan regulates cytokines and other inflammatory substances and influence inflammatory cell recruitment and chemotaxis. The actions of hyaluronan are dependent on its interacting proteins and cells. Hyaluronan binds to an array of proteins to elicit its biological roles. Hyaluronan-CD44 interactions have long been suggested important in leukocyte homing and recruitment. The recent demonstration of hyaluronan-Toll like receptor interactions provides the molecular insight into the mechanisms of hyaluronan signaling in inflammatory cells as well as in epithelial cells. Hyaluronan and its binding proteins play a role in the pathogenesis of many human diseases and in numerous experimental conditions. Therapeutic developments targeting hyaluronan and its binding proteins are developing. Understanding the role of hyaluronan and its binding proteins in the pathobiology of disease will facilitate the development of novel therapeutics for many critical diseases.

Figure 3. HAS2-deficient mice.

Figure 3

Disruption of hyaluronan synthase 2 resulted in the defect in heart formation. Representative wild-type (A) and Has2−/− (B) embryos at E9.5. Note the diminished size, the bloodless heart, and distorted somites of the Has2−/− embryo. E9.5 wild-type (C) and Has2−/− (D) embryos stained for the endothelial marker PECAM. Note the absence of an organized vascular network expressing PECAM in the Has2−/− embryo. P, pericardium; E, endoderm; M, mesoderm; OpP, optic placode; OtP, otic placode; first and secondpharyngeal pouches are numbered. From Camenisch et al., J. Clin. Invest. 106:349–360 (2000), with permission.

FIGURE 4. HA FRAGMENT.

FIGURE 4

Purified HA fragments but not HMW-HAinduced NF-KB DNA binding activity in MH-S cells. (A) Densitometric scanning demonstrating the molecular weights of HMW-HA (top), and purified HA fragments (bottom). (B) Electrophoretic mobility shift assay of nuclear extracts prepared from MH-S cell stimulated for 2 h with either serum-free media (unstim), HMW-HA, HA-fragment, or LPS. HA fragments induced NF-KB DNA binding activity. From Noble et al., J. Exp. Med. 183:2373–2378 (1996), with permission.

FIGURE 5. CD44 NULL HA STAINING.

FIGURE 5

Accumulation of HA after bleomycin treatment. Lung tissue stained for HA at day10 through 14. Wild-type (A) and CD44-deficient mice (B). From Teder et al., Science 296, 155 (2002), with permission.

FIGURE 6. HA FRAGMENT ACCUMULATION IN CD44 MICE.

FIGURE 6

Accumulation of HA after bleomycin treatment. (A) HA content was measured by an HA-specific enzyme-linked immunosorbent assay in lungs of wild-type (open square) and CD44-deficient mice (closed square). (B) HA MW at day 7 in lungs of saline-treatedwild-type mouse (top panel, MWaverage 1.443106), bleomycin-treated wild-type mouse (middle panel, MW average 0.54 3 106), and bleomycin-treatedCD44-de?cient mouse (bottom panel, MW averagesfrom left to right 2.10 3 106, 1.41 3 106, 0.16 3 106, and,0.02 3 106). From Teder et al., Science 296, 155 (2002), with permission.

FIGURE 7. CD44 AND HA SIGNALING.

FIGURE 7

HA signaling is independent on the presence of CD44. Chemokine CXCL2 expression by peritoneal macrophages was not affected by the deficiency of CD44. From Jiang et al., Nat Med, 11:1173 (2005), with permission.

FIGURE 8. TLR AND HA SIGNALING.

FIGURE 8

HA signaling is independent on the presence of CD44. Chemokine CXCL2 expression by peritoneal macrophages was not affected by the deficiency of CD44. From Jiang et al., Nat Med, 11:1173 (2005), with permission.

Figure 9. TLR and lung injury.

Figure 9

TLR2 and TLR4 double deficient mice were more susceptible (A), while mice overexpressing HAS2 on epithelial cells were more resistant, to bleomycin induced lung injury (B).

Figure 10. TLR, HA, and NF-kB.

Figure 10

TLR2 and TLR4 double deficient mice displayed lower basal NK-kB activity (A). The epithelial cells from TLR2−/− TLR4−/− mice expressed reduced cell surface HA (B).

Figure 11. Analogous signaling of HA and LPS.

Figure 11

Analogous signaling of HA and LPS. LPS uses TLR4, CD14 and MD2 to transduce its signal, while HA interacts with TLR4, CD44 and MD2 to elicit its signaling.

TABLE 1.

Selected genes induced by HA fragments.

Category Gene/protein Cell type References
Chemokines CCL3 Macrophages (265)
CCL4 Macrophages (265)
CXCL2 Macrophages (17, 175)
CCL5 Macrophages (265)
CCL2 Renal tubular epithelial cells (22, 265)
CXCL10 Macrophage (149, 265)
CXCL9 Macrophage (149)
CXCL1 Endothelial cells (411)
CCL5 Macrophages (265)
IL-8 Endothelial cells, epithelial cells (255, 417, 451)
CXCL1 Macrophages (148, 175)
Cytokines IL-12 Macropahges, dentritic cells (144, 425)
TNFα Dentritic cells (425)
IL-1β Dentritic cells (425)
Growth factors TGF-β2 Monocytes (418)
IGF-1 Macropahges (305)
Transcription factors IκBα Macrophages (306)
AP-1 Endothelial cells (74)
Rest Monocytes (418)
ECM MMP-10 Endothelial cells (417)
MMP-13 Monocytes, dentritic cells (97, 418)
PAI-1 Macrophages (150)
uPA Macrophages (150)
MME Macrophages (151)
MMP-9 Dentritic cells (97)
Collagen VIII Endothelial cells (354)
HSPG Syndecan-4 Endothelial cells (417)
Others iNOS Hepatocytes, endothelial, Kupffer, and stellate cells (353)
Cox2 Renal tubular epithelial cells (406)
MDR-1 Lymphocytes (436)
Trdn Monocytes (418)
Frk Monocytes (418)

Acknowledgments

The authors wish to thank the members of the Noble laboratory for their hard work in the areas of hyaluronan, innate immune and lung injury. This work has been supported by NIH grants HL060539, AI052201 and HL084917.

GRANTS

This work was supported by the National Institutes of Health grants R01-HL 060539, R01 AI052201, and RO1 HL 077291.

References

  • 1.Acharya PS, Majumdar S, Jacob M, Hayden J, Mrass P, Weninger W, Assoian RK, Pure E. Fibroblast migration is mediated by CD44-dependent TGF beta activation. J Cell Sci. 2008;121:1393–1402. doi: 10.1242/jcs.021683. [DOI] [PubMed] [Google Scholar]
  • 2.Aderem A, Ulevitch RJ. Toll-like receptors in the induction of the innate immune response. Nature. 2000;406:782–787. doi: 10.1038/35021228. [DOI] [PubMed] [Google Scholar]
  • 3.Agren UM, Tammi RH, Tammi MI. Reactive oxygen species contribute to epidermal hyaluronan catabolism in human skin organ culture. Free Radic Biol Med. 1997;23:996–1001. doi: 10.1016/s0891-5849(97)00098-1. [DOI] [PubMed] [Google Scholar]
  • 4.Akishima Y, Ito K, Zhang L, Ishikawa Y, Orikasa H, Kiguchi H, Akasaka Y, Komiyama K, Ishii T. Immunohistochemical detection of human small lymphatic vessels under normal and pathological conditions using the LYVE-1 antibody. Virchows Arch. 2004;444:153–157. doi: 10.1007/s00428-003-0950-8. [DOI] [PubMed] [Google Scholar]
  • 5.Al Qteishat A, Gaffney JJ, Krupinski J, Slevin M. Hyaluronan expression following middle cerebral artery occlusion in the rat. Neuroreport. 2006;17:1111–1114. doi: 10.1097/01.wnr.0000227986.69680.20. [DOI] [PubMed] [Google Scholar]
  • 6.Al’Qteishat A, Gaffney J, Krupinski J, Rubio F, West D, Kumar S, Kumar P, Mitsios N, Slevin M. Changes in hyaluronan production and metabolism following ischaemic stroke in man. Brain. 2006;129:2158–2176. doi: 10.1093/brain/awl139. [DOI] [PubMed] [Google Scholar]
  • 7.Alstergren P, Zhu B, Glogauer M, Mak TW, Ellen RP, Sodek J. Polarization and directed migration of murine neutrophils is dependent on cell surface expression of CD44. Cell Immunol. 2004;231:146–157. doi: 10.1016/j.cellimm.2005.01.007. [DOI] [PubMed] [Google Scholar]
  • 8.Ariel A, Lider O, Brill A, Cahalon L, Savion N, Varon D, Hershkoviz R. Induction of interactions between CD44 and hyaluronic acid by a short exposure of human T cells to diverse pro-inflammatory mediators. Immunology. 2000;100:345–351. doi: 10.1046/j.1365-2567.2000.00059.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Armstrong SE, Bell DR. Relationship between lymph and tissue hyaluronan in skin and skeletal muscle. Am J Physiol Heart Circ Physiol. 2002;283:H2485–2494. doi: 10.1152/ajpheart.00385.2002. [DOI] [PubMed] [Google Scholar]
  • 10.Aruffo A, Stamenkovic I, Melnick M, Underhill CB, Seed B. CD44 is the principal cell surface receptor for hyaluronate. Cell. 1990;61:1303–1313. doi: 10.1016/0092-8674(90)90694-a. [DOI] [PubMed] [Google Scholar]
  • 11.Asheim A, Lindblad G. Intra-articular treatment of arthritis in race-horses with sodium hyaluronate. Acta Vet Scand. 1976;17:379–394. doi: 10.1186/BF03547893. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Asher RA, Morgenstern DA, Fidler PS, Adcock KH, Oohira A, Braistead JE, Levine JM, Margolis RU, Rogers JH, Fawcett JW. Neurocan is upregulated in injured brain and in cytokine-treated astrocytes. J Neurosci. 2000;20:2427–2438. doi: 10.1523/JNEUROSCI.20-07-02427.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Asher RA, Morgenstern DA, Shearer MC, Adcock KH, Pesheva P, Fawcett JW. Versican is upregulated in CNS injury and is a product of oligodendrocyte lineage cells. J Neurosci. 2002;22:2225–2236. doi: 10.1523/JNEUROSCI.22-06-02225.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Atmuri V, Martin DC, Hemming R, Gutsol A, Byers S, Sahebjam S, Thliveris JA, Mort JS, Carmona E, Anderson JE, Dakshinamurti S, Triggs-Raine B. Hyaluronidase 3 (HYAL3) knockout mice do not display evidence of hyaluronan accumulation. Matrix Biol. 2008;27:653–660. doi: 10.1016/j.matbio.2008.07.006. [DOI] [PubMed] [Google Scholar]
  • 15.Avigdor A, Goichberg P, Shivtiel S, Dar A, Peled A, Samira S, Kollet O, Hershkoviz R, Alon R, Hardan I, Ben-Hur H, Naor D, Nagler A, Lapidot T. CD44 and hyaluronic acid cooperate with SDF-1 in the trafficking of human CD34+ stem/progenitor cells to bone marrow. Blood. 2004;103:2981–2989. doi: 10.1182/blood-2003-10-3611. [DOI] [PubMed] [Google Scholar]
  • 16.Back SA, Tuohy TM, Chen H, Wallingford N, Craig A, Struve J, Luo NL, Banine F, Liu Y, Chang A, Trapp BD, Bebo BF, Jr, Rao MS, Sherman LS. Hyaluronan accumulates in demyelinated lesions and inhibits oligodendrocyte progenitor maturation. Nat Med. 2005;11:966–972. doi: 10.1038/nm1279. [DOI] [PubMed] [Google Scholar]
  • 17.Bai KJ, Spicer AP, Mascarenhas MM, Yu L, Ochoa CD, Garg HG, Quinn DA. The role of hyaluronan synthase 3 in ventilator-induced lung injury. Am J Respir Crit Care Med. 2005;172:92–98. doi: 10.1164/rccm.200405-652OC. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Baier C, Baader SL, Jankowski J, Gieselmann V, Schilling K, Rauch U, Kappler J. Hyaluronan is organized into fiber-like structures along migratory pathways in the developing mouse cerebellum. Matrix Biol. 2007;26:348–358. doi: 10.1016/j.matbio.2007.02.002. [DOI] [PubMed] [Google Scholar]
  • 19.Banerji S, Ni J, Wang SX, Clasper S, Su J, Tammi R, Jones M, Jackson DG. LYVE-1, a new homologue of the CD44 glycoprotein, is a lymph-specific receptor for hyaluronan. J Cell Biol. 1999;144:789–801. doi: 10.1083/jcb.144.4.789. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.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:R592–599. doi: 10.1152/ajpregu.00412.2003. [DOI] [PubMed] [Google Scholar]
  • 21.Bates EJ, Harper GS, Lowther DA, Preston BN. Effect of oxygen-derived reactive species on cartilage proteoglycan-hyaluronate aggregates. Biochem Int. 1984;8:629–637. [PubMed] [Google Scholar]
  • 22.Beck-Schimmer B, Oertli B, Pasch T, Wuthrich RP. Hyaluronan induces monocyte chemoattractant protein-1 expression in renal tubular epithelial cells. J Am Soc Nephrol. 1998;9:2283–2290. doi: 10.1681/ASN.V9122283. [DOI] [PubMed] [Google Scholar]
  • 23.Bertheim U, Engstrom-Laurent A, Hofer PA, Hallgren P, Asplund J, Hellstrom S. Loss of hyaluronan in the basement membrane zone of the skin correlates to the degree of stiff hands in diabetic patients. Acta Derm Venereol. 2002;82:329–334. doi: 10.1080/000155502320624041. [DOI] [PubMed] [Google Scholar]
  • 24.Bharadwaj A, Ghosh I, Sengupta A, Cooper TG, Weinbauer GF, Brinkworth MH, Nieschlag E, Datta K. Stage-specific expression of proprotein form of hyaluronan binding protein 1 (HABP1) during spermatogenesis in rat. Mol Reprod Dev. 2002;62:223–232. doi: 10.1002/mrd.10135. [DOI] [PubMed] [Google Scholar]
  • 25.Bitterman PB, Rennard SI, Keogh BA, Wewers MD, Adelberg S, Crystal RG. Familial idiopathic pulmonary fibrosis. Evidence of lung inflammation in unaffected family members. N Engl J Med. 1986;314:1343–1347. doi: 10.1056/NEJM198605223142103. [DOI] [PubMed] [Google Scholar]
  • 26.Bjermer L, Eklund A, Blaschke E. Bronchoalveolar lavage fibronectin in patients with sarcoidosis: correlation to hyaluronan and disease activity. Eur Respir J. 1991;4:965–971. [PubMed] [Google Scholar]
  • 27.Bjermer L, Lundgren R, Hallgren R. Hyaluronan and type III procollagen peptide concentrations in bronchoalveolar lavage fluid in idiopathic pulmonary fibrosis. Thorax. 1989;44:126–131. doi: 10.1136/thx.44.2.126. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Bjork J, Kleinau S, Tengblad A, Smedegard G. Elevated levels of serum hyaluronate and correlation with disease activity in experimental models of arthritis. Arthritis Rheum. 1989;32:306–311. doi: 10.1002/anr.1780320312. [DOI] [PubMed] [Google Scholar]
  • 29.Bjorkbacka H, Kunjathoor VV, Moore KJ, Koehn S, Ordija CM, Lee MA, Means T, Halmen K, Luster AD, Golenbock DT, Freeman MW. Reduced atherosclerosis in MyD88-null mice links elevated serum cholesterol levels to activation of innate immunity signaling pathways. Nat Med. 2004;10:416–421. doi: 10.1038/nm1008. [DOI] [PubMed] [Google Scholar]
  • 30.Black PN, Young PG, Skinner SJ. Response of airway smooth muscle cells to TGF-beta 1: effects on growth and synthesis of glycosaminoglycans. Am J Physiol. 1996;271:L910–917. doi: 10.1152/ajplung.1996.271.6.L910. [DOI] [PubMed] [Google Scholar]
  • 31.Blaschke E, Eklund A, Hernbrand R. Extracellular matrix components in bronchoalveolar lavage fluid in sarcoidosis and their relationship to signs of alveolitis. Am Rev Respir Dis. 1990;141:1020–1025. doi: 10.1164/ajrccm/141.4_Pt_1.1020. [DOI] [PubMed] [Google Scholar]
  • 32.Blundell CD, Mahoney DJ, Cordell MR, Almond A, Kahmann JD, Perczel A, Taylor JD, Campbell ID, Day AJ. Determining the molecular basis for the pH-dependent interaction between the link module of human TSG-6 and hyaluronan. J Biol Chem. 2007;282:12976–12988. doi: 10.1074/jbc.M611713200. [DOI] [PubMed] [Google Scholar]
  • 33.Bockle BC, Solder E, Kind S, Romani N, Sepp NT. DC-SIGN+ CD163+ Macrophages Expressing Hyaluronan Receptor LYVE-1 Are Located within Chorion Villi of the Placenta. Placenta. 2008;29:187–192. doi: 10.1016/j.placenta.2007.11.003. [DOI] [PubMed] [Google Scholar]
  • 34.Bollyky PL, Lord JD, Masewicz SA, Evanko SP, Buckner JH, Wight TN, Nepom GT. Cutting edge: high molecular weight hyaluronan promotes the suppressive effects of CD4+CD25+ regulatory T cells. J Immunol. 2007;179:744–747. doi: 10.4049/jimmunol.179.2.744. [DOI] [PubMed] [Google Scholar]
  • 35.Bost F, Diarra-Mehrpour M, Martin JP. Inter-alpha-trypsin inhibitor proteoglycan family--a group of proteins binding and stabilizing the extracellular matrix. Eur J Biochem. 1998;252:339–346. doi: 10.1046/j.1432-1327.1998.2520339.x. [DOI] [PubMed] [Google Scholar]
  • 36.Boucher WS, Letourneau R, Huang M, Kempuraj D, Green M, Sant GR, Theoharides TC. Intravesical sodium hyaluronate inhibits the rat urinary mast cell mediator increase triggered by acute immobilization stress. J Urol. 2002;167:380–384. [PubMed] [Google Scholar]
  • 37.Bourguignon LY, Gilad E, Brightman A, Diedrich F, Singleton P. Hyaluronan-CD44 interaction with leukemia-associated RhoGEF and epidermal growth factor receptor promotes Rho/Ras co-activation, phospholipase C epsilon-Ca2+ signaling, and cytoskeleton modification in head and neck squamous cell carcinoma cells. J Biol Chem. 2006;281:14026–14040. doi: 10.1074/jbc.M507734200. [DOI] [PubMed] [Google Scholar]
  • 38.Bourguignon LY, Gilad E, Peyrollier K, Brightman A, Swanson RA. Hyaluronan-CD44 interaction stimulates Rac1 signaling and PKN gamma kinase activation leading to cytoskeleton function and cell migration in astrocytes. J Neurochem. 2007;101:1002–1017. doi: 10.1111/j.1471-4159.2007.04485.x. [DOI] [PubMed] [Google Scholar]
  • 39.Bousquet J, Chanez P, Lacoste JY, Enander I, Venge P, Peterson C, Ahlstedt S, Michel FB, Godard P. Indirect evidence of bronchial inflammation assessed by titration of inflammatory mediators in BAL fluid of patients with asthma. J Allergy Clin Immunol. 1991;88:649–660. doi: 10.1016/0091-6749(91)90159-l. [DOI] [PubMed] [Google Scholar]
  • 40.Bracke KR, Dentener MA, Papakonstantinou E, Vernooy JH, Demoor T, Pauwels NS, Cleutjens J, van Suylen R, Joos GF, Brusselle GG, Wouters EF. Enhanced Deposition of Low Weight Hyaluronan in Lungs of Cigarette Smoke-Exposed Mice. Am J Respir Cell Mol Biol. 2009 doi: 10.1165/rcmb.2008-0424OC. [DOI] [PubMed] [Google Scholar]
  • 41.Brakebusch C, Seidenbecher CI, Asztely F, Rauch U, Matthies H, Meyer H, Krug M, Bockers TM, Zhou X, Kreutz MR, Montag D, Gundelfinger ED, Fassler R. Brevican-deficient mice display impaired hippocampal CA1 long-term potentiation but show no obvious deficits in learning and memory. Mol Cell Biol. 2002;22:7417–7427. doi: 10.1128/MCB.22.21.7417-7427.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Busse WW, Lemanske RF., Jr Asthma. N Engl J Med. 2001;344:350–362. doi: 10.1056/NEJM200102013440507. [DOI] [PubMed] [Google Scholar]
  • 43.Calabro A, Oken MM, Hascall VC, Masellis AM. Characterization of hyaluronan synthase expression and hyaluronan synthesis in bone marrow mesenchymal progenitor cells: predominant expression of HAS1 mRNA and up-regulated hyaluronan synthesis in bone marrow cells derived from multiple myeloma patients. Blood. 2002;100:2578–2585. doi: 10.1182/blood-2002-01-0030. [DOI] [PubMed] [Google Scholar]
  • 44.Camenisch TD, Schroeder JA, Bradley J, Klewer SE, McDonald JA. Heart-valve mesenchyme formation is dependent on hyaluronan-augmented activation of ErbB2-ErbB3 receptors. Nat Med. 2002;8:850–855. doi: 10.1038/nm742. [DOI] [PubMed] [Google Scholar]
  • 45.Camenisch TD, Spicer AP, Brehm-Gibson T, Biesterfeldt J, Augustine ML, Calabro A, Jr, Kubalak S, Klewer SE, McDonald JA. Disruption of hyaluronan synthase-2 abrogates normal cardiac morphogenesis and hyaluronan-mediated transformation of epithelium to mesenchyme. J Clin Invest. 2000;106:349–360. doi: 10.1172/JCI10272. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Cantor JO. Potential therapeutic applications of hyaluronan in the lung. Int J Chron Obstruct Pulmon Dis. 2007;2:283–288. [PMC free article] [PubMed] [Google Scholar]
  • 47.Cantor JO, Cerreta JM, Ochoa M, Ma S, Chow T, Grunig G, Turino GM. Aerosolized hyaluronan limits airspace enlargement in a mouse model of cigarette smoke-induced pulmonary emphysema. Exp Lung Res. 2005;31:417–430. doi: 10.1080/01902140590918669. [DOI] [PubMed] [Google Scholar]
  • 48.Cantor JO, Shteyngart B, Cerreta JM, Liu M, Armand G, Turino GM. The effect of hyaluronan on elastic fiber injury in vitro and elastase-induced airspace enlargement in vivo. Proc Soc Exp Biol Med. 2000;225:65–71. doi: 10.1046/j.1525-1373.2000.22508.x. [DOI] [PubMed] [Google Scholar]
  • 49.Cantor JO, Turino GM. Can exogenously administered hyaluronan improve respiratory function in patients with pulmonary emphysema? Chest. 2004;125:288–292. doi: 10.1378/chest.125.1.288. [DOI] [PubMed] [Google Scholar]
  • 50.Carter GR. Improved hemagglutination test for identifying type A strains of Pasteurella multocida. Appl Microbiol. 1972;24:162–163. doi: 10.1128/am.24.1.162-163.1972. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Casalino-Matsuda SM, Monzon ME, Forteza RM. Epidermal growth factor receptor activation by epidermal growth factor mediates oxidant-induced goblet cell metaplasia in human airway epithelium. Am J Respir Cell Mol Biol. 2006;34:581–591. doi: 10.1165/rcmb.2005-0386OC. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Cechowska-Pasko M, Palka J, Bankowski E. Decreased biosynthesis of glycosaminoglycans in the skin of rats with chronic diabetes mellitus. Exp Toxicol Pathol. 1999;51:239–243. doi: 10.1016/S0940-2993(99)80105-5. [DOI] [PubMed] [Google Scholar]
  • 53.Chai S, Chai Q, Danielsen CC, Hjorth P, Nyengaard JR, Ledet T, Yamaguchi Y, Rasmussen LM, Wogensen L. Overexpression of hyaluronan in the tunica media promotes the development of atherosclerosis. Circ Res. 2005;96:583–591. doi: 10.1161/01.RES.0000158963.37132.8b. [DOI] [PubMed] [Google Scholar]
  • 54.Chajara A, Raoudi M, Delpech B, Courel M, Leroy M, Basuyau JP, Levesque H. Effects of diabetes and insulin treatment of diabetic rats on hyaluronan and hyaluronectin production in injured aorta. J Vasc Res. 1999;36:209–221. doi: 10.1159/000025644. [DOI] [PubMed] [Google Scholar]
  • 55.Chajara A, Raoudi M, Delpech B, Leroy M, Basuyau JP, Levesque H. Increased hyaluronan and hyaluronidase production and hyaluronan degradation in injured aorta of insulin-resistant rats. Arterioscler Thromb Vasc Biol. 2000;20:1480–1487. doi: 10.1161/01.atv.20.6.1480. [DOI] [PubMed] [Google Scholar]
  • 56.Chang EJ, Kim HJ, Ha J, Kim HJ, Ryu J, Park KH, Kim UH, Lee ZH, Kim HM, Fisher DE, Kim HH. Hyaluronan inhibits osteoclast differentiation via Toll-like receptor 4. J Cell Sci. 2007;120:166–176. doi: 10.1242/jcs.03310. [DOI] [PubMed] [Google Scholar]
  • 57.Chen D, McKallip RJ, Zeytun A, Do Y, Lombard C, Robertson JL, Mak TW, Nagarkatti PS, Nagarkatti M. CD44-deficient mice exhibit enhanced hepatitis after concanavalin A injection: evidence for involvement of CD44 in activation-induced cell death. J Immunol. 2001;166:5889–5897. doi: 10.4049/jimmunol.166.10.5889. [DOI] [PubMed] [Google Scholar]
  • 58.Choudhary M, Zhang X, Stojkovic P, Hyslop L, Anyfantis G, Herbert M, Murdoch AP, Stojkovic M, Lako M. Putative role of hyaluronan and its related genes, HAS2 and RHAMM, in human early preimplantation embryogenesis and embryonic stem cell characterization. Stem Cells. 2007;25:3045–3057. doi: 10.1634/stemcells.2007-0296. [DOI] [PubMed] [Google Scholar]
  • 59.Clark RA, Lin F, Greiling D, An J, Couchman JR. Fibroblast invasive migration into fibronectin/fibrin gels requires a previously uncharacterized dermatan sulfate-CD44 proteoglycan. J Invest Dermatol. 2004;122:266–277. doi: 10.1046/j.0022-202X.2004.22205.x. [DOI] [PubMed] [Google Scholar]
  • 60.Craig EA, Parker P, Camenisch TD. Size-dependent regulation of Snail2 by hyaluronan: its role in cellular invasion. Glycobiology. 2009;19:890–898. doi: 10.1093/glycob/cwp064. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Cristino S, Grassi F, Toneguzzi S, Piacentini A, Grigolo B, Santi S, Riccio M, Tognana E, Facchini A, Lisignoli G. Analysis of mesenchymal stem cells grown on a three-dimensional HYAFF 11-based prototype ligament scaffold. J Biomed Mater Res A. 2005;73:275–283. doi: 10.1002/jbm.a.30261. [DOI] [PubMed] [Google Scholar]
  • 62.Csoka AB, Frost GI, Stern R. The six hyaluronidase-like genes in the human and mouse genomes. Matrix Biol. 2001;20:499–508. doi: 10.1016/s0945-053x(01)00172-x. [DOI] [PubMed] [Google Scholar]
  • 63.Csoka AB, Frost GI, Wong T, Stern R. Purification and microsequencing of hyaluronidase isozymes from human urine. FEBS Lett. 1997;417:307–310. doi: 10.1016/s0014-5793(97)01309-4. [DOI] [PubMed] [Google Scholar]
  • 64.Csoka AB, Scherer SW, Stern R. Expression analysis of six paralogous human hyaluronidase genes clustered on chromosomes 3p21 and 7q31. Genomics. 1999;60:356–361. doi: 10.1006/geno.1999.5876. [DOI] [PubMed] [Google Scholar]
  • 65.Cuff CA, Kothapalli D, Azonobi I, Chun S, Zhang Y, Belkin R, Yeh C, Secreto A, Assoian RK, Rader DJ, Pure E. The adhesion receptor CD44 promotes atherosclerosis by mediating inflammatory cell recruitment and vascular cell activation. J Clin Invest. 2001;108:1031–1040. doi: 10.1172/JCI12455. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.D’Souza M, Datta K. A novel glycoprotein that binds to hyaluronic acid. Biochem Int. 1986;13:79–88. [PubMed] [Google Scholar]
  • 67.D’Souza M, Datta K. Evidence for naturally occurring hyaluronic acid binding protein in rat liver. Biochem Int. 1985;10:43–51. [PubMed] [Google Scholar]
  • 68.Davies JE, Tang X, Denning JW, Archibald SJ, Davies SJ. Decorin suppresses neurocan, brevican, phosphacan and NG2 expression and promotes axon growth across adult rat spinal cord injuries. Eur J Neurosci. 2004;19:1226–1242. doi: 10.1111/j.1460-9568.2004.03184.x. [DOI] [PubMed] [Google Scholar]
  • 69.Day AJ, de la Motte CA. Hyaluronan cross-linking: a protective mechanism in inflammation? Trends Immunol. 2005;26:637–643. doi: 10.1016/j.it.2005.09.009. [DOI] [PubMed] [Google Scholar]
  • 70.Day AJ, Sheehan JK. Hyaluronan: polysaccharide chaos to protein organisation. Curr Opin Struct Biol. 2001;11:617–622. doi: 10.1016/s0959-440x(00)00256-6. [DOI] [PubMed] [Google Scholar]
  • 71.DeAngelis PL. Hyaluronan synthases: fascinating glycosyltransferases from vertebrates, bacterial pathogens, and algal viruses. Cell Mol Life Sci. 1999;56:670–682. doi: 10.1007/s000180050461. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.DeAngelis PL, Papaconstantinou J, Weigel PH. Isolation of a Streptococcus pyogenes gene locus that directs hyaluronan biosynthesis in acapsular mutants and in heterologous bacteria. J Biol Chem. 1993;268:14568–14571. [PubMed] [Google Scholar]
  • 73.DeAngelis PL, Papaconstantinou J, Weigel PH. Molecular cloning, identification, and sequence of the hyaluronan synthase gene from group A Streptococcus pyogenes. J Biol Chem. 1993;268:19181–19184. [PubMed] [Google Scholar]
  • 74.Deed R, Rooney P, Kumar P, Norton JD, Smith J, Freemont AJ, Kumar S. Early-response gene signalling is induced by angiogenic oligosaccharides of hyaluronan in endothelial cells. Inhibition by non-angiogenic, high-molecular-weight hyaluronan. Int J Cancer. 1997;71:251–256. doi: 10.1002/(sici)1097-0215(19970410)71:2<251::aid-ijc21>3.0.co;2-j. [DOI] [PubMed] [Google Scholar]
  • 75.Deepa SS, Carulli D, Galtrey C, Rhodes K, Fukuda J, Mikami T, Sugahara K, Fawcett JW. Composition of perineuronal net extracellular matrix in rat brain: a different disaccharide composition for the net-associated proteoglycans. J Biol Chem. 2006;281:17789–17800. doi: 10.1074/jbc.M600544200. [DOI] [PubMed] [Google Scholar]
  • 76.DeGrendele HC, Estess P, Picker LJ, Siegelman MH. CD44 and its ligand hyaluronate mediate rolling under physiologic flow: a novel lymphocyte-endothelial cell primary adhesion pathway. J Exp Med. 1996;183:1119–1130. doi: 10.1084/jem.183.3.1119. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.DeGrendele HC, Estess P, Siegelman MH. Requirement for CD44 in activated T cell extravasation into an inflammatory site. Science. 1997;278:672–675. doi: 10.1126/science.278.5338.672. [DOI] [PubMed] [Google Scholar]
  • 78.DeGrendele HC, Kosfiszer M, Estess P, Siegelman MH. CD44 activation and associated primary adhesion is inducible via T cell receptor stimulation. J Immunol. 1997;159:2549–2553. [PubMed] [Google Scholar]
  • 79.del Fresno C, Otero K, Gomez-Garcia L, Gonzalez-Leon MC, Soler-Ranger L, Fuentes-Prior P, Escoll P, Baos R, Caveda L, Garcia F, Arnalich F, Lopez-Collazo E. Tumor cells deactivate human monocytes by up-regulating IL-1 receptor associated kinase-M expression via CD44 and TLR4. J Immunol. 2005;174:3032–3040. doi: 10.4049/jimmunol.174.5.3032. [DOI] [PubMed] [Google Scholar]
  • 80.Dentener MA, Vernooy JH, Hendriks S, Wouters EF. Enhanced levels of hyaluronan in lungs of patients with COPD: relationship with lung function and local inflammation. Thorax. 2005;60:114–119. doi: 10.1136/thx.2003.020842. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Do Y, Nagarkatti PS, Nagarkatti M. Role of CD44 and hyaluronic acid (HA) in activation of alloreactive and antigen-specific T cells by bone marrow-derived dendritic cells. J Immunother. 2004;27:1–12. doi: 10.1097/00002371-200401000-00001. [DOI] [PubMed] [Google Scholar]
  • 82.Duff B, Weigel JA, Bourne P, Weigel PH, McGary CT. Endothelium in hepatic cavernous hemangiomas does not express the hyaluronan receptor for endocytosis. Hum Pathol. 2002;33:265–269. doi: 10.1053/hupa.2002.32223. [DOI] [PubMed] [Google Scholar]
  • 83.Edward M, Fitzgerald L, Thind C, Leman J, Burden AD. Cutaneous mucinosis associated with dermatomyositis and nephrogenic fibrosing dermopathy: fibroblast hyaluronan synthesis and the effect of patient serum. Br J Dermatol. 2007;156:473–479. doi: 10.1111/j.1365-2133.2006.07652.x. [DOI] [PubMed] [Google Scholar]
  • 84.Eggli PS, Graber W. Cytochemical localization of hyaluronan in rat and human skin mast cell granules. J Invest Dermatol. 1993;100:121–125. doi: 10.1111/1523-1747.ep12462777. [DOI] [PubMed] [Google Scholar]
  • 85.Eklund A, Tornling G, Blaschke E, Curstedt T. Extracellular matrix components in bronchoalveolar lavage fluid in quartz exposed rats. Br J Ind Med. 1991;48:776–782. doi: 10.1136/oem.48.11.776. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Eklund A, van Hage-Hamsten M, Skold CM, Johansson SG. Elevated levels of tryptase in bronchoalveolar lavage fluid from patients with sarcoidosis. Sarcoidosis. 1993;10:12–17. [PubMed] [Google Scholar]
  • 87.Elias JA, Krol RC, Freundlich B, Sampson PM. Regulation of human lung fibroblast glycosaminoglycan production by recombinant interferons, tumor necrosis factor, and lymphotoxin. J Clin Invest. 1988;81:325–333. doi: 10.1172/JCI113324. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Emlen W, Niebur J, Flanders G, Rutledge J. Measurement of serum hyaluronic acid in patients with rheumatoid arthritis: correlation with disease activity. J Rheumatol. 1996;23:974–978. [PubMed] [Google Scholar]
  • 89.Engstrom-Laurent A, Hallgren R. Circulating hyaluronate in rheumatoid arthritis: relationship to inflammatory activity and the effect of corticosteroid therapy. Ann Rheum Dis. 1985;44:83–88. doi: 10.1136/ard.44.2.83. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Esnault S, Malter JS. Hyaluronic acid or TNF-alpha plus fibronectin triggers granulocyte macrophage-colony-stimulating factor mRNA stabilization in eosinophils yet engages differential intracellular pathways and mRNA binding proteins. J Immunol. 2003;171:6780–6787. doi: 10.4049/jimmunol.171.12.6780. [DOI] [PubMed] [Google Scholar]
  • 91.Estess P, Nandi A, Mohamadzadeh M, Siegelman MH. Interleukin 15 induces endothelial hyaluronan expression in vitro and promotes activated T cell extravasation through a CD44-dependent pathway in vivo. J Exp Med. 1999;190:9–19. doi: 10.1084/jem.190.1.9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Evanko SP, Angello JC, Wight TN. Formation of hyaluronan- and versican-rich pericellular matrix is required for proliferation and migration of vascular smooth muscle cells. Arterioscler Thromb Vasc Biol. 1999;19:1004–1013. doi: 10.1161/01.atv.19.4.1004. [DOI] [PubMed] [Google Scholar]
  • 93.Falkowski M, Schledzewski K, Hansen B, Goerdt S. Expression of stabilin-2, a novel fasciclin-like hyaluronan receptor protein, in murine sinusoidal endothelia, avascular tissues, and at solid/liquid interfaces. Histochem Cell Biol. 2003;120:361–369. doi: 10.1007/s00418-003-0585-5. [DOI] [PubMed] [Google Scholar]
  • 94.Fan H, Hu Y, Zhang C, Li X, Lv R, Qin L, Zhu R. Cartilage regeneration using mesenchymal stem cells and a PLGA-gelatin/chondroitin/hyaluronate hybrid scaffold. Biomaterials. 2006;27:4573–4580. doi: 10.1016/j.biomaterials.2006.04.013. [DOI] [PubMed] [Google Scholar]
  • 95.Fan J, Li Y, Vodovotz Y, Billiar TR, Wilson MA. Hemorrhagic shock-activated neutrophils augment TLR4 signaling-induced TLR2 upregulation in alveolar macrophages: role in hemorrhage-primed lung inflammation. Am J Physiol Lung Cell Mol Physiol. 2006;290:L738–L746. doi: 10.1152/ajplung.00280.2005. [DOI] [PubMed] [Google Scholar]
  • 96.Feusi E, Sun L, Sibalic A, Beck-Schimmer B, Oertli B, Wuthrich RP. Enhanced hyaluronan synthesis in the MRL-Fas(lpr) kidney: role of cytokines. Nephron. 1999;83:66–73. doi: 10.1159/000045475. [DOI] [PubMed] [Google Scholar]
  • 97.Fieber C, Baumann P, Vallon R, Termeer C, Simon JC, Hofmann M, Angel P, Herrlich P, Sleeman JP. Hyaluronan-oligosaccharide-induced transcription of metalloproteases. J Cell Sci. 2004;117:359–367. doi: 10.1242/jcs.00831. [DOI] [PubMed] [Google Scholar]
  • 98.Firan M, Dhillon S, Estess P, Siegelman MH. Suppressor activity and potency among regulatory T cells is discriminated by functionally active CD44. Blood. 2006;107:619–627. doi: 10.1182/blood-2005-06-2277. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Fischer JW, Schror K. Regulation of hyaluronan synthesis by vasodilatory prostaglandins. Implications for atherosclerosis. Thromb Haemost. 2007;98:287–295. [PubMed] [Google Scholar]
  • 100.Fiszer-Szafarz B, Czartoryska B, Tylki-Szymanska A. Serum hyaluronidase aberrations in metabolic and morphogenetic disorders. Glycoconj J. 2005;22:395–400. doi: 10.1007/s10719-005-1390-2. [DOI] [PubMed] [Google Scholar]
  • 101.Fletcher E, Jacobs JH, Markham RL. Viscosity studies on hyaluronic acid of synovial fluid in rheumatoid arthritis and osteoarthritis. Clin Sci (Lond) 1955;14:653–660. [PubMed] [Google Scholar]
  • 102.Fukui M, Whittlesey K, Metcalfe DD, Dastych J. Human mast cells express the hyaluronic-acid-binding isoform of CD44 and adhere to hyaluronic acid. Clin Immunol. 2000;94:173–178. doi: 10.1006/clim.1999.4830. [DOI] [PubMed] [Google Scholar]
  • 103.Fulop C, Kamath RV, Li Y, Otto JM, Salustri A, Olsen BR, Glant TT, Hascall VC. Coding sequence, exon-intron structure and chromosomal localization of murine TNF-stimulated gene 6 that is specifically expressed by expanding cumulus cell-oocyte complexes. Gene. 1997;202:95–102. doi: 10.1016/s0378-1119(97)00459-9. [DOI] [PubMed] [Google Scholar]
  • 104.Fulop C, Szanto S, Mukhopadhyay D, Bardos T, Kamath RV, Rugg MS, Day AJ, Salustri A, Hascall VC, Glant TT, Mikecz K. Impaired cumulus mucification and female sterility in tumor necrosis factor-induced protein-6 deficient mice. Development. 2003;130:2253–2261. doi: 10.1242/dev.00422. [DOI] [PubMed] [Google Scholar]
  • 105.Gale NW, Prevo R, Espinosa J, Ferguson DJ, Dominguez MG, Yancopoulos GD, Thurston G, Jackson DG. Normal lymphatic development and function in mice deficient for the lymphatic hyaluronan receptor LYVE-1. Mol Cell Biol. 2007;27:595–604. doi: 10.1128/MCB.01503-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Gao F, Koenitzer JR, Tobolewski JM, Jiang D, Liang J, Noble PW, Oury TD. Extracellular superoxide dismutase inhibits inflammation by preventing oxidative fragmentation of hyaluronan. J Biol Chem. 2008;283:6058–6066. doi: 10.1074/jbc.M709273200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Garantziotis S, Zudaire E, Trempus CS, Hollingsworth JW, Jiang D, Lancaster LH, Richardson E, Zhuo L, Cuttitta F, Brown KK, Noble PW, Kimata K, Schwartz DA. Serum inter-alpha-trypsin inhibitor and matrix hyaluronan promote angiogenesis in fibrotic lung injury. Am J Respir Crit Care Med. 2008;178:939–947. doi: 10.1164/rccm.200803-386OC. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Gary SC, Zerillo CA, Chiang VL, Gaw JU, Gray G, Hockfield S. cDNA cloning, chromosomal localization, and expression analysis of human BEHAB/brevican, a brain specific proteoglycan regulated during cortical development and in glioma. Gene. 2000;256:139–147. doi: 10.1016/s0378-1119(00)00362-0. [DOI] [PubMed] [Google Scholar]
  • 109.George J, Stern R. Serum hyaluronan and hyaluronidase: very early markers of toxic liver injury. Clin Chim Acta. 2004;348:189–197. doi: 10.1016/j.cccn.2004.05.018. [DOI] [PubMed] [Google Scholar]
  • 110.Getting SJ, Mahoney DJ, Cao T, Rugg MS, Fries E, Milner CM, Perretti M, Day AJ. The link module from human TSG-6 inhibits neutrophil migration in a hyaluronan- and inter-alpha -inhibitor-independent manner. J Biol Chem. 2002;277:51068–51076. doi: 10.1074/jbc.M205121200. [DOI] [PubMed] [Google Scholar]
  • 111.Ghebrehiwet B, Lim BL, Peerschke EI, Willis AC, Reid KB. Isolation, cDNA cloning, and overexpression of a 33-kD cell surface glycoprotein that binds to the globular “heads” of C1q. J Exp Med. 1994;179:1809–1821. doi: 10.1084/jem.179.6.1809. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Ghosh I, Bharadwaj A, Datta K. Reduction in the level of hyaluronan binding protein 1 (HABP1) is associated with loss of sperm motility. J Reprod Immunol. 2002;53:45–54. doi: 10.1016/s0165-0378(01)00095-x. [DOI] [PubMed] [Google Scholar]
  • 113.Goetinck PF, Stirpe NS, Tsonis PA, Carlone D. The tandemly repeated sequences of cartilage link protein contain the sites for interaction with hyaluronic acid. J Cell Biol. 1987;105:2403–2408. doi: 10.1083/jcb.105.5.2403. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Goldberg RL, Huff JP, Lenz ME, Glickman P, Katz R, Thonar EJ. Elevated plasma levels of hyaluronate in patients with osteoarthritis and rheumatoid arthritis. Arthritis Rheum. 1991;34:799–807. doi: 10.1002/art.1780340704. [DOI] [PubMed] [Google Scholar]
  • 115.Goldberg RL, Rubin AS. Serum hyaluronate as a marker for disease severity in the Lactobacillus casei model of arthritis in the rat. J Rheumatol. 1989;16:92–96. [PubMed] [Google Scholar]
  • 116.Goto M, Hanyu T, Yoshio T, Matsuno H, Shimizu M, Murata N, Shiozawa S, Matsubara T, Yamana S, Matsuda T. Intra-articular injection of hyaluronate (SI-6601D) improves joint pain and synovial fluid prostaglandin E2 levels in rheumatoid arthritis: a multicenter clinical trial. Clin Exp Rheumatol. 2001;19:377–383. [PubMed] [Google Scholar]
  • 117.Goueffic Y, Guilluy C, Guerin P, Patra P, Pacaud P, Loirand G. Hyaluronan induces vascular smooth muscle cell migration through RHAMM-mediated PI3K-dependent Rac activation. Cardiovasc Res. 2006;72:339–348. doi: 10.1016/j.cardiores.2006.07.017. [DOI] [PubMed] [Google Scholar]
  • 118.Grootveld M, Henderson EB, Farrell A, Blake DR, Parkes HG, Haycock P. Oxidative damage to hyaluronate and glucose in synovial fluid during exercise of the inflamed rheumatoid joint. Detection of abnormal low-molecular-mass metabolites by proton-n.m.r. spectroscopy. Biochem J. 1991;273(Pt 2):459–467. doi: 10.1042/bj2730459. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Guechot J, Loria A, Serfaty L, Giral P, Giboudeau J, Poupon R. Serum hyaluronan as a marker of liver fibrosis in chronic viral hepatitis C: effect of alpha-interferon therapy. J Hepatol. 1995;22:22–26. doi: 10.1016/0168-8278(95)80255-x. [DOI] [PubMed] [Google Scholar]
  • 120.Guechot J, Poupon RE, Giral P, Balkau B, Giboudeau J, Poupon R. Relationship between procollagen III aminoterminal propeptide and hyaluronan serum levels and histological fibrosis in primary biliary cirrhosis and chronic viral hepatitis C. J Hepatol. 1994;20:388–393. doi: 10.1016/s0168-8278(94)80013-8. [DOI] [PubMed] [Google Scholar]
  • 121.Guechot J, Serfaty L, Bonnand AM, Chazouilleres O, Poupon RE, Poupon R. Prognostic value of serum hyaluronan in patients with compensated HCV cirrhosis. J Hepatol. 2000;32:447–452. doi: 10.1016/s0168-8278(00)80396-7. [DOI] [PubMed] [Google Scholar]
  • 122.Hall CL, Yang B, Yang X, Zhang S, Turley M, Samuel S, Lange LA, Wang C, Curpen GD, Savani RC, Greenberg AH, Turley EA. Overexpression of the hyaluronan receptor RHAMM is transforming and is also required for H-ras transformation. Cell. 1995;82:19–26. doi: 10.1016/0092-8674(95)90048-9. [DOI] [PubMed] [Google Scholar]
  • 123.Hallgren R, Engstrom-Laurent A, Nisbeth U. Circulating hyaluronate. A potential marker of altered metabolism of the connective tissue in uremia. Nephron. 1987;46:150–154. doi: 10.1159/000184331. [DOI] [PubMed] [Google Scholar]
  • 124.Hallgren R, Gerdin B, Tengblad A, Tufveson G. Accumulation of hyaluronan (hyaluronic acid) in myocardial interstitial tissue parallels development of transplantation edema in heart allografts in rats. J Clin Invest. 1990;85:668–673. doi: 10.1172/JCI114490. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Hallgren R, Samuelsson T, Laurent TC, Modig J. Accumulation of hyaluronan (hyaluronic acid) in the lung in adult respiratory distress syndrome. Am Rev Respir Dis. 1989;139:682–687. doi: 10.1164/ajrccm/139.3.682. [DOI] [PubMed] [Google Scholar]
  • 126.Hamann KJ, Dowling TL, Neeley SP, Grant JA, Leff AR. Hyaluronic acid enhances cell proliferation during eosinopoiesis through the CD44 surface antigen. J Immunol. 1995;154:4073–4080. [PubMed] [Google Scholar]
  • 127.Hamerman D, Schuster H. Hyaluronate in normal human synovial fluid. J Clin Invest. 1958;37:57–64. doi: 10.1172/JCI103585. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Han CY, Subramanian S, Chan CK, Omer M, Chiba T, Wight TN, Chait A. Adipocyte-derived serum amyloid A3 and hyaluronan play a role in monocyte recruitment and adhesion. Diabetes. 2007;56:2260–2273. doi: 10.2337/db07-0218. [DOI] [PubMed] [Google Scholar]
  • 129.Hardwick C, Hoare K, Owens R, Hohn HP, Hook M, Moore D, Cripps V, Austen L, Nance DM, Turley EA. Molecular cloning of a novel hyaluronan receptor that mediates tumor cell motility. J Cell Biol. 1992;117:1343–1350. doi: 10.1083/jcb.117.6.1343. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Harris EN, Kyosseva SV, Weigel JA, Weigel PH. Expression, processing, and glycosaminoglycan binding activity of the recombinant human 315-kDa hyaluronic acid receptor for endocytosis (HARE) J Biol Chem. 2007;282:2785–2797. doi: 10.1074/jbc.M607787200. [DOI] [PubMed] [Google Scholar]
  • 131.Harris EN, Weigel JA, Weigel PH. Endocytic function, glycosaminoglycan specificity, and antibody sensitivity of the recombinant human 190-kDa hyaluronan receptor for endocytosis (HARE) J Biol Chem. 2004;279:36201–36209. doi: 10.1074/jbc.M405322200. [DOI] [PubMed] [Google Scholar]
  • 132.Hart SP, Dougherty GJ, Haslett C, Dransfield I. CD44 regulates phagocytosis of apoptotic neutrophil granulocytes, but not apoptotic lymphocytes, by human macrophages. J Immunol. 1997;159:919–925. [PubMed] [Google Scholar]
  • 133.Haylock DN, Nilsson SK. The role of hyaluronic acid in hemopoietic stem cell biology. Regen Med. 2006;1:437–445. doi: 10.2217/17460751.1.4.437. [DOI] [PubMed] [Google Scholar]
  • 134.Haynes BF, Hale LP, Patton KL, Martin ME, McCallum RM. Measurement of an adhesion molecule as an indicator of inflammatory disease activity. Up-regulation of the receptor for hyaluronate (CD44) in rheumatoid arthritis. Arthritis Rheum. 1991;34:1434–1443. doi: 10.1002/art.1780341115. [DOI] [PubMed] [Google Scholar]
  • 135.Hegewald AA, Ringe J, Bartel J, Kruger I, Notter M, Barnewitz D, Kaps C, Sittinger M. Hyaluronic acid and autologous synovial fluid induce chondrogenic differentiation of equine mesenchymal stem cells: a preliminary study. Tissue Cell. 2004;36:431–438. doi: 10.1016/j.tice.2004.07.003. [DOI] [PubMed] [Google Scholar]
  • 136.Heldermon C, DeAngelis PL, Weigel PH. Topological organization of the hyaluronan synthase from Streptococcus pyogenes. J Biol Chem. 2001;276:2037–2046. doi: 10.1074/jbc.M002276200. [DOI] [PubMed] [Google Scholar]
  • 137.Heldin P, Asplund T, Ytterberg D, Thelin S, Laurent TC. Characterization of the molecular mechanism involved in the activation of hyaluronan synthetase by platelet-derived growth factor in human mesothelial cells. Biochem J. 1992;283 (Pt 1):165–170. doi: 10.1042/bj2830165. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Hemming R, Martin DC, Slominski E, Nagy JI, Halayko AJ, Pind S, Triggs-Raine B. Mouse Hyal3 encodes a 45- to 56-kDa glycoprotein whose overexpression increases hyaluronidase 1 activity in cultured cells. Glycobiology. 2008;18:280–289. doi: 10.1093/glycob/cwn006. [DOI] [PubMed] [Google Scholar]
  • 139.Henke CA, Roongta U, Mickelson DJ, Knutson JR, McCarthy JB. CD44-related chondroitin sulfate proteoglycan, a cell surface receptor implicated with tumor cell invasion, mediates endothelial cell migration on fibrinogen and invasion into a fibrin matrix. J Clin Invest. 1996;97:2541–2552. doi: 10.1172/JCI118702. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Henriksen JH, Bentsen KD, Laurent TC. Splanchnic and renal extraction of circulating hyaluronan in patients with alcoholic liver disease. J Hepatol. 1988;6:158–166. doi: 10.1016/s0168-8278(88)80027-8. [DOI] [PubMed] [Google Scholar]
  • 141.Hernnas J, Nettelbladt O, Bjermer L, Sarnstrand B, Malmstrom A, Hallgren R. Alveolar accumulation of fibronectin and hyaluronan precedes bleomycin-induced pulmonary fibrosis in the rat. Eur Respir J. 1992;5:404–410. [PubMed] [Google Scholar]
  • 142.Herrera MB, Bussolati B, Bruno S, Morando L, Mauriello-Romanazzi G, Sanavio F, Stamenkovic I, Biancone L, Camussi G. Exogenous mesenchymal stem cells localize to the kidney by means of CD44 following acute tubular injury. Kidney Int. 2007;72:430–441. doi: 10.1038/sj.ki.5002334. [DOI] [PubMed] [Google Scholar]
  • 143.Hershon LE. Elaboration of hyaluronidase and chondroitin sulfatase by microorganisms inhabiting the gingival sulcus: evaluation of a screening method for periodontal disease. J Periodontol. 1971;42:34–36. doi: 10.1902/jop.1971.42.1.34. [DOI] [PubMed] [Google Scholar]
  • 144.Hodge-Dufour J, Noble PW, Horton MR, Bao C, Wysoka M, Burdick MD, Strieter RM, Trinchieri G, Pure E. Induction of IL-12 and chemokines by hyaluronan requires adhesion-dependent priming of resident but not elicited macrophages. J Immunol. 1997;159:2492–2500. [PubMed] [Google Scholar]
  • 145.Hofmann M, Assmann V, Fieber C, Sleeman JP, Moll J, Ponta H, Hart IR, Herrlich P. Problems with RHAMM: a new link between surface adhesion and oncogenesis? Cell. 1998;95:591–592. doi: 10.1016/s0092-8674(00)81628-1. author reply 592–593. [DOI] [PubMed] [Google Scholar]
  • 146.Honkanen E, Froseth B, Gronhagen-Riska C. Serum hyaluronic acid and procollagen III amino terminal propeptide in chronic renal failure. Am J Nephrol. 1991;11:201–206. doi: 10.1159/000168304. [DOI] [PubMed] [Google Scholar]
  • 147.Honore B, Madsen P, Rasmussen HH, Vandekerckhove J, Celis JE, Leffers H. Cloning and expression of a cDNA covering the complete coding region of the P32 subunit of human pre-mRNA splicing factor SF2. Gene. 1993;134:283–287. doi: 10.1016/0378-1119(93)90108-f. [DOI] [PubMed] [Google Scholar]
  • 148.Horton MR, Burdick MD, Strieter RM, Bao C, Noble PW. Regulation of hyaluronan-induced chemokine gene expression by IL-10 and IFN-gamma in mouse macrophages. J Immunol. 1998;160:3023–3030. [PubMed] [Google Scholar]
  • 149.Horton MR, McKee CM, Bao C, Liao F, Farber JM, Hodge-DuFour J, Pure E, Oliver BL, Wright TM, Noble PW. Hyaluronan fragments synergize with interferon-gamma to induce the C-X-C chemokines mig and interferon-inducible protein-10 in mouse macrophages. J Biol Chem. 1998;273:35088–35094. doi: 10.1074/jbc.273.52.35088. [DOI] [PubMed] [Google Scholar]
  • 150.Horton MR, Olman MA, Bao C, White KE, Choi AM, Chin BY, Noble PW, Lowenstein CJ. Regulation of plasminogen activator inhibitor-1 and urokinase by hyaluronan fragments in mouse macrophages. Am J Physiol Lung Cell Mol Physiol. 2000;279:L707–715. doi: 10.1152/ajplung.2000.279.4.L707. [DOI] [PubMed] [Google Scholar]
  • 151.Horton MR, Olman MA, Noble PW. Hyaluronan fragments induce plasminogen activator inhibitor-1 and inhibit urokinase activity in mouse alveolar macrophages: a potential mechanism for impaired fibrinolytic activity in acute lung injury. Chest. 1999;116:17S. [PubMed] [Google Scholar]
  • 152.Horton MR, Shapiro S, Bao C, Lowenstein CJ, Noble PW. Induction and regulation of macrophage metalloelastase by hyaluronan fragments in mouse macrophages. J Immunol. 1999;162:4171–4176. [PubMed] [Google Scholar]
  • 153.Hou S, Xu Q, Tian W, Cui F, Cai Q, Ma J, Lee IS. The repair of brain lesion by implantation of hyaluronic acid hydrogels modified with laminin. J Neurosci Methods. 2005;148:60–70. doi: 10.1016/j.jneumeth.2005.04.016. [DOI] [PubMed] [Google Scholar]
  • 154.Huang L, Yoneda M, Kimata K. A serum-derived hyaluronan-associated protein (SHAP) is the heavy chain of the inter alpha-trypsin inhibitor. J Biol Chem. 1993;268:26725–26730. [PubMed] [Google Scholar]
  • 155.Huang SS, Liu IH, Smith T, Shah MR, Johnson FE, Huang JS. CRSBP-1/LYVE-l-null mice exhibit identifiable morphological and functional alterations of lymphatic capillary vessels. FEBS Lett. 2006;580:6259–6268. doi: 10.1016/j.febslet.2006.10.028. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Huebener P, Abou-Khamis T, Zymek P, Bujak M, Ying X, Chatila K, Haudek S, Thakker G, Frangogiannis NG. CD44 Is Critically Involved in Infarct Healing by Regulating the Inflammatory and Fibrotic Response. J Immunol. 2008;180:2625–2633. doi: 10.4049/jimmunol.180.4.2625. [DOI] [PubMed] [Google Scholar]
  • 157.Itano N, Atsumi F, Sawai T, Yamada Y, Miyaishi O, Senga T, Hamaguchi M, Kimata K. Abnormal accumulation of hyaluronan matrix diminishes contact inhibition of cell growth and promotes cell migration. Proc Natl Acad Sci U S A. 2002;99:3609–3614. doi: 10.1073/pnas.052026799. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Itano N, Kimata K. Expression cloning and molecular characterization of HAS protein, a eukaryotic hyaluronan synthase. J Biol Chem. 1996;271:9875–9878. doi: 10.1074/jbc.271.17.9875. [DOI] [PubMed] [Google Scholar]
  • 159.Itano N, Kimata K. Mammalian hyaluronan synthases. IUBMB Life. 2002;54:195–199. doi: 10.1080/15216540214929. [DOI] [PubMed] [Google Scholar]
  • 160.Itano N, Sawai T, Yoshida M, Lenas P, Yamada Y, Imagawa M, Shinomura T, Hamaguchi M, Yoshida Y, Ohnuki Y, Miyauchi S, Spicer AP, McDonald JA, Kimata K. Three isoforms of mammalian hyaluronan synthases have distinct enzymatic properties. J Biol Chem. 1999;274:25085–25092. doi: 10.1074/jbc.274.35.25085. [DOI] [PubMed] [Google Scholar]
  • 161.Iwasaki S, Minamisawa S, Yokoyama U, Akaike T, Quan H, Nagashima Y, Nishimaki S, Ishikawa Y, Yokota S. Interleukin-15 inhibits smooth muscle cell proliferation and hyaluronan production in rat ductus arteriosus. Pediatr Res. 2007;62:392–398. doi: 10.1203/PDR.0b013e31813c9339. [DOI] [PubMed] [Google Scholar]
  • 162.Jackson DG. Biology of the lymphatic marker LYVE-1 and applications in research into lymphatic trafficking and lymphangiogenesis. Apmis. 2004;112:526–538. doi: 10.1111/j.1600-0463.2004.apm11207-0811.x. [DOI] [PubMed] [Google Scholar]
  • 163.Jackson DG, Prevo R, Clasper S, Banerji S. LYVE-1, the lymphatic system and tumor lymphangiogenesis. Trends Immunol. 2001;22:317–321. doi: 10.1016/s1471-4906(01)01936-6. [DOI] [PubMed] [Google Scholar]
  • 164.Jacobson A, Brinck J, Briskin MJ, Spicer AP, Heldin P. Expression of human hyaluronan synthases in response to external stimuli. Biochem J. 2000;348(Pt 1):29–35. [PMC free article] [PubMed] [Google Scholar]
  • 165.Jameson JM, Cauvi G, Sharp LL, Witherden DA, Havran WL. {gamma}{delta} T cell-induced hyaluronan production by epithelial cells regulates inflammation. J Exp Med. 2005;201:1269–1279. doi: 10.1084/jem.20042057. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 166.Jaworski DM, Kelly GM, Hockfield S. Intracranial injury acutely induces the expression of the secreted isoform of the CNS-specific hyaluronan-binding protein BEHAB/brevican. Exp Neurol. 1999;157:327–337. doi: 10.1006/exnr.1999.7062. [DOI] [PubMed] [Google Scholar]
  • 167.Jaworski DM, Kelly GM, Piepmeier JM, Hockfield S. BEHAB (brain enriched hyaluronan binding) is expressed in surgical samples of glioma and in intracranial grafts of invasive glioma cell lines. Cancer Res. 1996;56:2293–2298. [PubMed] [Google Scholar]
  • 168.Jederstrom G, Andersson A, Grasjo J, Sjoholm I. Formulating insulin for oral administration: preparation of hyaluronan-insulin complex. Pharm Res. 2004;21:2040–2047. doi: 10.1023/b:pham.0000048195.69304.ff. [DOI] [PubMed] [Google Scholar]
  • 169.Jederstrom G, Grasjo, Nordin A, Sjoholm I, Andersson A. Blood glucose-lowering activity of a hyaluronan-insulin complex after oral administration to rats with diabetes. Diabetes Technol Ther. 2005;7:948–957. doi: 10.1089/dia.2005.7.948. [DOI] [PubMed] [Google Scholar]
  • 170.Jenkins RH, Thomas GJ, Williams JD, Steadman R. Myofibroblastic differentiation leads to hyaluronan accumulation through reduced hyaluronan turnover. J Biol Chem. 2004;279:41453–41460. doi: 10.1074/jbc.M401678200. [DOI] [PubMed] [Google Scholar]
  • 171.Jha BK, Mitra N, Rana R, Surolia A, Salunke DM, Datta K. pH and cation-induced thermodynamic stability of human hyaluronan binding protein 1 regulates its hyaluronan affinity. J Biol Chem. 2004;279:23061–23072. doi: 10.1074/jbc.M310676200. [DOI] [PubMed] [Google Scholar]
  • 172.Jha BK, Salunke DM, Datta K. Disulfide bond formation through Cys186 facilitates functionally relevant dimerization of trimeric hyaluronan-binding protein 1 (HABP1)/p32/gC1qR. Eur J Biochem. 2002;269:298–306. doi: 10.1046/j.0014-2956.2001.02654.x. [DOI] [PubMed] [Google Scholar]
  • 173.Jha BK, Salunke DM, Datta K. Structural flexibility of multifunctional HABP1 may be important for regulating its binding to different ligands. J Biol Chem. 2003;278:27464–27472. doi: 10.1074/jbc.M206696200. [DOI] [PubMed] [Google Scholar]
  • 174.Ji RC, Kurihara K, Kato S. Lymphatic vascular endothelial hyaluronan receptor (LYVE)-1- and CCL21-positive lymphatic compartments in the diabetic thymus. Anat Sci Int. 2006;81:201–209. doi: 10.1111/j.1447-073X.2006.00145.x. [DOI] [PubMed] [Google Scholar]
  • 175.Jiang D, Liang J, Fan J, Yu S, Chen S, Luo Y, Prestwich GD, Mascarenhas MM, Garg HG, Quinn DA, Homer RJ, Goldstein DR, Bucala R, Lee PJ, Medzhitov R, Noble PW. Regulation of lung injury and repair by Toll-like receptors and hyaluronan. Nat Med. 2005;11:1173–1179. doi: 10.1038/nm1315. [DOI] [PubMed] [Google Scholar]
  • 176.Jiang D, Liang J, Li Y, Noble PW. The role of Toll-like receptors in non-infectious lung injury. Cell Res. 2006;16:693–701. doi: 10.1038/sj.cr.7310085. [DOI] [PubMed] [Google Scholar]
  • 177.Jiang D, Liang J, Noble PW. Hyaluronan in tissue injury and repair. Annu Rev Cell Dev Biol. 2007;23:435–461. doi: 10.1146/annurev.cellbio.23.090506.123337. [DOI] [PubMed] [Google Scholar]
  • 178.Johnson LA, Clasper S, Holt AP, Lalor PF, Baban D, Jackson DG. An inflammation-induced mechanism for leukocyte transmigration across lymphatic vessel endothelium. J Exp Med. 2006;203:2763–2777. doi: 10.1084/jem.20051759. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 179.Johnson LA, Prevo R, Clasper S, Jackson DG. Inflammation-induced uptake and degradation of the lymphatic endothelial hyaluronan receptor LYVE-1. J Biol Chem. 2007;282:33671–33680. doi: 10.1074/jbc.M702889200. [DOI] [PubMed] [Google Scholar]
  • 180.Jones ES. Hyaluronidase activity in the skin, rheumatic disease, and salicylates. Ann Rheum Dis. 1950;9:137–148. doi: 10.1136/ard.9.2.137. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 181.Jones LL, Margolis RU, Tuszynski MH. The chondroitin sulfate proteoglycans neurocan, brevican, phosphacan, and versican are differentially regulated following spinal cord injury. Exp Neurol. 2003;182:399–411. doi: 10.1016/s0014-4886(03)00087-6. [DOI] [PubMed] [Google Scholar]
  • 182.Juhlin L. Hyaluronan in skin. J Intern Med. 1997;242:61–66. doi: 10.1046/j.1365-2796.1997.00175.x. [DOI] [PubMed] [Google Scholar]
  • 183.Jun Z, Hill PA, Lan HY, Foti R, Mu W, Atkins RC, Nikolic-Paterson DJ. CD44 and hyaluronan expression in the development of experimental crescentic glomerulonephritis. Clin Exp Immunol. 1997;108:69–77. doi: 10.1046/j.1365-2249.1997.d01-977.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 184.Juul SE, Kinsella MG, Jackson JC, Truog WE, Standaert TA, Hodson WA. Changes in hyaluronan deposition during early respiratory distress syndrome in premature monkeys. Pediatr Res. 1994;35:238–243. doi: 10.1203/00006450-199402000-00024. [DOI] [PubMed] [Google Scholar]
  • 185.Juul SE, Krueger RC, Jr, Scofield L, Hershenson MB, Schwartz NB. Hyperoxia alone causes changes in lung proteoglycans and hyaluronan in neonatal rat pups. Am J Respir Cell Mol Biol. 1995;13:629–638. doi: 10.1165/ajrcmb.13.6.7576700. [DOI] [PubMed] [Google Scholar]
  • 186.Kallestrup EB, Jorgensen SS, Nordling J, Hald T. Treatment of interstitial cystitis with Cystistat: a hyaluronic acid product. Scand J Urol Nephrol. 2005;39:143–147. doi: 10.1080/00365590410015876-1. [DOI] [PubMed] [Google Scholar]
  • 187.Kamal A, Datta K. Upregulation of hyaluronan binding protein 1 (HABP1/p32/gC1qR) is associated with Cisplatin induced apoptosis. Apoptosis. 2006;11:861–874. doi: 10.1007/s10495-006-5396-4. [DOI] [PubMed] [Google Scholar]
  • 188.Karvinen S, Pasonen-Seppanen S, Hyttinen JM, Pienimaki JP, Torronen K, Jokela TA, Tammi MI, Tammi R. Keratinocyte growth factor stimulates migration and hyaluronan synthesis in the epidermis by activation of keratinocyte hyaluronan synthases 2 and 3. J Biol Chem. 2003;278:49495–49504. doi: 10.1074/jbc.M310445200. [DOI] [PubMed] [Google Scholar]
  • 189.Katoh S, Matsumoto N, Kawakita K, Tominaga A, Kincade PW, Matsukura S. A role for CD44 in an antigen-induced murine model of pulmonary eosinophilia. J Clin Invest. 2003;111:1563–1570. doi: 10.1172/JCI16583. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 190.Katoh S, Taniguchi H, Matsubara Y, Matsumoto N, Fukushima K, Kadota J, Matsukura S, Kohno S. Overexpression of CD44 on alveolar eosinophils with high concentrations of soluble CD44 in bronchoalveolar lavage fluid in patients with eosinophilic pneumonia. Allergy. 1999;54:1286–1292. doi: 10.1034/j.1398-9995.1999.00277.x. [DOI] [PubMed] [Google Scholar]
  • 191.Kavalkovich KW, Boynton RE, Murphy JM, Barry F. Chondrogenic differentiation of human mesenchymal stem cells within an alginate layer culture system. In Vitro Cell Dev Biol Anim. 2002;38:457–466. doi: 10.1290/1071-2690(2002)038<0457:cdohms>2.0.co;2. [DOI] [PubMed] [Google Scholar]
  • 192.Kaya G, Rodriguez I, Jorcano JL, Vassalli P, Stamenkovic I. Selective suppression of CD44 in keratinocytes of mice bearing an antisense CD44 transgene driven by a tissue-specific promoter disrupts hyaluronate metabolism in the skin and impairs keratinocyte proliferation. Genes Dev. 1997;11:996–1007. doi: 10.1101/gad.11.8.996. [DOI] [PubMed] [Google Scholar]
  • 193.Khaldoyanidi S, Moll J, Karakhanova S, Herrlich P, Ponta H. Hyaluronate-enhanced hematopoiesis: two different receptors trigger the release of interleukin-1beta and interleukin-6 from bone marrow macrophages. Blood. 1999;94:940–949. [PubMed] [Google Scholar]
  • 194.Khaldoyanidi S, Sikora L, Orlovskaya I, Matrosova V, Kozlov V, Sriramarao P. Correlation between nicotine-induced inhibition of hematopoiesis and decreased CD44 expression on bone marrow stromal cells. Blood. 2001;98:303–312. doi: 10.1182/blood.v98.2.303. [DOI] [PubMed] [Google Scholar]
  • 195.Khan AI, Kerfoot SM, Heit B, Liu L, Andonegui G, Ruffell B, Johnson P, Kubes P. Role of CD44 and hyaluronan in neutrophil recruitment. J Immunol. 2004;173:7594–7601. doi: 10.4049/jimmunol.173.12.7594. [DOI] [PubMed] [Google Scholar]
  • 196.Kida D, Yoneda M, Miyaura S, Ishimaru T, Yoshida Y, Ito T, Ishiguro N, Iwata H, Kimata K. The SHAP-HA complex in sera from patients with rheumatoid arthritis and osteoarthritis. J Rheumatol. 1999;26:1230–1238. [PubMed] [Google Scholar]
  • 197.Kiechl S, Lorenz E, Reindl M, Wiedermann CJ, Oberhollenzer F, Bonora E, Willeit J, Schwartz DA. Toll-like receptor 4 polymorphisms and atherogenesis. N Engl J Med. 2002;347:185–192. doi: 10.1056/NEJMoa012673. [DOI] [PubMed] [Google Scholar]
  • 198.Kim E, Baba D, Kimura M, Yamashita M, Kashiwabara S, Baba T. Identification of a hyaluronidase, Hyal5, involved in penetration of mouse sperm through cumulus mass. Proc Natl Acad Sci U S A. 2005;102:18028–18033. doi: 10.1073/pnas.0506825102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 199.Kim HM, Park BS, Kim JI, Kim SE, Lee J, Oh SC, Enkhbayar P, Matsushima N, Lee H, Yoo OJ, Lee JO. Crystal structure of the TLR4-MD-2 complex with bound endotoxin antagonist Eritoran. Cell. 2007;130:906–917. doi: 10.1016/j.cell.2007.08.002. [DOI] [PubMed] [Google Scholar]
  • 200.Knudson CB, Knudson W. Hyaluronan and CD44: modulators of chondrocyte metabolism. Clin Orthop Relat Res. 2004:S152–162. [PubMed] [Google Scholar]
  • 201.Knudson W, Chow G, Knudson CB. CD44-mediated uptake and degradation of hyaluronan. Matrix Biol. 2002;21:15–23. doi: 10.1016/s0945-053x(01)00186-x. [DOI] [PubMed] [Google Scholar]
  • 202.Kobayashi K, Hernandez LD, Galan JE, Janeway CA, Jr, Medzhitov R, Flavell RA. IRAK-M is a negative regulator of Toll-like receptor signaling. Cell. 2002;110:191–202. doi: 10.1016/s0092-8674(02)00827-9. [DOI] [PubMed] [Google Scholar]
  • 203.Kofoed JA, Barcelo AC. The synovial fluid hyaluronic acid in rheumatoid arthritis. Experientia. 1978;34:1545–1546. doi: 10.1007/BF02034662. [DOI] [PubMed] [Google Scholar]
  • 204.Kohda D, Morton CJ, Parkar AA, Hatanaka H, Inagaki FM, Campbell ID, Day AJ. Solution structure of the link module: a hyaluronan-binding domain involved in extracellular matrix stability and cell migration. Cell. 1996;86:767–775. doi: 10.1016/s0092-8674(00)80151-8. [DOI] [PubMed] [Google Scholar]
  • 205.Kopp S, Akerman S, Nilner M. Short-term effects of intra-articular sodium hyaluronate, glucocorticoid, and saline injections on rheumatoid arthritis of the temporomandibular joint. J Craniomandib Disord. 1991;5:231–238. [PubMed] [Google Scholar]
  • 206.Koprunner M, Mullegger J, Lepperdinger G. Synthesis of hyaluronan of distinctly different chain length is regulated by differential expression of Xhas1 and 2 during early development of Xenopus laevis. Mech Dev. 2000;90:275–278. doi: 10.1016/s0925-4773(99)00238-5. [DOI] [PubMed] [Google Scholar]
  • 207.Korner T, Kropf J, Kosche B, Kristahl H, Jaspersen D, Gressner AM. Improvement of prognostic power of the Child-Pugh classification of liver cirrhosis by hyaluronan. J Hepatol. 2003;39:947–953. doi: 10.1016/s0168-8278(03)00431-8. [DOI] [PubMed] [Google Scholar]
  • 208.Krainer AR, Mayeda A, Kozak D, Binns G. Functional expression of cloned human splicing factor SF2: homology to RNA-binding proteins, U1 70K, and Drosophila splicing regulators. Cell. 1991;66:383–394. doi: 10.1016/0092-8674(91)90627-b. [DOI] [PubMed] [Google Scholar]
  • 209.Kredel M, Muellenbach RM, Brock RW, Wilckens HH, Brederlau J, Roewer N, Wunder C. Liver dysfunction after lung recruitment manoeuvres during pressure-controlled ventilation in experimental acute respiratory distress. Crit Care. 2007;11:R13. doi: 10.1186/cc5674. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 210.Kreil G. Hyaluronidases--a group of neglected enzymes. Protein Sci. 1995;4:1666–1669. doi: 10.1002/pro.5560040902. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 211.Kubo M, Kikuchi K, Yazawa N, Fujimoto M, Tamaki T, Tamaki K. Increased serum concentration of hyaluronate in dermatomyositis patients. Arch Dermatol Res. 1998;290:579–581. doi: 10.1007/s004030050355. [DOI] [PubMed] [Google Scholar]
  • 212.Kultti A, Rilla K, Tiihonen R, Spicer AP, Tammi RH, Tammi MI. Hyaluronan synthesis induces microvillus-like cell surface protrusions. J Biol Chem. 2006;281:15821–15828. doi: 10.1074/jbc.M512840200. [DOI] [PubMed] [Google Scholar]
  • 213.Kumar R, Choudhury NR, Salunke DM, Datta K. Evidence for clustered mannose as a new ligand for hyaluronan- binding protein (HABP1) from human fibroblasts. J Biosci. 2001;26:325–332. doi: 10.1007/BF02703741. [DOI] [PubMed] [Google Scholar]
  • 214.Kunz LI, van Rensen EL, Sterk PJ. Inhaled hyaluronic acid against exercise-induced bronchoconstriction in asthma. Pulm Pharmacol Ther. 2006;19:286–291. doi: 10.1016/j.pupt.2005.04.011. [DOI] [PubMed] [Google Scholar]
  • 215.Kuppusamy UR, Das NP. Protective effects of tannic acid and related natural compounds on Crotalus adamenteus subcutaneous poisoning in mice. Pharmacol Toxicol. 1993;72:290–295. doi: 10.1111/j.1600-0773.1993.tb01652.x. [DOI] [PubMed] [Google Scholar]
  • 216.Kuppusamy UR, Khoo HE, Das NP. Structure-activity studies of flavonoids as inhibitors of hyaluronidase. Biochem Pharmacol. 1990;40:397–401. doi: 10.1016/0006-2952(90)90709-t. [DOI] [PubMed] [Google Scholar]
  • 217.Lai KN, Szeto CC, Lam CW, Lai KB, Wong TY, Leung JC. Increased ascitic level of hyaluronan in liver cirrhosis. J Lab Clin Med. 1998;131:354–359. doi: 10.1016/s0022-2143(98)90186-x. [DOI] [PubMed] [Google Scholar]
  • 218.Land W. Innate alloimmunity: history and current knowledge. Exp Clin Transplant. 2007;5:575–584. [PubMed] [Google Scholar]
  • 219.Lataillade JJ, Domenech J, Le Bousse-Kerdiles MC. Stromal cell-derived factor-1 (SDF-1)\CXCR4 couple plays multiple roles on haematopoietic progenitors at the border between the old cytokine and new chemokine worlds: survival, cell cycling and trafficking. Eur Cytokine Netw. 2004;15:177–188. [PubMed] [Google Scholar]
  • 220.Lathrop WF, Carmichael EP, Myles DG, Primakoff P. cDNA cloning reveals the molecular structure of a sperm surface protein, PH-20, involved in sperm-egg adhesion and the wide distribution of its gene among mammals. J Cell Biol. 1990;111:2939–2949. doi: 10.1083/jcb.111.6.2939. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 221.Laurent TC, Fraser JR. Hyaluronan. FASEB J. 1992;6:2397–2404. [PubMed] [Google Scholar]
  • 222.Laurent TC, Lilja K, Brunnberg L, Engstrom-Laurent A, Laurent UB, Lindqvist U, Murata K, Ytterberg D. Urinary excretion of hyaluronan in man. Scand J Clin Lab Invest. 1987;47:793–799. [PubMed] [Google Scholar]
  • 223.Laurent UB, Laurent TC, Hellsing LK, Persson L, Hartman M, Lilja K. Hyaluronan in human cerebrospinal fluid. Acta Neurol Scand. 1996;94:194–206. doi: 10.1111/j.1600-0404.1996.tb07052.x. [DOI] [PubMed] [Google Scholar]
  • 224.Lawrence T, Bebien M, Liu GY, Nizet V, Karin M. IKKalpha limits macrophage NF-kappaB activation and contributes to the resolution of inflammation. Nature. 2005;434:1138–1143. doi: 10.1038/nature03491. [DOI] [PubMed] [Google Scholar]
  • 225.Lebensztejn DM, Skiba E, Sobaniec-Lotowska ME, Kaczmarski M. Serum hyaluronan and laminin level in children with chronic hepatitis B during long-term lamivudine treatment. Hepatogastroenterology. 2007;54:834–838. [PubMed] [Google Scholar]
  • 226.Lee JH, Zhou HY, Cho SY, Kim YS, Lee YS, Jeong CS. Anti-inflammatory mechanisms of apigenin: inhibition of cyclooxygenase-2 expression, adhesion of monocytes to human umbilical vein endothelial cells, and expression of cellular adhesion molecules. Arch Pharm Res. 2007;30:1318–1327. doi: 10.1007/BF02980273. [DOI] [PubMed] [Google Scholar]
  • 227.Lee JY, Spicer AP. Hyaluronan: a multifunctional, megaDalton, stealth molecule. Curr Opin Cell Biol. 2000;12:581–586. doi: 10.1016/s0955-0674(00)00135-6. [DOI] [PubMed] [Google Scholar]
  • 228.Lee TH, Wisniewski HG, Vilcek J. A novel secretory tumor necrosis factor-inducible protein (TSG-6) is a member of the family of hyaluronate binding proteins, closely related to the adhesion receptor CD44. J Cell Biol. 1992;116:545–557. doi: 10.1083/jcb.116.2.545. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 229.Lepperdinger G, Strobl B, Kreil G. HYAL2, a human gene expressed in many cells, encodes a lysosomal hyaluronidase with a novel type of specificity. J Biol Chem. 1998;273:22466–22470. doi: 10.1074/jbc.273.35.22466. [DOI] [PubMed] [Google Scholar]
  • 230.Lesley J, Howes N, Perschl A, Hyman R. Hyaluronan binding function of CD44 is transiently activated on T cells during an in vivo immune response. J Exp Med. 1994;180:383–387. doi: 10.1084/jem.180.1.383. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 231.Lesley J, Hyman R. CD44 can be activated to function as an hyaluronic acid receptor in normal murine T cells. Eur J Immunol. 1992;22:2719–2723. doi: 10.1002/eji.1830221036. [DOI] [PubMed] [Google Scholar]
  • 232.Lesley J, Hyman R, Kincade PW. CD44 and its interaction with extracellular matrix. Adv Immunol. 1993;54:271–335. doi: 10.1016/s0065-2776(08)60537-4. [DOI] [PubMed] [Google Scholar]
  • 233.Levesque H, Delpech B, Le Loet X, Deshayes P. Serum hyaluronate in rheumatoid arthritis: study by affino-immunoenzymatic assay. Br J Rheumatol. 1988;27:445–449. doi: 10.1093/rheumatology/27.6.445. [DOI] [PubMed] [Google Scholar]
  • 234.Li L, Heldin CH, Heldin P. Inhibition of platelet-derived growth factor-BB-induced receptor activation and fibroblast migration by hyaluronan activation of CD44. J Biol Chem. 2006;281:26512–26519. doi: 10.1074/jbc.M605607200. [DOI] [PubMed] [Google Scholar]
  • 235.Li Y, Rahmanian M, Widstrom C, Lepperdinger G, Frost GI, Heldin P. Irradiation-induced expression of hyaluronan (HA) synthase 2 and hyaluronidase 2 genes in rat lung tissue accompanies active turnover of HA and induction of types I and III collagen gene expression. Am J Respir Cell Mol Biol. 2000;23:411–418. doi: 10.1165/ajrcmb.23.3.4102. [DOI] [PubMed] [Google Scholar]
  • 236.Li Y, Toole BP, Dealy CN, Kosher RA. Hyaluronan in limb morphogenesis. Dev Biol. 2007;305:411–420. doi: 10.1016/j.ydbio.2007.02.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 237.Liang J, Jiang D, Griffith J, Yu S, Fan J, Zhao X, Bucala R, Noble PW. CD44 is a negative regulator of acute pulmonary inflammation and lipopolysaccharide-TLR signaling in mouse macrophages. J Immunol. 2007;178:2469–2475. doi: 10.4049/jimmunol.178.4.2469. [DOI] [PubMed] [Google Scholar]
  • 238.Lin F, Ren XD, Doris G, Clark RA. Three-dimensional migration of human adult dermal fibroblasts from collagen lattices into fibrin/fibronectin gels requires syndecan-4 proteoglycan. J Invest Dermatol. 2005;124:906–913. doi: 10.1111/j.0022-202X.2005.23740.x. [DOI] [PubMed] [Google Scholar]
  • 239.Lipkin GW, Forbes MA, Cooper EH, Turney JH. Hyaluronic acid metabolism and its clinical significance in patients treated by continuous ambulatory peritoneal dialysis. Nephrol Dial Transplant. 1993;8:357–360. [PubMed] [Google Scholar]
  • 240.Lokeshwar VB, Selzer MG. Differences in hyaluronic acid-mediated functions and signaling in arterial, microvessel, and vein-derived human endothelial cells. J Biol Chem. 2000;275:27641–27649. doi: 10.1074/jbc.M003084200. [DOI] [PubMed] [Google Scholar]
  • 241.Lowther DA, Rogers HJ. Biosynthesis of hyaluronate. Nature. 1955;175:435. doi: 10.1038/175435a0. [DOI] [PubMed] [Google Scholar]
  • 242.Luke HJ, Prehm P. Synthesis and shedding of hyaluronan from plasma membranes of human fibroblasts and metastatic and non-metastatic melanoma cells. Biochem J. 1999;343(Pt 1):71–75. [PMC free article] [PubMed] [Google Scholar]
  • 243.Maclennan AP. The production of capsules, hyaluronic acid and hyaluronidase by 25 strains of group C streptococci. J Gen Microbiol. 1956;15:485–491. doi: 10.1099/00221287-15-3-485. [DOI] [PubMed] [Google Scholar]
  • 244.Mahadevan P, Larkins RG, Fraser JR, Dunlop ME. Effect of prostaglandin E2 and hyaluronan on mesangial cell proliferation. A potential contribution to glomerular hypercellularity in diabetes. Diabetes. 1996;45:44–50. doi: 10.2337/diab.45.1.44. [DOI] [PubMed] [Google Scholar]
  • 245.Mahadevan P, Larkins RG, Fraser JR, Fosang AJ, Dunlop ME. Increased hyaluronan production in the glomeruli from diabetic rats: a link between glucose-induced prostaglandin production and reduced sulphated proteoglycan. Diabetologia. 1995;38:298–305. doi: 10.1007/BF00400634. [DOI] [PubMed] [Google Scholar]
  • 246.Majumdar M, Meenakshi J, Goswami SK, Datta K. Hyaluronan binding protein 1 (HABP1)/C1QBP/p32 is an endogenous substrate for MAP kinase and is translocated to the nucleus upon mitogenic stimulation. Biochem Biophys Res Commun. 2002;291:829–837. doi: 10.1006/bbrc.2002.6491. [DOI] [PubMed] [Google Scholar]
  • 247.Manicourt DH, Triki R, Fukuda K, Devogelaer JP, Nagant de Deuxchaisnes C, Thonar EJ. Levels of circulating tumor necrosis factor alpha and interleukin-6 in patients with rheumatoid arthritis. Relationship to serum levels of hyaluronan and antigenic keratan sulfate. Arthritis Rheum. 1993;36:490–499. doi: 10.1002/art.1780360409. [DOI] [PubMed] [Google Scholar]
  • 248.Manzanares D, Monzon ME, Savani RC, Salathe M. Apical oxidative hyaluronan degradation stimulates airway ciliary beating via RHAMM and RON. Am J Respir Cell Mol Biol. 2007;37:160–168. doi: 10.1165/rcmb.2006-0413OC. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 249.Margolis RK, Rauch U, Maurel P, Margolis RU. Neurocan and phosphacan: two major nervous tissue-specific chondroitin sulfate proteoglycans. Perspect Dev Neurobiol. 1996;3:273–290. [PubMed] [Google Scholar]
  • 250.Maroko PR, Davidson DM, Libby P, Hagan AD, Braunwald E. Effects of hyaluronidase administration on myocardial ischemic injury in acute infarction. A preliminary study in 24 patients. Ann Intern Med. 1975;82:516–520. doi: 10.7326/0003-4819-82-4-516. [DOI] [PubMed] [Google Scholar]
  • 251.Maroko PR, Hillis LD, Muller JE, Tavazzi L, Heyndrickx GR, Ray M, Chiariello M, Distante A, Askenazi J, Salerno J, Carpentier J, Reshetnaya NI, Radvany P, Libby P, Raabe DS, Chazov EI, Bobba P, Braunwald E. Favorable effects of hyaluronidase on electrocardiographic evidence of necrosis in patients with acute myocardial infarction. N Engl J Med. 1977;296:898–903. doi: 10.1056/NEJM197704212961603. [DOI] [PubMed] [Google Scholar]
  • 252.Martin C, Papazian L, Payan MJ, Saux P, Gouin F. Pulmonary fibrosis correlates with outcome in adult respiratory distress syndrome. A study in mechanically ventilated patients. Chest. 1995;107:196–200. doi: 10.1378/chest.107.1.196. [DOI] [PubMed] [Google Scholar]
  • 253.Martinez FJ, Safrin S, Weycker D, Starko KM, Bradford WZ, King TE, Jr, Flaherty KR, Schwartz DA, Noble PW, Raghu G, Brown KK. The clinical course of patients with idiopathic pulmonary fibrosis. Ann Intern Med. 2005;142:963–967. doi: 10.7326/0003-4819-142-12_part_1-200506210-00005. [DOI] [PubMed] [Google Scholar]
  • 254.Marutsuka K, Woodcock-Mitchell J, Sakamoto T, Sobel BE, Fujii S. Pathogenetic implications of hyaluronan-induced modification of vascular smooth muscle cell fibrinolysis in diabetes. Coron Artery Dis. 1998;9:177–184. doi: 10.1097/00019501-199809040-00002. [DOI] [PubMed] [Google Scholar]
  • 255.Mascarenhas MM, Day RM, Ochoa CD, Choi WI, Yu L, Ouyang B, Garg HG, Hales CA, Quinn DA. Low molecular weight hyaluronan from stretched lung enhances interleukin-8 expression. Am J Respir Cell Mol Biol. 2004;30:51–60. doi: 10.1165/rcmb.2002-0167OC. [DOI] [PubMed] [Google Scholar]
  • 256.Matrosova VY, Orlovskaya IA, Serobyan N, Khaldoyanidi SK. Hyaluronic acid facilitates the recovery of hematopoiesis following 5-fluorouracil administration. Stem Cells. 2004;22:544–555. doi: 10.1634/stemcells.22-4-544. [DOI] [PubMed] [Google Scholar]
  • 257.Matsuda M, Shikata K, Shimizu F, Suzuki Y, Miyasaka M, Kawachi H, Kawashima H, Wada J, Sugimoto H, Shikata Y, Ogawa D, Tojo SJ, Akima K, Makino H. Therapeutic effect of sulphated hyaluronic acid, a potential selectin-blocking agent, on experimental progressive mesangial proliferative glomerulonephritis. J Pathol. 2002;198:407–414. doi: 10.1002/path.1209. [DOI] [PubMed] [Google Scholar]
  • 258.Matsumoto K, Li Y, Jakuba C, Sugiyama Y, Sayo T, Okuno M, Dealy CN, Toole BP, Takeda J, Yamaguchi Y, Kosher RA. Conditional inactivation of Has2 reveals a crucial role for hyaluronan in skeletal growth, patterning, chondrocyte maturation and joint formation in the developing limb. Development. 2009;136:2825–2835. doi: 10.1242/dev.038505. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 259.Maula SM, Luukkaa M, Grenman R, Jackson D, Jalkanen S, Ristamaki R. Intratumoral lymphatics are essential for the metastatic spread and prognosis in squamous cell carcinomas of the head and neck region. Cancer Res. 2003;63:1920–1926. [PubMed] [Google Scholar]
  • 260.McDonald JA, Camenisch TD. Hyaluronan: genetic insights into the complex biology of a simple polysaccharide. Glycoconj J. 2002;19:331–339. doi: 10.1023/A:1025369004783. [DOI] [PubMed] [Google Scholar]
  • 261.McDonald TO, Gerrity RG, Jen C, Chen HJ, Wark K, Wight TN, Chait A, O’Brien KD. Diabetes and arterial extracellular matrix changes in a porcine model of atherosclerosis. J Histochem Cytochem. 2007;55:1149–1157. doi: 10.1369/jhc.7A7221.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 262.McGary CT, Weigel JA, Weigel PH. Study of hyaluronan-binding proteins and receptors using iodinated hyaluronan derivatives. Methods Enzymol. 2003;363:354–365. doi: 10.1016/S0076-6879(03)01064-4. [DOI] [PubMed] [Google Scholar]
  • 263.McKallip RJ, Fisher M, Gunthert U, Szakal AK, Nagarkatti PS, Nagarkatti M. Role of CD44 and its v7 isoform in staphylococcal enterotoxin B-induced toxic shock: CD44 deficiency on hepatic mononuclear cells leads to reduced activation-induced apoptosis that results in increased liver damage. Infect Immun. 2005;73:50–61. doi: 10.1128/IAI.73.1.50-61.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 264.McKee CM, Lowenstein CJ, Horton MR, Wu J, Bao C, Chin BY, Choi AM, Noble PW. Hyaluronan fragments induce nitric-oxide synthase in murine macrophages through a nuclear factor kappaB-dependent mechanism. J Biol Chem. 1997;272:8013–8018. doi: 10.1074/jbc.272.12.8013. [DOI] [PubMed] [Google Scholar]
  • 265.McKee CM, Penno MB, Cowman M, Burdick MD, Strieter RM, Bao C, Noble PW. Hyaluronan (HA) fragments induce chemokine gene expression in alveolar macrophages. The role of HA size and CD44. J Clin Invest. 1996;98:2403–2413. doi: 10.1172/JCI119054. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 266.McKeon RJ, Jurynec MJ, Buck CR. The chondroitin sulfate proteoglycans neurocan and phosphacan are expressed by reactive astrocytes in the chronic CNS glial scar. J Neurosci. 1999;19:10778–10788. doi: 10.1523/JNEUROSCI.19-24-10778.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 267.Meenakshi J, Anupama, Goswami SK, Datta K. Constitutive expression of hyaluronan binding protein 1 (HABP1/p32/gC1qR) in normal fibroblast cells perturbs its growth characteristics and induces apoptosis. Biochem Biophys Res Commun. 2003;300:686–693. doi: 10.1016/s0006-291x(02)02788-2. [DOI] [PubMed] [Google Scholar]
  • 268.Mendez E, Fernandez-Luna JL, Grubb A, Leyva-Cobian F. Human protein HC and its IgA complex are inhibitors of neutrophil chemotaxis. Proc Natl Acad Sci U S A. 1986;83:1472–1475. doi: 10.1073/pnas.83.5.1472. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 269.Mendez-Ferrer S, Frenette PS. Hematopoietic stem cell trafficking: regulated adhesion and attraction to bone marrow microenvironment. Ann N Y Acad Sci. 2007;1116:392–413. doi: 10.1196/annals.1402.086. [DOI] [PubMed] [Google Scholar]
  • 270.Meran S, Thomas DW, Stephens P, Enoch S, Martin J, Steadman R, Phillips AO. Hyaluronan facilitates transforming growth factor-beta1-mediated fibroblast proliferation. J Biol Chem. 2008;283:6530–6545. doi: 10.1074/jbc.M704819200. [DOI] [PubMed] [Google Scholar]
  • 271.Meyer K, Palmer JW. The Polysaccharide of the vitreous humor. J Biol Chem. 1934;107:629–634. [Google Scholar]
  • 272.Mikecz K, Brennan FR, Kim JH, Glant TT. Anti-CD44 treatment abrogates tissue oedema and leukocyte infiltration in murine arthritis. Nat Med. 1995;1:558–563. doi: 10.1038/nm0695-558. [DOI] [PubMed] [Google Scholar]
  • 273.Milner CM, Day AJ. TSG-6: a multifunctional protein associated with inflammation. J Cell Sci. 2003;116:1863–1873. doi: 10.1242/jcs.00407. [DOI] [PubMed] [Google Scholar]
  • 274.Mindrescu C, Thorbecke GJ, Klein MJ, Vilcek J, Wisniewski HG. Amelioration of collagen-induced arthritis in DBA/1J mice by recombinant TSG-6, a tumor necrosis factor/interleukin-1-inducible protein. Arthritis Rheum. 2000;43:2668–2677. doi: 10.1002/1529-0131(200012)43:12<2668::AID-ANR6>3.0.CO;2-E. [DOI] [PubMed] [Google Scholar]
  • 275.Mine S, Okada Y, Kawahara C, Tabata T, Tanaka Y. Serum hyaluronan concentration as a marker of angiopathy in patients with diabetes mellitus. Endocr J. 2006;53:761–766. doi: 10.1507/endocrj.k05-119. [DOI] [PubMed] [Google Scholar]
  • 276.Mjaatvedt CH, Yamamura H, Capehart AA, Turner D, Markwald RR. The Cspg2 gene, disrupted in the hdf mutant, is required for right cardiac chamber and endocardial cushion formation. Dev Biol. 1998;202:56–66. doi: 10.1006/dbio.1998.9001. [DOI] [PubMed] [Google Scholar]
  • 277.Mohamadzadeh M, DeGrendele H, Arizpe H, Estess P, Siegelman M. Proinflammatory stimuli regulate endothelial hyaluronan expression and CD44/HA-dependent primary adhesion. J Clin Invest. 1998;101:97–108. doi: 10.1172/JCI1604. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 278.Monslow J, Williams JD, Fraser DJ, Michael DR, Foka P, Kift-Morgan AP, Luo DD, Fielding CA, Craig KJ, Topley N, Jones SA, Ramji DP, Bowen T. Sp1 and Sp3 mediate constitutive transcription of the human hyaluronan synthase 2 gene. J Biol Chem. 2006;281:18043–18050. doi: 10.1074/jbc.M510467200. [DOI] [PubMed] [Google Scholar]
  • 279.Monzon ME, Casalino-Matsuda SM, Forteza RM. Identification of glycosaminoglycans in human airway secretions. Am J Respir Cell Mol Biol. 2006;34:135–141. doi: 10.1165/rcmb.2005-0256OC. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 280.Morisada T, Oike Y, Yamada Y, Urano T, Akao M, Kubota Y, Maekawa H, Kimura Y, Ohmura M, Miyamoto T, Nozawa S, Koh GY, Alitalo K, Suda T. Angiopoietin-1 promotes LYVE-1-positive lymphatic vessel formation. Blood. 2005;105:4649–4656. doi: 10.1182/blood-2004-08-3382. [DOI] [PubMed] [Google Scholar]
  • 281.Moseley R, Waddington RJ, Embery G. Degradation of glycosaminoglycans by reactive oxygen species derived from stimulated polymorphonuclear leukocytes. Biochim Biophys Acta. 1997;1362:221–231. doi: 10.1016/s0925-4439(97)00083-5. [DOI] [PubMed] [Google Scholar]
  • 282.Mouta Carreira C, Nasser SM, di Tomaso E, Padera TP, Boucher Y, Tomarev SI, Jain RK. LYVE-1 is not restricted to the lymph vessels: expression in normal liver blood sinusoids and down-regulation in human liver cancer and cirrhosis. Cancer Res. 2001;61:8079–8084. [PubMed] [Google Scholar]
  • 283.Mummert ME, Mohamadzadeh M, Mummert DI, Mizumoto N, Takashima A. Development of a peptide inhibitor of hyaluronan-mediated leukocyte trafficking. J Exp Med. 2000;192:769–779. doi: 10.1084/jem.192.6.769. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 284.Murawaki Y, Ikuta Y, Koda M, Nishimura Y, Kawasaki H. Clinical significance of serum hyaluronan in patients with chronic viral liver disease. J Gastroenterol Hepatol. 1996;11:459–465. doi: 10.1111/j.1440-1746.1996.tb00291.x. [DOI] [PubMed] [Google Scholar]
  • 285.Mustafa A, McKallip RJ, Fisher M, Duncan R, Nagarkatti PS, Nagarkatti M. Regulation of interleukin-2-induced vascular leak syndrome by targeting CD44 using hyaluronic acid and anti-CD44 antibodies. J Immunother (1997) 2002;25:476–488. doi: 10.1097/00002371-200211000-00004. [DOI] [PubMed] [Google Scholar]
  • 286.Muto J, Yamasaki K, Taylor KR, Gallo RL. Engagement of CD44 by hyaluronan suppresses TLR4 signaling and the septic response to LPS. Mol Immunol. 2009 doi: 10.1016/j.molimm.2009.08.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 287.Mylona E, Jones KA, Mills ST, Pavlath GK. CD44 regulates myoblast migration and differentiation. J Cell Physiol. 2006;209:314–321. doi: 10.1002/jcp.20724. [DOI] [PubMed] [Google Scholar]
  • 288.Nadkarni PP, Kulkarni GS, Cerreta JM, Ma S, Cantor JO. Dichotomous effect of aerosolized hyaluronan in a hamster model of endotoxin-induced lung injury. Exp Lung Res. 2005;31:807–818. doi: 10.1080/01902140600574942. [DOI] [PubMed] [Google Scholar]
  • 289.Nairn AV, Kinoshita-Toyoda A, Toyoda H, Xie J, Harris K, Dalton S, Kulik M, Pierce JM, Toida T, Moremen KW, Linhardt RJ. Glycomics of proteoglycan biosynthesis in murine embryonic stem cell differentiation. J Proteome Res. 2007;6:4374–4387. doi: 10.1021/pr070446f. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 290.Nakamura H, Kitazawa K, Honda H, Sugisaki T. Roles of and correlation between alpha-smooth muscle actin, CD44, hyaluronic acid and osteopontin in crescent formation in human glomerulonephritis. Clin Nephrol. 2005;64:401–411. doi: 10.5414/cnp64401. [DOI] [PubMed] [Google Scholar]
  • 291.Nakamura M, Sato N, Chikama TI, Hasegawa Y, Nishida T. Hyaluronan facilitates corneal epithelial wound healing in diabetic rats. Exp Eye Res. 1997;64:1043–1050. doi: 10.1006/exer.1997.0302. [DOI] [PubMed] [Google Scholar]
  • 292.Nandi A, Estess P, Siegelman M. Bimolecular complex between rolling and firm adhesion receptors required for cell arrest; CD44 association with VLA-4 in T cell extravasation. Immunity. 2004;20:455–465. doi: 10.1016/s1074-7613(04)00077-9. [DOI] [PubMed] [Google Scholar]
  • 293.Natowicz MR, Short MP, Wang Y, Dickersin GR, Gebhardt MC, Rosenthal DI, Sims KB, Rosenberg AE. Clinical and biochemical manifestations of hyaluronidase deficiency. N Engl J Med. 1996;335:1029–1033. doi: 10.1056/NEJM199610033351405. [DOI] [PubMed] [Google Scholar]
  • 294.Nedvetzki S, Gonen E, Assayag N, Reich R, Williams RO, Thurmond RL, Huang JF, Neudecker BA, Wang FS, Turley EA, Naor D. RHAMM, a receptor for hyaluronan-mediated motility, compensates for CD44 in inflamed CD44-knockout mice: a different interpretation of redundancy. Proc Natl Acad Sci U S A. 2004;101:18081–18086. doi: 10.1073/pnas.0407378102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 295.Nettelbladt O, Scheynius A, Bergh J, Tengblad A, Hallgren R. Alveolar accumulation of hyaluronan and alveolar cellular response in bleomycin-induced alveolitis. Eur Respir J. 1991;4:407–414. [PubMed] [Google Scholar]
  • 296.Nettelbladt O, Tengblad A, Hallgren R. High-dose corticosteroids during bleomycin-induced alveolitis in the rat do not suppress the accumulation of hyaluronan (hyaluronic acid) in lung tissue. Eur Respir J. 1990;3:421–428. [PubMed] [Google Scholar]
  • 297.Neudecker BA, Stern R, Connolly MK. Aberrant serum hyaluronan and hyaluronidase levels in scleroderma. Br J Dermatol. 2004;150:469–476. doi: 10.1046/j.1365-2133.2004.05805.x. [DOI] [PubMed] [Google Scholar]
  • 298.Neudecker BA, Stern R, Mark LA, Steinberg S. Scleromyxedema-like lesions of patients in renal failure contain hyaluronan: a possible pathophysiological mechanism. J Cutan Pathol. 2005;32:612–615. doi: 10.1111/j.0303-6987.2005.00415.x. [DOI] [PubMed] [Google Scholar]
  • 299.Nieuwdorp M, Holleman F, de Groot E, Vink H, Gort J, Kontush A, Chapman MJ, Hutten BA, Brouwer CB, Hoekstra JB, Kastelein JJ, Stroes ES. Perturbation of hyaluronan metabolism predisposes patients with type 1 diabetes mellitus to atherosclerosis. Diabetologia. 2007;50:1288–1293. doi: 10.1007/s00125-007-0666-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 300.Nikolic-Paterson DJ, Jun Z, Tesch GH, Lan HY, Foti R, Atkins RC. De novo CD44 expression by proliferating mesangial cells in rat anti-Thy-1 nephritis. J Am Soc Nephrol. 1996;7:1006–1014. doi: 10.1681/ASN.V771006. [DOI] [PubMed] [Google Scholar]
  • 301.Nilsson SK, Haylock DN, Johnston HM, Occhiodoro T, Brown TJ, Simmons PJ. Hyaluronan is synthesized by primitive hemopoietic cells, participates in their lodgment at the endosteum following transplantation, and is involved in the regulation of their proliferation and differentiation in vitro. Blood. 2003;101:856–862. doi: 10.1182/blood-2002-05-1344. [DOI] [PubMed] [Google Scholar]
  • 302.Nishikawa K, Andres G, Bhan AK, McCluskey RT, Collins AB, Stow JL, Stamenkovic I. Hyaluronate is a component of crescents in rat autoimmune glomerulonephritis. Lab Invest. 1993;68:146–153. [PubMed] [Google Scholar]
  • 303.Noble PW. Hyaluronan and its catabolic products in tissue injury and repair. Matrix Biol. 2002;21:25–29. doi: 10.1016/s0945-053x(01)00184-6. [DOI] [PubMed] [Google Scholar]
  • 304.Noble PW. Idiopathic pulmonary fibrosis: natural history and prognosis. Clin Chest Med. 2006;27:S11–16. v. doi: 10.1016/j.ccm.2005.08.003. [DOI] [PubMed] [Google Scholar]
  • 305.Noble PW, Lake FR, Henson PM, Riches DW. Hyaluronate activation of CD44 induces insulin-like growth factor-1 expression by a tumor necrosis factor-alpha-dependent mechanism in murine macrophages. J Clin Invest. 1993;91:2368–2377. doi: 10.1172/JCI116469. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 306.Noble PW, McKee CM, Cowman M, Shin HS. Hyaluronan fragments activate an NF-kappa B/I-kappa B alpha autoregulatory loop in murine macrophages. J Exp Med. 1996;183:2373–2378. doi: 10.1084/jem.183.5.2373. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 307.Nonaka H, Tanaka M, Suzuki K, Miyajima A. Development of murine hepatic sinusoidal endothelial cells characterized by the expression of hyaluronan receptors. Dev Dyn. 2007;236:2258–2267. doi: 10.1002/dvdy.21227. [DOI] [PubMed] [Google Scholar]
  • 308.Nyberg A, Lindqvist U, Engstrom-Laurent A. Serum hyaluronan and aminoterminal propeptide of type III procollagen in primary biliary cirrhosis: relation to clinical symptoms, liver histopathology and outcome. J Intern Med. 1992;231:485–491. doi: 10.1111/j.1365-2796.1992.tb00964.x. [DOI] [PubMed] [Google Scholar]
  • 309.O’Neill LA. TLRs play good cop, bad cop in the lung. Nat Med. 2005;11:1161–1162. doi: 10.1038/nm1105-1161. [DOI] [PubMed] [Google Scholar]
  • 310.Oertli B, Beck-Schimmer B, Fan X, Wuthrich RP. Mechanisms of hyaluronan-induced up-regulation of ICAM-1 and VCAM-1 expression by murine kidney tubular epithelial cells: hyaluronan triggers cell adhesion molecule expression through a mechanism involving activation of nuclear factor-kappa B and activating protein-1. J Immunol. 1998;161:3431–3437. [PubMed] [Google Scholar]
  • 311.Ogata T, Okuda K, Ueno T, Saito N, Aoyagi S. Serum hyaluronan as a predictor of hepatic regeneration after hepatectomy in humans. Eur J Clin Invest. 1999;29:780–785. doi: 10.1046/j.1365-2362.1999.00513.x. [DOI] [PubMed] [Google Scholar]
  • 312.Ogawa D, Shikata K, Matsuda M, Akima K, Iwahashi M, Okada S, Tsuchiyama Y, Shikata Y, Wada J, Makino H. Sulfated hyaluronic acid, a potential selectin inhibitor, ameliorates experimentally induced crescentic glomerulonephritis. Nephron Exp Nephrol. 2005;99:e26–32. doi: 10.1159/000081795. [DOI] [PubMed] [Google Scholar]
  • 313.Ohkawara Y, Tamura G, Iwasaki T, Tanaka A, Kikuchi T, Shirato K. Activation and transforming growth factor-beta production in eosinophils by hyaluronan. Am J Respir Cell Mol Biol. 2000;23:444–451. doi: 10.1165/ajrcmb.23.4.3875. [DOI] [PubMed] [Google Scholar]
  • 314.Ohto U, Satow Y. Crystal twinning of human MD-2 recognizing endotoxin cores of lipopolysaccharide. J Synchrotron Radiat. 2008;15:262–265. doi: 10.1107/S0909049507056531. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 315.Okayama N, Fujimura K, Suehiro Y, Hamanaka Y, Fujiwara M, Matsubara T, Maekawa T, Hazama S, Oka M, Nohara H, Kayano K, Okita K, Hinoda Y. Simple genotype analysis of the Asp299Gly polymorphism of the Toll-like receptor-4 gene that is associated with lipopolysaccharide hyporesponsiveness. J Clin Lab Anal. 2002;16:56–58. doi: 10.1002/jcla.2075. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 316.Panasyuk A, Frati E, Ribault D, Mitrovic D. Effect of reactive oxygen species on the biosynthesis and structure of newly synthesized proteoglycans. Free Radic Biol Med. 1994;16:157–167. doi: 10.1016/0891-5849(94)90139-2. [DOI] [PubMed] [Google Scholar]
  • 317.Panes J, Gerritsen ME, Anderson DC, Miyasaka M, Granger DN. Apigenin inhibits tumor necrosis factor-induced intercellular adhesion molecule-1 upregulation in vivo. Microcirculation. 1996;3:279–286. doi: 10.3109/10739689609148302. [DOI] [PubMed] [Google Scholar]
  • 318.Papakonstantinou E, Karakiulakis G, Roth M, Block LH. Platelet-derived growth factor stimulates the secretion of hyaluronic acid by proliferating human vascular smooth muscle cells. Proc Natl Acad Sci U S A. 1995;92:9881–9885. doi: 10.1073/pnas.92.21.9881. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 319.Papakonstantinou E, Karakiulakis G, Tamm M, Perruchoud AP, Roth M. Hypoxia modifies the effect of PDGF on glycosaminoglycan synthesis by primary human lung cells. Am J Physiol Lung Cell Mol Physiol. 2000;279:L825–834. doi: 10.1152/ajplung.2000.279.5.L825. [DOI] [PubMed] [Google Scholar]
  • 320.Papakonstantinou E, Roth M, Block LH, Mirtsou-Fidani V, Argiriadis P, Karakiulakis G. The differential distribution of hyaluronic acid in the layers of human atheromatic aortas is associated with vascular smooth muscle cell proliferation and migration. Atherosclerosis. 1998;138:79–89. doi: 10.1016/s0021-9150(98)00006-9. [DOI] [PubMed] [Google Scholar]
  • 321.Park BS, Song DH, Kim HM, Choi BS, Lee H, Lee JO. The structural basis of lipopolysaccharide recognition by the TLR4-MD-2 complex. Nature. 2009;458:1191–1195. doi: 10.1038/nature07830. [DOI] [PubMed] [Google Scholar]
  • 322.Parkar AA, Day AJ. Overlapping sites on the Link module of human TSG-6 mediate binding to hyaluronan and chrondroitin-4-sulphate. FEBS Lett. 1997;410:413–417. doi: 10.1016/s0014-5793(97)00621-2. [DOI] [PubMed] [Google Scholar]
  • 323.Peck D, Isacke CM. CD44 phosphorylation regulates melanoma cell and fibroblast migration on, but not attachment to, a hyaluronan substratum. Curr Biol. 1996;6:884–890. doi: 10.1016/s0960-9822(02)00612-7. [DOI] [PubMed] [Google Scholar]
  • 324.Pecoits-Filho R, Heimburger O, Barany P, Suliman M, Fehrman-Ekholm I, Lindholm B, Stenvinkel P. Associations between circulating inflammatory markers and residual renal function in CRF patients. Am J Kidney Dis. 2003;41:1212–1218. doi: 10.1016/s0272-6386(03)00353-6. [DOI] [PubMed] [Google Scholar]
  • 325.Peterson RS, Andhare RA, Rousche KT, Knudson W, Wang W, Grossfield JB, Thomas RO, Hollingsworth RE, Knudson CB. CD44 modulates Smad1 activation in the BMP-7 signaling pathway. J Cell Biol. 2004;166:1081–1091. doi: 10.1083/jcb.200402138. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 326.Petrigni G, Allegra L. Aerosolised hyaluronic acid prevents exercise-induced bronchoconstriction, suggesting novel hypotheses on the correction of matrix defects in asthma. Pulm Pharmacol Ther. 2006;19:166–171. doi: 10.1016/j.pupt.2005.03.002. [DOI] [PubMed] [Google Scholar]
  • 327.Pienimaki JP, Rilla K, Fulop C, Sironen RK, Karvinen S, Pasonen S, Lammi MJ, Tammi R, Hascall VC, Tammi MI. Epidermal growth factor activates hyaluronan synthase 2 in epidermal keratinocytes and increases pericellular and intracellular hyaluronan. J Biol Chem. 2001;276:20428–20435. doi: 10.1074/jbc.M007601200. [DOI] [PubMed] [Google Scholar]
  • 328.Pilarski LM, Pruski E, Wizniak J, Paine D, Seeberger K, Mant MJ, Brown CB, Belch AR. Potential role for hyaluronan and the hyaluronan receptor RHAMM in mobilization and trafficking of hematopoietic progenitor cells. Blood. 1999;93:2918–2927. [PubMed] [Google Scholar]
  • 329.Politz O, Gratchev A, McCourt PA, Schledzewski K, Guillot P, Johansson S, Svineng G, Franke P, Kannicht C, Kzhyshkowska J, Longati P, Velten FW, Johansson S, Goerdt S. Stabilin-1 and -2 constitute a novel family of fasciclin-like hyaluronan receptor homologues. Biochem J. 2002;362:155–164. doi: 10.1042/0264-6021:3620155. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 330.Pontinha N, Pessegueiro H, Barros H. Serum hyaluronan as a marker of liver fibrosis in asymptomatic chronic viral hepatitis B. Scand J Clin Lab Invest. 1999;59:343–347. doi: 10.1080/00365519950185535. [DOI] [PubMed] [Google Scholar]
  • 331.Poulsom R. CD44 and hyaluronan help mesenchymal stem cells move to a neighborhood in need of regeneration. Kidney Int. 2007;72:389–390. doi: 10.1038/sj.ki.5002398. [DOI] [PubMed] [Google Scholar]
  • 332.Prehm P. Hyaluronate is synthesized at plasma membranes. Biochem J. 1984;220:597–600. doi: 10.1042/bj2200597. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 333.Prehm P. Release of hyaluronate from eukaryotic cells. Biochem J. 1990;267:185–189. doi: 10.1042/bj2670185. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 334.Prehm P. Synthesis of hyaluronate in differentiated teratocarcinoma cells. Characterization of the synthase. Biochem J. 1983;211:181–189. doi: 10.1042/bj2110181. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 335.Prehm P. Synthesis of hyaluronate in differentiated teratocarcinoma cells. Mechanism of chain growth. Biochem J. 1983;211:191–198. doi: 10.1042/bj2110191. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 336.Premaratne S, Watanabe BI, LaPenna WF, McNamara JJ. Effects of hyaluronidase on reducing myocardial infarct size in a baboon model of ischemia-reperfusion injury. J Surg Res. 1995;58:205–210. doi: 10.1006/jsre.1995.1032. [DOI] [PubMed] [Google Scholar]
  • 337.Prevo R, Banerji S, Ferguson DJ, Clasper S, Jackson DG. Mouse LYVE-1 is an endocytic receptor for hyaluronan in lymphatic endothelium. J Biol Chem. 2001;276:19420–19430. doi: 10.1074/jbc.M011004200. [DOI] [PubMed] [Google Scholar]
  • 338.Radhakrishnamurthy B. Consequences of prolonged inhalation of ozone on F344/N rats: collaborative studies. Part III: Effects on complex carbohydrates of lung connective tissue. Res Rep Health Eff Inst. 1994:3–14. discussion 15–23. [PubMed] [Google Scholar]
  • 339.Rafi-Janajreh AQ, Chen D, Schmits R, Mak TW, Grayson RL, Sponenberg DP, Nagarkatti M, Nagarkatti PS. Evidence for the involvement of CD44 in endothelial cell injury and induction of vascular leak syndrome by IL-2. J Immunol. 1999;163:1619–1627. [PubMed] [Google Scholar]
  • 340.Ragan C, Meyer K. The Hyaluronic Acid of Synovial Fluid in Rheumatoid Arthritis. J Clin Invest. 1949;28:56–59. doi: 10.1172/JCI102053. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 341.Rai SK, Duh FM, Vigdorovich V, Danilkovitch-Miagkova A, Lerman MI, Miller AD. Candidate tumor suppressor HYAL2 is a glycosylphosphatidylinositol (GPI)-anchored cell-surface receptor for jaagsiekte sheep retrovirus, the envelope protein of which mediates oncogenic transformation. Proc Natl Acad Sci U S A. 2001;98:4443–4448. doi: 10.1073/pnas.071572898. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 342.Rakoff-Nahoum S, Paglino J, Eslami-Varzaneh F, Edberg S, Medzhitov R. Recognition of commensal microflora by toll-like receptors is required for intestinal homeostasis. Cell. 2004;118:229–241. doi: 10.1016/j.cell.2004.07.002. [DOI] [PubMed] [Google Scholar]
  • 343.Ramadori G, Zohrens G, Manns M, Rieder H, Dienes HP, Hess G, Meyer KH, Buschenfelde Z. Serum hyaluronate and type III procollagen aminoterminal propeptide concentration in chronic liver disease. Relationship to cirrhosis and disease activity. Eur J Clin Invest. 1991;21:323–330. doi: 10.1111/j.1365-2362.1991.tb01377.x. [DOI] [PubMed] [Google Scholar]
  • 344.Rao CM, Salotra P, Datta K. Possible role of the 34-kilodalton hyaluronic acid-binding protein in visceral Leishmaniasis. J Parasitol. 1999;85:682–687. [PubMed] [Google Scholar]
  • 345.Rao DA, Tracey KJ, Pober JS. IL-1alpha and IL-1beta are endogenous mediators linking cell injury to the adaptive alloimmune response. J Immunol. 2007;179:6536–6546. doi: 10.4049/jimmunol.179.10.6536. [DOI] [PubMed] [Google Scholar]
  • 346.Rauch U. Extracellular matrix components associated with remodeling processes in brain. Cell Mol Life Sci. 2004;61:2031–2045. doi: 10.1007/s00018-004-4043-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 347.Rauch U, Clement A, Retzler C, Frohlich L, Fassler R, Gohring W, Faissner A. Mapping of a defined neurocan binding site to distinct domains of tenascin-C. J Biol Chem. 1997;272:26905–26912. doi: 10.1074/jbc.272.43.26905. [DOI] [PubMed] [Google Scholar]
  • 348.Rauch U, Hirakawa S, Oohashi T, Kappler J, Roos G. Cartilage link protein interacts with neurocan, which shows hyaluronan binding characteristics different from CD44 and TSG-6. Matrix Biol. 2004;22:629–639. doi: 10.1016/j.matbio.2003.11.007. [DOI] [PubMed] [Google Scholar]
  • 349.Regan C, Meyer K. Hyaluronic acid-hyaluronidase and the rheumatic diseases. Ann N Y Acad Sci. 1950;52:1108–1111. doi: 10.1111/j.1749-6632.1950.tb54012.x. [DOI] [PubMed] [Google Scholar]
  • 350.Reitinger S, Laschober GT, Fehrer C, Greiderer B, Lepperdinger G. Mouse testicular hyaluronidase-like proteins SPAM1 and HYAL5 but not HYALP1 degrade hyaluronan. Biochem J. 2007;401:79–85. doi: 10.1042/BJ20060598. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 351.Rilla K, Siiskonen H, Spicer AP, Hyttinen JM, Tammi MI, Tammi RH. Plasma membrane residence of hyaluronan synthase is coupled to its enzymatic activity. J Biol Chem. 2005;280:31890–31897. doi: 10.1074/jbc.M504736200. [DOI] [PubMed] [Google Scholar]
  • 352.Roberts R, Braunwald E, Muller JE, Croft C, Gold HK, Hartwell TD, Jaffe AS, Mullin SM, Parker C, Passamani ER, et al. Effect of hyaluronidase on mortality and morbidity in patients with early peaking of plasma creatine kinase MB and non-transmural ischaemia. Multicentre investigation for the limitation of infarct size (MILIS) Br Heart J. 1988;60:290–298. doi: 10.1136/hrt.60.4.290. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 353.Rockey DC, Chung JJ, McKee CM, Noble PW. Stimulation of inducible nitric oxide synthase in rat liver by hyaluronan fragments. Hepatology. 1998;27:86–92. doi: 10.1002/hep.510270115. [DOI] [PubMed] [Google Scholar]
  • 354.Rooney P, Wang M, Kumar P, Kumar S. Angiogenic oligosaccharides of hyaluronan enhance the production of collagens by endothelial cells. J Cell Sci. 1993;105 (Pt 1):213–218. doi: 10.1242/jcs.105.1.213. [DOI] [PubMed] [Google Scholar]
  • 355.Roth A, Mollenhauer J, Wagner A, Fuhrmann R, Straub A, Venbrocks RA, Petrow P, Brauer R, Schubert H, Ozegowski J, Peschel G, Muller PJ, Kinne RW. Intra-articular injections of high-molecular-weight hyaluronic acid have biphasic effects on joint inflammation and destruction in rat antigen-induced arthritis. Arthritis Res Ther. 2005;7:R677–686. doi: 10.1186/ar1725. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 356.Rouschop KM, Roelofs JJ, Claessen N, da Costa Martins P, Zwaginga JJ, Pals ST, Weening JJ, Florquin S. Protection against renal ischemia reperfusion injury by CD44 disruption. J Am Soc Nephrol. 2005;16:2034–2043. doi: 10.1681/ASN.2005010054. [DOI] [PubMed] [Google Scholar]
  • 357.Saari H, Konttinen YT. Determination of synovial fluid hyaluronate concentration and polymerisation by high performance liquid chromatography. Ann Rheum Dis. 1989;48:565–570. doi: 10.1136/ard.48.7.565. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 358.Sahu S, Lynn WS. Hyaluronic acid in the pulmonary secretions of patients with asthma. Biochem J. 1978;173:565–568. doi: 10.1042/bj1730565. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 359.Salathe M, Forteza R, Conner GE. Post-secretory fate of host defence components in mucus. Novartis Found Symp. 2002;248:20–26. discussion 27–37, 277–282. [PubMed] [Google Scholar]
  • 360.Salmen S, Hoechstetter J, Kasbauer C, Paper DH, Bernhardt G, Buschauer A. Sulphated oligosaccharides as inhibitors of hyaluronidases from bovine testis, bee venom and Streptococcus agalactiae. Planta Med. 2005;71:727–732. doi: 10.1055/s-2005-871255. [DOI] [PubMed] [Google Scholar]
  • 361.Sanchez-Rodriguez A, Criado M, Flores O, Olveira-Martin A, Martin-Oterino JA, Esteller A. Correlation of high levels of hyaluronan and cytokines (IL-1beta, IL-6, and TGF-beta) in ascitic fluid of cirrhotic patients. Dig Dis Sci. 2000;45:2229–2232. doi: 10.1023/a:1026648805129. [DOI] [PubMed] [Google Scholar]
  • 362.Sano N, Kitazawa K, Sugisaki T. Localization and roles of CD44, hyaluronic acid and osteopontin in IgA nephropathy. Nephron. 2001;89:416–421. doi: 10.1159/000046113. [DOI] [PubMed] [Google Scholar]
  • 363.Sant GR, Kempuraj D, Marchand JE, Theoharides TC. The mast cell in interstitial cystitis: role in pathophysiology and pathogenesis. Urology. 2007;69:34–40. doi: 10.1016/j.urology.2006.08.1109. [DOI] [PubMed] [Google Scholar]
  • 364.Sato H, Takahashi T, Ide H, Fukushima T, Tabata M, Sekine F, Kobayashi K, Negishi M, Niwa Y. Antioxidant activity of synovial fluid, hyaluronic acid, and two subcomponents of hyaluronic acid. Synovial fluid scavenging effect is enhanced in rheumatoid arthritis patients. Arthritis Rheum. 1988;31:63–71. doi: 10.1002/art.1780310110. [DOI] [PubMed] [Google Scholar]
  • 365.Savani RC, Cao G, Pooler PM, Zaman A, Zhou Z, DeLisser HM. Differential involvement of the hyaluronan (HA) receptors CD44 and receptor for HA-mediated motility in endothelial cell function and angiogenesis. J Biol Chem. 2001;276:36770–36778. doi: 10.1074/jbc.M102273200. [DOI] [PubMed] [Google Scholar]
  • 366.Savani RC, Wang C, Yang B, Zhang S, Kinsella MG, Wight TN, Stern R, Nance DM, Turley EA. Migration of bovine aortic smooth muscle cells after wounding injury. The role of hyaluronan and RHAMM. J Clin Invest. 1995;95:1158–1168. doi: 10.1172/JCI117764. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 367.Sbaa-Ketata E, Courel MN, Delpech B, Vannier JP. Hyaluronan-derived oligosaccharides enhance SDF-1-dependent chemotactic effect on peripheral blood hematopoietic CD34(+) cells. Stem Cells. 2002;20:585–587. doi: 10.1634/stemcells.20-6-585. [DOI] [PubMed] [Google Scholar]
  • 368.Scheibner KA, Lutz MA, Boodoo S, Fenton MJ, Powell JD, Horton MR. Hyaluronan fragments act as an endogenous danger signal by engaging TLR2. J Immunol. 2006;177:1272–1281. doi: 10.4049/jimmunol.177.2.1272. [DOI] [PubMed] [Google Scholar]
  • 369.Schenck P, Schneider S, Miehlke R, Prehm P. Synthesis and degradation of hyaluronate by synovia from patients with rheumatoid arthritis. J Rheumatol. 1995;22:400–405. [PubMed] [Google Scholar]
  • 370.Schledzewski K, Falkowski M, Moldenhauer G, Metharom P, Kzhyshkowska J, Ganss R, Demory A, Falkowska-Hansen B, Kurzen H, Ugurel S, Geginat G, Arnold B, Goerdt S. Lymphatic endothelium-specific hyaluronan receptor LYVE-1 is expressed by stabilin-1+, F4/80+, CD11b+ macrophages in malignant tumours and wound healing tissue in vivo and in bone marrow cultures in vitro: implications for the assessment of lymphangiogenesis. J Pathol. 2006;209:67–77. doi: 10.1002/path.1942. [DOI] [PubMed] [Google Scholar]
  • 371.Schmits R, Filmus J, Gerwin N, Senaldi G, Kiefer F, Kundig T, Wakeham A, Shahinian A, Catzavelos C, Rak J, Furlonger C, Zakarian A, Simard JJ, Ohashi PS, Paige CJ, Gutierrez-Ramos JC, Mak TW. CD44 regulates hematopoietic progenitor distribution, granuloma formation, and tumorigenicity. Blood. 1997;90:2217–2233. [PubMed] [Google Scholar]
  • 372.Schober A, Karshovska E, Zernecke A, Weber C. SDF-1alpha-mediated tissue repair by stem cells: a promising tool in cardiovascular medicine? Trends Cardiovasc Med. 2006;16:103–108. doi: 10.1016/j.tcm.2006.01.006. [DOI] [PubMed] [Google Scholar]
  • 373.Sconocchia G, Campagnano L, Adorno D, Iacona A, Cococcetta NY, Boffo V, Amadori S, Casciani CU. CD44 ligation on peripheral blood polymorphonuclear cells induces interleukin-6 production. Blood. 2001;97:3621–3627. doi: 10.1182/blood.v97.11.3621. [DOI] [PubMed] [Google Scholar]
  • 374.Scott JE, Heatley F. Hyaluronan forms specific stable tertiary structures in aqueous solution: a 13C NMR study. Proc Natl Acad Sci U S A. 1999;96:4850–4855. doi: 10.1073/pnas.96.9.4850. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 375.Scuri M, Abraham WM. Hyaluronan blocks human neutrophil elastase (HNE)-induced airway responses in sheep. Pulm Pharmacol Ther. 2003;16:335–340. doi: 10.1016/S1094-5539(03)00089-0. [DOI] [PubMed] [Google Scholar]
  • 376.Seike M, Ikeda M, Matsumoto M, Hamada R, Takeya M, Kodama H. Hyaluronan forms complexes with low density lipoprotein while also inducing foam cell infiltration in the dermis. J Dermatol Sci. 2006;41:197–204. doi: 10.1016/j.jdermsci.2005.10.008. [DOI] [PubMed] [Google Scholar]
  • 377.Sengupta A, Banerjee B, Tyagi RK, Datta K. Golgi localization and dynamics of hyaluronan binding protein 1 (HABP1/p32/C1QBP) during the cell cycle. Cell Res. 2005;15:183–186. doi: 10.1038/sj.cr.7290284. [DOI] [PubMed] [Google Scholar]
  • 378.Seyfried NT, McVey GF, Almond A, Mahoney DJ, Dudhia J, Day AJ. Expression and purification of functionally active hyaluronan-binding domains from human cartilage link protein, aggrecan and versican: formation of ternary complexes with defined hyaluronan oligosaccharides. J Biol Chem. 2005;280:5435–5448. doi: 10.1074/jbc.M411297200. [DOI] [PubMed] [Google Scholar]
  • 379.Shen L, Zhuo L, Okumura A, Ishikawa T, Miyachi M, Owa Y, Ishizawa T, Sugiura N, Nagata Y, Nonami T, Kakumu S, Kimata K. The SHAP-hyaluronan complex in serum from patients with chronic liver diseases caused by hepatitis virus infection. Hepatol Res. 2006;34:178–186. doi: 10.1016/j.hepres.2005.12.008. [DOI] [PubMed] [Google Scholar]
  • 380.Shen ZJ, Esnault S, Malter JS. The peptidyl-prolyl isomerase Pin1 regulates the stability of granulocyte-macrophage colony-stimulating factor mRNA in activated eosinophils. Nat Immunol. 2005;6:1280–1287. doi: 10.1038/ni1266. [DOI] [PubMed] [Google Scholar]
  • 381.Sherman L, Sleeman J, Herrlich P, Ponta H. Hyaluronate receptors: key players in growth, differentiation, migration and tumor progression. Curr Opin Cell Biol. 1994;6:726–733. doi: 10.1016/0955-0674(94)90100-7. [DOI] [PubMed] [Google Scholar]
  • 382.Shirai M, Imanaka-Yoshida K, Schneider MD, Schwartz RJ, Morisaki T. T-box 2, a mediator of Bmp-Smad signaling, induced hyaluronan synthase 2 and Tgf{beta}2 expression and endocardial cushion formation. Proc Natl Acad Sci U S A. 2009 doi: 10.1073/pnas.0900635106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 383.Shyjan AM, Heldin P, Butcher EC, Yoshino T, Briskin MJ. Functional cloning of the cDNA for a human hyaluronan synthase. J Biol Chem. 1996;271:23395–23399. doi: 10.1074/jbc.271.38.23395. [DOI] [PubMed] [Google Scholar]
  • 384.Sibalic V, Fan X, Loffing J, Wuthrich RP. Upregulated renal tubular CD44, hyaluronan, and osteopontin in kdkd mice with interstitial nephritis. Nephrol Dial Transplant. 1997;12:1344–1353. doi: 10.1093/ndt/12.7.1344. [DOI] [PubMed] [Google Scholar]
  • 385.Siegelman MH, Stanescu D, Estess P. The CD44-initiated pathway of T-cell extravasation uses VLA-4 but not LFA-1 for firm adhesion. J Clin Invest. 2000;105:683–691. doi: 10.1172/JCI8692. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 386.Singleton PA, Dudek SM, Ma SF, Garcia JG. Transactivation of Sphingosine 1-Phosphate Receptors Is Essential for Vascular Barrier Regulation: NOVEL ROLE FOR HYALURONAN AND CD44 RECEPTOR FAMILY. J Biol Chem. 2006;281:34381–34393. doi: 10.1074/jbc.M603680200. [DOI] [PubMed] [Google Scholar]
  • 387.Singleton PA, Salgia R, Moreno-Vinasco L, Moitra J, Sammani S, Mirzapoiazova T, Garcia JG. CD44 regulates hepatocyte growth factor-mediated vascular integrity. Role of c-Met, Tiam1/Rac1, dynamin 2, and cortactin. J Biol Chem. 2007;282:30643–30657. doi: 10.1074/jbc.M702573200. [DOI] [PubMed] [Google Scholar]
  • 388.Slevin M, Krupinski J, Gaffney J, Matou S, West D, Delisser H, Savani RC, Kumar S. Hyaluronan-mediated angiogenesis in vascular disease: uncovering RHAMM and CD44 receptor signaling pathways. Matrix Biol. 2007;26:58–68. doi: 10.1016/j.matbio.2006.08.261. [DOI] [PubMed] [Google Scholar]
  • 389.Slevin M, Krupinski J, Kumar S, Gaffney J. Angiogenic oligosaccharides of hyaluronan induce protein tyrosine kinase activity in endothelial cells and activate a cytoplasmic signal transduction pathway resulting in proliferation. Lab Invest. 1998;78:987–1003. [PubMed] [Google Scholar]
  • 390.Slevin M, Kumar S, Gaffney J. Angiogenic oligosaccharides of hyaluronan induce multiple signaling pathways affecting vascular endothelial cell mitogenic and wound healing responses. J Biol Chem. 2002;277:41046–41059. doi: 10.1074/jbc.M109443200. [DOI] [PubMed] [Google Scholar]
  • 391.Soderberg M, Bjermer L, Hallgren R, Lundgren R. Increased hyaluronan (hyaluronic acid) levels in bronchoalveolar lavage fluid after histamine inhalation. Int Arch Allergy Appl Immunol. 1989;88:373–376. doi: 10.1159/000234719. [DOI] [PubMed] [Google Scholar]
  • 392.Soltes L, Mendichi R, Kogan G, Schiller J, Stankovska M, Arnhold J. Degradative action of reactive oxygen species on hyaluronan. Biomacromolecules. 2006;7:659–668. doi: 10.1021/bm050867v. [DOI] [PubMed] [Google Scholar]
  • 393.Song WD, Zhang AC, Pang YY, Liu LH, Zhao JY, Deng SH, Zhang SY. Fibronectin and hyaluronan in bronchoalveolar lavage fluid from young patients with chronic obstructive pulmonary diseases. Respiration. 1995;62:125–129. doi: 10.1159/000196406. [DOI] [PubMed] [Google Scholar]
  • 394.Spicer AP, Augustine ML, McDonald JA. Molecular cloning and characterization of a putative mouse hyaluronan synthase. J Biol Chem. 1996;271:23400–23406. doi: 10.1074/jbc.271.38.23400. [DOI] [PubMed] [Google Scholar]
  • 395.Spicer AP, McDonald JA. Characterization and molecular evolution of a vertebrate hyaluronan synthase gene family. J Biol Chem. 1998;273:1923–1932. doi: 10.1074/jbc.273.4.1923. [DOI] [PubMed] [Google Scholar]
  • 396.Spicer AP, Seldin MF, Olsen AS, Brown N, Wells DE, Doggett NA, Itano N, Kimata K, Inazawa J, McDonald JA. Chromosomal localization of the human and mouse hyaluronan synthase genes. Genomics. 1997;41:493–497. doi: 10.1006/geno.1997.4696. [DOI] [PubMed] [Google Scholar]
  • 397.Spicer AP, Tien JL, Joo A, Bowling RA., Jr Investigation of hyaluronan function in the mouse through targeted mutagenesis. Glycoconj J. 2002;19:341–345. doi: 10.1023/A:1025321105691. [DOI] [PubMed] [Google Scholar]
  • 398.Spicer AP, Tien JY. Hyaluronan and morphogenesis. Birth Defects Res C Embryo Today. 2004;72:89–108. doi: 10.1002/bdrc.20006. [DOI] [PubMed] [Google Scholar]
  • 399.Stenvinkel P, Heimburger O, Wang T, Lindholm B, Bergstrom J, Elinder CG. High serum hyaluronan indicates poor survival in renal replacement therapy. Am J Kidney Dis. 1999;34:1083–1088. doi: 10.1016/S0272-6386(99)70014-4. [DOI] [PubMed] [Google Scholar]
  • 400.Stern R, Asari AA, Sugahara KN. Hyaluronan fragments: an information-rich system. Eur J Cell Biol. 2006;85:699–715. doi: 10.1016/j.ejcb.2006.05.009. [DOI] [PubMed] [Google Scholar]
  • 401.Stoop R, Gal I, Glant TT, McNeish JD, Mikecz K. Trafficking of CD44-deficient murine lymphocytes under normal and inflammatory conditions. Eur J Immunol. 2002;32:2532–2542. doi: 10.1002/1521-4141(200209)32:9<2532::AID-IMMU2532>3.0.CO;2-A. [DOI] [PubMed] [Google Scholar]
  • 402.Stuhlmeier KM, Pollaschek C. Differential effect of transforming growth factor beta (TGF-beta) on the genes encoding hyaluronan synthases and utilization of the p38 MAPK pathway in TGF-beta-induced hyaluronan synthase 1 activation. J Biol Chem. 2004;279:8753–8760. doi: 10.1074/jbc.M303945200. [DOI] [PubMed] [Google Scholar]
  • 403.Stylli SS, Kaye AH, Novak U. Induction of CD44 expression in stab wounds of the brain: long term persistence of CD44 expression. J Clin Neurosci. 2000;7:137–140. doi: 10.1054/jocn.1999.0187. [DOI] [PubMed] [Google Scholar]
  • 404.Sugahara KN, Hirata T, Hayasaka H, Stern R, Murai T, Miyasaka M. Tumor cells enhance their own CD44 cleavage and motility by generating hyaluronan fragments. J Biol Chem. 2006;281:5861–5868. doi: 10.1074/jbc.M506740200. [DOI] [PubMed] [Google Scholar]
  • 405.Sugahara KN, Murai T, Nishinakamura H, Kawashima H, Saya H, Miyasaka M. Hyaluronan oligosaccharides induce CD44 cleavage and promote cell migration in CD44-expressing tumor cells. J Biol Chem. 2003;278:32259–32265. doi: 10.1074/jbc.M300347200. [DOI] [PubMed] [Google Scholar]
  • 406.Sun LK, Beck-Schimmer B, Oertli B, Wuthrich RP. Hyaluronan-induced cyclooxygenase-2 expression promotes thromboxane A2 production by renal cells. Kidney Int. 2001;59:190–196. doi: 10.1046/j.1523-1755.2001.00479.x. [DOI] [PubMed] [Google Scholar]
  • 407.Sussmann M, Sarbia M, Meyer-Kirchrath J, Nusing RM, Schror K, Fischer JW. Induction of hyaluronic acid synthase 2 (HAS2) in human vascular smooth muscle cells by vasodilatory prostaglandins. Circ Res. 2004;94:592–600. doi: 10.1161/01.RES.0000119169.87429.A0. [DOI] [PubMed] [Google Scholar]
  • 408.Svee K, White J, Vaillant P, Jessurun J, Roongta U, Krumwiede M, Johnson D, Henke C. Acute lung injury fibroblast migration and invasion of a fibrin matrix is mediated by CD44. J Clin Invest. 1996;98:1713–1727. doi: 10.1172/JCI118970. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 409.Szanto S, Bardos T, Gal I, Glant TT, Mikecz K. Enhanced neutrophil extravasation and rapid progression of proteoglycan-induced arthritis in TSG-6-knockout mice. Arthritis Rheum. 2004;50:3012–3022. doi: 10.1002/art.20655. [DOI] [PubMed] [Google Scholar]
  • 410.Tabata T, Mine S, Okada Y, Tanaka Y. Low molecular weight hyaluronan increases the Uptaking of oxidized LDL into monocytes. Endocr J. 2007;54:685–693. doi: 10.1507/endocrj.k05-120. [DOI] [PubMed] [Google Scholar]
  • 411.Takahashi Y, Li L, Kamiryo M, Asteriou T, Moustakas A, Yamashita H, Heldin P. Hyaluronan fragments induce endothelial cell differentiation in a CD44- and CXCL1/GRO1-dependent manner. J Biol Chem. 2005;280:24195–24204. doi: 10.1074/jbc.M411913200. [DOI] [PubMed] [Google Scholar]
  • 412.Takazoe K, Tesch GH, Hill PA, Hurst LA, Jun Z, Lan HY, Atkins RC, Nikolic-Paterson DJ. CD44-mediated neutrophil apoptosis in the rat. Kidney Int. 2000;58:1920–1930. doi: 10.1111/j.1523-1755.2000.00364.x. [DOI] [PubMed] [Google Scholar]
  • 413.Takeda K, Kaisho T, Akira S. Toll-like receptors. Annu Rev Immunol. 2003;21:335–376. doi: 10.1146/annurev.immunol.21.120601.141126. [DOI] [PubMed] [Google Scholar]
  • 414.Takeuchi O, Hoshino K, Kawai T, Sanjo H, Takada H, Ogawa T, Takeda K, Akira S. Differential roles of TLR2 and TLR4 in recognition of gram-negative and gram-positive bacterial cell wall components. Immunity. 1999;11:443–451. doi: 10.1016/s1074-7613(00)80119-3. [DOI] [PubMed] [Google Scholar]
  • 415.Tammi R, Pasonen-Seppanen S, Kolehmainen E, Tammi M. Hyaluronan synthase induction and hyaluronan accumulation in mouse epidermis following skin injury. J Invest Dermatol. 2005;124:898–905. doi: 10.1111/j.0022-202X.2005.23697.x. [DOI] [PubMed] [Google Scholar]
  • 416.Tammi R, Ripellino JA, Margolis RU, Tammi M. Localization of epidermal hyaluronic acid using the hyaluronate binding region of cartilage proteoglycan as a specific probe. J Invest Dermatol. 1988;90:412–414. doi: 10.1111/1523-1747.ep12456530. [DOI] [PubMed] [Google Scholar]
  • 417.Taylor KR, Trowbridge JM, Rudisill JA, Termeer CC, Simon JC, Gallo RL. Hyaluronan fragments stimulate endothelial recognition of injury through TLR4. J Biol Chem. 2004;279:17079–17084. doi: 10.1074/jbc.M310859200. [DOI] [PubMed] [Google Scholar]
  • 418.Taylor KR, Yamasaki K, Radek KA, Di Nardo A, Goodarzi H, Golenbock D, Beutler B, Gallo RL. Recognition of hyaluronan released in sterile injury involves a unique receptor complex dependent on Toll-like receptor 4, CD44, and MD-2. J Biol Chem. 2007;282:18265–18275. doi: 10.1074/jbc.M606352200. [DOI] [PubMed] [Google Scholar]
  • 419.Teder P, Heldin P. Mechanism of impaired local hyaluronan turnover in bleomycin-induced lung injury in rat. Am J Respir Cell Mol Biol. 1997;17:376–385. doi: 10.1165/ajrcmb.17.3.2698. [DOI] [PubMed] [Google Scholar]
  • 420.Teder P, Nettelbladt O, Heldin P. Characterization of the mechanism involved in bleomycin-induced increased hyaluronan production in rat lung. Am J Respir Cell Mol Biol. 1995;12:181–189. doi: 10.1165/ajrcmb.12.2.7532420. [DOI] [PubMed] [Google Scholar]
  • 421.Teder P, Vandivier RW, Jiang D, Liang J, Cohn L, Pure E, Henson PM, Noble PW. Resolution of lung inflammation by CD44. Science. 2002;296:155–158. doi: 10.1126/science.1069659. [DOI] [PubMed] [Google Scholar]
  • 422.Teriete P, Banerji S, Noble M, Blundell CD, Wright AJ, Pickford AR, Lowe E, Mahoney DJ, Tammi MI, Kahmann JD, Campbell ID, Day AJ, Jackson DG. Structure of the regulatory hyaluronan binding domain in the inflammatory leukocyte homing receptor CD44. Mol Cell. 2004;13:483–496. doi: 10.1016/s1097-2765(04)00080-2. [DOI] [PubMed] [Google Scholar]
  • 423.Termeer C, Benedix F, Sleeman J, Fieber C, Voith U, Ahrens T, Miyake K, Freudenberg M, Galanos C, Simon JC. Oligosaccharides of Hyaluronan activate dendritic cells via toll-like receptor 4. J Exp Med. 2002;195:99–111. doi: 10.1084/jem.20001858. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 424.Termeer C, Sleeman JP, Simon JC. Hyaluronan--magic glue for the regulation of the immune response? Trends Immunol. 2003;24:112–114. doi: 10.1016/s1471-4906(03)00029-2. [DOI] [PubMed] [Google Scholar]
  • 425.Termeer CC, Hennies J, Voith U, Ahrens T, Weiss JM, Prehm P, Simon JC. Oligosaccharides of hyaluronan are potent activators of dendritic cells. J Immunol. 2000;165:1863–1870. doi: 10.4049/jimmunol.165.4.1863. [DOI] [PubMed] [Google Scholar]
  • 426.Tesar BM, Jiang D, Liang J, Palmer SM, Noble PW, Goldstein DR. The role of hyaluronan degradation products as innate alloimmune agonists. Am J Transplant. 2006;6:2622–2635. doi: 10.1111/j.1600-6143.2006.01537.x. [DOI] [PubMed] [Google Scholar]
  • 427.Thakur SC, Kumar V, Ghosh I, Bharadwaj A, Datta K. Appearance of hyaluronan binding protein 1 proprotein in pachytene spermatocytes and round spermatids correlates with spermatogenesis. J Androl. 2006;27:604–610. doi: 10.2164/jandrol.05142. [DOI] [PubMed] [Google Scholar]
  • 428.Tian WM, Hou SP, Ma J, Zhang CL, Xu QY, Lee IS, Li HD, Spector M, Cui FZ. Hyaluronic acid-poly-D-lysine-based three-dimensional hydrogel for traumatic brain injury. Tissue Eng. 2005;11:513–525. doi: 10.1089/ten.2005.11.513. [DOI] [PubMed] [Google Scholar]
  • 429.Tien JY, Spicer AP. Three vertebrate hyaluronan synthases are expressed during mouse development in distinct spatial and temporal patterns. Dev Dyn. 2005;233:130–141. doi: 10.1002/dvdy.20328. [DOI] [PubMed] [Google Scholar]
  • 430.Tolg C, Hamilton SR, Nakrieko KA, Kooshesh F, Walton P, McCarthy JB, Bissell MJ, Turley EA. Rhamm−/− fibroblasts are defective in CD44-mediated ERK1,2 motogenic signaling, leading to defective skin wound repair. J Cell Biol. 2006;175:1017–1028. doi: 10.1083/jcb.200511027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 431.Tolg C, Poon R, Fodde R, Turley EA, Alman BA. Genetic deletion of receptor for hyaluronan-mediated motility (Rhamm) attenuates the formation of aggressive fibromatosis (desmoid tumor) Oncogene. 2003;22:6873–6882. doi: 10.1038/sj.onc.1206811. [DOI] [PubMed] [Google Scholar]
  • 432.Toole BP. Hyaluronan is not just a goo! J Clin Invest. 2000;106:335–336. doi: 10.1172/JCI10706. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 433.Toole BP. Hyaluronan: from extracellular glue to pericellular cue. Nat Rev Cancer. 2004;4:528–539. doi: 10.1038/nrc1391. [DOI] [PubMed] [Google Scholar]
  • 434.Toole BP, Zoltan-Jones A, Misra S, Ghatak S. Hyaluronan: a critical component of epithelial-mesenchymal and epithelial-carcinoma transitions. Cells Tissues Organs. 2005;179:66–72. doi: 10.1159/000084510. [DOI] [PubMed] [Google Scholar]
  • 435.Torii S, Bashey R. High content of hyaluronic acid in normal human heart valves. Nature. 1966;209:506–507. doi: 10.1038/209506b0. [DOI] [PubMed] [Google Scholar]
  • 436.Tsujimura S, Saito K, Kohno K, Tanaka Y. Fragmented hyaluronan induces transcriptional up-regulation of the multidrug resistance-1 gene in CD4+ T cells. J Biol Chem. 2006 Dec 8;281:38089–38097. doi: 10.1074/jbc.M601030200. [DOI] [PubMed] [Google Scholar]
  • 437.Tsukahara S, Ikeda R, Goto S, Yoshida K, Mitsumori R, Sakamoto Y, Tajima A, Yokoyama T, Toh S, Furukawa K, Inoue I. Tumour necrosis factor alpha-stimulated gene-6 inhibits osteoblastic differentiation of human mesenchymal stem cells induced by osteogenic differentiation medium and BMP-2. Biochem J. 2006;398:595–603. doi: 10.1042/BJ20060027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 438.Turino GM, Cantor JO. Hyaluronan in respiratory injury and repair. Am J Respir Crit Care Med. 2003;167:1169–1175. doi: 10.1164/rccm.200205-449PP. [DOI] [PubMed] [Google Scholar]
  • 439.Turley EA. The role of a cell-associated hyaluronan-binding protein in fibroblast behaviour. Ciba Found Symp. 1989;143:121–133. doi: 10.1002/9780470513774.ch8. discussion 133–127, 281–125. [DOI] [PubMed] [Google Scholar]
  • 440.Turney JH, Davison AM, Forbes MA, Cooper EH. Hyaluronic acid in end-stage renal failure treated by haemodialysis: clinical correlates and implications. Nephrol Dial Transplant. 1991;6:566–570. doi: 10.1093/ndt/6.8.566. [DOI] [PubMed] [Google Scholar]
  • 441.Tzircotis G, Thorne RF, Isacke CM. Chemotaxis towards hyaluronan is dependent on CD44 expression and modulated by cell type variation in CD44-hyaluronan binding. J Cell Sci. 2005;118:5119–5128. doi: 10.1242/jcs.02629. [DOI] [PubMed] [Google Scholar]
  • 442.Underhill DM, Ozinsky A, Hajjar AM, Stevens A, Wilson CB, Bassetti M, Aderem A. The Toll-like receptor 2 is recruited to macrophage phagosomes and discriminates between pathogens. Nature. 1999;401:811–815. doi: 10.1038/44605. [DOI] [PubMed] [Google Scholar]
  • 443.Vazquez JR, Short B, Findlow AH, Nixon BP, Boulton AJ, Armstrong DG. Outcomes of hyaluronan therapy in diabetic foot wounds. Diabetes Res Clin Pract. 2003;59:123–127. doi: 10.1016/s0168-8227(02)00197-3. [DOI] [PubMed] [Google Scholar]
  • 444.Ventura C, Branzi A. Autocrine and intracrine signaling for cardiogenesis in embryonic stem cells: a clue for the development of novel differentiating agents. Handb Exp Pharmacol. 2006:123–146. [PubMed] [Google Scholar]
  • 445.Ventura C, Cantoni S, Bianchi F, Lionetti V, Cavallini C, Scarlata I, Foroni L, Maioli M, Bonsi L, Alviano F, Fossati V, Bagnara GP, Pasquinelli G, Recchia FA, Perbellini A. Hyaluronan mixed esters of butyric and retinoic Acid drive cardiac and endothelial fate in term placenta human mesenchymal stem cells and enhance cardiac repair in infarcted rat hearts. J Biol Chem. 2007;282:14243–14252. doi: 10.1074/jbc.M609350200. [DOI] [PubMed] [Google Scholar]
  • 446.Ventura C, Maioli M, Asara Y, Santoni D, Scarlata I, Cantoni S, Perbellini A. Butyric and retinoic mixed ester of hyaluronan. A novel differentiating glycoconjugate affording a high throughput of cardiogenesis in embryonic stem cells. J Biol Chem. 2004;279:23574–23579. doi: 10.1074/jbc.M401869200. [DOI] [PubMed] [Google Scholar]
  • 447.Vetr H, Gebhard W. Structure of the human alpha 1-microglobulin-bikunin gene. Biol Chem Hoppe Seyler. 1990;371:1185–1196. doi: 10.1515/bchm3.1990.371.2.1185. [DOI] [PubMed] [Google Scholar]
  • 448.Vigetti D, Viola M, Karousou E, Rizzi M, Moretto P, Genasetti A, Clerici M, Hascall VC, De Luca G, Passi A. Hyaluronan-CD44-ERK1/2 Regulate Human Aortic Smooth Muscle Cell Motility during Aging. J Biol Chem. 2008;283:4448–4458. doi: 10.1074/jbc.M709051200. [DOI] [PubMed] [Google Scholar]
  • 449.Vignola AM, Chanez P, Campbell AM, Souques F, Lebel B, Enander I, Bousquet J. Airway inflammation in mild intermittent and in persistent asthma. Am J Respir Crit Care Med. 1998;157:403–409. doi: 10.1164/ajrccm.157.2.96-08040. [DOI] [PubMed] [Google Scholar]
  • 450.Vivers S, Dransfield I, Hart SP. Role of macrophage CD44 in the disposal of inflammatory cell corpses. Clin Sci (Lond) 2002;103:441–449. doi: 10.1042/cs1030441. [DOI] [PubMed] [Google Scholar]
  • 451.Voelcker V, Gebhardt C, Averbeck M, Saalbach A, Wolf V, Weih F, Sleeman J, Anderegg U, Simon J. Hyaluronan fragments induce cytokine and metalloprotease upregulation in human melanoma cells in part by signalling via TLR4. Exp Dermatol. 2008;17:100–107. doi: 10.1111/j.1600-0625.2007.00638.x. [DOI] [PubMed] [Google Scholar]
  • 452.Wang H, Zhan Y, Xu L, Feuerstein GZ, Wang X. Use of suppression subtractive hybridization for differential gene expression in stroke: discovery of CD44 gene expression and localization in permanent focal stroke in rats. Stroke. 2001;32:1020–1027. doi: 10.1161/01.str.32.4.1020. [DOI] [PubMed] [Google Scholar]
  • 453.Wang MJ, Kuo JS, Lee WW, Huang HY, Chen WF, Lin SZ. Translational event mediates differential production of tumor necrosis factor-alpha in hyaluronan-stimulated microglia and macrophages. J Neurochem. 2006;97:857–871. doi: 10.1111/j.1471-4159.2006.03776.x. [DOI] [PubMed] [Google Scholar]
  • 454.Wang Q, Teder P, Judd NP, Noble PW, Doerschuk CM. CD44 deficiency leads to enhanced neutrophil migration and lung injury in Escherichia coli pneumonia in mice. Am J Pathol. 2002;161:2219–2228. doi: 10.1016/S0002-9440(10)64498-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 455.Watanabe K, Yamaguchi Y. Molecular identification of a putative human hyaluronan synthase. J Biol Chem. 1996;271:22945–22948. doi: 10.1074/jbc.271.38.22945. [DOI] [PubMed] [Google Scholar]
  • 456.Webber J, Jenkins RH, Meran S, Phillips A, Steadman R. Modulation of TGFbeta1-dependent myofibroblast differentiation by hyaluronan. Am J Pathol. 2009;175:148–160. doi: 10.2353/ajpath.2009.080837. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 457.Webber J, Meran S, Steadman R, Phillips A. Hyaluronan orchestrates transforming growth factor-beta1-dependent maintenance of myofibroblast phenotype. J Biol Chem. 2009;284:9083–9092. doi: 10.1074/jbc.M806989200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 458.Weigel JA, Raymond RC, McGary C, Singh A, Weigel PH. A blocking antibody to the hyaluronan receptor for endocytosis (HARE) inhibits hyaluronan clearance by perfused liver. J Biol Chem. 2003;278:9808–9812. doi: 10.1074/jbc.m211462200. [DOI] [PubMed] [Google Scholar]
  • 459.Weigel JA, Weigel PH. Characterization of the recombinant rat 175-kDa hyaluronan receptor for endocytosis (HARE) J Biol Chem. 2003;278:42802–42811. doi: 10.1074/jbc.M307201200. [DOI] [PubMed] [Google Scholar]
  • 460.Weigel PH. Functional characteristics and catalytic mechanisms of the bacterial hyaluronan synthases. IUBMB Life. 2002;54:201–211. doi: 10.1080/15216540214931. [DOI] [PubMed] [Google Scholar]
  • 461.Weigel PH, Hascall VC, Tammi M. Hyaluronan synthases. J Biol Chem. 1997;272:13997–14000. doi: 10.1074/jbc.272.22.13997. [DOI] [PubMed] [Google Scholar]
  • 462.Weiss L, Slavin S, Reich S, Cohen P, Shuster S, Stern R, Kaganovsky E, Okon E, Rubinstein AM, Naor D. Induction of resistance to diabetes in non-obese diabetic mice by targeting CD44 with a specific monoclonal antibody. Proc Natl Acad Sci U S A. 2000;97:285–290. doi: 10.1073/pnas.97.1.285. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 463.Weissmann B, Meyer K. The structure of hyalobiuronic acid and of hyaluronic acid from umbilical cord. J Am Chem Soc. 1954;76:1753–1757. [Google Scholar]
  • 464.Weissmann B, Meyer K, Sampson P, Linker A. Isolation of oligosaccharides enzymatically produced from hyaluronic acid. J Biol Chem. 1954;208:417–429. [PubMed] [Google Scholar]
  • 465.West DC, Hampson IN, Arnold F, Kumar S. Angiogenesis induced by degradation products of hyaluronic acid. Science. 1985;228:1324–1326. doi: 10.1126/science.2408340. [DOI] [PubMed] [Google Scholar]
  • 466.West DC, Yaqoob M. Serum hyaluronan levels follow disease activity in vasculitis. Clin Nephrol. 1997;48:9–15. [PubMed] [Google Scholar]
  • 467.Westergren-Thorsson G, Chakir J, Lafreniere-Allard MJ, Boulet LP, Tremblay GM. Correlation between airway responsiveness and proteoglycan production by bronchial fibroblasts from normal and asthmatic subjects. Int J Biochem Cell Biol. 2002;34:1256–1267. doi: 10.1016/s1357-2725(02)00058-4. [DOI] [PubMed] [Google Scholar]
  • 468.Westergren-Thorsson G, Sime P, Jordana M, Gauldie J, Sarnstrand B, Malmstrom A. Lung fibroblast clones from normal and fibrotic subjects differ in hyaluronan and decorin production and rate of proliferation. Int J Biochem Cell Biol. 2004;36:1573–1584. doi: 10.1016/j.biocel.2004.01.009. [DOI] [PubMed] [Google Scholar]
  • 469.Wheatley SC, Isacke CM. Induction of a hyaluronan receptor, CD44, during embryonal carcinoma and embryonic stem cell differentiation. Cell Adhes Commun. 1995;3:217–230. doi: 10.3109/15419069509081288. [DOI] [PubMed] [Google Scholar]
  • 470.Wight TN, Merrilees MJ. Proteoglycans in atherosclerosis and restenosis: key roles for versican. Circ Res. 2004;94:1158–1167. doi: 10.1161/01.RES.0000126921.29919.51. [DOI] [PubMed] [Google Scholar]
  • 471.Wilkinson TS, Bressler SL, Evanko SP, Braun KR, Wight TN. Overexpression of hyaluronan synthases alters vascular smooth muscle cell phenotype and promotes monocyte adhesion. J Cell Physiol. 2006;206:378–385. doi: 10.1002/jcp.20468. [DOI] [PubMed] [Google Scholar]
  • 472.Wilkinson TS, Potter-Perigo S, Tsoi C, Altman LC, Wight TN. Pro- and anti-inflammatory factors cooperate to control hyaluronan synthesis in lung fibroblasts. Am J Respir Cell Mol Biol. 2004;31:92–99. doi: 10.1165/rcmb.2003-0380OC. [DOI] [PubMed] [Google Scholar]
  • 473.Wisniewski HG, Hua JC, Poppers DM, Naime D, Vilcek J, Cronstein BN. TNF/IL-1-inducible protein TSG-6 potentiates plasmin inhibition by inter-alpha-inhibitor and exerts a strong anti-inflammatory effect in vivo. J Immunol. 1996;156:1609–1615. [PubMed] [Google Scholar]
  • 474.Wisniewski HG, Maier R, Lotz M, Lee S, Klampfer L, Lee TH, Vilcek J. TSG-6: a TNF-, IL-1-, and LPS-inducible secreted glycoprotein associated with arthritis. J Immunol. 1993;151:6593–6601. [PubMed] [Google Scholar]
  • 475.Woodrow G, Turney JH, Davison AM, Cooper EH. Serum hyaluronan concentrations predict survival in patients with chronic renal failure on maintenance haemodialysis. Nephrol Dial Transplant. 1996;11:98–100. [PubMed] [Google Scholar]
  • 476.Wrobel T, Dziegiel P, Mazur G, Zabel M, Kuliczkowski K, Szuba A. LYVE-1 expression on high endothelial venules (HEVs) of lymph nodes. Lymphology. 2005;38:107–110. [PubMed] [Google Scholar]
  • 477.Wu GD, Nolta JA, Jin YS, Barr ML, Yu H, Starnes VA, Cramer DV. Migration of mesenchymal stem cells to heart allografts during chronic rejection. Transplantation. 2003;75:679–685. doi: 10.1097/01.TP.0000048488.35010.95. [DOI] [PubMed] [Google Scholar]
  • 478.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. doi: 10.1172/JCI31008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 479.Xu H, Edwards J, Banerji S, Prevo R, Jackson DG, Athanasou NA. Distribution of lymphatic vessels in normal and arthritic human synovial tissues. Ann Rheum Dis. 2003;62:1227–1229. doi: 10.1136/ard.2003.005876. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 480.Yamada H, Fredette B, Shitara K, Hagihara K, Miura R, Ranscht B, Stallcup WB, Yamaguchi Y. The brain chondroitin sulfate proteoglycan brevican associates with astrocytes ensheathing cerebellar glomeruli and inhibits neurite outgrowth from granule neurons. J Neurosci. 1997;17:7784–7795. doi: 10.1523/JNEUROSCI.17-20-07784.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 481.Yang B, Yang BL, Savani RC, Turley EA. Identification of a common hyaluronan binding motif in the hyaluronan binding proteins RHAMM, CD44 and link protein. Embo J. 1994;13:286–296. doi: 10.1002/j.1460-2075.1994.tb06261.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 482.Yang B, Zhang L, Turley EA. Identification of two hyaluronan-binding domains in the hyaluronan receptor RHAMM. J Biol Chem. 1993;268:8617–8623. [PubMed] [Google Scholar]
  • 483.Yasuda T, Ribeiro LG, Maroko PR. Effect of hyaluronidase on experimental cerebral infarct size and mortality. Lab Invest. 1982;46:400–404. [PubMed] [Google Scholar]
  • 484.Ye L, Mora R, Akhayani N, Haudenschild CC, Liau G. Growth factor and cytokine-regulated hyaluronan-binding protein TSG-6 is localized to the injury-induced rat neointima and confers enhanced growth in vascular smooth muscle cells. Circ Res. 1997;81:289–296. doi: 10.1161/01.res.81.3.289. [DOI] [PubMed] [Google Scholar]
  • 485.Yingprasertchai S, Bunyasrisawat S, Ratanabanangkoon K. Hyaluronidase inhibitors (sodium cromoglycate and sodium auro-thiomalate) reduce the local tissue damage and prolong the survival time of mice injected with Naja kaouthia and Calloselasma rhodostoma venoms. Toxicon. 2003;42:635–646. doi: 10.1016/j.toxicon.2003.09.001. [DOI] [PubMed] [Google Scholar]
  • 486.Yingsung W, Zhuo L, Morgelin M, Yoneda M, Kida D, Watanabe H, Ishiguro N, Iwata H, Kimata K. Molecular heterogeneity of the SHAP-hyaluronan complex. Isolation and characterization of the complex in synovial fluid from patients with rheumatoid arthritis. J Biol Chem. 2003;278:32710–32718. doi: 10.1074/jbc.M303658200. [DOI] [PubMed] [Google Scholar]
  • 487.Yoneda M, Suzuki S, Kimata K. Hyaluronic acid associated with the surfaces of cultured fibroblasts is linked to a serum-derived 85-kDa protein. J Biol Chem. 1990;265:5247–5257. [PubMed] [Google Scholar]
  • 488.Yoshida M, Itano N, Yamada Y, Kimata K. In vitro synthesis of hyaluronan by a single protein derived from mouse HAS1 gene and characterization of amino acid residues essential for the activity. J Biol Chem. 2000;275:497–506. doi: 10.1074/jbc.275.1.497. [DOI] [PubMed] [Google Scholar]
  • 489.Yung S, Thomas GJ, Davies M. Induction of hyaluronan metabolism after mechanical injury of human peritoneal mesothelial cells in vitro. Kidney Int. 2000;58:1953–1962. doi: 10.1111/j.1523-1755.2000.00367.x. [DOI] [PubMed] [Google Scholar]
  • 490.Yung S, Tsang RC, Leung JK, Chan TM. Increased mesangial cell hyaluronan expression in lupus nephritis is mediated by anti-DNA antibody-induced IL-1beta. Kidney Int. 2006;69:272–280. doi: 10.1038/sj.ki.5000042. [DOI] [PubMed] [Google Scholar]
  • 491.Zaman A, Cui Z, Foley JP, Zhao H, Grimm PC, Delisser HM, Savani RC. Expression and role of the hyaluronan receptor RHAMM in inflammation after bleomycin injury. Am J Respir Cell Mol Biol. 2005;33:447–454. doi: 10.1165/rcmb.2004-0333OC. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 492.Zhang X, Shan P, Jiang G, Cohn L, Lee PJ. Toll-like receptor 4 deficiency causes pulmonary emphysema. J Clin Invest. 2006;116:3050–3059. doi: 10.1172/JCI28139. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 493.Zhang X, Shan P, Qureshi S, Homer R, Medzhitov R, Noble PW, Lee PJ. Cutting Edge: TLR4 Deficiency Confers Susceptibility to Lethal Oxidant Lung Injury. J Immunol. 2005;175:4834–4838. doi: 10.4049/jimmunol.175.8.4834. [DOI] [PubMed] [Google Scholar]
  • 494.Zhao M, Yoneda M, Ohashi Y, Kurono S, Iwata H, Ohnuki Y, Kimata K. Evidence for the covalent binding of SHAP, heavy chains of inter-alpha-trypsin inhibitor, to hyaluronan. J Biol Chem. 1995;270:26657–26663. doi: 10.1074/jbc.270.44.26657. [DOI] [PubMed] [Google Scholar]
  • 495.Zhou B, McGary CT, Weigel JA, Saxena A, Weigel PH. Purification and molecular identification of the human hyaluronan receptor for endocytosis. Glycobiology. 2003;13:339–349. doi: 10.1093/glycob/cwg029. [DOI] [PubMed] [Google Scholar]
  • 496.Zhou B, Oka JA, Singh A, Weigel PH. Purification and subunit characterization of the rat liver endocytic hyaluronan receptor. J Biol Chem. 1999;274:33831–33834. doi: 10.1074/jbc.274.48.33831. [DOI] [PubMed] [Google Scholar]
  • 497.Zhou B, Weigel JA, Fauss L, Weigel PH. Identification of the hyaluronan receptor for endocytosis (HARE) J Biol Chem. 2000;275:37733–37741. doi: 10.1074/jbc.M003030200. [DOI] [PubMed] [Google Scholar]
  • 498.Zhou B, Weigel JA, Saxena A, Weigel PH. Molecular cloning and functional expression of the rat 175-kDa hyaluronan receptor for endocytosis. Mol Biol Cell. 2002;13:2853–2868. doi: 10.1091/mbc.02-03-0048. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 499.Zhou XH, Brakebusch C, Matthies H, Oohashi T, Hirsch E, Moser M, Krug M, Seidenbecher CI, Boeckers TM, Rauch U, Buettner R, Gundelfinger ED, Fassler R. Neurocan is dispensable for brain development. Mol Cell Biol. 2001;21:5970–5978. doi: 10.1128/MCB.21.17.5970-5978.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 500.Zhu H, Mitsuhashi N, Klein A, Barsky LW, Weinberg K, Barr ML, Demetriou A, Wu GD. The role of the hyaluronan receptor CD44 in mesenchymal stem cell migration in the extracellular matrix. Stem Cells. 2006;24:928–935. doi: 10.1634/stemcells.2005-0186. [DOI] [PubMed] [Google Scholar]
  • 501.Zhuo L, Kanamori A, Kannagi R, Itano N, Wu J, Hamaguchi M, Ishiguro N, Kimata K. SHAP potentiates the CD44-mediated leukocyte adhesion to the hyaluronan substratum. J Biol Chem. 2006;281:20303–20314. doi: 10.1074/jbc.M506703200. [DOI] [PubMed] [Google Scholar]
  • 502.Zhuo L, Salustri A, Kimata K. A physiological function of serum proteoglycan bikunin: the chondroitin sulfate moiety plays a central role. Glycoconj J. 2002;19:241–247. doi: 10.1023/A:1025331929373. [DOI] [PubMed] [Google Scholar]
  • 503.Zhuo L, Yoneda M, Zhao M, Yingsung W, Yoshida N, Kitagawa Y, Kawamura K, Suzuki T, Kimata K. Defect in SHAP-hyaluronan complex causes severe female infertility. A study by inactivation of the bikunin gene in mice. J Biol Chem. 2001;276:7693–7696. doi: 10.1074/jbc.C000899200. [DOI] [PubMed] [Google Scholar]
  • 504.Zohar R, Sodek J, McCulloch CA. Characterization of stromal progenitor cells enriched by flow cytometry. Blood. 1997;90:3471–3481. [PubMed] [Google Scholar]
  • 505.Zoltan-Jones A, Huang L, Ghatak S, Toole BP. Elevated hyaluronan production induces mesenchymal and transformed properties in epithelial cells. J Biol Chem. 2003;278:45801–45810. doi: 10.1074/jbc.M308168200. [DOI] [PubMed] [Google Scholar]

RESOURCES