Skip to main content
BioMed Research International logoLink to BioMed Research International
. 2014 Aug 4;2014:103923. doi: 10.1155/2014/103923

Hyaluronan and RHAMM in Wound Repair and the “Cancerization” of Stromal Tissues

Cornelia Tolg 1, James B McCarthy 2, Arjang Yazdani 3, Eva A Turley 1,3,*
PMCID: PMC4137499  PMID: 25157350

Abstract

Tumors and wounds share many similarities including loss of tissue architecture, cell polarity and cell differentiation, aberrant extracellular matrix (ECM) remodeling (Ballard et al., 2006) increased inflammation, angiogenesis, and elevated cell migration and proliferation. Whereas these changes are transient in repairing wounds, tumors do not regain tissue architecture but rather their continued progression is fueled in part by loss of normal tissue structure. As a result tumors are often described as wounds that do not heal. The ECM component hyaluronan (HA) and its receptor RHAMM have both been implicated in wound repair and tumor progression. This review highlights the similarities and differences in their roles during these processes and proposes that RHAMM-regulated wound repair functions may contribute to “cancerization” of the tumor microenvironment.

1. Introduction

Tumors have often been compared to chronic wounds that do not heal. The tumor microenvironment, which is a critical but incompletely understood factor in promoting tumor progression, exhibits tissue remodeling characteristics similar to wounds. These include loss of cell polarity/tissue architecture and remodeling (degradation/resynthesis and reorganization) of the ECM [1], as well as cell dedifferentiation, migration, and proliferation [27]. A prolonged and episodic remodeling of adult tissue that results in loss of architecture is also associated with an increased susceptibility for tumor initiation. For example, gestation and involution in breast tissue, which are two periods of prolonged and repeated mammary tissue remodeling, are both linked to increased breast cancer (BCa) susceptibility [815].

Most adult wounds heal by fibrosis, which is characterized by an inflammatory response, changes in the composition of ECM, accumulation of biologically active ECM fragments, and scarring [1620]. There are also accompanying changes in the cellular content of the wound environment that include the differentiation of myofibroblasts, which contribute to wound closure, the formation of a microvasculature, collagen I deposition, and scarring [17, 2126]. Finally, there is an infiltration of circulating fibroblasts [2729] and innate immune cells [30] that synthesize and ultimately contribute to repair completion and restoration of tissue architecture. Aspects of this fibrotic milieu provide a protumorigenic microenvironment that enhances both tumor initiation and expansion [3134]. For example, the presence of high density or fibrotic regions in breast, often resulting from radiation treatment, are sites commonly associated with tumor recurrence [35, 36]. This observation and others suggest a model for tumor initiation that is associated with the chronic or frequent (e.g., episodic) loss of normal tissue architecture and wound-like ECM remodeling, which enhances rogue behavior of mutant cells by providing a “cancerized” microenvironment (Figure 1) [37, 38]. Once tumors are initiated, molecular mechanisms associated with malignant progression function in a dynamic and reciprocal manner with host cells to sustain and enhance this protumorigenic wound-like microenvironment. It should therefore be no surprise that gene signatures and transcriptomes of tumors are enriched in wound repair profiles and that these profiles are associated with or prognostic of poor outcome [3944].

Figure 1.

Figure 1

Schematic summarizing wound and tumor microenvironment remodeling in skin. The normal tissue architecture of skin is well-organized in both the epidermis, which consists of differentiated cohesive keratinocytes, and the dermis, which is composed of fibroblasts, blood vessels, and well-organized collagen fibrils amongst other ECM components. Tissue injury results in temporary changes in tissue architecture as keratinocytes dedifferentiate and migrate across wound gaps, proinflammatory macrophages migrate into the dermis, angiogenesis is promoted, and subpopulations of fibroblasts differentiate into myofibroblasts that organize collagen fibrils, which contribute to scar tissue. Tumor initiation also results in dedifferentiation, proliferation and migration/invasion of keratinocytes, influx of macrophages, differentiation of fibroblasts into myofibroblasts that increase deposition and scar like organization of collagen fibrils, and formation of new immature blood vessels. However, this disorganized tissue architecture is not transient as it is in wound repair but increases with tumor progression.

Quite often discussion on the importance of ECM remodeling in wound repair and protumorigenic stroma focuses upon alterations in the synthesis and fragmentation of ECM proteins [4547]. However, a consideration of the tissue polysaccharide HA is usually not included in these discussions, despite the fact that elevated HA production is essential for tissue repair, is required for tumor progression in numerous experimental models, and is linked to poor outcome in many cancers including BCa [3, 30, 48]. Therefore the first part of this review will focus on HA metabolism as it relates to wound healing and BCa initiation/malignant progression. There is also clear and convincing evidence that HA receptors such as cluster designation 44 (CD44), receptor for hyaluronan mediated motility [49], and toll-like receptors 2,4 (TLR2,4) (to name a few) are all important contributors to malignant progression and outcome in BCa patients. There are many excellent reviews on the functions associated with these and other HA receptors in tissue homeostasis, wound repair, and tumor 4 progression [3, 30, 48, 5053]. However, this review will focus on the multifunctional HA receptor, RHAMM (gene name HMMR), because of its clear roles in fibrotic wound repair that are apparently relevant to BCa initiation and progression. For example, the expression levels of HMMR/RHAMM are frequently increased in BCa and linked to poor clinical outcome [54] and considerable in vivo evidence links RHAMM expression levels to mesenchymal response-to-injury [5560]. The disparate functions of RHAMM are related to its complex subcellular localization. RHAMM was originally described in the context of one of these functions, which is to facilitate HA mediated cell motility [48, 6164] but was more recently shown to affect centrosomal function and mitotic spindle integrity. The purpose of this review is to highlight the roles of HA and RHAMM that converge in wound repair and BCa progression.

2. Hyaluronan Background

HA is a polysaccharide belonging to the glycosaminoglycan family of macromolecules. This biopolymer consists of repeating disaccharides of N-acetylglucosamine and β-glucuronic acid, the latter of which confers anionic properties to HA [51, 53, 65]. There are three known hyaluronan synthases (HAS1-3) that produce HA and these are differentially expressed during wound repair and in tumors [48, 6678]. HAS enzymes are proteins that contain multiple membrane spanning sequences and to date they are the only known glycosyltransferases imbedded in the plasma membrane. “Activated” UDP-sugars are sequentially added to the catalytic portion of the enzyme located on the inner face of the plasma membrane. The synthesized polymer is then extruded through the plasma membrane, possibly through channels created by oligomerization of the synthases [76, 79].

HA is a ubiquitous component of tissue ECM but is found in particularly high concentrations as a native homeostatic form within hydrated tissues such as the vitreous of the eye, articular cartilage, and lymphatics and skin. It is particularly enriched in the epidermis, where it is important for maintaining the hydration of this tissue so that it can form a more effective barrier to the environment [8082]. During embryonic development HA is a crucial component of cardiac jelly and is absolutely required for heart development where it provides a migration-supporting environment for cardiac cushion cells [83]. Genetic deletion of HAS2, the HA synthase that is responsible for HA synthesis during the embryonic developmental period when heart development occurs, results in embryonic lethality as a result of defective cardiac development [84]. Genetic deletions of HAS1 or 3 do not have the same developmental consequences emphasizing that these synthases are expressed differentially. They also exhibit distinct spatial distributions in tissues [68, 8587]. The biological significance of having 3 unique but closely related HA isoforms is not completely understood, but in addition to differential mechanisms regulating their expression, they also synthesize HA biopolymers of dissimilar average sizes. Such differences in polymer size are linked to distinct HA functions [48, 65, 8890]. Thus, it is not surprising that there are temporal, spatial, and cell-type specific differences in HAS1-3 expression during tissue repair and in tumor progression. For example, during repair of excisional wounds, keratinocyte migration is associated with elevated HAS2 and 3 expression while peritoneal mesothelial cells upregulate only HAS2 following mechanical injury [69, 72]. There is surprisingly little information about HAS expression and HA production during tissue remodeling that is not associated with injury. For example, although branching morphogenesis in general is known to be HA dependent, to our knowledge its role in mammary gland morphogenesis has not been reported. A great deal more is known about the roles of HA and HAS isoforms in the initiation and progression of tumors from breast tissue. HAS1 and HAS2 expression are commonly upregulated in BCa. Their elevated expression and HA accumulation are linked to ominous features of malignant progression. This includes epithelial-to-mesenchymal transformation (EMT) [9193] and increased invasion [3, 54, 9496], providing a partial explanation for HA's relationship to poor outcome [48, 74, 93, 97].

Glycosaminoglycans such as HA bind, concentrate, present, and prevent diffusion of growth factors in tissues. HA is fundamentally important for both maintaining tissue homeostasis and orchestrating the inflammatory, fibrotic, and stem cell renewal responses in damaged tissues. High molecular weight HA is typically produced by homeostatic tissues and can reach up to 2,000 kDa. High molecular weight HA performs ECM and growth factor presenting/scaffolding functions and is critical for tissue hydration [81, 98]. By providing a relatively loose matrix, HA supports cell migration, acts as scavenger for reactive oxygen and nitrogen species, and is typically anti-inflammatory [30, 50, 65, 99]. Importantly, recent findings also suggest that high molecular weight HA exhibits tumor suppressive functions in skin [100]. By sharp contrast, the fragmentation of polymeric HA, which occurs during both normal and disease-associated tissue remodeling, drastically alters the functions of HA. HA fragments are typically proinflammatory and promigratory and promote proliferation (Figure 2). This change in function as a result of depolymerization is similar in principle to the fragmentation of certain embryonic morphogens and extracellular matrikines which have distinct biological properties compared to the intact molecules, and which occur during tissue remodeling [98]. HA fragmentation can occur as a result of hyaluronidase activity (e.g., Hyal1 and Hyal2) or the presence of reactive oxygen and nitrogen species [81, 98]. Furthermore, HA size heterogeneity within a wound can also result from altered expression levels of distinct HAS isoforms [48]. Localized changes in HA synthesis and fragmentation may therefore represent a type of “on-off switch,” which is important for providing early warning signals of tissue damage, requiring host cells to respond appropriately to restore tissue function and architecture [50, 98, 101105]. It needs to be emphasized, however, that the majority of these studies have been performed in culture with very few analyses determining the extent of HA fragmentation in intact tissues during tissue remodeling in vivo. Therefore aspects of these in vitro studies that are applicable to tissue remodeling in vivo have yet to be fully defined.

Figure 2.

Figure 2

HA functions are molecular weight dependent. HA occurs as a large native polymer in homeostatic tissues but is degraded following tissue injury by free radicals and hyaluronidases. The resulting fragments have different bioactivity than the native polymer depending upon their size. For example, intermediate fragments can stimulate cell proliferation while smaller fragments have been reported to only promote cell migration.

The development of new techniques [106, 107] that facilitate the isolation and determination of HA fragments size distributions from complex tissues will ultimately provide an important framework for understanding the biological importance of size heterogeneity of this biopolymer. While in vivo analysis of these fragmentation patterns is just the beginning, there is mounting evidence in vitro that the biological impact of HA size heterogeneity is related to the ability of different HA receptors to bind high molecular weight polymeric versus HA fragments. As an example, while higher molecular weight HA binds to CD44, smaller fragments of HA bind to RHAMM and even smaller fragments bind to TLR 2,4. There is evidence that these HA fragments can also function to inhibit the binding of higher molecular weight HA to CD44 [108, 109]. Thus, fragmented HA could impact tissues by either directly binding specific receptors or antagonizing the binding of larger HA polymers to their cognate receptors. These complex interactions control a variety of signaling pathways that regulate cell adhesion/motility, mitotic spindle formation, and transcriptomes. Clearly much work is needed to dissect the complex roles of native HA versus fragments in normal and diseased tissues, with longer term impact of providing specific targets associated with pathologies linked to altered HA metabolism.

3. Hyaluronan and Tissue Remodelling

The importance of the synthesis/fragmentation of high molecular weight HA has been most extensively studied in models of tissue repair. HA is considered a “keystone” or central molecule in regulating response to tissue stress since rapid alterations in HA production, macromolecular organization, and size within tissues are among the earliest changes that can be detected following injury including those resulting from exposure to ionizing radiation [30, 81, 98, 110, 111]. Like other ECM components (e.g., collagen), HA is fragmented during wound repair into small proinflammatory oligosaccharides by oxygen/nitrogen free radicals that accumulate in the stressed tissue [112, 113]. These wound fragments constitute an early “danger signal” that is part of the damage-associated molecular process (DAMP). DAMP stimulates the innate immune response, which, when not resolved, directly contributes to chronic inflammation and tissue fibrosis [106, 114117]. Experimental models of tissue injury have documented the contribution of HA to DAMP and to the repair of excisional skin wounds [81, 118], vascular response to injury [119, 120], and induced lung injury [30]. The paradigm for the functions of HA in all of these injured tissues is similar, and recent reviews have summarized the literature in both vascular and lung injury models [110, 121]. Our studies on the identification and characterization of HA size heterogeneity in excisional wound healing are among the first to attempt to shed light on these issues in skin wounds in vivo [50, 106]. These studies have demonstrated an association between HA fragmentation and cellular infiltration into wounded tissues [56, 122, 123]. Furthermore they have shown that this cellular infiltration is defective in animals that are embryonic null for RHAMM [55, 56].

HA occurs in large amounts in skin and it is a key factor in its homeostasis since it controls both fibroblast differentiation and epidermal activation and renewal [118, 124]. As in other tissues, native, high molecular weight HA suppresses skin inflammation [30, 118, 125], myofibroblast differentiation [125], and fibrosis [118]. Similar to other injured tissues, fragmentation of HA, which results from ionizing radiation and other damage, promotes inflammation and tissue fibrosis. Homeostatic skin contains the largest depot of high molecular weight, anti-inflammatory (native) HA in the body and this is primarily organized into extracellular macromolecular complexes in the dermal ECM. It is also detected as pericellular coats, which are particularly noticeable around keratinocytes [126128]. These HA coats, which are sustained on keratinocytes surfaces via the HA receptor CD44 [129], are required for maintenance of barrier/permeability functions of skin as well as for keratinocyte renewal, proliferation, and differentiation [81, 130133]. Because of their exposure to the environment, a key additional homeostatic function of keratinocyte-associated coats is that they protect against ionizing radiation-induced DNA damage. For example, production of native HA protects against DNA damage resulting from either UVB or gamma radiation [81]. Native HA in skin is also linked to reduced risk of cancer (skin) and metastasis [92, 100]. Importantly, loss of keratinocyte HA coats is linked to epidermal atrophy and other pathologic lesions including increased dermal fibrosis [124]. In addition to its effects on keratinocytes, skin HA metabolism also controls proinflammatory immune cell influx [30] and TGFβ-1 induced myofibroblast differentiation, which is largely responsible for fibrosis [118, 125]. Reepithelialization of injured skin begins early after injury and is controlled by EGF signaling, which stimulates HA production and regulates promigratory/proliferation signaling through HA and CD44 [134]. TGFβ-1 is strongly antiproliferative in these cells [135].

Reepithelialization serves as a critical function in maintaining the integrity of the dermal layers due to crosstalk between these two skin layers. Although not a great deal is known about this function, TGFβ-1/SMAD 3 signaling plays a key role. Thus, when wound repair is deregulated and reepithelialization is prevented, dermal fibrosis is enhanced [136]. The epithelium in transgenic mice, which are engineered to suppress SMAD 3, exhibits accelerated reepithelialization, reduced inflammation, and reduced dermal fibrosis [137]. Reepithelialization of wounds normally coincides with and may instruct removal of dermal myofibroblasts by apoptosis. Blocking reepithelialization prevents myofibroblast apoptosis and results in hypertrophic scars or chronic tissue fibrosis [124]. HA, which as noted above is necessary for keratinocyte proliferation and migration in response to epidermal injury, is also a key regulator of TGFβ-1 functions in fibrosis and myofibroblast differentiation [118, 125].

Fibroblast differentiation into myofibroblasts is controlled by two cooperating pathways, TGFβ-R/SMAD and HA mediated RHAMM:CD44:EGFR signaling complexes, both of which must be activated to induce myofibroblast differentiation [138]. Thus, the TGFβ-1 signaling pathway promotes HAS2 dependent HA synthesis and pericellular HA coat formation. This results in increased levels of endogenous TGFβ-1, which maintain myofibroblast differentiation via an autocrine loop consisting of HA:TGFβ-1 production [139]. TGFβ-1 and HA dependent signaling promote CD44 and EGFR interaction in lipid rafts resulting in the activation of ERK1,2 and calmodulin kinase II activation [138]. Intriguingly, increased accumulation of native extracellular HA [118] such as what occurs during repair of embryonic tissues [140, 141] or disruption of HA:RHAMM:CD44 [106] complexes in wounds can negatively regulate these signaling pathways so that myofibroblast differentiation is reduced and, conversely, reepithelialization is promoted. The role of HA in wound myofibroblast differentiation is important since these cells closely resemble cancer associated fibroblasts that are intimately involved in development of a cancerized microenvironment that facilitate the progression of BCa (e.g., [142]).

Studies implicating differential roles for high molecular weight versus fragmented HA in the repair of other tissues prompted us to further investigate the potential relationship between HA synthesis and size heterogeneity in excisional wound repair [30, 98]. As anticipated, homeostatic adult mouse skin contains mainly high molecular weight HA (>5,000 kDa) while fragmentation can be detected in wounds as rapidly as 24 h, peaking at 7 days after injury [106]. Importantly high molecular weight polymers (>5,000 kDa) as well as intermediate-to-small (30–500 kDa) and very small (<10 kDa) HA fragments coexist in these wounds. This fragmentation pattern is consistent with an important role for HA size heterogeneity in promoting excisional dermal wound repair. HA fragments signal as “on” switches to both circulating and resident cells, which promote inflammation and angiogenesis while coexisting high molecular weight polymers signal as “off” switches that limit responses to fragments. In support of this notion, it has been shown that either forced HAS1 expression or supplementing the wounds with high molecular weight HA limits the fibrotic response such that the healing of these wounds more closely resembles those observed in embryos. The effect of adding excess high molecular weight HA to these wounds leads to reduced levels of TGFβ-1, attenuated inflammation, and a reduction in biomechanical stress [22, 143, 144]. A number of studies have established that small to intermediate size HA fragments promote the influx of immune cells. Such fragments also activate the proinflammatory functions of these infiltrating cells, which include stimulating expression of chemokines such as MIF-1a and MCP and increasing the expression of profibrotic growth factors such as TGFβ-1 [30]. This size range of HA fragments also promotes branching morphogenesis of wound associated blood vessels [119] and activation of fibroblasts [56, 145].

The availability of precisely or at least restricted sizes of small HA polymers produced by recombinant technology has allowed studies that suggest a much more complex functional repertoire and interplay of high molecular weight HA versus HA fragments than previously suspected. For example, a range of HA fragment sizes (e.g., HA-12 (12 saccharides) and HA-880 (880 saccharides) and native, high molecular weight HA) all activate ERK1,2, Akt, and P38 signaling cascades and all increase expression of ECM remodeling proteinases such as MMP1,3 [122, 146150]. However, HA-12 and native HA selectively promote expression of collagen III and TGFβ-3 and HA-12 solely promotes TIMP1 expression by dermal fibroblasts in culture [56, 149]. HA-6 (6 saccharides) but not HA-8, HA-10, 40 kDa, or native HA stimulates wound closure and increases wound macrophages and TGFβ-1 levels. In spite of stimulating TGFβ-1, HA-6 does not increase myofibroblast differentiation suggesting requirement of additional stimuli and possibly other HA fragment sizes [56]. These studies serve to emphasize the enormous amount of information that is generated by differential fragmentation of HA associated with tissue injury. The mechanism by which this information is transduced to the cell is currently not well understood. For example, are the trafficking/display patterns [54] of HA receptors that have differential binding properties for discrete sizes of HA fragments involved in signal transduction of the HA fragment pool as a whole? Furthermore, is there a temporal or spatial relationship between HA fragmentation and the response of specific cellular subpopulations expressing various HA receptors? Despite the complexity, an important first step is to document the presence and kinetics of HA fragmentation patterns in order to address these and other questions in the future. These analyses of HA functions in excisional wounds and other models of tissue injury have uncovered an information-rich mechanism for finely regulating the key processes of inflammation, angiogenesis, and fibroblast activation that are all essential for efficient wound repair [98]. These same processes in the tumor microenvironment appear to be required for the initiation and progression of BCa.

4. Stromal Hyaluronan and Breast Cancer

In BCa it is clear that tumors progress more aggressively in a HA-rich microenvironment and that stromal HA affects both host and tumor cells to accelerate progression. Many of the functions of HA during BCa progression have been summarized in several recent excellent reviews [3, 48, 78, 97, 125]. The importance of stromal HA in mediating host responses, which support BCa progression, is the focus in this section.

Both tumor parenchyma and host cells in the tumor microenvironment express HAS isoforms and produce HA, which then accumulates in tumor parenchyma and in the peritumor stromal tissues [48, 97]. Clinical studies suggest that HA accumulating in either the tumor parenchyma or surrounding peritumor stroma is tightly linked to BCa progression and both are independent prognostic indicators of poor outcome [97]. Stromal cells, in particular cancer associated fibroblasts, express all HAS isoforms. Increased HAS expression by cancer associated fibroblasts correlates with increased HA accumulation, increased stromal CD44 expression, high relapse rate, and short overall survival [74]. Furthermore, high stromal HA accumulation is significantly associated with the appearance of a tumor reactive stroma, which associates with tumor cell positive lymph nodes, high tumor grade, and lymphatic tumor emboli [151]. The increased expression of extracellular HA binding proteins such as versican has also been reported [152]. Analyses of head and neck tumors reveal previously unappreciated stromal cancer associated fibroblast heterogeneity involving HA: a subpopulation of cancer associated fibroblasts produces high levels of HA and promotes local tissue invasion by cancer cells [153].

Bigenic expression of Neu and HAS2 in ductal epithelium using the MMTV promoter in mice results in marked changes in the peritumor stroma resembling those observed clinically in tumor reactive stroma [154]. Furthermore, these studies show that HA produced by the tumor parenchyma by itself can enhance stromal cell recruitment and the formation of a tumor reactive stroma. Notable phenotypic changes include increased formation of intratumoral HA-rich stroma, accumulation of ECM components, such as versican, collagen 1, and fibronectin, neovascularization, and the infiltration of immature mesenchymal cells. Cytokine analyses suggest that the increased accumulation of these stromal cells stimulates neoangiogenesis. Coinjection experiments with cancer associated fibroblasts that are derived from these bigenic tumors reveal that the cancer associated fibroblasts are responsible for the observed increase in tumor growth, reactive stroma formation, angiogenesis, and lymphangiogenesis [155]. Others have shown that tumor-associated HA supports tumor cell epithelial-mesenchymal transformation that also enables the growth and spread of tumor cells [154157]. Collectively these studies indicate that production of HA by tumor cells favors recruitment of mesenchymal cells that remodel the peritumor stroma to create a tumor friendly microenvironment.

In addition, microenvironments rich in HA provide mitogenic and motogenic signals for tumor cells. For example, we have shown that human BCa lines are heterogeneous in their ability to bind to HA and that exposure to this glycosaminoglycan promotes specific subpopulations within these lines to divide rapidly while stimulating other subpopulations to invade aggressively but proliferate slowly [54]. This type of functional heterogeneity may be partly responsible for the relationship between HA-rich tumor microenvironments and relapse/poor outcome reported by a number of clinical studies. Elevated stromal HA is linked to HER2 positivity and several key clinicopathologic features including poor prognosis factors such as tumor size, lymph node positivity, hormone receptor negativity, increased relapse rate, and shortened survival [158]. The mechanisms by which stromal HA effects BCa progression and the roles of native versus fragmented HA are not yet well understood. However, considering the evidence from multiple model systems (discussed above) it is clear that these effects are mediated through HA receptors. Multiple HA receptors (e.g., CD44, RHAMM and LYVE-1, and TLR2,3 among others) are known to be involved in BCa progression [3, 48, 156, 159161]. The role of RHAMM in progression of this disease and in wound repair will be considered here because RHAMM is unique in the ways in which it converts HA “signaling” into multiple key aspects of cellular functions that are relevant to response to injury and to tumor progression.

5. RHAMM Background

Studies using RHAMM null animals have clearly established an important role for this protein in tissue response to injury [55, 57]. Furthermore, a number of studies have linked RHAMM expression to BCa since it is frequently elevated in breast and other cancers and is associated with poor outcome [54, 61, 162]. RHAMM is a largely hydrophilic helical protein (Figure 3) that was originally isolated from conditioned medium from chicken heart explant cultures exhibiting high HA production and increased cell migration [163]. It binds to HA via positively charged amino acid clusters in the carboxyl terminus that are structurally distinct from the link module responsible for HA binding to CD44 [164]. RHAMM also binds to microtubules via sequences located in its N- and carboxyl termini. It directly binds to ERK1 via a sequence with homology to a D-box MAP kinase interaction site [95, 165, 166]. The protein contains several leucine zippers and these together with its potential for forming a coiled coil predict that it can self-associate as dimers or trimers (Figure 3).

Figure 3.

Figure 3

RHAMM isoforms, protein secondary structure, and posttranslational modification (a) RHAMM exon structure is shown as black boxes. Lines underneath this diagram show the known isoform structures. The full-length protein (85 kDa in human) is largely associated with interphase microtubules and the mitotic spindle during the cell cycle. Three isoforms are generated by alternative splicing of exon 4, 5, or 13. Loss of exon 4 sequence disrupts stable association with interphase microtubules and results in the appearance of RHAMM in the interphase cell nucleus. N-terminal truncations that may be generated by posttranslational mechanism or alternative start codon usage are very transiently expressed during early tissue injury but are constitutively present in some aggressive breast cancer cell lines and tumors. These accumulate in the nucleus and on the cell surface. (b) RHAMM protein is predicted to be largely a-helical, with unstructured sequences at the extreme N and C-termini. The orange P at the carboxyl terminus indicates an AURKA and ERK1,2 phosphoacceptor site. RHAMM also contains approximately 30 putative protein kinase C phosphoacceptor sites (not shown). This posttranslational modification is associated with the nuclear accumulation of RHAMM. Although RHAMM is phosphorylated by protein kinase C, the acceptor sites have not yet been reported. Protein kinase C modification of RHAMM is linked to interphase centrosomal placement. (c) The secondary structure predictions shown in (b) indicate that RHAMM proteins can self-associate to form random coiled coils.

In vivo, RHAMM expression is tightly regulated: it is poorly expressed in most homeostatic adult tissues with the exception of ovaries, testes, and ciliated epithelium of the respiratory tract in which elevated RHAMM mRNA levels are detected [167169]. However, RHAMM mRNA and protein expression are strongly but transiently increased in response to injury. A number of mechanisms have been identified that either promote or suppress RHAMM expression. Promoting factors include TGFβ-1, RON, and the YAP-HIPPO pathway while tumor suppressors such as p53 and BRCA1 reduce its expression [49, 61, 96, 162, 170176]. Analyses of RHAMM knockout mice [57, 58, 177] show that it is perhaps surprisingly not required for embryogenesis or homeostatic adult functions, the latter predicted by its low or absent expression in most tissues [177]. However it is essential for a variety of tissue repair processes that like embryogenesis require cell migration, invasion, and ECM remodeling. Since elevated levels of RHAMM are associated with poor prognosis in human cancers, it would appear that tumor cells usurp these wound repair functions of RHAMM to facilitate their survival and progression. The restricted expression of RHAMM makes it a potential target for cancer and wound repair therapy with low toxicity. Indeed, RHAMM peptides that are currently being tested in phase II clinical trials for multiple myeloma and myelodysplastic syndrome show efficacy and low toxicity in patients [178, 179]. The biological functions of RHAMM are complex. It is one of the first proteins to be identified for which its extracellular and intracellular functions differ markedly.

6. RHAMM Signaling

The signaling functions of RHAMM are multifaceted and context dependent as might be expected by its complex subcellular compartmentalization. RHAMM is a cytoskeletal, centrosomal, mitotic spindle, and nuclear protein [49, 61, 166, 180], which is exported to the cell surface by unconventional mechanisms during wounding [181] by cytokines such as TGFβ-1 [173]. Cell surface RHAMM associates with several integral protein tyrosine kinase and nonprotein tyrosine kinase receptors including PDGFR [182], TGFβ Receptor-1 [170], CD44 [55, 64], CD44-EGFR complexes [183, 184], bFGFR [185], and RON [171]. RHAMM impacts upon the signaling competency of these receptors in response to their cognate ligands (Figure 4). Cell surface RHAMM:CD44 complexes, in association with one or more of the above growth factor receptors, promotes random cell motility in a protein tyrosine and ERK1,2 kinase dependent manner [186188] (Figure 4). This random motility function does not require intracellular RHAMM proteins and immobilized recombinant cell surface RHAMM isoform (70 kDa) added to RHAMM−/−:CD44−/− fibroblasts is sufficient restore fibroblast motility speed to that of wild type or RHAMM-rescued fibroblasts [55]. Cell surface RHAMM also likely participates in functions required for wound repair such as cell division fidelity, mitotic spindle integrity, and cell cycle progression that were originally thought to be HA-independent functions of intracellular RHAMM proteins. For example, blocking cell surface RHAMM signaling reduces cell cycle progression of fibroblasts through G2M, a stage in the cell cycle where RHAMM and HAS2 mRNA are transiently elevated and for which HA production is necessary to facilitate cell rounding [189]. Exogenous HA also promotes the association of microtubule-associated protein homolog (TPX2) with nuclear RHAMM and phosphorylation of AURORA Kinase A (AURKA) to stimulate progression through the cell cycle [183]. The details of these signaling pathways as they are regulated by cell surface and intracellular RHAMM protein forms have been recently reviewed in detail [48, 61, 62]. The coordinated and separate signaling functions of intracellular and cell surface RHAMM in wound repair and in BCa remain to be resolved.

Figure 4.

Figure 4

Model of RHAMM signaling. Model summarizes the known signaling functions of cell surface and intracellular RHAMM. Cell surface RHAMM interacts with CD44, HA, and growth factors to activate protein tyrosine kinase signaling cascades that activate the ERK1,2 MAP kinase cascade. This signaling can increase random motility in the absence of intracellular RHAMM. Intracellular RHAMM also binds to a number of protein partners that mediate its functions as a regulator of microtubule dynamics, centrosome structure/function, and gene expression. For example, during interphase, cytoplasmic RHAMM:protein partner interactions (MEK1/ERK1,2 shown) contribute to the dynamic properties of interphase microtubules and the number, placement, and structure of centrosomes, which affect cell polarity and direct cell migration. Nuclear RHAMM:MEK1:ERK1,2 complexes also control expression of genes involved in cell motility such as PAI-1 and MMP-9. During the cell cycle, RHAMM:TPX2 complexes contribute to mitotic spindle integrity and cell cycle progression through G2M while RHAMM:supervillin complexes promote cytokinesis.

A simplified model of the proposed coordinated extracellular and intracellular RHAMM signaling functions is depicted in Figure 4. It is intriguing that both share the ability to regulate activation and subcellular localization of components of the MAP kinase (ERK1,2) cascade [48, 61, 62]. HA stimulation of cell surface RHAMM has consistently been shown to control the duration of ERK1,2 activity [55]. Intracellular RHAMM proteins form complexes with MEK1 and ERK1,2 and target these kinases to the cytoskeleton [165] and nucleus [55]. These signaling functions are required for random motility, mitotic spindle integrity, progression through the cell cycle, and gene expression (e.g., PAI-1 [170] and MMP9 [190]). RHAMM:ERK1,2 complexes are also likely to be important to centrosomal function since both RHAMM and ERK1,2 are required for microtubule nucleation [133, 191] and both are functionally linked to key centrosomal proteins such as TPX2 and AURKA [49, 183, 192]. In addition to microtubules, intracellular RHAMM partners with cortical actin proteins such as supervillin [193]. Supervillin, a membrane bound actin binding protein that participates in myosin II mediated contractility, interacts with calponin and regulates the activity of another RHAMM binding partner, ERK1,2 [194]. Supervillin coordinates processes that require dynamic cytoskeleton and membrane turnover including cell migration and cytokinesis [193, 195]. Indeed, RHAMM−/− cells often undergo aberrant cytokinesis causing the formation of multinucleated cells [165]. To date, the signaling functions of RHAMM can therefore be roughly divided into those that require (1) intracellular RHAMM and cell surface RHAMM (e.g., to control microtubule dynamics/nucleation and gene expression), (2) only cell surface RHAMM (e.g., to control random motility speed), and (3) only intracellular RHAMM (e.g., possibly cytokinesis).

While cell surface RHAMM controls the kinetics of ERK1,2 activation, intracellular RHAMM appears to target MEK1/ERK1,2 complexes to microtubules thus contributing to the dynamic turnover of interphase microtubules [196] and mitotic spindles [165]. Mitotic spindle formation is complex and includes key proteins such as AURKA and TPX2, which is a regulator kinase of AURKA. RHAMM-regulated ERK1,2 activity is required for bipolar spindle formation and loss of RHAMM can be compensated for by mutant active MEK1 in this function [165]. Intracellular RHAMM:TPX2 interactions and additional function interactions with BRCA1/BARD1 [49] also regulate the number, structure, and placement of centrosomes, in part through regulating AURKA activity. However, Hatano and colleagues have shown that nuclear RHAMM:TPX2 colocalization only occurs during metaphase. This group further showed that addition of HA stimulates both an association of RHAMM with TPX2 and an increase in the phosphorylation of the TPX2 regulator kinase, AURKA [183]. Since endogenous HA levels are high at G2M, it is likely that RHAMM:TPX2 interactions noted in other studies [49] are also controlled by HA. This centrosomal function of intracellular RHAMM is required for cell division fidelity in vascular response to injury, mitotic spindle integrity, progression through G2M, and basal-apical polarity of breast epithelial cells [57, 59, 197199]. Nuclear RHAMM may play a further role in sequestering TPX2 [49] to these compartments to prevent premature changes in microtubules/mitotic spindle assembly and to facilitate repair of DNA aberrations caused, for example, by ionizing radiation [200]. In addition, RHAMM together with CD44 or TGFβR1, and possibly intracellular HA, may directly affect the transcription of genes controlling cell migration and proliferation [170, 190, 201] (e.g., PAI-1, MMP-9, Figure 4).

7. RHAMM Subcellular Compartmentalization

The mechanisms responsible for the complex subcellular targeting of RHAMM are still incompletely understood but are likely to be contributed to by isoform structure and posttranslational modification. RHAMM is subject to mRNA splicing and is phosphorylated by a variety of serine threonine kinases (Figure 3). Several RHAMM mRNA splice variants have been identified in breast and other cancers including a 48 bp deletion in exon 4 (RHAMM-48), a 346 bp deletion in exon 5, and a 147 bp deletion in exon 13 although the presence of these forms in wound repair has not been reported [202205]. RHAMM is phosphorylated by protein kinase C, AURKA, and ERK1,2 [49, 197, 206]. Additionally, smaller than full-length N-terminal truncated RHAMM proteins have been reported during wound repair and in breast cancer cell lines [95]. Expression of these RHAMM isoforms and production of phosphoprotein specific antibodies have been utilized to identify isoform-specific subcellular targeting. These studies have shown that full-length RHAMM is largely associated with the cytoskeleton in interphase cells and in particular binds to microtubules [165, 207]. Targeting the nucleus is achieved by either alternative splicing of the full-length form in exon 5 [207], truncation of N-terminal sequence [208], or phosphorylation at T703 (human, but evolutionarily conserved [49]). The phosphorylation of RHAMM by protein kinase Cα is required for rear-polarization of the microtubule organizing center (MTOC) of migrating neointimal smooth muscle cells [59, 197]. Cell surface labeling of cultured cells reveals a predominance of N-terminal truncated RHAMM proteins [209]. These small truncated proteins, which are generated by as yet unknown mechanisms, are less prominent on the cytoskeleton than the full-length protein, and FRAP studies show they are more mobile within the cell, accumulating in the nucleus and at the cell surface [61]. Since the multiple functions of RHAMM appear to be dictated by subcellular location it is likely that the various isoforms perform different functions and regulate distinct signaling pathways. Although isoform expression levels and interplay have been linked to tumor progression, their roles in response to injury and the differential manner in which they regulate specific functions are still poorly understood. The subcellular compartmentalization and signaling functions of RHAMM isoforms are critical for efficient repair of adult tissues and appear to provide some tumor cell types with growth, survival, and invasive advantages.

8. RHAMM and Tissue Remodelling

RHAMM mRNA and protein expression are coordinately and transiently upregulated following tissue injury. RHAMM protein expression is detected at the site of excisional skin wounds 24 hrs after injury, peaks at 3 days, and disappears by day 7 [55]. A similar rapid and transient increase in RHAMM expression is observed following scratch wound injury of fibroblasts and smooth muscle cells in cell culture [181]. Analyses of response to injury processes in RHAMM knockout mice or following functional blockade of RHAMM protein in wild type animals show that it regulates HA mediated ECM remodeling, polarized cell migration, cell division fidelity, and mesenchymal differentiation. These functions have been particularly well studied in vascular, lung, and excisional skin injury models [55, 57, 106, 119, 181, 197, 209, 210].

RHAMM was first shown to be required for smooth muscle cell migration into scratch wounds in the mid-1990s [181] and then later demonstrated to be required for endothelial cell signaling and migration during vessel morphogenesis in culture [123, 170, 185, 211]. More recently, the role of RHAMM was studied following vessel damage using RHAMM−/− mice [57, 119, 181, 197]. Cell culture studies comparing RHAMM−/− and wild type smooth muscle cells and blocking RHAMM function with antibodies show that RHAMM:HA interactions mediate smooth muscle cell adhesion and contraction of collagen gels. In vivo, loss of RHAMM increases vessel lumen size and reduces the size of adventitia and collagen deposition within the artery wall [57]. These results suggest that cell surface RHAMM:HA interactions promote lumen constriction and blocking this function of RHAMM may be clinically useful in preventing restenosis. The role of RHAMM in smooth muscle proliferation following balloon injury of rat carotid arteries has also been reported [197]. These studies show that the rapid proliferation of neointimal smooth muscle cells is RHAMM mediated. In this injury setting, RHAMM:dynein complexes localized to the mitotic spindle are required to promote mitotic fidelity by controlling centrosome placement. Furthermore, intracellular RHAMM phosphorylated by protein kinase Cα is required for correct placement of centrosomes and directed migration of smooth muscle cells into wounds [59, 197]. In lung, RHAMM:HA fragment interactions are required for macrophage chemotaxis in surfactant-stimulated and bleomycin injured lungs [209]. RHAMM expression was first linked to skin wound fibroplasia and fibrosis in vivo using a transplantation model comparing incisional and excisional wounds [212]. The former heal without fibroplasia or scarring while repair of the latter is accompanied with extensive fibroplasia and scar formation. Only the excisional wounds exhibit increased CD44 and RHAMM expression. Later studies using RHAMM−/− mice, function blocking antibodies, and RHAMM mimetic peptide antagonists have established that HA:RHAMM interactions are critical for macrophage influx into the wound, as well as for fibroplasia and angiogenesis [55, 106]. Thus, blocking RHAMM function or deleting RHAMM expression results in a reduction of both M1 and M2 macrophages. Furthermore, loss of RHAMM function reduces the level of wound TGFβ-1, causes reduced fibroblast migration into wounds, and inhibits their differentiation into myofibroblasts. There is also a reduction in collagen 1 accumulation and in the number of wound blood vessels.

Evidence from these studies and others [123] suggests that RHAMM binds to fragmented HA and that these interactions may be important in stimulating a RHAMM mediated “danger signal” to cells within injured tissues. Importantly, the binding of RHAMM to HA fragments is surprisingly size specific: a mixture of 4–20 saccharides promotes endothelial cell migration through RHAMM but HA-6 present in this mixture uniquely promotes wound closure, M1 and M2 macrophage influx into wounds, and TGFβ-1 production through RHAMM and CD44 [56, 123]. In wound dermal cells, RHAMM:CD44 appears to cooperate to activate ERK1,2 and FAK. These results emphasize that HA-receptor interactions in healing wounds are complex and that multiple HA receptors can collaborate to control important aspects of wound repair. Similar RHAMM-regulated signaling appears to be at play in BCa progression.

9. RHAMM and Breast Cancer

To date, the role of RHAMM in BCa and other tumors have focused upon tumor cell parenchyma. However, it is likely that RHAMM expressed either by tumor or host cells directly or indirectly facilitates tumor progression. Blocking RHAMM in certain tumor cells inhibits tumor proliferation and migration/invasion while in others it primarily affects migration and invasion [48, 62]. RHAMM mRNA and protein expression are increased in most tumors and these high levels are positively associated with aggressive tumors. However, in a few tumor types (e.g., malignant peripheral nerve sheath tumors) knockdown of RHAMM levels actually enhances tumor aggression. Intriguingly, loss of RHAMM in these tumors is associated with increased AURKA activity and enhanced sensitivity to AURKA inhibitors [206, 213, 214]. RHAMM is also implicated in promoting the self-renewal and tumorigenic potential of tumor stem cells in cancers such as glioblastoma [215]. Despite the complexity of its functions in tumors, levels/distribution of RHAMM isoforms have diagnostic or prognostic value such as identifying which tumor types are sensitive to targeted therapy (e.g., AURKA inhibitors). Considering its involvement in many of the critical driver pathways important for malignant progression, the targeting of RHAMM may also have therapeutic value in some cancers including BCa (see below). For example, RHAMM silencing blocked the self-renewing capability of glioblastoma stem cells, and loss of RHAMM in malignant peripheral nerve sheath tumors or multiple myeloma enhances the sensitivity of tumor cells to AURKA inhibitors. RHAMM hyperexpression occurs in castration-resistant prostate cancer and is also associated with the likelihood of biochemical recurrence in prostate cancer patients with intermediate grade (Gleason grade 7) prostate tumors [206, 215217].

Data bank indicates that hyperexpression of RHAMM mRNA expression is common in BCa and these elevated levels are often linked to poor clinical outcome [54]. Common genetic mutations at the low penetrance susceptibility RHAMM/HMMR locus enhance breast cancer risk in BRCA-1 mutation carriers [49]. Furthermore, in a large BCa patient cohort, RHAMM hyperexpression in breast tumor cell subsets predicts poor clinical outcome and is associated with elevated risk of peripheral metastases [218]. Other studies demonstrate that RHAMM expression is linked to increased BCa cell invasion and metastases [48, 62]. RHAMM transcription, which is regulated by mevalonate and HIPPO pathways, is required for ERK1,2-controlled BCa cell line migration and invasion, with relatively little impact on proliferation [96]. Similarly, RHAMM is an essential part of an autocrine motility mechanism in aggressive BCa lines for sustaining motility and invasion that requires HA production, ERK1,2 activation, and CD44 display [95]. Human BCa lines of all subtypes are heterogeneous in their ability to bind to fluorescent-HA probes. Subpopulations of tumor cells sorted according to their ability to bind to HA exhibit very different phenotypes. Cells that bind high levels of HA display both CD44 and RHAMM, proliferate slowly but are highly invasive in culture-based assay and in vivo, and are metastatic in vivo. In contrast, cell subpopulations, which bind low or no HA, express only CD44, proliferate rapidly but are poorly invasive in both culture based assays and in vivo, and are poorly metastatic [54]. These studies predict that major RHAMM functions in BCa are to support invasion and metastasis and that coordinated HA:cell surface RHAMM: intracellular RHAMM signaling contributes to BCa metastases in multiple but as yet incompletely understood ways [95, 165].

10. Conclusions

Wound repair and tumor progression are two complex but similar biological processes that share many molecular mechanisms for controlling cell migration, invasion, survival, and proliferation. HA and its receptors control essential functions in these two processes and this effect appears to be controlled in part by its binding to RHAMM. RHAMM is also upregulated during both processes where it appears to be similarly involved in the control of cell migration, invasion, proliferation, and differentiation. RHAMM is a multifunctional protein that signals through the ERK1,2 and TPX2 pathways at multiple steps. Its action on these pathways appears to be coordinately important to the initiation and progression of BCa and normal response to injury. The selective expression of RHAMM during times of tissue remodeling makes it a promising marker and target for diagnosis and therapy of disease involving aberrant wound repair and cancer.

Abbreviations

AURKA:

AURORA Kinase A

BCa:

Breast cancer

CD44:

Cluster designation 44

ECM:

Extracellular matrix

ERK1,2:

Extracellular regulated kinase 1,2

HA:

Hyaluronan

HAS:

Hyaluronan synthase

Hyal:

Hyaluronidase

HMMR:

Human HA mediated motility receptor

MMP-9:

Metalloproteinase-9

PAI-1:

Plasminogen activation inhibitor 1

RHAMM:

Receptor for HA mediated motility

TLR2,4:

Toll-like Receptor 2,4

TPX2:

Microtubule-associated protein homolog.

Conflict of Interests

The authors declare that there is no conflict of interests regarding the publication of this paper.

References

  • 1.Ballard PL, Gonzales LW, Godinez RI, et al. Surfactant composition and function in a primate model of infant chronic lung disease: Effects of inhaled nitric oxide. Pediatric Research. 2006;59(1):157–162. doi: 10.1203/01.pdr.0000190664.69081.f1. [DOI] [PubMed] [Google Scholar]
  • 2.Schäfer M, Werner S. Cancer as an overhealing wound: an old hypothesis revisited. Nature Reviews Molecular Cell Biology. 2008;9(8):628–638. doi: 10.1038/nrm2455. [DOI] [PubMed] [Google Scholar]
  • 3.Veiseh M, Turley EA. Hyaluronan metabolism in remodeling extracellular matrix: probes for imaging and therapy of breast cancer. Integrative Biology. 2011;3(4):304–315. doi: 10.1039/c0ib00096e. [DOI] [PubMed] [Google Scholar]
  • 4.Byun JS, Gardner K. Wounds that will not heal: Pervasive cellular reprogramming in cancer. American Journal of Pathology. 2013;182(4):1055–1064. doi: 10.1016/j.ajpath.2013.01.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Meng X, Zhong J, Liu S, Murray M, Gonzalez-Angulo AM. A new hypothesis for the cancer mechanism. Cancer and Metastasis Reviews. 2012;31(1-2):247–268. doi: 10.1007/s10555-011-9342-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Mukhtar RA, Nseyo O, Campbell MJ, Esserman LJ. Tumor-associated macrophages in breast cancer as potential biomarkers for new treatments and diagnostics. Expert Review of Molecular Diagnostics. 2011;11(1):91–100. doi: 10.1586/erm.10.97. [DOI] [PubMed] [Google Scholar]
  • 7.Dieci MV, Arnedos M, Andre F, Soria JC. Fibroblast growth factor receptor inhibitors as a cancer treatment: from a biologic rationale to medical perspectives. Cancer Discovery. 2013;3(3):264–279. doi: 10.1158/2159-8290.CD-12-0362. [DOI] [PubMed] [Google Scholar]
  • 8.Polyak K. Pregnancy and breast cancer: the other side of the coin. Cancer Cell. 2006;9(3):151–153. doi: 10.1016/j.ccr.2006.02.026. [DOI] [PubMed] [Google Scholar]
  • 9.Radisky DC, Hartmann LC. Mammary involution and breast cancer risk: transgenic models and clinical studies. Journal of Mammary Gland Biology and Neoplasia. 2009;14(2):181–191. doi: 10.1007/s10911-009-9123-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Kobayashi S, Sugiura H, Ando Y, et al. Reproductive history and breast cancer risk. Breast Cancer. 2012;19(4):302–308. doi: 10.1007/s12282-012-0384-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Ercan C, van Diest PJ, Vooijs M. Mammary development and breast cancer: the role of stem cells. Current Molecular Medicine. 2011;11(4):270–285. doi: 10.2174/156652411795678007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Stein T, Salomonis N, Nuyten DS, van de Vijver MJ, Gusterson BA. A mouse mammary gland involution mRNA signature identifies biological pathways potentially associated with breast cancer metastasis. Journal of Mammary Gland Biology and Neoplasia. 2009;14(2):99–116. doi: 10.1007/s10911-009-9120-1. [DOI] [PubMed] [Google Scholar]
  • 13.Monks J, Henson PM. Differentiation of the mammary epithelial cell during involution: Implications for breast cancer. Journal of Mammary Gland Biology and Neoplasia. 2009;14(2):159–170. doi: 10.1007/s10911-009-9121-0. [DOI] [PubMed] [Google Scholar]
  • 14.Lanigan F, O'Connor D, Martin F, Gallagher WM. Molecular links between mammary gland development and breast cancer. Cellular and Molecular Life Sciences. 2007;64(24):3159–3184. doi: 10.1007/s00018-007-7386-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Schedin P, O'Brien J, Rudolph M, Stein T, Borges V. Microenvironment of the involuting mammary gland mediates mammary cancer progression. Journal of Mammary Gland Biology and Neoplasia. 2007;12(1):71–82. doi: 10.1007/s10911-007-9039-3. [DOI] [PubMed] [Google Scholar]
  • 16.Penn JW, Grobbelaar AO, Rolfe KJ. The role of the TGF-beta family in wound healing, burns and scarring: a review. International Journal of Burns and Trauma. 2012;2(1):18–28. [PMC free article] [PubMed] [Google Scholar]
  • 17.Phan SH. Genesis of the myofibroblast in lung injury and fibrosis. Proceedings of the American Thoracic Society. 2012;9(3):148–152. doi: 10.1513/pats.201201-011AW. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Gurtner GC, Werner S, Barrandon Y, Longaker MT. Wound repair and regeneration. Nature. 2008;453(7193):314–321. doi: 10.1038/nature07039. [DOI] [PubMed] [Google Scholar]
  • 19.Rhett JM, Ghatnekar GS, Palatinus JA, O'Quinn M, Yost MJ, Gourdie RG. Novel therapies for scar reduction and regenerative healing of skin wounds. Trends in Biotechnology. 2008;26(4):173–180. doi: 10.1016/j.tibtech.2007.12.007. [DOI] [PubMed] [Google Scholar]
  • 20.Diegelmann RF, Evans MC. Wound healing: an overview of acute, fibrotic and delayed healing. Frontiers in Bioscience. 2004;9:283–289. doi: 10.2741/1184. [DOI] [PubMed] [Google Scholar]
  • 21.Rybinski B, Franco-Barraza J, Cukierman E. The wound healing, chronic fibrosis, and cancer progression triad. Physiological Genomics. 2014;46(7):223–244. doi: 10.1152/physiolgenomics.00158.2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Namazi MR, Fallahzadeh MK, Schwartz RA. Strategies for prevention of scars: what can we learn from fetal skin? International Journal of Dermatology. 2011;50(1):85–93. doi: 10.1111/j.1365-4632.2010.04678.x. [DOI] [PubMed] [Google Scholar]
  • 23.Van De Water L, Varney S, Tomasek JJ. Mechanoregulation of the myofibroblast in wound contraction, scarring, and fibrosis: opportunities for new therapeutic intervention. Advances in Wound Care. 2013;2(4):122–141. doi: 10.1089/wound.2012.0393. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Mann J, Mann DA. Epigenetic regulation of wound healing and fibrosis. Current Opinion in Rheumatology. 2013;25(1):101–107. doi: 10.1097/BOR.0b013e32835b13e1. [DOI] [PubMed] [Google Scholar]
  • 25.Duffield JS, Lupher M, Thannickal VJ, Wynn TA. Host responses in tissue repair and fibrosis. Annual Review of Pathology: Mechanisms of Disease. 2013;8:241–276. doi: 10.1146/annurev-pathol-020712-163930. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Klingberg F, Hinz B, White ES. The myofibroblast matrix: Implications for tissue repair andfibrosis. Journal of Pathology. 2013;229(2):298–309. doi: 10.1002/path.4104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Fan X, Liang H. Criculating fibrocytes: a potent cell population in antigen-presenting and wound healing. Chinese Journal of Traumatology. 2010;13(2):111–116. [PubMed] [Google Scholar]
  • 28.Keeley EC, Mehrad B, Strieter RM. The role of circulating mesenchymal progenitor cells (fibrocytes) in the pathogenesis of fibrotic disorders. Thrombosis and Haemostasis. 2009;101(4):613–618. [PMC free article] [PubMed] [Google Scholar]
  • 29.Eyden B. The myofibroblast: phenotypic characterization as a prerequisite to understanding its functions in translational medicine. Journal of Cellular and Molecular Medicine. 2008;12(1):22–37. doi: 10.1111/j.1582-4934.2007.00213.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Jiang D, Liang J, Noble PW. Hyaluronan as an immune regulator in human diseases. Physiological Reviews. 2011;91(1):221–264. doi: 10.1152/physrev.00052.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Barron DA, Rowley DR. The reactive stroma microenvironment and prostate cancer progression. Endocrine-Related Cancer. 2012;19(6):R187–R204. doi: 10.1530/ERC-12-0085. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Togo S, Polanska UM, Horimoto Y, Orimo A. Carcinoma-associated fibroblasts are a promising therapeutic target. Cancers. 2013;5(1):149–169. doi: 10.3390/cancers5010149. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Mao Y, Keller ET, Garfield DH, Shen K, Wang J. Stromal cells in tumor microenvironment and breast cancer. Cancer and Metastasis Reviews. 2013;32(1-2):303–315. doi: 10.1007/s10555-012-9415-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Hanahan D, Coussens LM. Accessories to the crime: functions of cells recruited to the tumor microenvironment. Cancer Cell. 2012;21(3):309–322. doi: 10.1016/j.ccr.2012.02.022. [DOI] [PubMed] [Google Scholar]
  • 35.Huo CW, Chew GL, Britt KL. Mammographic density: a review on the current understanding of its association with breast cancer. Breast Cancer Research and Treatment. 2014;144(3):479–502. doi: 10.1007/s10549-014-2901-2. [DOI] [PubMed] [Google Scholar]
  • 36.Ghosh K, Brandt KR, Reynolds C, et al. Tissue composition of mammographically dense and non-dense breast tissue. Breast Cancer Research and Treatment. 2012;131(1):267–275. doi: 10.1007/s10549-011-1727-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Bissell MJ, Hines WC. Why don't we get more cancer? A proposed role of the microenvironment in restraining cancer progression. Nature Medicine. 2011;17(3):320–329. doi: 10.1038/nm.2328. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Keely PJ. Mechanisms by which the extracellular matrix and integrin signaling act to regulate the switch between tumor suppression and tumor promotion. Journal of Mammary Gland Biology and Neoplasia. 2011;16(3):205–219. doi: 10.1007/s10911-011-9226-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Paulsson J, Micke P. Prognostic relevance of cancer-associated fibroblasts in human cancer. Seminars in Cancer Biology. 2014;25:61–68. doi: 10.1016/j.semcancer.2014.02.006. [DOI] [PubMed] [Google Scholar]
  • 40.Troester MA, Lee MH, Carter M, et al. Activation of host wound responses in breast cancer microenvironment. Clinical Cancer Research. 2009;15(22):7020–7028. doi: 10.1158/1078-0432.CCR-09-1126. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Manjili MH, Najarian K, Wang X. Signatures of tumor-immune interactions as biomarkers for breast cancer prognosis. Future Oncology. 2012;8(6):703–711. doi: 10.2217/fon.12.57. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Ahn S, Cho J, Sung J, et al. The prognostic significance of tumor-associated stroma in invasive breast carcinoma. Tumor Biology. 2012;33(5):1573–1580. doi: 10.1007/s13277-012-0411-6. [DOI] [PubMed] [Google Scholar]
  • 43.Räsänen K, Vaheri A. Activation of fibroblasts in cancer stroma. Experimental Cell Research. 2010;316(17):2713–2722. doi: 10.1016/j.yexcr.2010.04.032. [DOI] [PubMed] [Google Scholar]
  • 44.Xouri G, Christian S. Origin and function of tumor stroma fibroblasts. Seminars in Cell and Developmental Biology. 2010;21(1):40–46. doi: 10.1016/j.semcdb.2009.11.017. [DOI] [PubMed] [Google Scholar]
  • 45.Monboisse JC, Oudart JB, Ramont L, Brassart-Pasco S, Maquart FX. Matrikines from basement membrane collagens: a new anti-cancer strategy. Biochimica et Biophysica Acta. 2014;1840(8):2589–2598. doi: 10.1016/j.bbagen.2013.12.029. [DOI] [PubMed] [Google Scholar]
  • 46.Miles FL, Sikes RA. Insidious changes in stromal matrix fuel cancer progression. Molecular Cancer Research. 2014;12(3):297–312. doi: 10.1158/1541-7786.MCR-13-0535. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Samples J, Willis M, Klauber-Demore N. Targeting angiogenesis and the tumor microenvironment. Surgical Oncology Clinics of North America. 2013;22(4):629–639. doi: 10.1016/j.soc.2013.06.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Heldin P, Basu K, Olofsson B, Porsch H, Kozlova I, Kahata K. Deregulation of hyaluronan synthesis, degradation and binding promotes breast cancer. The Journal of Biochemistry. 2013;154(5):395–408. doi: 10.1093/jb/mvt085. [DOI] [PubMed] [Google Scholar]
  • 49.Maxwell CA, Benítez J, Gómez-Baldó L, et al. Interplay between BRCA1 and RHAMM regulates epithelial apicobasal polarization and may influence risk of breast cancer. PLoS Biology. 2011;9(11) doi: 10.1371/journal.pbio.1001199.e1001199 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Petrey AC, de la Motte CA. Hyaluronan, a crucial regulator of inflammation. Frontiers in Immunology. 2014;5:p. 101. doi: 10.3389/fimmu.2014.00101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Vigetti D, Karousou E, Viola M, Deleonibus S, De Luca G, Passi A. Hyaluronan: biosynthesis and signaling. Biochimica et Biophysica Acta. 2014;1840(8):2452–2459. doi: 10.1016/j.bbagen.2014.02.001. [DOI] [PubMed] [Google Scholar]
  • 52.Toole BP, Slomiany MG. Hyaluronan: a constitutive regulator of chemoresistance and malignancy in cancer cells. Seminars in Cancer Biology. 2008;18(4):244–250. doi: 10.1016/j.semcancer.2008.03.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Dicker KT, Gurski LA, Pradhan-Bhatt S, Witt RL, Farach-Carson MC, Jia X. Hyaluronan: a simple polysaccharide with diverse biological functions. Acta Biomaterialia. 2014;10(4):1558–1570. doi: 10.1016/j.actbio.2013.12.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Veiseha M, Kwona DH, Borowsky AD, et al. Cellular heterogeneity profiling by hyaluronan probes reveals an invasive but slow-growing breast tumor subset. Proceedings of the National Academy of Sciences of the United States of America. 2014;111(17):E1731–E1739. doi: 10.1073/pnas.1402383111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Tolg C, Hamilton SR, Nakrieko K, et al. Rhamm−/− fibroblasts are defective in CD44-mediated ERK1,2 motogenic signaling, leading to defective skin wound repair. Journal of Cell Biology. 2006;175(6):1017–1028. doi: 10.1083/jcb.200511027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Tolg C, Telmer P, Turley E. Specific sizes of hyaluronan oligosaccharides stimulate fibroblast migration and excisional wound repair. PLoS ONE. 2014;9(2) doi: 10.1371/journal.pone.0088479.e88479 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Ma X, Pearce JD, Wilson DB, English WP, Edwards MS, Geary RL. Loss of the hyaluronan receptor RHAMM prevents constrictive artery wall remodeling. Journal of Vascular Surgery. 2014;59(3):804–813. doi: 10.1016/j.jvs.2013.03.047. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Foley JP, Lam D, Jiang H, et al. Toll-like receptor 2 (TLR2), transforming growth factor-β, hyaluronan (HA), and receptor for HA-mediated motility (RHAMM) are required for surfactant protein a-stimulated macrophage chemotaxis. Journal of Biological Chemistry. 2012;287(44):37406–37419. doi: 10.1074/jbc.M112.360982. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Silverman-Gavrila R, Silverman-Gavrila L, Bendeck MP. Cell division fidelity is altered during the vascular response to injury: its novel role in atherosclerosis progression. The American Journal of Pathology. 2013;182(3):628–639. doi: 10.1016/j.ajpath.2012.11.007. [DOI] [PubMed] [Google Scholar]
  • 60.Nedvetzki S, Gonen E, Assayag N, et al. RHAMM, a receptor for hyaluronan-mediated motility, compensates for CD44 in inflamed CD44-knockout mice: A different interpretation of redundancy. Proceedings of the National Academy of Sciences of the United States of America. 2004;101(52):18081–18086. doi: 10.1073/pnas.0407378102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Maxwell CA, McCarthy J, Turley E. Cell-surface and mitotic-spindle RHAMM: moonlighting or dual oncogenic functions? Journal of Cell Science. 2008;121(7):925–932. doi: 10.1242/jcs.022038. [DOI] [PubMed] [Google Scholar]
  • 62.Shigeishi H, Higashikawa K, Takechi M. Role of receptor for hyaluronan-mediated motility (RHAMM) in human head and neck cancers. Journal of Cancer Research and Clinical Oncology. 2014 doi: 10.1007/s00432-014-1653-z. [DOI] [PubMed] [Google Scholar]
  • 63.Karbownik MS, Nowak JZ. Hyaluronan: towards novel anti-cancer therapeutics. Pharmacological Reports. 2013;65(5):1056–1074. doi: 10.1016/s1734-1140(13)71465-8. [DOI] [PubMed] [Google Scholar]
  • 64.Nikitovic D, Kouvidi K, Karamanos NK, Tzanakakis GN. The roles of hyaluronan/RHAMM/CD44 and their respective interactions along the insidious pathways of fibrosarcoma progression. BioMed Research International. 2013;2013:12 pages. doi: 10.1155/2013/929531.929531 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Toole BP. Hyaluronan: from extracellular glue to pericellular cue. Nature Reviews Cancer. 2004;4(7):528–539. doi: 10.1038/nrc1391. [DOI] [PubMed] [Google Scholar]
  • 66.Jiang D, Liang J, Noble PW. Hyaluronan in tissue injury and repair. Annual Review of Cell and Developmental Biology. 2007;23:435–461. doi: 10.1146/annurev.cellbio.23.090506.123337. [DOI] [PubMed] [Google Scholar]
  • 67.Tammi R, Pasonen-Seppänen S, Kolehmainen E, Tammi M. Hyaluronan synthase induction and hyaluronan accumulation in mouse epidermis following skin injury. Journal of Investigative Dermatology. 2005;124(5):898–905. doi: 10.1111/j.0022-202X.2005.23697.x. [DOI] [PubMed] [Google Scholar]
  • 68.Bai KJ, Spicer AP, Mascarenhas MM, et al. The role of hyaluronan synthase 3 in ventilator-induced lung injury. American Journal of Respiratory and Critical Care Medicine. 2005;172(1):92–98. doi: 10.1164/rccm.200405-652OC. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Karvinen S, Pasonen-Seppänen S, Hyttinen JMT, et al. Keratinocyte growth factor stimulates migration and hyaluronan synthesis in the epidermis by activation of keratinocyte hyaluronan synthases 2 and 3. The Journal of Biological Chemistry. 2003;278(49):49495–49504. doi: 10.1074/jbc.M310445200. [DOI] [PubMed] [Google Scholar]
  • 70.Kuroda K, Utani A, Hamasaki Y, Shinkai H. Up-regulation of putative hyaluronan synthase mRNA by basic fibroblast growth factor and insulin-like growth factor-1 in human skin fibroblasts. Journal of Dermatological Science. 2001;26(2):156–160. doi: 10.1016/s0923-1811(00)00155-9. [DOI] [PubMed] [Google Scholar]
  • 71.Pienimäki JP, Rilla K, Fülöp C, et al. Epidermal growth factor activates hyaluronan synthase 2 in epidermal keratinocytes and increases pericellular and intracellular hyaluronan. The Journal of Biological Chemistry. 2001;276(23):20428–20435. doi: 10.1074/jbc.M007601200. [DOI] [PubMed] [Google Scholar]
  • 72.Yung S, Thomas GJ, Davies M. Induction of hyaluronan metabolism after mechanical injury of human peritoneal mesothelial cells in vitro. Kidney International. 2000;58(5):1953–1962. doi: 10.1111/j.1523-1755.2000.00367.x. [DOI] [PubMed] [Google Scholar]
  • 73.Kennedy CI, Diegelmann RF, Haynes JH, Yager DR. Proinflammatory cytokines differentially regulate hyaluronan synthase isoforms in fetal and adult fibroblasts. Journal of Pediatric Surgery. 2000;35(6):874–879. doi: 10.1053/jpsu.2000.6869. [DOI] [PubMed] [Google Scholar]
  • 74.Auvinen P, Rilla K, Tumelius R, et al. Hyaluronan synthases (HAS1-3) in stromal and malignant cells correlate with breast cancer grade and predict patient survival. Breast Cancer Research and Treatment. 2014;143(2):277–286. doi: 10.1007/s10549-013-2804-7. [DOI] [PubMed] [Google Scholar]
  • 75.Adamia S, Pilarski PM, Belch AR, Pilarski LM. Aberrant splicing, hyaluronan synthases and intracellular hyaluronan as drivers of oncogenesis and potential drug targets. Current Cancer Drug Targets. 2013;13(4):347–361. doi: 10.2174/1568009611313040001. [DOI] [PubMed] [Google Scholar]
  • 76.Tammi RH, Passi AG, Rilla K, et al. Transcriptional and post-translational regulation of hyaluronan synthesis. FEBS Journal. 2011;278(9):1419–1428. doi: 10.1111/j.1742-4658.2011.08070.x. [DOI] [PubMed] [Google Scholar]
  • 77.Nykopp TK, Rilla K, Sironen R, et al. Expression of hyaluronan synthases (HAS1-3) and hyaluronidases (HYAL1-2) in serous ovarian carcinomas: inverse correlation between HYAL1 and hyaluronan content. BMC Cancer. 2009;9, article 143 doi: 10.1186/1471-2407-9-143. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Itano N, Kimata K. Altered hyaluronan biosynthesis in cancer progression. Seminars in Cancer Biology. 2008;18(4):268–274. doi: 10.1016/j.semcancer.2008.03.006. [DOI] [PubMed] [Google Scholar]
  • 79.Pummill PE, DeAngelis PL. Evaluation of critical structural elements of UDP-sugar substrates and certain cysteine residues of a vertebrate hyaluronan synthase. The Journal of Biological Chemistry. 2002;277(24):21610–21616. doi: 10.1074/jbc.M202456200. [DOI] [PubMed] [Google Scholar]
  • 80.Bonté F. Skin moisturization mechanisms: new data. Annales Pharmaceutiques Francaises. 2011;69(3):135–141. doi: 10.1016/j.pharma.2011.01.004. [DOI] [PubMed] [Google Scholar]
  • 81.Stern R, Maibach HI. Hyaluronan in skin: aspects of aging and its pharmacologic modulation. Clinics in Dermatology. 2008;26(2):106–122. doi: 10.1016/j.clindermatol.2007.09.013. [DOI] [PubMed] [Google Scholar]
  • 82.Verdier-Sévrain S, Bonté F. Skin hydration: a review on its molecular mechanisms. Journal of Cosmetic Dermatology. 2007;6(2):75–82. doi: 10.1111/j.1473-2165.2007.00300.x. [DOI] [PubMed] [Google Scholar]
  • 83.Camenisch TD, Biesterfeldt J, Brehm-Gibson T, Bradley J, McDonald JA. Regulation of cardiac cushion development by hyaluronan. Experimental & Clinical Cardiology. 2001;6(1):4–10. [PMC free article] [PubMed] [Google Scholar]
  • 84.Camenisch TD, Spicer AP, Brehm-Gibson T, et al. Disruption of hyaluronan synthase-2 abrogates normal cardiac morphogenesis and hyaluronan-mediated transformation of epithelium to mesenchyme. Journal of Clinical Investigation. 2000;106(3):349–360. doi: 10.1172/JCI10272. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Goncharova V, Serobyan N, Iizuka S, et al. Hyaluronan expressed by the hematopoietic microenvironment is required for bone marrow hematopoiesis. The Journal of Biological Chemistry. 2012;287(30):25419–25433. doi: 10.1074/jbc.M112.376699. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Törrönen K, Nikunen K, Kärnä R, Tammi M, Tammi R, Rilla K. Tissue distribution and subcellular localization of hyaluronan synthase isoenzymes. Histochemistry and Cell Biology. 2014;141(1):17–31. doi: 10.1007/s00418-013-1143-4. [DOI] [PubMed] [Google Scholar]
  • 87.Tien JYL, Spicer AP. Three vertebrate hyaluronan synthases are expressed during mouse development in distinct spatial and temporal patterns. Developmental Dynamics. 2005;233(1):130–141. doi: 10.1002/dvdy.20328. [DOI] [PubMed] [Google Scholar]
  • 88.Vigetti D, Viola M, Karousou E, de Luca G, Passi A. Metabolic control of hyaluronan synthases. Matrix Biology. 2014;35:8–13. doi: 10.1016/j.matbio.2013.10.002. [DOI] [PubMed] [Google Scholar]
  • 89.DeAngelis PL. Glycosaminoglycan polysaccharide biosynthesis and production: today and tomorrow. Applied Microbiology and Biotechnology. 2012;94(2):295–305. doi: 10.1007/s00253-011-3801-6. [DOI] [PubMed] [Google Scholar]
  • 90.Weigel PH, DeAngelis PL. Hyaluronan synthases: a decade-plus of novel glycosyltransferases. The Journal of Biological Chemistry. 2007;282(51):36777–36781. doi: 10.1074/jbc.R700036200. [DOI] [PubMed] [Google Scholar]
  • 91.Inai K, Burnside JL, Hoffman S, Toole BP, Sugi Y. BMP-2 induces versican and hyaluronan that contribute to post-EMT AV cushion cell migration. PLoS ONE. 2013;8(10) doi: 10.1371/journal.pone.0077593.e77593 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Hirose Y, Saijou E, Sugano Y, et al. Inhibition of Stabilin-2 elevates circulating hyaluronic acid levels and prevents tumor metastasis. Proceedings of the National Academy of Sciences of the United States of America. 2012;109(11):4263–4268. doi: 10.1073/pnas.1117560109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Lien HC, Lee YH, Jeng YM, Lin CH, Lu YS, Yao YT. Differential expression of hyaluronan synthase 2 in breast carcinoma and its biological significance. Histopathology. 2014 doi: 10.1111/his.12390. [DOI] [PubMed] [Google Scholar]
  • 94.Kung CI, Chen C, Yang C, Lin C, Chen T, Wang H. Enhanced membrane-type 1 matrix metalloproteinase expression by hyaluronan oligosaccharides in breast cancer cells facilitates CD44 cleavage and tumor cell migration. Oncology Reports. 2012;28(5):1808–1814. doi: 10.3892/or.2012.1993. [DOI] [PubMed] [Google Scholar]
  • 95.Hamilton SR, Fard SF, Paiwand FF, et al. The hyaluronan receptors CD44 and Rhamm (CD168) form complexes with ERK1,2 that sustain high basal motility in breast cancer cells. The Journal of Biological Chemistry. 2007;282(22):16667–16680. doi: 10.1074/jbc.M702078200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Wang Z, Wu Y, Wang H, et al. Interplay of mevalonate and Hippo pathways regulates RHAMM transcription via YAP to modulate breast cancer cell motility. Proceedings of the National Academy of Sciences of the United States of America. 2014;111(1):E89–E98. doi: 10.1073/pnas.1319190110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Tammi RH, Kultti A, Kosma V, Pirinen R, Auvinen P, Tammi MI. Hyaluronan in human tumors: pathobiological and prognostic messages from cell-associated and stromal hyaluronan. Seminars in Cancer Biology. 2008;18(4):288–295. doi: 10.1016/j.semcancer.2008.03.005. [DOI] [PubMed] [Google Scholar]
  • 98.Stern R, Asari AA, Sugahara KN. Hyaluronan fragments: an information-rich system. European Journal of Cell Biology. 2006;85(8):699–715. doi: 10.1016/j.ejcb.2006.05.009. [DOI] [PubMed] [Google Scholar]
  • 99.Bollyky PL, Bogdani M, Bollyky JB, Hull RL, Wight TN. The role of hyaluronan and the extracellular matrix in islet inflammation and immune regulation. Current Diabetes Reports. 2012;12(5):471–480. doi: 10.1007/s11892-012-0297-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Tian X, Azpurua J, Hine C, et al. High-molecular-mass hyaluronan mediates the cancer resistance of the naked mole rat. Nature. 2013;499(7458):346–349. doi: 10.1038/nature12234. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Esser PR, Wölfle U, Dürr C, et al. Contact sensitizers induce skin inflammation via ROS production and hyaluronic acid degradation. PLoS ONE. 2012;7(7) doi: 10.1371/journal.pone.0041340.e41340 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Monzon ME, Fregien N, Schmid N, et al. Reactive oxygen species and hyaluronidase 2 regulate airway epithelial hyaluronan fragmentation. Journal of Biological Chemistry. 2010;285(34):26126–26134. doi: 10.1074/jbc.M110.135194. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Gao F, Koenitzer JR, Tobolewski JM, et al. Extracellular superoxide dismutase inhibits inflammation by preventing oxidative fragmentation of hyaluronan. The Journal of Biological Chemistry. 2008;283(10):6058–6066. doi: 10.1074/jbc.M709273200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Rees MD, Hawkins CL, Davies MJ. Hypochlorite and superoxide radicals can act synergistically to induce fragmentation of hyaluronan and chondroitin sulphates. Biochemical Journal. 2004;381, part 1:175–184. doi: 10.1042/BJ20040148. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Termeer CC, Hennies J, Voith U, et al. Oligosaccharides of hyaluronan are potent activators of dendritic cells. Journal of Immunology. 2000;165(4):1863–1870. doi: 10.4049/jimmunol.165.4.1863. [DOI] [PubMed] [Google Scholar]
  • 106.Tolg C, Hamilton SR, Zalinska E, et al. A RHAMM mimetic peptide blocks hyaluronan signaling and reduces inflammation and fibrogenesis in excisional skin wounds. The American Journal of Pathology. 2012;181(4):1250–1270. doi: 10.1016/j.ajpath.2012.06.036. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Yuan H, Tank M, Alsofyani A. Molecular mass dependence of hyaluronan detection by sandwich ELISA-like assay and membrane blotting using biotinylated hyaluronan binding protein. Glycobiology. 2013;23(11):1270–1280. doi: 10.1093/glycob/cwt064. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Ruppert SM, Hawn TR, Arrigoni A, Wight TN, Bollyky PL. Tissue integrity signals communicated by high-molecular weight hyaluronan and the resolution of inflammation. Immunologic Research. 2014;58(2-3):186–192. doi: 10.1007/s12026-014-8495-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Urakawa H, Nishida Y, Knudson W, et al. Therapeutic potential of hyaluronan oligosaccharides for bone metastasis of breast cancer. Journal of Orthopaedic Research. 2012;30(4):662–672. doi: 10.1002/jor.21557. [DOI] [PubMed] [Google Scholar]
  • 110.Lennon FE, Singleton PA. Hyaluronan regulation of vascular integrity. American Journal of Cardiovascular Disease. 2011;1(3):200–213. [PMC free article] [PubMed] [Google Scholar]
  • 111.Price RD, Myers S, Leigh IM, Navsaria HA. The role of hyaluronic acid in wound healing: assessment of clinical evidence. American Journal of Clinical Dermatology. 2005;6(6):393–402. doi: 10.2165/00128071-200506060-00006. [DOI] [PubMed] [Google Scholar]
  • 112.Hrabarova E, Juranek I, Soltes L. Pro-oxidative effect of peroxynitrite regarding biological systems: a special focus on high-molar-mass hyaluronan degradation. General Physiology and Biophysics. 2011;30(3):223–238. doi: 10.4149/gpb_2011_03_223. [DOI] [PubMed] [Google Scholar]
  • 113.Šoltés L, Mendichi R, Kogan G, Schiller J, Stankovská M, Arnhold J. Degradative action of reactive oxygen species on hyaluronan. Biomacromolecules. 2006;7(3):659–668. doi: 10.1021/bm050867v. [DOI] [PubMed] [Google Scholar]
  • 114.Scheibner KA, Lutz MA, Boodoo S, Fenton MJ, Powell JD, Horton MR. Hyaluronan fragments act as an endogenous danger signal by engaging TLR2. The Journal of Immunology. 2006;177(2):1272–1281. doi: 10.4049/jimmunol.177.2.1272. [DOI] [PubMed] [Google Scholar]
  • 115.de la Motte CA. Hyaluronan in intestinal homeostasis and inflammation: implications for fibrosis. American Journal of Physiology—Gastrointestinal and Liver Physiology. 2011;301(6):G945–G949. doi: 10.1152/ajpgi.00063.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Kuipers MT, van der Poll T, Schultz MJ, Wieland CW. Bench-to-bedside review: damage-associated molecular patterns in the onset of ventilator-induced lung injury. Critical Care. 2011;15(6, article 235) doi: 10.1186/cc10437. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Powell JD, Horton MR. Threat matrix: low-molecular-weight hyaluronan (HA) as a danger signal. Immunologic Research. 2005;31(3):207–218. doi: 10.1385/IR:31:3:207. [DOI] [PubMed] [Google Scholar]
  • 118.Kishi K, Okabe K, Shimizu R, Kubota Y. Fetal skin possesses the ability to regenerate completely: Complete regeneration of skin. Keio Journal of Medicine. 2012;61(4):101–108. doi: 10.2302/kjm.2011-0002-ir. [DOI] [PubMed] [Google Scholar]
  • 119.Slevin M, Krupinski J, Gaffney J, et al. Hyaluronan-mediated angiogenesis in vascular disease: uncovering RHAMM and CD44 receptor signaling pathways. Matrix Biology. 2007;26(1):58–68. doi: 10.1016/j.matbio.2006.08.261. [DOI] [PubMed] [Google Scholar]
  • 120.Wight TN, Potter-Perigo S. The extracellular matrix: an active or passive player in fibrosis? The American Journal of Physiology - Gastrointestinal and Liver Physiology. 2011;301(6):G950–G955. doi: 10.1152/ajpgi.00132.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Jiang D, Liang J, Noble PW. Regulation of non-infectious lung injury, inflammation, and repair by the extracellular matrix glycosaminoglycan hyaluronan. Anatomical Record. 2010;293(6):982–985. doi: 10.1002/ar.21102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Gao F, Liu Y, He Y, et al. Hyaluronan oligosaccharides promote excisional wound healing through enhanced angiogenesis. Matrix Biology. 2010;29(2):107–116. doi: 10.1016/j.matbio.2009.11.002. [DOI] [PubMed] [Google Scholar]
  • 123.Gao F, Yang CX, Mo W, Liu YW, He YQ. Hyaluronan oligosaccharides are potential stimulators to angiogenesis via RHAMM mediated signal pathway in wound healing. Clinical and Investigative Medicine. 2008;31(3):E106–E116. doi: 10.25011/cim.v31i3.3467. [DOI] [PubMed] [Google Scholar]
  • 124.Anderegg U, Simon JC, Averbeck M. More than just a filler—the role of hyaluronan for skin homeostasis. Experimental Dermatology. 2014;23(5):295–303. doi: 10.1111/exd.12370. [DOI] [PubMed] [Google Scholar]
  • 125.Moustakas A, Heldin P. TGFβ and matrix-regulated epithelial to mesenchymal transition. Biochimica et Biophysica Acta. 2014;1840(8):2621–2634. doi: 10.1016/j.bbagen.2014.02.004. [DOI] [PubMed] [Google Scholar]
  • 126.Tammi R, Agren UM, Tuhkanen AL, Tammi M. Hyaluronan metabolism in skin. Progress in Histochemistry and Cytochemistry. 1994;29(2):1–81. doi: 10.1016/s0079-6336(11)80023-9. [DOI] [PubMed] [Google Scholar]
  • 127.Wang B. Photoaging: a review of current concepts of pathogenesis. Journal of Cutaneous Medicine and Surgery. 2011;15(supplement 1):S374–S377. doi: 10.2310/7750.2011.00006. [DOI] [PubMed] [Google Scholar]
  • 128.Leung A, Crombleholme TM, Keswani SG. Fetal wound healing: Implications for minimal scar formation. Current Opinion in Pediatrics. 2012;24(3):371–378. doi: 10.1097/MOP.0b013e3283535790. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Pasonen-Seppänen S, Hyttinen JMT, Rilla K, et al. Role of CD44 in the organization of keratinocyte pericellular hyaluronan. Histochemistry and Cell Biology. 2012;137(1):107–120. doi: 10.1007/s00418-011-0883-2. [DOI] [PubMed] [Google Scholar]
  • 130.Malaisse J, Bourguignon V, de Vuyst E, et al. Hyaluronan Metabolism in Human Keratinocytes and Atopic Dermatitis Skin Is Driven by a Balance of Hyaluronan Synthases 1 and 3. Journal of Investigative Dermatology. 2014;134(8):2174–2182. doi: 10.1038/jid.2014.147. [DOI] [PubMed] [Google Scholar]
  • 131.Kage M, Tokudome Y, Matsunaga Y, Hariya T, Hashimoto F. Effect of hyaluronan tetrasaccharides on epidermal differentiation in normal human epidermal keratinocytes. International Journal of Cosmetic Science. 2014;36(1):109–115. doi: 10.1111/ics.12105. [DOI] [PubMed] [Google Scholar]
  • 132.Choi H, Kang Y, Na J, et al. Oligosaccharides of hyaluronic acid increased epidermal cell stemness by modulation of integrin expression. Journal of Cosmetic Dermatology. 2012;11(4):290–296. doi: 10.1111/jocd.12009. [DOI] [PubMed] [Google Scholar]
  • 133.Colello D, Mathew S, Ward R, Pumiglia K, LaFlamme SE. Integrins regulate microtubule nucleating activity of centrosome through mitogen-activated protein kinase/extracellular signal-regulated kinase kinase/extracellular signal-regulated kinase (MEK/ERK) signaling. The Journal of Biological Chemistry. 2012;287(4):2520–2530. doi: 10.1074/jbc.M111.254128. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Tammi RH, Tammi MI. Hyaluronan accumulation in wounded epidermis: a mediator of keratinocyte activation. The Journal of Investigative Dermatology. 2009;129(8):1858–1860. doi: 10.1038/jid.2009.91. [DOI] [PubMed] [Google Scholar]
  • 135.Xu Y, Baker D, Quan T, Baldassare JJ, Voorhees JJ, Fisher GJ. Receptor type protein tyrosine phosphatase-κ mediates cross-talk between transforming growth factor-β and epidermal growth factor receptor signaling pathways in human keratinocytes. Molecular Biology of the Cell. 2010;21(1):29–35. doi: 10.1091/mbc.E09-08-0710. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Mancini ML, Sonis ST. Mechanisms of cellular fibrosis associated with cancer regimen-related toxicities. Frontiers in Pharmacology. 2014;5:p. 51. doi: 10.3389/fphar.2014.00051. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Ashcroft GS, Yang X, Glick AB, et al. Mice lacking Smad3 show accelerated wound healing and an impaired local inflammatory response. Nature Cell Biology. 1999;1(5):260–266. doi: 10.1038/12971. [DOI] [PubMed] [Google Scholar]
  • 138.Midgley AC, Rogers M, Hallett MB, et al. Transforming growth factor-β1 (TGF-β1)-stimulated fibroblast to myofibroblast differentiation is mediated by hyaluronan (HA)-facilitated epidermal growth factor receptor (EGFR) and CD44 co-localization in lipid rafts. Journal of Biological Chemistry. 2013;288(21):14824–14838. doi: 10.1074/jbc.M113.451336. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139.Webber J, Jenkins RH, Meran S, Phillips A, Steadman R. Modulation of TGFβ1-dependent myofibroblast differentiation by hyaluronan. The American Journal of Pathology. 2009;175(1):148–160. doi: 10.2353/ajpath.2009.080837. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Hu MS, Maan ZN, Wu JC, et al. Tissue engineering and regenerative repair in wound healing. Annals of Biomedical Engineering. 2014;42(7):1494–1507. doi: 10.1007/s10439-014-1010-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Zgheib C, Xu J, Liechty KW. Targeting inflammatory cytokines and extracellular matrix composition to promote wound regeneration. Advances in Wound Care. 2014;3(4):344–355. doi: 10.1089/wound.2013.0456. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Orimo A, Weinberg RA. Stromal fibroblasts in cancer: a novel tumor-promoting cell type. Cell Cycle. 2006;5(15):1597–1601. doi: 10.4161/cc.5.15.3112. [DOI] [PubMed] [Google Scholar]
  • 143.Caskey RC, Allukian M, Lind RC, et al. Lentiviral-mediated over-expression of hyaluronan synthase-1 (HAS-1) decreases the cellular inflammatory response and results in regenerative wound repair. Cell and Tissue Research. 2013;351(1):117–125. doi: 10.1007/s00441-012-1504-7. [DOI] [PubMed] [Google Scholar]
  • 144.Prosdocimi M, Bevilacqua C. Exogenous hyaluronic acid and wound healing: an updated vision. Panminerva Medica. 2012;54(2):129–135. [PubMed] [Google Scholar]
  • 145.Campo GM, Avenoso A, D'Ascola A, et al. Inhibition of hyaluronan synthesis reduced inflammatory response in mouse synovial fibroblasts subjected to collagen-induced arthritis. Archives of Biochemistry and Biophysics. 2012;518(1):42–52. doi: 10.1016/j.abb.2011.12.005. [DOI] [PubMed] [Google Scholar]
  • 146.Ghazi K, Deng-Pichon U, Warnet J, Rat P. Hyaluronan fragments improve wound healing on in vitro cutaneous model through P2X7 purinoreceptor basal activation: role of molecular weight. PLoS ONE. 2012;7(11) doi: 10.1371/journal.pone.0048351.e48351 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.de La Motte C, Nigro J, Vasanji A, et al. Platelet-derived hyaluronidase 2 cleaves hyaluronan into fragments that trigger monocyte-mediated production of proinflammatory cytokines. American Journal of Pathology. 2009;174(6):2254–2264. doi: 10.2353/ajpath.2009.080831. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Ferguson EL, Roberts JL, Moseley R, Griffiths PC, Thomas DW. Evaluation of the physical and biological properties of hyaluronan and hyaluronan fragments. International Journal of Pharmaceutics. 2011;420(1):84–92. doi: 10.1016/j.ijpharm.2011.08.031. [DOI] [PubMed] [Google Scholar]
  • 149.David-Raoudi M, Tranchepain F, Deschrevel B, et al. Differential effects of hyaluronan and its fragments on fibroblasts: relation to wound healing. Wound Repair and Regeneration. 2008;16(2):274–287. doi: 10.1111/j.1524-475X.2007.00342.x. [DOI] [PubMed] [Google Scholar]
  • 150.Kaya G, Tran C, Sorg O, et al. Hyaluronate fragments reverse skin atrophy by a CD44-dependent mechanism. PLoS Medicine. 2006;3(12, article e493):2291–2303. doi: 10.1371/journal.pmed.0030493. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151.Wernicke M, Piñeiro LC, Caramutti D, et al. Breast cancer stromal myxoid changes are associated with tumor invasion and metastasis: a central role for hyaluronan. Modern Pathology. 2003;16(2):99–107. doi: 10.1097/01.MP.0000051582.75890.2D. [DOI] [PubMed] [Google Scholar]
  • 152.Takahashi Y, Kuwabara H, Yoneda M, Isogai Z, Tanigawa N, Shibayama Y. Versican G1 and G3 domains are upregulated and latent transforming growth factor-β binding protein-4 is downregulated in breast cancer stroma. Breast Cancer. 2012;19(1):46–53. doi: 10.1007/s12282-011-0264-7. [DOI] [PubMed] [Google Scholar]
  • 153.Costea DE, Hills A, Osman AH, et al. Identification of two distinct carcinoma-associated fibroblast subtypes with differential tumor-promoting abilities in oral squamous cell carcinoma. Cancer Research. 2013;73(13):3888–3901. doi: 10.1158/0008-5472.CAN-12-4150. [DOI] [PubMed] [Google Scholar]
  • 154.Koyama H, Hibi T, Isogai Z, et al. Hyperproduction of hyaluronan in Neu-induced mammary tumor accelerates angiogenesis through stromal cell recruitment: Possible involvement of versican/PG-M. The American Journal of Pathology. 2007;170(3):1086–1099. doi: 10.2353/ajpath.2007.060793. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Koyama H, Kobayashi N, Harada M, et al. Significance of tumor-associated stroma in promotion of intratumoral lymphangiogenesis: pivotal role of a hyaluronan-rich tumor microenvironment. American Journal of Pathology. 2008;172(1):179–193. doi: 10.2353/ajpath.2008.070360. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Bono P, Wasenius V, Heikkilä P, Lundin J, Jackson DG, Joensuu H. High LYVE-1-positive lymphatic vessel numbers are associated with poor outcome in breast cancer. Clinical Cancer Research. 2004;10(21):7144–7149. doi: 10.1158/1078-0432.CCR-03-0826. [DOI] [PubMed] [Google Scholar]
  • 157.Kobayashi N, Miyoshi S, Mikami T, et al. Hyaluronan deficiency in tumor stroma impairs macrophage trafficking and tumor neovascularization. Cancer Research. 2010;70(18):7073–7083. doi: 10.1158/0008-5472.CAN-09-4687. [DOI] [PubMed] [Google Scholar]
  • 158.Auvinen P, Tammi R, Kosma V, et al. Increased hyaluronan content and stromal cell CD44 associate with HER2 positivity and poor prognosis in human breast cancer. International Journal of Cancer. 2013;132(3):531–539. doi: 10.1002/ijc.27707. [DOI] [PubMed] [Google Scholar]
  • 159.Lopez JI, Camenisch TD, Stevens MV, Sands BJ, McDonald J, Schroeder JA. CD44 attenuates metastatic invasion during breast cancer progression. Cancer Research. 2005;65(15):6755–6763. doi: 10.1158/0008-5472.CAN-05-0863. [DOI] [PubMed] [Google Scholar]
  • 160.Louderbough JMV, Schroeder JA. Understanding the dual nature of CD44 in breast cancer progression. Molecular Cancer Research. 2011;9(12):1573–1586. doi: 10.1158/1541-7786.MCR-11-0156. [DOI] [PubMed] [Google Scholar]
  • 161.Montgomery N, Hill A, McFarlane S, et al. CD44 enhances invasion of basal-like breast cancer cells by upregulating serine protease and collagen-degrading enzymatic expression and activity. Breast Cancer Research. 2012;14(3, article R84) doi: 10.1186/bcr3199. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162.Joukov V, Groen AC, Prokhorova T, et al. The BRCA1/BARD1 heterodimer modulates ran-dependent mitotic spindle assembly. Cell. 2006;127(3):539–552. doi: 10.1016/j.cell.2006.08.053. [DOI] [PubMed] [Google Scholar]
  • 163.Hardwick C, Hoare K, Owens R, et al. Molecular cloning of a novel hyaluronan receptor that mediates tumor cell motility. Journal of Cell Biology. 1992;117(6):1343–1350. doi: 10.1083/jcb.117.6.1343. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 164.Yang B, Zhang L, Turley EA. Identification of two hyaluronan-binding domains in the hyaluronan receptor RHAMM. The Journal of Biological Chemistry. 1993;268(12):8617–8623. [PubMed] [Google Scholar]
  • 165.Tolg C, Hamilton SR, Morningstar L, et al. RHAMM promotes interphase microtubule instability and mitotic spindle integrity through MEK1/ERK1/2 activity. The Journal of Biological Chemistry. 2010;285(34):26461–26474. doi: 10.1074/jbc.M110.121491. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 166.Maxwell CA, Keats JJ, Belch AR, Pilarski LM, Reiman T. Receptor for hyaluronan-mediated motility correlates with centrosome abnormalities in multiple myeloma and maintains mitotic integrity. Cancer Research. 2005;65(3):850–860. [PubMed] [Google Scholar]
  • 167.Forteza R, Lieb T, Aoki T, Savani RC, Conner GE, Salathe M. Hyaluronan serves a novel role in airway mucosal host defense. The FASEB Journal. 2001;15(12):2179–2186. doi: 10.1096/fj.01-0036com. [DOI] [PubMed] [Google Scholar]
  • 168.Schoenfelder M, Einspanier R. Expression of hyaluronan synthases and corresponding hyaluronan receptors is differentially regulated during oocyte maturation in cattle. Biology of Reproduction. 2003;69(1):269–277. doi: 10.1095/biolreprod.102.011577. [DOI] [PubMed] [Google Scholar]
  • 169.Kornovski BS, McCoshen J, Kredentser J, Turley E. The regulation of sperm motility by a novel hyaluronan receptor. Fertility and Sterility. 1994;61(5):935–940. doi: 10.1016/s0015-0282(16)56709-0. [DOI] [PubMed] [Google Scholar]
  • 170.Park D, Kim Y, Kim H, et al. Hyaluronic acid promotes angiogenesis by inducing RHAMM-TGFβ receptor interaction via CD44-PKCδ . Molecules and Cells. 2012;33(6):563–574. doi: 10.1007/s10059-012-2294-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 171.Manzanares D, Monzon M, Savani RC, Salathe M. Apical oxidative hyaluronan degradation stimulates airway ciliary beating via RHAMM and RON. The American Journal of Respiratory Cell and Molecular Biology. 2007;37(2):160–168. doi: 10.1165/rcmb.2006-0413OC. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 172.Amara FM, Entwistle J, Kuschak TI, Turley EA, Wright JA. Transforming growth factor-β1 stimulates multiple protein interactions at a unique cis-element in the 3′-untranslated region of the hyaluronan receptor RHAMM mRNA. The Journal of Biological Chemistry. 1996;271(25):15279–15284. doi: 10.1074/jbc.271.25.15279. [DOI] [PubMed] [Google Scholar]
  • 173.Samuel SK, Hurta RAR, Spearman MA, Wright JA, Turley EA, Greenberg AH. TGF-β1 stimulation of cell locomotion utilizes the hyaluronan receptor RHAMM and hyaluronan. Journal of Cell Biology. 1993;123(3):749–758. doi: 10.1083/jcb.123.3.749. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 174.Buganim Y, Rotter V. RHAMM in the complex p53 cell cycle network. Cell Cycle. 2008;7(21):3287–3291. [Google Scholar]
  • 175.Godar S, Weinber RA. Filling the mosaic of p53 actions: p53 represses RHAMM expression. Cell Cycle. 2008;7(22):p. 3479. doi: 10.4161/cc.7.22.7320. [DOI] [PubMed] [Google Scholar]
  • 176.Sohr S, Engeland K. RHAMM is differentially expressed in the cell cycle and downregulated by the tumor suppressor p53. Cell Cycle. 2008;7(21):3448–3460. doi: 10.4161/cc.7.21.7014. [DOI] [PubMed] [Google Scholar]
  • 177.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(44):6873–6882. doi: 10.1038/sj.onc.1206811. [DOI] [PubMed] [Google Scholar]
  • 178.Greiner J, Schmitt A, Giannopoulos K, et al. High-dose RHAMM-R3 peptide vaccination for patients with acute myeloid leukemia, myelodysplastic syndrome and multiple myeloma. Haematologica. 2010;95(7):1191–1197. doi: 10.3324/haematol.2009.014704. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 179.Schmitt M, Schmitt A, Rojewski MT, et al. RHAMM-R3 peptide vaccination in patients with acute myeloid leukemia, myelodysplastic syndrome, and multiple myeloma elicits immunologic and clinical responses. Blood. 2008;111(3):1357–1365. doi: 10.1182/blood-2007-07-099366. [DOI] [PubMed] [Google Scholar]
  • 180.Smith TC, Fang Z, Luna EJ. Novel interactors and a role for Supervillin in early cytokinesis. Cytoskeleton. 2010;67(6):346–364. doi: 10.1002/cm.20449. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 181.Savani RC, Wang C, Yang B, et al. Migration of bovine aortic smooth muscle cells after wounding injury. The role of hyaluronan and RHAMM. Journal of Clinical Investigation. 1995;95(3):1158–1168. doi: 10.1172/JCI117764. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 182.Zhang S, Chang MCY, Zylka D, Turley S, Harrison R, Turley EA. The hyaluronan receptor RHAMM regulates extracellular-regulated kinase. Journal of Biological Chemistry. 1998;273(18):11342–11348. doi: 10.1074/jbc.273.18.11342. [DOI] [PubMed] [Google Scholar]
  • 183.Hatano H, Shigeishi H, Kudo Y, et al. RHAMM/ERK interaction induces proliferative activities of cementifying fibroma cells through a mechanism based on the CD44-EGFR. Laboratory Investigation. 2011;91(3):379–391. doi: 10.1038/labinvest.2010.176. [DOI] [PubMed] [Google Scholar]
  • 184.Du YN, Chou C, Klimstra DS, Varmus H. Receptor for hyaluronan-mediated motility isoform B promotes liver metastasis in a mouse model of multistep tumorigenesis and a tail vein assay for metastasis. Proceedings of the National Academy of Sciences of the United States of America. 2011;108(40):16753–16758. doi: 10.1073/pnas.1114022108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 185.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. Journal of Biological Chemistry. 2001;276(39):36770–36778. doi: 10.1074/jbc.M102273200. [DOI] [PubMed] [Google Scholar]
  • 186.Lokeshwar VB, Selzer MG. Differences in hyaluronic acid-mediated functions and signaling in arterial, microvessel, and vein-derived human endothelial cells. Journal of Biological Chemistry. 2000;275(36):27641–27649. doi: 10.1074/jbc.M003084200. [DOI] [PubMed] [Google Scholar]
  • 187.Hall CL, Wang FS, Turley EA. Src-/- fibroblasts are defective in their ability to disassemble focal adhesions in response to phorbol ester/hyaluronan treatment. Cell Communication and Adhesion. 2002;9(5-6):273–283. doi: 10.1080/15419060216306. [DOI] [PubMed] [Google Scholar]
  • 188.Kouvidi K, Berdiaki A, Nikitovic D, et al. Role of receptor for hyaluronic acid-mediated motility (RHAMM) in low molecular weight hyaluronan (LMWHA)-mediated fibrosarcoma cell adhesion. Journal of Biological Chemistry. 2011;286(44):38509–38520. doi: 10.1074/jbc.M111.275875. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 189.Mohapatra S, Yang X, Wright JA, Turley EA, Greenberg AH. Soluble hyaluronan receptor RHAMM induces mitotic arrest by suppressing Cdc2 and cyclin B1 expression. The Journal of Experimental Medicine. 1996;183(4):1663–1668. doi: 10.1084/jem.183.4.1663. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 190.Telmer PEA. RHAMM is required for MMP-9 expression and release from podosomes 10. submitted, 2014. [Google Scholar]
  • 191.Groen AC, Cameron LA, Coughlin M, Miyamoto DT, Mitchison TJ, Ohi R. XRHAMM functions in ran-dependent microtubule nucleation and pole formation during anastral spindle assembly. Current Biology. 2004;14(20):1801–1811. doi: 10.1016/j.cub.2004.10.002. [DOI] [PubMed] [Google Scholar]
  • 192.Chen H, Mohan P, Jiang J, et al. Spatial regulation of Aurora A activity during mitotic spindle assembly requires RHAMM to correctly localize TPX2. Cell Cycle. 2014;13(14):2248–2261. doi: 10.4161/cc.29270. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 193.Smith TC, Fridy PC, Li Y, et al. Supervillin binding to myosin II and synergism with anillin are required for cytokinesis. Molecular Biology of the Cell. 2013;24(23):3603–3619. doi: 10.1091/mbc.E12-10-0714. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 194.Gangopadhyay SS, Takizawa N, Gallant C, et al. Smooth muscle archvillin: a novel regulator of signaling and contractility in vascular smooth muscle. Journal of Cell Science. 2004;117(21):5043–5057. doi: 10.1242/jcs.01378. [DOI] [PubMed] [Google Scholar]
  • 195.Hasegawa H, Hyodo T, Asano E, et al. The role of PLK1-phosphorylated SVIL in myosin II activation and cytokinetic furrowing. Journal of Cell Science. 2013;126(16):3627–3637. doi: 10.1242/jcs.124818. [DOI] [PubMed] [Google Scholar]
  • 196.Harrison RE, Turley EA. Active erk regulates microtubule stability in H-ras-transformed cells. Neoplasia. 2001;3(5):385–394. doi: 10.1038/sj.neo.7900180. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 197.Silverman-Gavrila R, Silverman-Gavrila L, Hou G, Zhang M, Charlton M, Bendeck MP. Rear polarization of the microtubule-organizing center in neointimal smooth muscle cells depends on PKCα, ARPC5, and RHAMM. The American Journal of Pathology. 2011;178(2):895–910. doi: 10.1016/j.ajpath.2010.10.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 198.Gouëffic Y, Guilluy C, Guérin P, Patra P, Pacaud P, Loirand G. Hyaluronan induces vascular smooth muscle cell migration through RHAMM-mediated PI3K-dependent Rac activation. Cardiovascular Research. 2006;72(2):339–348. doi: 10.1016/j.cardiores.2006.07.017. [DOI] [PubMed] [Google Scholar]
  • 199.Evanko SP, Parks T, Wight TN. Intracellular hyaluronan in arterial smooth muscle cells: association with microtubules, RHAMM, and the mitotic spindle. Journal of Histochemistry and Cytochemistry. 2004;52(12):1525–1535. doi: 10.1369/jhc.4A6356.2004. [DOI] [PubMed] [Google Scholar]
  • 200.Bhatia P, Menigatti M, Brocard M, Morley S, Ferrari S. Mitotic DNA damage targets the Aurora A/TPX2 complex. Cell Cycle. 2010;9(22):4592–4599. doi: 10.4161/cc.9.22.13665. [DOI] [PubMed] [Google Scholar]
  • 201.Telmer PG, Tolg C, McCarthy JB, Turley EA. How does a protein with dual mitotic spindle and extracellular matrix receptor functions affect tumor susceptibility and progression? Communicative and Integrative Biology. 2011;4(2):182–185. doi: 10.4161/cib.4.2.14270. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 202.Maxwell CA, Rasmussen E, Zhan F, et al. RHAMM expression and isoform balance predict aggressive disease and poor survival in multiple myeloma. Blood. 2004;104(4):1151–1158. doi: 10.1182/blood-2003-11-4079. [DOI] [PubMed] [Google Scholar]
  • 203.Zhou R, Wu X, Skalli O. The hyaluronan receptor RHAMM/IHABP in astrocytoma cells: Expression of a tumor-specific variant and association with microtubules. Journal of Neuro-Oncology. 2002;59(1):15–26. doi: 10.1023/a:1016373015569. [DOI] [PubMed] [Google Scholar]
  • 204.Lynn BD, Li X, Cattini PA, Turley EA, Nagy JI. Identification of sequence, protein isoforms, and distribution of the hyaluronan-binding protein RHAMM in adult and developing rat brain. Journal of Comparative Neurology. 2001;439(3):315–330. doi: 10.1002/cne.1353. [DOI] [PubMed] [Google Scholar]
  • 205.Crainie M, Belch AR, Mant MJ, Pilarski LM. Overexpression of the receptor for hyaluronan-mediated motility (RHAMM) characterizes the malignant clone in multiple myeloma: identification of three distinct RHAMM variants. Blood. 1999;93(5):1684–1696. [PubMed] [Google Scholar]
  • 206.Mohan P, Castellsague J, Jiang J, et al. Genomic imbalance of HMMR/RHAMM regulates the sensitivity and response of malignant peripheral nerve sheath tumour cells to aurora kinase inhibition. Oncotarget. 2013;4(1):80–93. doi: 10.18632/oncotarget.793. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 207.Assmann V, Jenkinson D, Marshall JF, Hart IR. The intracellular hyaluronan receptor RHAMM/IHABP interacts with microtubules and actin filaments. Journal of Cell Science. 1999;112(part 22):3943–3954. doi: 10.1242/jcs.112.22.3943. [DOI] [PubMed] [Google Scholar]
  • 208.Dunsch AK, Hammond D, Lloyd J, Schermelleh L, Gruneberg U, Barr FA. Dynein light chain 1 and a spindle-associated adaptor promote dynein asymmetry and spindle orientation. Journal of Cell Biology. 2012;198(6):1039–1054. doi: 10.1083/jcb.201202112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 209.Zaman A, Cui Z, Foley JP, et al. Expression and role of the hyaluronan receptor RHAMM in inflammation after bleomycin injury. The American Journal of Respiratory Cell and Molecular Biology. 2005;33(5):447–454. doi: 10.1165/rcmb.2004-0333OC. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 210.Lennon FE, Singleton PA. Role of hyaluronan and hyaluronan-binding proteins in lung pathobiology. American Journal of Physiology—Lung Cellular and Molecular Physiology. 2011;301(2):L137–L147. doi: 10.1152/ajplung.00071.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 211.Matou-Nasri S, Gaffney J, Kumar S, Slevin M. Oligosaccharides of hyaluronan induce angiogenesis through distinct CD44 and RHAMM-mediated signalling pathways involving Cdc2 and γ-adducin. International Journal of Oncology. 2009;35(4):761–773. doi: 10.3892/ijo_00000389. [DOI] [PubMed] [Google Scholar]
  • 212.Lovvorn HN, III, Cass DL, Sylvester KG, et al. Hyaluronan receptor expression increases in fetal excisional skin wounds and correlates with fibroplasia. Journal of Pediatric Surgery. 1998;33(7):1062–1070. doi: 10.1016/s0022-3468(98)90532-2. [DOI] [PubMed] [Google Scholar]
  • 213.Mantripragada KK, Spurlock G, Kluwe L, et al. High-resolution DNA copy number profiling of malignant peripheral nerve sheath tumors using targeted microarray-based comparative genomic hybridization. Clinical Cancer Research. 2008;14(4):1015–1024. doi: 10.1158/1078-0432.CCR-07-1305. [DOI] [PubMed] [Google Scholar]
  • 214.Shi Y, Reiman T, Li W, et al. Targeting aurora kinases as therapy in multiple myeloma. Blood. 2007;109(9):3915–3921. doi: 10.1182/blood-2006-07-037671. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 215.Tilghman J, Wu H, Sang Y, et al. HMMR maintains the stemness and tumorigenicity of glioblastoma stem-like cells. Cancer Research. 2014;74(11):3168–3179. doi: 10.1158/0008-5472.CAN-13-2103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 216.Rizzardi AE, Rosener NK, Koopmeiners JS, et al. Evaluation of protein biomarkers of prostate cancer aggressiveness. BMC Cancer. 2014;14(1, article 244) doi: 10.1186/1471-2407-14-244. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 217.Rizzardi AE, Vogel R, Koopmeiners JS, et al. Elevated hyaluronan and hyaluronan-mediated motility receptor are associated with biochemical failure in patients with intermediate-grade prostate tumors. Cancer. 2014;120(12):1800–1809. doi: 10.1002/cncr.28646. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 218.Wang C, Thor AD, Moore DH, II, et al. The overexpression of RHAMM, a hyaluronan-binding protein that regulates ras signaling, correlates with overexpression of mitogen-activated protein kinase and is a significant parameter in breast cancer progression. Clinical Cancer Research. 1998;4(3):567–576. [PubMed] [Google Scholar]

Articles from BioMed Research International are provided here courtesy of Wiley

RESOURCES