Hyaluronan metabolism in remodeling extracellular matrix: probes for imaging and therapy of breast cancer

M. Veiseh *a and E. A. Turley *b
aLife Sciences Division, Lawrence Berkeley National Laboratories, Berkeley, CA US. E-mail: mveiseh@lbl.gov; Fax: +1 510-486-5586; Tel: +1 510-486-4368
bLondon Regional Cancer Program, Depts. Oncology and Biochemistry, University of Western Ontario, London, ON CA. E-mail: eva.turley@lhsc.on.ca; Fax: +1 519-685-8616; Tel: +1 519-685-8600 ×53677

Received 3rd September 2010 , Accepted 13th December 2010

First published on 24th January 2011


Abstract

Clinical and experimental evidence increasingly support the concept of cancer as a disease that emulates a component of wound healing, in particular abnormal stromal extracellular matrix remodeling. Here we review the biology and function of one remodeling process, hyaluronan (HA) metabolism, which is essential for wound resolution but closely linked to breast cancer (BCA) progression. Components of the HA metabolic cycle (HAS2, SPAM1 and HA receptors CD44, RHAMM/HMMR and TLR2) are discussed in terms of their known functions in wound healing and in breast cancer progression. Finally, we discuss recent advances in the use of HA-based platforms for developing nanoprobes to image areas of active HA metabolism and for therapeutics in breast cancer.



                  M. Veiseh

M. Veiseh

Mandana Veiseh is a Ruth L. Kirschste National Research Service Award fellow in the laboratory of Dr Mina J. Bissell at Lawrence Berkeley National Laboratory (LBNL). She has received a dual PhD in Materials Science & Engineering and Nanotechnology from the University of Washington. She was an Advanced Light Source doctoral fellow at LBNL prior to pursuing a postdoctoral study at the Fred Hutchinson Cancer Research Center. Her research experiences include: synthesis and characterization of cancer-targeted imaging probes using microenvironmental ligands; surface molecular engineering on bio-micro-electro-mechanical systems for micro/nano patterning; device development for cell-based biosensing, and multiplexed delivery/screening of cancer therapeutics.


                  E. A. Turley

E. A. Turley

Eva Turley is a Distinguished Oncology Scientist and Professor at the London Health Sciences center, London Regional Cancer Program and Depts. Biochemistry and Oncology at the University of Western Ontario. She received her PhD at the University of British Columbia in cell biology and conducted postdoctoral training at Johns Hopkins University in the area of glycan biology. She pioneered the area of hyaluronan signaling and cloned/characterized RHAMM/HMMR. Her current research interests include defining the roles of hyaluronan receptors, and the signaling pathways they regulate, in controlling breast tumor cell migration and genomic stability. She and her collaborators are actively involved in developing HA-based probes for imaging and treating sites of aberrant hyaluronan metabolism.



Insight, innovation, integration

Progressing tumors actively remodel their microenvironment like repairing wounds. The mechanisms by which these remodeling processes contribute to tumor progression are still poorly understood. This article reviews the evidence that hyaluronan metabolism, which is one remodeling process that is essential for wound resolution and contributes to breast cancer progression. The role that the breakdown of hyaluronan into bioactive fragments plays in innate immunity and in neoangiogenesis and the impact they have on properties of tumor cells (e.g. resistance to apoptosis and enhanced migration) is discussed. The innovative translation of this basic cell biology knowledge to the development of clinically useful hyaluronan-based co-polymers, pro-drugs, nanoprobes and hydrogels is discussed. The application of these novel reagents to imaging and therapeutics in breast cancer is described.

Introduction

The extracellular matrix (ECM) is an important component of tissue microenvironments, which provide biophysical and biochemical cues that maintain tissue architecture integrity.1 Modification of these signals appears to be an integral factor in promoting the progression of cancers arising from oncogenic mutations. For instance, forced expression of stromelysin I, an extracellular collagenase, is sufficient to initiate mammary tumors.2 In some cases, the effects of the microenvironment on tumor progression can even dominate genetic aberrations in controlling aggressiveness of cancer cells.3 Examples include blocking signaling through β-1 integrin extracellular matrix receptors or HA receptors,4 such as receptor for hyaluronan mediated motility (RHAMM), which cause reversion of tumor cells to a non-tumorigenic state.4c These and other studies (reviewed in ref. 5) have provided compelling experimental evidence that an aberrant tumor ECM is a factor in sustaining tumor initiation and progression. Clinical evidence also supports this concept: epidemiological studies have linked chronic inflammation to cancer initiation while gene signatures derived from either isolated stromal cells (e.g. activated fibroblasts and macrophages) or tumor-associated stroma have predicted poor patient outcome and relapse.5e,f,6 These studies additionally raise the possibility that a tumor microenvironment exhibits properties of wounds that do not heal,6f,g one aspect of which is a constitutively remodeling ECM.

Prognostic stromal transcriptomes typically contain high levels of ECM genes including collagens, collagenases, integrins, HA synthases, and hyaluronidases.5f,6a,7 HA receptors as well as factors that regulate expression of ECM genes, such as transforming growth factor beta (TGFB's) and insulin-like growth factor 1 (IGF-1), are also commonly modified in prognostic stromal gene signatures.6b,7a In general, stromal environments that actively promote cancer include those that exhibit aspects of response-to-injury processes including chronic inflammation,7a,8 repair following chronic radiation,9 and tissue undergoing involution (e.g. involuting mammary gland).7a,10 Therapeutics targeting specific wound healing processes, such as angiogenesis, have already shown promise in treatment of some cancers.5e,11 Nevertheless, the key ECM remodeling events responsible for promoting cancer aggression and progression are still not well understood. One process common to wound repair, breast cancer progression, and that of many other cancers is an elevated ability to metabolize HA.12

We have developed a method for detecting cells that are actively metabolizing HA and have shown that this property is shared by both normal but injured cells and aggressive BCA cell lines (Fig. 1, and ref. 13). HA metabolism was detected and quantified as the amount of fluorescent, high molecular weight (HMW) HA—referred to as Texas Red-HA (TR-HA)—that can bind to and be taken up by these cultured cells. Quiescent fibroblast monolayers showed little uptake while scratch-wounding these monolayers resulted in high TR-HA uptake at the wound edge (Fig. 1). We verified that the TR-HA probe accurately detected areas of high HA metabolism by demonstrating its rapid and HA-receptor dependent uptake in both liver, which is perhaps the most active of all tissues in metabolizing HA,14 and injured carotid arteries (vs. uninjured contra-lateral carotid artery used as a control), which also exhibited increased HA metabolism.13,15 Using this TR-HA probe as a marker for high HA metabolism, we showed that aggressive BCA lines such as MDA-MB-231 metabolized HA more actively than less aggressive BCA lines such as MCF-7 (Fig. 1). This finding is consistent with previous evidence that MDA-MB-231 tumor cells produced larger amounts of HA, expressed higher levels of injury-induced HA receptors such as cluster designation 44 (CD44) and RHAMM,16 and expressed higher levels of hyaluronidases. These results imply that the TR-HA probe is functional because it detects and accurately reports sites of active HA metabolism.



          HA metabolism in scratch wounds (left images) and in breast cancer cell lines (right images). A. Endogenous HA accumulates at the wound edge (arrows). B. Texas Red-HA is also largely taken up at the wound edge, suggesting that sites of endogenous accumulation also are sites of active HA uptake/metabolism (arrow in right image of black/white panel). A similarly increased uptake of Texas Red-HA is observed in aggressive human BCA lines (e.g. MDA-MB-231 tumor cells), while much lower amounts of HA are taken up by less aggressive human BCA such as MCF-7 tumor cells. Magnification bar = 10 μm. Red color is Texas Red and blue is DAPI.
Fig. 1 HA metabolism in scratch wounds (left images) and in breast cancer cell lines (right images). A. Endogenous HA accumulates at the wound edge (arrows). B. Texas Red-HA is also largely taken up at the wound edge, suggesting that sites of endogenous accumulation also are sites of active HA uptake/metabolism (arrow in right image of black/white panel). A similarly increased uptake of Texas Red-HA is observed in aggressive human BCA lines (e.g. MDA-MB-231 tumor cells), while much lower amounts of HA are taken up by less aggressive human BCA such as MCF-7 tumor cells. Magnification bar = 10 μm. Red color is Texas Red and blue is DAPI.

A number of studies indicate that HA metabolism is a transient but essential process for wound resolution.17 However, our results and those of others (see “Clinical uses of HA metabolic activity” in this review) indicate that it is aberrantly and constitutively active in aggressive breast cancer and likely contributes to progression of this disease rather than its resolution.12a,18 We are currently exploring the flexibility of the HA polymer for developing multi-modality imaging strategies to detect active HA metabolism in remodeling microenvironments of injured and neoplastic tissues in vivo. Here we review evidence that the constitutive activation of an HA metabolic cycle is utilized by breast tumors to establish a microenvironment that is permissive for progression. We also discuss potential uses of functional HA-based platforms such as HA nanoprobes for imaging and characterizing breast cancer cells, particularly with respect to properties that are necessary for disease progression.

The HA metabolic cycle in homeostasis and injury

The metabolism of HA is a complicated, multistep and multifunctional process involving coordinated synthesis of HA by one of three HA synthases (HAS1-3) that produce an extracellular HMW polymer. HA uptake is mediated by HA receptors, and the internalized polymer is degraded by lysosomal hyaluronidases.19 During homeostasis, HMW HA produced in tissues is filtered through lymphatic tissue and then enters the blood vasculature. A small amount of HA produced in tissues is endocytosed and degraded in situ, but most accumulates within lymphatic sinuses. Some of this lymphatic HA binds to and is taken up by HA receptors, lymphatic vessel endothelial hyaluronan receptor (LYVE1), Stabilin1 (STAB1) and Stabilin2 (STAB2), which are located on lymph node sinus endothelial cells.19d,20 The majority of HA escapes from lymphatic sinuses into blood vessels and is rapidly removed by liver endothelial cells as a result of binding to STAB1 and 2 (also known as HA receptor for endocytosis [HARE]) receptors. Kidney tissue participates in HA metabolism but this contribution is minor compared to the amount metabolized by liver tissue.14,20 The homeostatic HA metabolism scenario is markedly modified when tissues are injured and begin to actively remodel their microenvironment.
Table 1 mRNA expression of genes involved in HA metabolism that are upregulated in BCA
Gene name mRNA expression increased in BCA vs. normal (p < 0.05) Expression related to tumor grade and/or poor outcome (p < 0.05)
+ to +++++ is a semiquantitative assessment of the percentage of data sets in which elevated gene expression is related to poor outcome. Poor outcome parameters include relapse, appearance of metastases after treatment, high tumor grade, and death. All data were obtained from Oncomine (www.oncomine.com).
Hyaluronan synthases
HAS2 16% +++
Hyaluronan receptors
RHAMM/HMMR 78% +++++
CD44 11% +++
TLR2 11% ++
Hyaluronidase
SPAM No change but increased within BCA groups +


During response-to-injury, HMW HA polymers are rapidly metabolized within the injured tissue.12a,c,15b,18,21 Local HA production and accumulation are increased as a result of elevated HAS expression in injured cells, and HA is then rapidly fragmented by extracellular hyaluronidases, reactive oxygen species (ROS) and mechanical shearing.19a,22 During wound repair, these LMW HA fragments serve multiple functions including control of innate immune cell function,15b,21a promotion of angiogenesis,17e and regulation of wound cell proliferation, migration and differentiation.23 The functional effects of HA fragments and their clearance from the repairing tissue are the consequences of their binding to HA receptors such as CD44, RHAMM and Toll-Like Receptors 2,4 (TLR2,4).5d,12c,14,17d,18,19c,21b,23b,24 Internalized LMW HA is targeted to the lysosome and further degraded by hyaluronidase 1 and 2 (HYAL1, HYAL2), which are intracellular endoglycosidases, into tetra and hexasaccharides.22b,25 HYAL1 and HYAL2 are the major hyaluronidases expressed in humans and can equally depolymerize HA, chondroitin, and chondroitin sulfate.19d The oligosaccharides produced by HYAL1 and HYAL2 are then further degraded by two lysosomal exoglycosidases, β-glucuronidase (GUSB) and β-N-acetyl hexosaminidase (HEXA).19d,22b,25 A similar process of this injury-specific HA metabolism is constitutively activated in BCA and other tumors.

Functions of HA, HA receptors and HYAL enzymes in homeostasis and tissue repair

The central player of this metabolic cycle, HA, is a relatively simple polysaccharide belonging to the family of glycosaminoglycans that also include chondroitin sulfates, heparins/heparan sulfates, and keratin sulfate.17c,19b,26 The HA polymer is composed of repeating disaccharide units formed by N-acetyl-glucosamine and glucuronic acid. These disaccharide units are linked together by β(1–4) bonds (Fig. 2). The HMW HA polymer is synthesized by dual action plasma membrane glycosyl transferases (HAS1-3, named for their order of discovery)26 that bind both uridine diphospho (UDP)-glucuronic acid and UDP-N-acetylglucosamine at the inner membrane surface. The mechanisms for export of growing HA chains to the microenvironment is not understood, but possibilities include extrusion through intra-protein pores formed from synthase oligomers or via transporters such as the adenosine-5′-triphosphate binding cassette (ABC) system that are known to export other polysaccharides.19b,26,27 Mammalian HA synthases are 55–70% identical but only HAS2 is necessary for embryonic development.19b,c,26 Although the tissue expression patterns of HAS1-3 often overlap, they differ in their rate of synthesis, the size of polymer that they produce, and the type of cancer in which they are aberrantly expressed.19b,28 For example, HAS2 expression is linked to BCA (Table 1),3c,29 while HAS2 and HAS3 expression play a role in prostate cancer aggression.30

            Schematic representation of injury-induced HA metabolic cycle in wounds. The figure shows the unit structure of HA (N-acetyl-glucosamine and β-glucuronic acid) and demonstrates the functions of HA metabolism in BCA and other cancers. HA binding and uptake are tracked by labeling a high molecular weight polymer with a fluorescent dye. HA fragments produced by ROS, shear and extracellular hyaluronidases bind to HA receptors thereby activating their signaling potential. This interaction also results in uptake of HA, its targeting to the lysosome, further degradation by lysosmal HYAL1,2 and ectoglycosidases, GUSB and HEXA. In injured tissues, expression of HA receptors and other genes involved in HA metabolism is down-regulated as tissue regains homeostasis. The cycle is constitutive in transformed cells.
Fig. 2 Schematic representation of injury-induced HA metabolic cycle in wounds. The figure shows the unit structure of HA (N-acetyl-glucosamine and β-glucuronic acid) and demonstrates the functions of HA metabolism in BCA and other cancers. HA binding and uptake are tracked by labeling a high molecular weight polymer with a fluorescent dye. HA fragments produced by ROS, shear and extracellular hyaluronidases bind to HA receptors thereby activating their signaling potential. This interaction also results in uptake of HA, its targeting to the lysosome, further degradation by lysosmal HYAL1,2 and ectoglycosidases, GUSB and HEXA. In injured tissues, expression of HA receptors and other genes involved in HA metabolism is down-regulated as tissue regains homeostasis. The cycle is constitutive in transformed cells.

The biological functions of HA are strictly size-dependant (for reviews, see ref. 17d,e,18,19b). HMW HA polymers (e.g. >200 kDa) carry out structural and other roles that contribute to tissue architecture and function during homeostasis. One of these is to provide a macromolecular template, which concentrates and organizes the assembly of other proteins in the extracellular and possibly intracellular space.19b–d LMW forms, which are generated during tissue repair, activate specific signaling cascades and transcription factors such as extracellular signal-regulated kinases 1 and 2 (ERK1,2), phospho-inositide-3-kinase (PI3K)/ protein kinase B (PKB) and nuclear factor kappa-light-chain enhanced activated B (NFKappaB) cells, and activating protein-1 transcription factor complex (AP1). These pathways control cell processes that are responsible for preserving stem cell compartments, repairing and re-establishing tissue architecture, controlling mesenchymal differentiation (e.g. angiogenesis), promoting cell migration/survival/proliferation, and sustaining innate immune function.12c,17d,18,23b

During response-to-injury, reduction of HMW HA to sizes that can bind to cellular receptors results from reactive oxygen species, mechanical shear, and extracellular hyaluronidase activity (e.g. serum HYAL1, cell surface HYAL2), although the contribution of each is currently not well understood.22a,25,31 Extracellular hyaluronidases 19a,22a,32 have been reported in injured tissues, although the pH of the microenvironment per se is presumably not optimal for their lytic activity. It is likely that cell surface HYAL2 may be active only within cell surface micro-domains that can maintain localized low pH levels as a result of high ion transport activity (Fig. 2).32a Regardless of the mechanisms for generating LMW, expression of HA receptors is required for response to these bioactive HA fragments, and this response is required for normal repair.

CD44, RHAMM, and TLR2,4 are the key HA receptors that are activated by LMW HA generated during response-to-injury.19b,d,33 Our understanding of the wound and, to a lesser extent, tumor functions of these receptors has been greatly enhanced by studying tissue injury in mice or cell lines that lack these receptors. A number of studies indicated that, while genetic deletion of either CD44 or RHAMM kept homeostatic functions intact, it resulted in altered innate immune function and repair, depending on the injury stimulus and responding tissue. For example, CD44 loss was associated with high levels of extracellular bioactive HA fragments, sustained accumulation of macrophages, and unremitting inflammation of bleomycin-injured lung tissue.17d On the other hand, genetic loss of RHAMM (but not CD44) resulted in delayed repair of excisional skin wounds and aberrant mesenchymal differentiation within the wound site.34 In an immune-privileged site such as brain tissue, deletion of CD44 reduced processes associated with the inflammation stage of tissue repair.17e Thus, following cerebral artery occlusion, infarct size was smaller and angiogenic response and neurological damage were less in CD44−/− than in wild type animals. In addition to aberrant response-to-injury, genetic deletion of these HA receptors also affected tumor susceptibility. For instance, loss of CD44 greatly enhanced disease progression in transgenic mice, which are susceptible to BCA due to conditional expression of polyoma middle T-antigen (PyMT) in mammary epithelium. In contrast, loss of RHAMM in a mutant adenomatous polyposis coli gene product (APC)-driven model of fibromatoses reduced tumor invasion.35 These experimental results predict that injury induced-metabolism of the HA homeostatic tissues is restricted because of the potent biological effects of extracellular HA fragments, which could damage normal tissue architecture, promote inappropriate inflammation, and render tissues susceptible to neoplastic conversion and progression.

HA metabolism in BCA

Clinically, high levels of HA within tumor cells or in the peri-tumor stroma can be observed in many cancers and are strong independent prognostic indicators of poor outcome in breast, ovarian, gastric, and colorectal cancers.12a,17f,18,19b,36 Normal breast, ovarian, gastric, and colorectal epithelia produce low levels of HA, although HA accumulation can be observed in the corresponding normal stroma.12a The percentage of tumors that accumulate higher than normal tissue HA levels is not 100%, but ranges between 50% and 80%, which suggests that HA may be an unfavorable factor in subgroups of cancers. Interestingly, neoplastic conversion of tissues that normally produce high levels of HA, such as the keratinocyte layer of the skin, is associated with loss rather than gain of HA accumulation.12a

High accumulation of HA in the tumor and peri-tumor stroma is particularly associated with breast tumor progression (for review, see ref. 12a). HA accumulation within BCA malignant stroma and tumor parenchyma is much higher in malignant than in benign lesions, and these levels are correlated with high tumor grade,37 auxiliary lymph node positivity, and shortened survival.38 HA accumulation is also correlated with BCA treatment resistance. In women <50 years, tumor HA levels predict occurrence of relapse.39 In experimental models, HA production by BCA tumor cells contributes to adjuvant therapy (Trastuzumab) resistance likely because it masks ErbB2 antigenic sites that are normally recognized by Trastuzumab.39b

Increased HA accumulation and metabolism are constitutive in BCA, and BCA tumors express elevated levels of HA synthases, receptors and hyaluronidases, as identified by both experimental analyses and data mining of data banks such as Oncomine (http://www.oncomine.com, Table 1).12aTable 1 summarizes the mRNA expression of HA metabolic genes that are most commonly increased in BCA. Results were obtained by querying the Oncomine database for genes involved in HA metabolism that were both significantly (p < 0.05) increased in breast tumor compared to normal tissue and associated with parameters of poor clinical outcome (e.g. recurrence of primary tumors, appearance of post-treatment metastases, and death). Using these criteria, HAS2 is a prominent factor in BCA. This conclusion is supported both by evidence that HAS2 is one of several genes that are commonly re-arranged in BCA cell lines and sporadic BCA40 and by reports showing that increased HAS2 expression is involved in BCA aggression.12a,38a The mRNA levels of RHAMM/HMMR in particular, but also CD44 and TLR2, are increased in BCA compared to normal breast. These mRNA increases link to poor clinical outcome, a finding that is also noted in experimental evidence.29,41,42 The relationship of CD44 expression with poor clinical outcome (Table 1) may be related to its prominent display on aggressive BCA progenitor subsets.41a,b,43

SPAM1 is a glycosylphosphatidylinositol (GPI)-linked cell surface hyaluronidase normally expressed on sperm surfaces and not in breast tissue44 but is, however, aberrantly expressed in human BCA.16b,45 Analyses of mRNA expression using Oncomine data banks also show that this hyaluronidase is more commonly increased and linked to poor clinical outcome parameters than other hyaluronidases (Table 1).

HA and HA receptor functions common to wounds and BCA

Since HMW HA is present in biologically active amounts in blood (ng l−1), it is one of the first ECM molecules to rapidly bathe injured tissues. Thus, it is an important participant in initiating remodeling processes necessary for tissue repair.19b,d Hypoxic conditions and cytokines released by adherent platelets stimulate further HA production by cells remaining at the wound site or by those adjacent to it.46 This results in an early but transient production/accumulation of HA within wound sites,17e,21b,46a and is a likely constitutive stimulus for HA production in BCA since tumor microenvironments are generally hypoxic. Evidence gleaned from use of modified HA as tissue grafts and wound healing promoters19c predict that the functions of HMW HA in wounds are to provide protection against ROS, hydrate tissues, and restrict passage of microbes. The HMW polymer also reduces antigenicity of ECM protein fragments, reduces angiogenesis and inflammation, and protects stem cell compartments.21b As repair processes are initiated, HA is steadily fragmented into bioactive LMW HA fragments. These attract and stimulate cells of the innate immune system to produce cytokines that propel wounds into the inflammatory stage of repair. Expression of CD44 is required for leukocyte recruitment to the wound and is also essential for clearing the wound of the immunogenic LMW HA fragments. This ultimately results in the blunting of inflammation so that the fibrogenesis/remodeling stage of the wound site can be initiated.5e,17b,20 HA fragments stimulate pro-inflammatory cytokine production by macrophages via TLR4 and also bind to and activate cell surface RHAMM, which complexes with CD44 and activates ERK1,2 signaling cascades.16a,34 These last interactions are required for migration and differentiation of wound fibroblasts into myofibroblasts and other mesenchymal cell types within the wound.34 RHAMM is also present in intracellular compartments, notably the mitotic spindle and nucleus. These intracellular forms of RHAMM appear to control proliferation and gene expression necessary for cell cycle progression,33,47 while extracellular forms appear to promote cell migration.34 Final clearance of HA fragments by injury-induced HA receptors and down-regulation of HA receptor display at the site of injury are necessary for wound resolution.17d

Elements of these wound repair processes have been hijacked by breast tumors, and de-regulated expression of HAS, HA receptors and hyaluronidases appear to promote rather than resolve this disease. Experimental tumor models in which these genes have been modified suggest a role for BCA cell HA to directly promote invasion (e.g. tumor cell migration) as well as proliferation,12a,16a,17f and contribute to stromal changes that promote tumor growth. For example, HA fragments increase host-derived angiogenesis, lymphangiogenesis, and recruitment of macrophages.19b,48 On the other hand, production of HA by stromal cells also contributes to BCA progression. For example, the development of HA-rich malignant stroma promotes the growth and invasion of BCA cells. It has been shown that combined growth of high HA-producing tumor-associated fibroblasts with MCF-7 BCA in immune-deficient mice strongly promotes tumor growth and this is accompanied by stromal lymphangiogenesis and a stromal reaction.48 As such, MCF-7 tumors grow slowly in the absence of these genetically modified fibroblasts, stimulate sparse lymphangiogenesis, and evoke a small stromal reaction.18,48 Similar results were obtained with knockdown of HAS2 in fibroblasts, which prevented stromal angiogenesis/lymphangiogenesis and macrophage recruitment into the stroma of MCF-7 xenografts. HA also participates in the development of BCA drug resistance. HA/CD44 interactions promote drug transporter expression and activity, and affect epitope display of key receptor tyrosine kinases (e.g. ErbB2), thereby limiting receptor-oriented therapy.18,39b

Collectively, these data indicate that HA performs multiple wound-like functions in BCA progression and that components of the tumor HA metabolic cycle are likely to be useful targets for both addressing drug resistance and designing microenvironment targeted therapeutics.

Clinical uses of the HA metabolic cycle

Because of the unique structural properties of HMW HA, it has long been used as a designer biomaterial for tissue engineering applications.49 Currently many preparations have been approved as tissue replacement supplements, cosmetic fillers, anti-adhesives to prevent surgical adhesions and drug delivery vehicles (e.g. treatment of skin lesions such as actinic keratoses and oral mucositis). Moreover, HA can act as a high-affinity probe for imaging and therapy due to its remarkable ability to target to sites where HA receptors are displayed (particularly CD44 and LYVE1).50 Traditional approaches for imaging and therapeutic applications have taken advantage of the altered vascular architecture (100–600 nm fenestration) and lymphatic drainage of solid tumors to selectively retain probes at the tumor site, a phenomenon called “enhanced permeability and retention (EPR)”.51 Although some of the accumulation of HA-based therapeutics is probably due to EPR, the enhanced ability of HA-imaging or HA-therapeutic probes to target and to accumulate in tumor tissue is at least partly due to both HA binding to tumor HA receptors and a concentration dependent-promotion of HA half-life in the circulation. For instance, an administered dose of 3 mg kg−1 HA has a T1/2 = 12 h in human subjects.52

Above and beyond these properties, the ability of HA probes to be processed by target cells as part of a metabolic cycle allows for the design of probes that report HA metabolic activity. For example, when fluorescently labeled HA was adsorbed to gold nanoparticles, the fluorescent signal was quenched. When the adsorbed HA was clipped by hyaluronidases, which were elevated in the BCA tumor microenvironment, a fluorescent signal could be detected.53 In another approach, a HMW tagged HA could accumulate within tumors as a result of being fragmented into LMW HA fragments, which was bound to and was endocytosed by HA receptors displayed on tumor cells (e.g.Fig. 2). These biological properties, combined with its non-antigenicity, excellent biocompatibility profile and hydrophilic/anionic nature that can be easily modified with a variety of functional moieties, have facilitated development of HA nanoscale particles and copolymers for imaging and therapeutic use in inflammatory and neoplastic diseases.54 Conjugation of HA to drugs improves the solubility of hydrophobic drugs, enables controlled release of these drugs, and increases their circulation time within the vasculature.19c,52,55 Ultimately, these HA-based platforms can also be used to enable the isolation and study of tumor cell compartments/subsets undergoing HA metabolism, which in turn will provide a better understanding of the process and function that HA metabolism carries out in BCA progression. To date, HA in the forms of drug-conjugates/prodrugs, nanogels/hydrogels, nano-liposomes, and nanoparticles have been developed for use in therapy or imaging (Table 2).

Table 2 Examples of HA-based platforms for therapy and imaging
Platform Coupling moiety Application Outcome of coupled moiety
a N-(2-hydroxypropyl)methacrylamide. b Poly(ethyleneglycol)-polycaprolactone. c Poly[lactic-co-(glycolic acid)]. d Reactive oxygen species. e Magnetic resonance imaging.
HA Doxorubicin-HPMAa Therapy56 Improved disease-targeted delivery
Sodium Butyrate Therapy57 Improved half-life and disease-targeted delivery
Taxol Therapy58 Improved disease-targeted delivery
Hydrogen peroxide Therapy59 Long acting radiosensitization of local tumors
HA nanogel 5-flurouracil, Therapy60 Controlled release
GFP-siRNA Therapy61 Controlled release and disease-targeted delivery
Cisplatin Therapy62 Improved disease-targeted delivery and release
Doxorubicin-PEG-PCLb Therapy63 Improved disease-targeted delivery and sustained release
Doxorubicin-PLGAc Therapy64 Improved disease-targeted delivery
HA + nanoliposome Doxorubicin Therapy65 Long-term circulation and disease-targeted delivery
Mitomyocin (MMC) Therapy66 Long-term circulation and disease-targeted delivery
HA + nanoparticle Gold nanoparticles Imaging53,67 Optical detection of ROSd and hyaluronidase
Fe2O3 nanoparticles Imaging68 Improved diagnostic MRIe imaging
Gadolinium nanoparticles Imaging69 Detection of hyaluronidase activity
Quantum dots Imaging54b,70,71 Improved disease-targeted optical imaging


Regardless of their physical structures and sizes, all of these site-specific platforms contain HA as a carrier or as a bioactive targeting probe. They are formed either through direct chemical conjugation to functionally active sites of polymer chain or by physical incorporation strategies (Fig. 3). Such HA-based platforms are particularly well suited for treatment/imaging of BCA. HA prodrugs injected directly into breast tumors are well-retained within tumor tissue, which is likely due to the high expression of HA receptors. For example, Kochi Oxydol-Radiation Therapy is a new radiosensitizer containing H2O2 that turns radiation-resistant breast tumors into a radiation-sensitive state following its injection into the tumor. Phase I clinical trials of this HA-based therapeutic have revealed no adverse effects.59 Direct injection of Cisplatin- or doxorubicin-HA nanoconjugates into the mammary fat pad or into nearby subcutaneous tissue has resulted in the accumulation of these conjugates, both in the tumor and in nearby lymphatics. Since aggressive BCA cells initially invade and extravasate via local lymphatic tissues, this property of HA nanoconjugates may allow for early treatment of invasive BCA.62a The lymphatic accumulation of HA nanoconjugates is likely the result of high LYVE1 expression.72 These studies collectively point to the specific targeting ability of HA. However, the use of HA for targeting, particularly if it is unmodified, has challenges. These include avoiding rapid clearance from the circulation by the liver as a result of HA endocytic receptors and, to a lesser extent, the reticular endothelial system (RES).73 Strategies that have been devised for addressing these issues include cross-linking of individual polymer chains, chemical modification of carboxyl groups, and pre-treatment of the liver with chondroitin sulfate before exposure to HA-based nanoprobes. Chondroitin sulfate binds to HA endocytic receptors resulting in their internalization and prolonged blocking of HA uptake by the liver.54a,73–75



            HA modification for preparation of probes. HA is modified for preparation as a therapeutic or an imaging probe through physical incorporation of moieties within HA polymers (black arrow) or chemical conjugation of HA polymers/HA nanoparticles to moieties (green arrow).
Fig. 3 HA modification for preparation of probes. HA is modified for preparation as a therapeutic or an imaging probe through physical incorporation of moieties within HA polymers (black arrow) or chemical conjugation of HA polymers/HA nanoparticles to moieties (green arrow).

An alternative approach to using the HA polymer for tumor-targeted imaging or therapy at sites of HA metabolism is to utilize HA receptors that are expressed by BCA cells. Anti-CD44 antibodies have been tested in phase I clinical trials in patients with head and neck cancers and have also been evaluated for imaging these tumors in pre-clinical studies.75d However, trials were terminated due to the death of an enrolled patient. Current efforts are focused on designing reagents that specifically block HA/CD44 interactions. As noted above, the relationship between CD44 expression and BCA progression appears to be complex, and use of these types of reagents is therefore limited by our still-rudimentary understanding of the basic biology of CD44/HA interactions. RHAMM is the most commonly over-expressed HA receptor in BCA (Table 1) and has been classified as a tumor marker in this neoplastic disease.76 However, the biological role of RHAMM in BCA is also not yet well understood,77 which limits design of therapeutic approaches to target this receptor. Nevertheless, RHAMM is also highly expressed in blood malignancies, and development of RHAMM vaccines currently shows promise for treatment of acute myeloid leukemia and multiple myeloma.78

Conclusions

Both clinical and experimental evidence indicate an involvement of HA, HA receptors, and hyaluronidases in BCA malignancy. Data mining confirms that elevated mRNA expression of genes involved in HA metabolism is common in BCA. This and experimental data suggest that HA performs multiple wound-like functions in BCA progression and predict that the components of the tumor HA metabolic cycle are useful therapeutic targets. In addition, HA-based nanoprobes that detect active HA metabolic cycling or directly target HA receptor-bearing cells may be useful tools for diagnostic imaging. However, effective use of these as therapeutics requires much more basic information about the functions of HA metabolism in BCA progression and the types of BCA that display active HA metabolism.

Table of abbreviations

ABCAdenosine-5′-triphosphate binding cassette
PKBProtein kinase B
AP1Activating protein-1 transcription factor complex
APCAdenomatous polyposis coli gene product
BCABreast cancer
CD44Cluster designation 44
ECMExtracellular matrix
EPREnhanced permeability and retention
ErbB2Human epidermal growth factor receptor 2
ERK1/2Extracellular signal-regulated kinases 1 and 2
GPIGlycosylphosphatidylinositol
GUSBβ-glucuronidase
HAHyaluronan, hyaluronic acid, hyaluronate
HAREHyaluronan(HA) receptor for endocytosis (liver, Stabilin 1,2)
HAS1-3HA synthases1,2,3
HEXAHexosaminidase
HMWHigh molecular weight
HPMA N-(2-hydroxypropyl)methacrylamide
HYAL1-4Hyaluronidases1-4
IGF-1Insulin-like growth factor 1
LMWLow molecular weight
LYVE1Lymphatic vessel endothelial hyaluronan receptor1
NFKappaBNuclear factor kappa-light-chain enhanced activated B cells
PI3KPhospho-innositide-3-kinase
PyMTPolyoma middle T antigen
RHAMM/HMMRReceptor for hyaluronan mediated motility (protein designation)
RESReticuloendothelial system
ROSReactive oxygen species
SPAM1Hyaluronidase PH-20
STAB1,2Stabilin1,2
TGFBTransforming growth factor beta
TLR2,4Toll-Like Receptors 2,4
TR-HATexas red-HA
UDPUridine diphospho [e.g. glucuronic acid, N-acetyl-glucosamine]

Acknowledgements

We thank Mina J. Bissell for her continuing enthusiasm and support of this work; we also thank Daniel H Kwon and Catlin Ward for administrative help. This work was supported by a DOD-BCRP IDEA award to Mina J Bissell and Eva A Turley (BC044087), by the National Institute of Health grants R37CA064786 and R01CA057621, Low Dose Radiation Program (contract no. DE-AC02-05CH1123) to MJB, the Canadian Breast Cancer Alliance and Cancer Research Society to EAT, and by a Distinguished Fellow Award to MJB. Mandana Veiseh was supported by a Ruth L. Kirschstein National Research Service Award (NRSA) F32 postdoctoral fellowship from National Cancer Institute of National Institute of Health (FCA132491A).

References

  1. (a) J. M. Lee, S. Dedhar, R. Kalluri and E. W. Thompson, The epithelial-mesenchymal transition: new insights in signaling, development, and disease, J. Cell Biol., 2006, 172(7), 973–81 CrossRef CAS; (b) M. J. Paszek and V. M. Weaver, The tension mounts: mechanics meets morphogenesis and malignancy, J. Mammary Gland Biol. Neoplasia, 2004, 9(4), 325–42 CrossRef; (c) R. Xu, A. Boudreau and M. J. Bissell, Tissue architecture and function: dynamic reciprocity via extra- and intra-cellular matrices, Cancer Metastasis Rev., 2009, 28(1–2), 167–76 CrossRef; (d) D. E. Ingber, Can cancer be reversed by engineering the tumor microenvironment?, Semin. Cancer Biol., 2008, 18(5), 356–64 CrossRef CAS; (e) C. Box, S. J. Rogers, M. Mendiola and S. A. Eccles, Tumour-microenvironmental interactions: paths to progression and targets for treatment, Semin. Cancer Biol., 2010, 20(3), 128–38 CrossRef CAS.
  2. M. D. Sternlicht, M. J. Bissell and Z. Werb, The matrix metalloproteinase stromelysin-1 acts as a natural mammary tumor promoter, Oncogene, 2000, 19(8), 1102–13 CrossRef CAS.
  3. (a) C. A. Maxwell, E. Rasmussen, F. Zhan, J. J. Keats, S. Adamia, E. Strachan, M. Crainie, R. Walker, A. R. Belch, L. M. Pilarski, B. Barlogie, J. Shaughnessy, Jr and T. Reiman, RHAMM expression and isoform balance predict aggressive disease and poor survival in multiple myeloma, Blood, 2004, 104(4), 1151–8 CrossRef CAS; (b) P. A. Kenny and M. J. Bissell, Tumor reversion: correction of malignant behavior by microenvironmental cues, Int. J. Cancer, 2003, 107(5), 688–95 CrossRef CAS; (c) Y. Li, L. Li, T. J. Brown and P. Heldin, Silencing of hyaluronan synthase 2 suppresses the malignant phenotype of invasive breast cancer cells, Int. J. Cancer, 2007, 120(12), 2557–67 CrossRef CAS; (d) B. P. Toole, Hyaluronan: from extracellular glue to pericellular cue, Nat. Rev. Cancer, 2004, 4(7), 528–39 CrossRef CAS.
  4. (a) F. Wang, R. K. Hansen, D. Radisky, T. Yoneda, M. H. Barcellos-Hoff, O. W. Petersen, E. A. Turley and M. J. Bissell, Phenotypic reversion or death of cancer cells by altering signaling pathways in three-dimensional contexts, J. Natl. Cancer Inst., 2002, 94(19), 1494–503 CrossRef CAS; (b) V. M. Weaver, O. W. Petersen, F. Wang, C. A. Larabell, P. Briand, C. Damsky and M. J. Bissell, Reversion of the malignant phenotype of human breast cells in three-dimensional culture and n vivo by integrin blocking antibodies, J. Cell Biol., 1997, 137(1), 231–45 CrossRef CAS; (c) C. L. Hall, B. Yang, X. Yang, S. Zhang, M. Turley, S. Samuel, L. A. Lange, C. Wang, G. D. Curpen, R. C. Savani, A. H. Greenberg and E. A. Turley, Overexpression of the hyaluronan receptor RHAMM is transforming and is also required for H-ras transformation, Cell, 1995, 82(1), 19–26 CrossRef CAS.
  5. (a) J. Condeelis and J. W. Pollard, Macrophages: obligate partners for tumor cell migration, invasion, and metastasis, Cell, 2006, 124(2), 263–6 CrossRef CAS; (b) L. M. Postovit, N. V. Margaryan, E. A. Seftor and M. J. Hendrix, Role of nodal signaling and the microenvironment underlying melanoma plasticity, Pigm. Cell Melanoma Res., 2008, 21(3), 348–57 Search PubMed; (c) M. J. Bissell, D. C. Radisky, A. Rizki, V. M. Weaver and O. W. Petersen, The organizing principle: microenvironmental influences n the normal and malignant breast, Differentiation, 2002, 70(9–10), 537–46 CrossRef; (d) E. A. Turley, M. Veiseh, D. C. Radisky and M. J. Bissell, Mechanisms of disease: epithelial-mesenchymal transition--does cellular plasticity fuel neoplastic progression?, Nat. Clin. Pract. Oncol., 2008, 5(5), 280–90 Search PubMed; (e) M. Egeblad, E. S. Nakasone and Z. Werb, Tumors as organs: complex tissues that interface with the entire organism, Dev. Cell, 2010, 18(6), 884–901 CrossRef CAS; (f) L. Vera-Ramirez, P. Sanchez-Rovira, C. L. Ramirez-Tortosa, J. L. Quiles, M. C. Ramirez-Tortosa, J. C. Alvarez, M. Fernandez-Navarro and J. A. Lorente, Gene-expression profiles, tumor microenvironment, and cancer stem cells in breast cancer: latest advances towards an integrated approach, Cancer Treat. Rev., 2010, 36(6), 477–84 CrossRef CAS.
  6. (a) A. H. Beck, I. Espinosa, C. B. Gilks, M. van de Rijn and R. B. West, The fibromatosis signature defines a robust stromal response in breast carcinoma, Lab. Invest., 2008, 88(6), 591–601 Search PubMed; (b) K. C. Flanders and L. M. Wakefield, Transforming growth factor-(beta)s and mammary gland involution; functional roles and implications for cancer progression, J. Mammary Gland Biol. Neoplasia, 2009, 14(2), 131–44 CrossRef; (c) T. S. Pandit, W. Kennette, L. Mackenzie, G. Zhang, W. Al-Katib, J. Andrews, S. A. Vantyghem, D. G. Ormond, A. L. Allan, D. I. Rodenhiser, A. F. Chambers and A. B. Tuck, Lymphatic metastasis of breast cancer cells is associated with differential gene expression profiles that predict cancer stem cell-like properties and the ability to survive, establish and grow in a foreign environment, Int. J. Oncol., 2009, 35(2), 297–308 CAS; (d) J. O'Brien and P. Schedin, Macrophages in breast cancer: do involution macrophages account for the poor prognosis of pregnancy-associated breast cancer?, J. Mammary Gland Biol. Neoplasia, 2009, 14(2), 145–57 CrossRef; (e) C. R. Acharya, D. S. Hsu, C. K. Anders, A. Anguiano, K. H. Salter, K. S. Walters, R. C. Redman, S. A. Tuchman, C. A. Moylan, S. Mukherjee, W. T. Barry, H. K. Dressman, G. S. Ginsburg, K. P. Marcom, K. S. Garman, G. H. Lyman, J. R. Nevins and A. Potti, Gene expression signatures, clinicopathological features, and individualized therapy in breast cancer, JAMA, J. Am. Med. Assoc., 2008, 299(13), 1574–87 CrossRef CAS; (f) H. F. Dvorak, Tumors: wounds that do not heal. Similarities between tumor stroma generation and wound healing, N. Engl. J. Med., 1986, 315(26), 1650–9 CrossRef CAS; (g) C. Derleth and I. A. Mayer, Antiangiogenic therapies in early-stage breast cancer, Clin. Breast Cancer, 2010, 10(Suppl 1), E23–31 Search PubMed.
  7. (a) P. Schedin, J. O'Brien, M. Rudolph, T. Stein and V. Borges, Microenvironment of the involuting mammary gland mediates mammary cancer progression, J. Mammary Gland Biol. Neoplasia, 2007, 12(1), 71–82 CrossRef; (b) E. A. McSherry, S. Donatello, A. M. Hopkins and S. McDonnell, Molecular basis of invasion in breast cancer, Cell. Mol. Life Sci., 2007, 64(24), 3201–18 CrossRef CAS.
  8. (a) L. M. Coussens and Z. Werb, Inflammatory cells and cancer: think different!, J. Exp. Med., 2001, 193(6), F23–6 CrossRef CAS; (b) J. A. Van Ginderachter, S. Meerschaut, Y. Liu, L. Brys, K. De. Groeve, G. Hassanzadeh Ghassabeh, G. Raes and P. De Baetselier, Peroxisome proliferator-activated receptor gamma (PPARgamma) ligands reverse CTL suppression by alternatively activated (M2) macrophages in cancer, Blood, 2006, 108(2), 525–35 CrossRef CAS.
  9. M. H. Barcellos-Hoff and R. J. Akhurst, Transforming growth factor-beta in breast cancer: too much, too late, Breast Cancer Res., 2009, 11(1), 202 CrossRef.
  10. T. Stein, N. Salomonis, D. S. Nuyten, M. J. van de Vijver and B. A. Gusterson, A mouse mammary gland involution mRNA signature identifies biological pathways potentially associated with breast cancer metastasis, J. Mammary Gland Biol. Neoplasia, 2009, 14(2), 99–116 CrossRef.
  11. (a) J. Y. Hsu and H. A. Wakelee, Monoclonal antibodies targeting vascular endothelial growth factor: current status and future challenges in cancer therapy, BioDrugs, 2009, 23(5), 289–304 Search PubMed; (b) E. A. Perez, A. Moreno-Aspitia, E. Aubrey Thompson and C. A. Andorfer, Adjuvant therapy of triple negative breast cancer, Breast Cancer Res. Treat., 2010, 120(2), 285–91 CrossRef CAS.
  12. (a) R. H. Tammi, A. Kultti, V. M. Kosma, R. Pirinen, P. Auvinen and M. I. Tammi, Hyaluronan in human tumors: pathobiological and prognostic messages from cell-associated and stromal hyaluronan, Semin. Cancer Biol., 2008, 18(4), 288–95 CrossRef CAS; (b) A. D. Theocharis, S. S. Skandalis, G. N. Tzanakakis and N. K. Karamanos, Proteoglycans in health and disease: novel roles for proteoglycans in malignancy and their pharmacological targeting, FEBS J., 2010, 277(19), 3904–23 CrossRef CAS; (c) L. Alaniz, M. Garcia, M. Rizzo, F. Piccioni and G. Mazzolini, Altered hyaluronan biosynthesis and cancer progression: an immunological perspective, Mini-Rev. Med. Chem., 2009, 9(13), 1538–46 CrossRef CAS; (d) I. O. Potapenko, V. D. Haakensen, T. Luders, A. Helland, I. Bukholm, T. Sorlie, V. N. Kristensen, O. C. Lingjaerde and A. L. Borresen-Dale, Glycan gene expression signatures in normal and malignant breast tissue; possible role in diagnosis and progression, Mol. Oncol., 2010, 4(2), 98–118 Search PubMed.
  13. M. Veiseh, J. Zhang, R. C. Savani, R. E. Harrison, D. Mikilus, L. Collis, J. Koropatnick, L. Luyt, M. J. Bissell and E. A. Turley, Imaging the Remodeling Microenvironment of Injured and Neoplastic Tissues with Hyaluronan-Based Functional Probes, in review, 2011 Search PubMed.
  14. P. H. Weigel and J. H. Yik, Glycans as endocytosis signals: the cases of the asialoglycoprotein and hyaluronan/chondroitin sulfate receptors, Biochim. Biophys. Acta, Gen. Subj., 2002, 1572(2–3), 341–63 CrossRef CAS.
  15. (a) R. C. Savani and E. A. Turley, The role of hyaluronan and its receptors in restenosis after balloon angioplasty: development of a potential therapy, Int. J. Tissue React., 1995, 17(4), 141–51 Search PubMed; (b) P. W. Noble and D. Jiang, Matrix regulation of lung injury, inflammation, and repair: the role of innate immunity, Proc. Am. Thorac. Soc., 2006, 3(5), 401–4 CrossRef CAS.
  16. (a) S. R. Hamilton, S. F. Fard, F. F. Paiwand, C. Tolg, M. Veiseh, C. Wang, J. B. McCarthy, M. J. Bissell, J. Koropatnick and E. A. Turley, The hyaluronan receptors CD44 and Rhamm (CD168) form complexes with ERK1,2 that sustain high basal motility in breast cancer cells, J. Biol. Chem., 2007, 282(22), 16667–80 CrossRef CAS; (b) L. P. Wang, X. M. Xu, H. Y. Ning, S. M. Yang, J. G. Chen, J. Y. Yu, H. Y. Ding, C. B. Underhill and L. R. Zhang, Expression of PH20 in primary and metastatic breast cancer and its pathological significance, Zhonghua Bing Li Xue Za Zhi, 2004, 33(4), 320–3 Search PubMed.
  17. (a) R. D. Price, M. G. Berry and H. A. Navsaria, Hyaluronic acid: the scientific and clinical evidence, J. Plast., Reconstr. Aesthetic Surg., 2007, 60(10), 1110–9 Search PubMed; (b) W. Y. Chen and G. Abatangelo, Functions of hyaluronan in wound repair, Wound Repair Regener., 1999, 7(2), 79–89 Search PubMed; (c) K. R. Taylor and R. L. Gallo, Glycosaminoglycans and their proteoglycans: host-associated molecular patterns for initiation and modulation of inflammation, FASEB J., 2006, 20(1), 9–22 CrossRef CAS; (d) D. Jiang, J. Liang and P. W. Noble, Hyaluronan in tissue injury and repair, Annu. Rev. Cell Dev. Biol., 2007, 23, 435–61 CrossRef CAS; (e) M. Slevin, J. Krupinski, J. Gaffney, S. Matou, D. West, H. Delisser, R. C. Savani and S. Kumar, Hyaluronan-mediated angiogenesis in vascular disease: uncovering RHAMM and CD44 receptor signaling pathways, Matrix Biol., 2007, 26(1), 58–68 CrossRef CAS; (f) M. I. Tammi, A. J. Day and E. A. Turley, Hyaluronan and homeostasis: a balancing act, J. Biol. Chem., 2002, 277(7), 4581–4 CrossRef CAS.
  18. B. P. Toole and M. G. Slomiany, Hyaluronan: a constitutive regulator of chemoresistance and malignancy in cancer cells, Semin. Cancer Biol., 2008, 18(4), 244–50 CrossRef CAS.
  19. (a) R. Stern, Hyaluronidases in cancer biology, Semin. Cancer Biol., 2008, 18(4), 275–80 CrossRef CAS; (b) N. Itano, Simple primary structure, complex turnover regulation and multiple roles of hyaluronan, J. Biochem., 2008, 144(2), 131–7 CrossRef CAS; (c) J. Gaffney, S. Matou-Nasri, M. Grau-Olivares and M. Slevin, Therapeutic applications of hyaluronan, Mol. BioSyst., 2010, 6(3), 437–43 RSC; (d) K. S. Girish and K. Kemparaju, The magic glue hyaluronan and its eraser hyaluronidase: a biological overview, Life Sci., 2007, 80(21), 1921–43 CrossRef CAS.
  20. D. G. Jackson, Immunological functions of hyaluronan and its receptors in the lymphatics, Immunol. Rev., 2009, 230(1), 216–31 CrossRef CAS.
  21. (a) J. A. Sloane, D. Blitz, Z. Margolin and T. Vartanian, A clear and present danger: endogenous ligands of Toll-like receptors, NeuroMol. Med., 2010, 12(2), 149–63 Search PubMed; (b) E. P. Buchanan, M. T. Longaker and H. P. Lorenz, Fetal skin wound healing, Adv. Clin. Chem., 2009, 48, 137–61 Search PubMed.
  22. (a) M. E. Monzon, N. Fregien, N. Schmid, N. S. Falcon, M. Campos, S. M. Casalino-Matsuda and R. M. Forteza, Reactive oxygen species and hyaluronidase 2 regulate airway epithelial hyaluronan fragmentation, J. Biol. Chem., 2010, 285(34), 26126–34 CrossRef CAS; (b) V. B. Lokeshwar, P. Gomez, M. Kramer, J. Knapp, M. A. McCornack, L. E. Lopez, N. Fregien, N. Dhir, S. Scherer, D. J. Klumpp, M. Manoharan, M. S. Soloway and B. L. Lokeshwar, Epigenetic regulation of HYAL-1 hyaluronidase expression. identification of HYAL-1 promoter, J. Biol. Chem., 2008, 283(43), 29215–27 CrossRef CAS.
  23. (a) B. P. Toole and M. G. Slomiany, Hyaluronan, CD44 and Emmprin: partners in cancer cell chemoresistance, Drug Resist. Updates, 2008, 11(3), 110–21 CrossRef CAS; (b) E. A. Turley, P. W. Noble and L. Y. Bourguignon, Signaling properties of hyaluronan receptors, J. Biol. Chem., 2002, 277(7), 4589–92 CrossRef CAS.
  24. R. Stern, A. A. Asari and K. N. Sugahara, Hyaluronan fragments: an information-rich system, Eur. J. Cell Biol., 2006, 85(8), 699–715 CrossRef CAS.
  25. R. Stern and H. I. Maibach, Hyaluronan in skin: aspects of aging and its pharmacologic modulation, Clin. Dermatol., 2008, 26(2), 106–22 CrossRef.
  26. P. H. Weigel and P. L. DeAngelis, Hyaluronan synthases: a decade-plus of novel glycosyltransferases, J. Biol. Chem., 2007, 282(51), 36777–81 CrossRef CAS.
  27. L. Cuthbertson, I. L. Mainprize, J. H. Naismith and C. Whitfield, Pivotal roles of the outer membrane polysaccharide export and polysaccharide copolymerase protein families in export of extracellular polysaccharides in gram-negative bacteria, Microbiol. Mol. Biol. Rev., 2009, 73(1), 155–77 CrossRef CAS.
  28. (a) M. A. Simpson and V. B. Lokeshwar, Hyaluronan and hyaluronidase in genitourinary tumors, Front. Biosci., 2008, 13, 5664–80 CrossRef CAS; (b) K. M. Bullard, H. R. Kim, M. A. Wheeler, C. M. Wilson, C. L. Neudauer, M. A. Simpson and J. B. McCarthy, Hyaluronan synthase-3 is upregulated in metastatic colon carcinoma cells and manipulation of expression alters matrix retention and cellular growth, Int. J. Cancer, 2003, 107(5), 739–46 CrossRef CAS.
  29. L. Udabage, G. R. Brownlee, S. K. Nilsson and T. J. Brown, The over-expression of HAS2, Hyal-2 and CD44 is implicated in the invasiveness of breast cancer, Exp. Cell Res., 2005, 310(1), 205–17 CrossRef CAS.
  30. M. A. Simpson, C. M. Wilson and J. B. McCarthy, Inhibition of prostate tumor cell hyaluronan synthesis impairs subcutaneous growth and vascularization in immunocompromised mice, Am. J. Pathol., 2002, 161(3), 849–57 CAS.
  31. (a) M. Eberlein, K. A. Scheibner, K. E. Black, S. L. Collins, Y. Chan-Li, J. D. Powell and M. R. Horton, Anti-oxidant inhibition of hyaluronan fragment-induced inflammatory gene expression, J. Inflammation, 2008, 5, 20 Search PubMed; (b) L. Soltes, R. Mendichi, G. Kogan, J. Schiller, M. Stankovska and J. Arnhold, Degradative action of reactive oxygen species on hyaluronan, Biomacromolecules, 2006, 7(3), 659–68 CrossRef CAS; (c) G. Mendoza, A. I. Alvarez, M. M. Pulido, A. J. Molina, G. Merino, R. Real, P. Fernandes and J. G. Prieto, Inhibitory effects of different antioxidants on hyaluronan depolymerization, Carbohydr. Res., 2007, 342(1), 96–102 CrossRef CAS; (d) R. Moseley, M. Walker, R. J. Waddington and W. Y. Chen, Comparison of the antioxidant properties of wound dressing materials—carboxymethylcellulose, hyaluronan benzyl ester and hyaluronan, towards polymorphonuclear leukocyte-derived reactive oxygen species, Biomaterials, 2003, 24(9), 1549–57 CrossRef CAS; (e) G. M. Campo, A. Avenoso, S. Campo, A. Ferlazzo, D. Altavilla, C. Micali and A. Calatroni, Aromatic trap analysis of free radicals production in experimental collagen-induced arthritis in the rat: protective effect of glycosaminoglycans treatment, Free Radical Res., 2003, 37(3), 257–68 CrossRef CAS.
  32. (a) L. Y. Bourguignon, P. A. Singleton, F. Diedrich, R. Stern and E. Gilad, CD44 interaction with Na+-H+ exchanger (NHE1) creates acidic microenvironments leading to hyaluronidase-2 and cathepsin B activation and breast tumor cell invasion, J. Biol. Chem., 2004, 279(26), 26991–7007 CrossRef CAS; (b) M. C. Gasingirwa, J. Thirion, J. Mertens-Strijthagen, S. Wattiaux-De Coninck, B. Flamion, R. Wattiaux and M. Jadot, Endocytosis of hyaluronidase-1 by the liver, Biochem. J., 2010, 430(2), 305–13 CrossRef CAS.
  33. C. A. Maxwell, J. McCarthy and E. Turley, Cell-surface and mitotic-spindle RHAMM: moonlighting or dual oncogenic functions?, J. Cell Sci., 2008, 121(7), 925–32 CrossRef CAS.
  34. C. Tolg, S. R. Hamilton, K. A. Nakrieko, F. Kooshesh, P. Walton, J. B. McCarthy, M. J. Bissell and E. A. Turley, Rhamm-/- fibroblasts are defective in CD44-mediated ERK1,2 motogenic signaling, leading to defective skin wound repair, J. Cell Biol., 2006, 175(6), 1017–28 CrossRef CAS.
  35. (a) J. I. Lopez, T. D. Camenisch, M. V. Stevens, B. J. Sands, J. McDonald and J. A. Schroeder, CD44 attenuates metastatic invasion during breast cancer progression, Cancer Res., 2005, 65(15), 6755–63 CrossRef CAS; (b) C. Tolg, R. Poon, R. Fodde, E. A. Turley and B. A. Alman, Genetic deletion of receptor for hyaluronan-mediated motility (Rhamm) attenuates the formation of aggressive fibromatosis (desmoid tumor), Oncogene, 2003, 22(44), 6873–82 CrossRef CAS.
  36. V. B. Lokeshwar and M. G. Selzer, Hyalurondiase: both a tumor promoter and suppressor, Semin. Cancer Biol., 2008, 18(4), 281–7 CrossRef CAS.
  37. P. K. Auvinen, J. J. Parkkinen, R. T. Johansson, U. M. Agren, R. H. Tammi, M. J. Eskelinen and V. M. Kosma, Expression of hyaluronan in benign and malignant breast lesions, Int. J. Cancer, 1997, 74(5), 477–81 CrossRef CAS.
  38. (a) P. Auvinen, R. Tammi, J. Parkkinen, M. Tammi, U. Agren, R. Johansson, P. Hirvikoski, M. Eskelinen and V. M. Kosma, Hyaluronan in peritumoral stroma and malignant cells associates with breast cancer spreading and predicts survival, Am. J. Pathol., 2000, 156(2), 529–36 CAS; (b) S. Suwiwat, C. Ricciardelli, R. Tammi, M. Tammi, P. Auvinen, V. M. Kosma, R. G. LeBaron, W. A. Raymond, W. D. Tilley and D. J. Horsfall, Expression of extracellular matrix components versican, chondroitin sulfate, tenascin, and hyaluronan, and their association with disease outcome in node-negative breast cancer, Clin. Cancer Res., 2004, 10(7), 2491–8 CrossRef CAS.
  39. (a) P. Casalini, M. L. Carcangiu, R. Tammi, P. Auvinen, V. M. Kosma, P. Valagussa, M. Greco, A. Balsari, S. Menard and E. Tagliabue, Two distinct local relapse subtypes in invasive breast cancer: effect on their prognostic impact, Clin. Cancer Res., 2008, 14(1), 25–31 CrossRef CAS; (b) Z. Palyi-Krekk, M. Barok, J. Isola, M. Tammi, J. Szollosi and P. Nagy, Hyaluronan-induced masking of ErbB2 and CD44-enhanced trastuzumab internalisation in trastuzumab resistant breast cancer, Eur. J. Cancer, 2007, 43(16), 2423–33 CrossRef CAS.
  40. K. Unger, J. Wienberg, A. Riches, L. Hieber, A. Walch, A. Brown, P. C. O'Brien, C. Briscoe, L. Gray, E. Rodriguez, G. Jackl, J. Knijnenburg, G. Tallini, M. Ferguson-Smith and H. Zitzelsberger, Novel gene rearrangements in transformed breast cells identified by high-resolution breakpoint analysis of chromosomal aberrations, Endocr. Relat. Cancer, 2010, 17(1), 87–98 CrossRef CAS.
  41. (a) L. R. Oliveira, S. S. Jeffrey and A. Ribeiro-Silva, Stem cells in human breast cancer, Histol. Histopathol., 2010, 25(3), 371–85 Search PubMed; (b) A. L. Stratford, K. Reipas, C. Maxwell and S. E. Dunn, Targeting tumour-initiating cells to improve the cure rates for triple-negative breast cancer, Expert Rev. Mol. Med., 2010, 12, e22 Search PubMed; (c) K. Kai, Y. Arima, T. Kamiya and H. Saya, Breast cancer stem cells, Breast Cancer, 2010, 17(2), 80–5 Search PubMed; (d) S. Chuthapisith, J. Eremin, M. El-Sheemey and O. Eremin, Breast cancer chemoresistance: emerging importance of cancer stem cells, Surg. Oncol., 2010, 19(1), 27–32 CrossRef.
  42. (a) W. Xie, Y. Huang, A. Guo and W. Wu, Bacteria peptidoglycan promoted breast cancer cell invasiveness and adhesiveness by targeting toll-like receptor 2 in the cancer cells, PLoS One, 2010, 5(5), e10850 CrossRef; (b) W. Xie, Y. Wang, Y. Huang, H. Yang, J. Wang and Z. Hu, Toll-like receptor 2 mediates invasion via activating NF-kappaB in MDA-MB-231 breast cancer cells, Biochem. Biophys. Res. Commun., 2009, 379(4), 1027–32 CrossRef CAS; (c) S. A. Armogida, N. M. Yannaras, A. L. Melton and M. D. Srivastava, Identification and quantification of innate immune system mediators in human breast milk, Allergy Asthma Proc., 2004, 25(5), 297–304 Search PubMed.
  43. B. Dave and J. Chang, Treatment resistance in stem cells and breast cancer, J. Mammary Gland Biol. Neoplasia, 2009, 14(1), 79–82 CrossRef.
  44. (a) P. A. Martin-DeLeon, Epididymal SPAM1 and its impact on sperm function, Mol. Cell. Endocrinol., 2006, 250(1–2), 114–21 CrossRef CAS; (b) A. B. Csoka, G. I. Frost and R. Stern, The six hyaluronidase-like genes in the human and mouse genomes, Matrix Biol., 2001, 20(8), 499–508 CrossRef CAS.
  45. D. J. Beech, A. K. Madan and N. Deng, Expression of PH-20 in normal and neoplastic breast tissue, J. Surg. Res., 2002, 103(2), 203–7 CrossRef CAS.
  46. (a) F. Gao, Y. Liu, Y. He, C. Yang, Y. Wang, X. Shi and G. Wei, Hyaluronan oligosaccharides promote excisional wound healing through enhanced angiogenesis, Matrix Biol., 2010, 29(2), 107–16 CrossRef CAS; (b) F. Gao, P. Okunieff, Z. Han, I. Ding, L. Wang, W. Liu, J. Zhang, S. Yang, J. Chen, C. B. Underhill, S. Kim and L. Zhang, Hypoxia-induced alterations in hyaluronan and hyaluronidase, Adv. Exp. Med. Biol., 2005, 566, 249–56 CrossRef CAS.
  47. C. Tolg, S. Hamilton, L. Morningstar, J. Zhang, K. Esguerra, P. Telmer, L. Luyt, R. Harrison, J. McCarthy and E. Turley, Rhamm promotes interphase microtubule instability and mitotic spindle integrity through Mek1/Erk1/2 activity, J. Biol. Chem., 2010, 285(34), 26461–74 CrossRef CAS.
  48. H. Koyama, N. Kobayashi, M. Harada, M. Takeoka, Y. Kawai, K. Sano, M. Fujimori, J. Amano, T. Ohhashi, R. Kannagi, K. Kimata, S. Taniguchi and N. Itano, Significance of tumor-associated stroma in promotion of intratumoral lymphangiogenesis: pivotal role of a hyaluronan-rich tumor microenvironment, Am. J. Pathol., 2008, 172(1), 179–93 CAS.
  49. (a) M. Veiseh, E. A. Turley and M. J. Bissell, A top-down Analysis of a Dynamic Environment: Extracellular Matrix Structure and Function, in Nanotechnology and Tissue Engineering: The Scaffold, ed. C. T. Laurencin and L. S. Nair, CRS Press/Taylor and Francis Group FL, 2008, pp. 6387–6392 Search PubMed; (b) D. D. Allison and K. J. Grande-Allen, Review. Hyaluronan: a powerful tissue engineering tool, Tissue Eng., 2006, 12(8), 2131–40 CrossRef CAS.
  50. I. De Stefano, A. Battaglia, G. F. Zannoni, M. G. Prisco, A. Fattorossi, D. Travaglia, S. Baroni, D. Renier, G. Scambia, C. Ferlini and D. Gallo, Hyaluronic acid-paclitaxel: effects of intraperitoneal administration against CD44(+) human ovarian cancer xenografts, Cancer Chemother. Pharmacol., 2010 Search PubMed.
  51. (a) Y. Matsumura and H. Maeda, A new concept for macromolecular therapeutics in cancer chemotherapy: mechanism of tumoritropic accumulation of proteins and the antitumor agent smancs, Cancer Res., 1986, 46(12 Pt 1), 6387–92 CAS; (b) H. Maeda, J. Wu, T. Sawa, Y. Matsumura and K. Hori, Tumor vascular permeability and the EPR effect in macromolecular therapeutics: a review, J. Controlled Release, 2000, 65(1–2), 271–84 CrossRef CAS.
  52. S. R. Hamilton, M. Veiseh, C. Tolg, R. Tirona, J. Richardson, R. Brown, M. Gonzalez, M. Vanzieleghem, P. Anderson, S. Asculai, F. Winnik, R. Savani, D. Freeman, L. Luyt, J. Koropatnick and E. A. Turley, Pharmacokinetics and Pharmacodynamics of hyaluronan infused into healthy human volunters, The Open Drug Metabolism Journal, 2009, 3, 43–55 Search PubMed.
  53. H. Lee, K. Lee, I. K. Kim and T. G. Park, Synthesis, characterization, and in vivo diagnostic applications of hyaluronic acid immobilized gold nanoprobes, Biomaterials, 2008, 29(35), 4709–18 CrossRef CAS.
  54. (a) H. Lee, K. Lee and T. G. Park, Hyaluronic acid-paclitaxel conjugate micelles: synthesis, characterization, and antitumor activity, Bioconjugate Chem., 2008, 19(6), 1319–25 CrossRef CAS; (b) E. J. Oh, K. Park, K. S. Kim, J. Kim, J. A. Yang, J. H. Kong, M. Y. Lee, A. S. Hoffman and S. K. Hahn, Target specific and long-acting delivery of protein, peptide, and nucleotide therapeutics using hyaluronic acid derivatives, J. Controlled Release, 2010, 141(1), 2–12 CrossRef CAS.
  55. (a) T. J. Brown, The development of hyaluronan as a drug transporter and excipient for chemotherapeutic drugs, Curr. Pharm. Biotechnol., 2008, 9(4), 253–60 CAS; (b) V. M. Platt and F. C. Szoka Jr, Anticancer therapeutics: targeting macromolecules and nanocarriers to hyaluronan or CD44, a hyaluronan receptor, Mol. Pharmaceutics, 2008, 5(4), 474–86 CrossRef CAS.
  56. Y. Luo, N. J. Bernshaw, Z. R. Lu, J. Kopecek and G. D. Prestwich, Targeted delivery of doxorubicin by HPMA copolymer-hyaluronan bioconjugates, Pharm. Res., 2002, 19(4), 396–402 CrossRef CAS.
  57. D. Coradini, C. Pellizzaro, G. Miglierini, M. G. Daidone and A. Perbellini, Hyaluronic acid as drug delivery for sodium butyrate: improvement of the anti-proliferative activity on a breast-cancer cell line, Int. J. Cancer, 1999, 81(3), 411–6 CrossRef CAS.
  58. (a) Y. Luo and G. D. Prestwich, Synthesis and selective cytotoxicity of a hyaluronic acid-antitumor bioconjugate, Bioconjugate Chem., 1999, 10(5), 755–63 CrossRef CAS; (b) Y. Luo, M. R. Ziebell and G. D. Prestwich, A hyaluronic acid-taxol antitumor bioconjugate targeted to cancer cells, Biomacromolecules, 2000, 1(2), 208–18 CrossRef CAS.
  59. Y. Ogawa, K. Kubota, H. Ue, Y. Kataoka, M. Tadokoro, K. Miyatake, K. Tsuzuki, T. Yamanishi, S. Itoh, J. Hitomi, N. Hamada, S. Kariya, M. Fukumoto, A. Nishioka and T. Inomata, Phase I study of a new radiosensitizer containing hydrogen peroxide and sodium hyaluronate for topical tumor injection: a new enzyme-targeting radiosensitization treatment, Kochi Oxydol-Radiation Therapy for Unresectable Carcinomas, Type II (KORTUC II), Int. J. Oncol., 2009, 34(3), 609–18 CAS.
  60. G. Pitarresi, E. F. Craparo, F. S. Palumbo, B. Carlisi and G. Giammona, Composite nanoparticles based on hyaluronic acid chemically cross-linked with alpha,beta-polyaspartylhydrazide, Biomacromolecules, 2007, 8(6), 1890–8 CrossRef CAS.
  61. H. Lee, H. Mok, S. Lee, Y. K. Oh and T. G. Park, Target-specific intracellular delivery of siRNA using degradable hyaluronic acid nanogels, J. Controlled Release, 2007, 119(2), 245–52 CrossRef CAS.
  62. (a) M. S. Cohen, S. Cai, Y. Xie and M. L. Forrest, A novel intralymphatic nanocarrier delivery system for cisplatin therapy in breast cancer with improved tumor efficacy and lower systemic toxicity in vivo, Am. J. Surg., 2009, 198(6), 781–6 CrossRef CAS; (b) Y. I. Jeong, S. T. Kim, S. G. Jin, H. H. Ryu, Y. H. Jin, T. Y. Jung, I. Y. Kim and S. Jung, Cisplatin-incorporated hyaluronic acid nanoparticles based on ion-complex formation, J. Pharm. Sci., 2008, 97(3), 1268–76 CrossRef CAS; (c) B. Rosenberg, Platinum Complexes for Treatment of Cancer, Interdisciplinary Science Reviews, 1978, 3(2), 134–147 CAS.
  63. A. K. Yadav, P. Mishra, S. Jain, A. K. Mishra and G. P. Agrawal, Preparation and characterization of HA-PEG-PCL intelligent core-corona nanoparticles for delivery of doxorubicin, J. Drug Targeting, 2008, 16(6), 464–78 CrossRef CAS.
  64. H. Lee, C. H. Ahn and T. G. Park, Poly[lactic-co-(glycolic acid)]-grafted hyaluronic acid copolymer micelle nanoparticles for target-specific delivery of doxorubicin, Macromol. Biosci., 2009, 9(4), 336–42 CrossRef CAS.
  65. D. Peer and R. Margalit, Tumor-targeted hyaluronan nanoliposomes increase the antitumor activity of liposomal Doxorubicin in syngeneic and human xenograft mouse tumor models, Neoplasia, 2004, 6(4), 343–53 CrossRef CAS.
  66. D. Peer and R. Margalit, Loading mitomycin C inside long circulating hyaluronan targeted nano-liposomes increases its antitumor activity in three mice tumor models, Int. J. Cancer, 2004, 108(5), 780–9 CrossRef CAS.
  67. H. Lee, K. Lee, I. K. Kim and T. G. Park, Fluorescent gold nanoprobe sensitive to intracellular reactive oxygen species, Adv. Funct. Mater., 2009, 19(12), 1884–1890 CrossRef CAS.
  68. Y. Lee, H. Lee, Y. B. Kim, J. Kim, T. Hyeon, H. Park, P. B. Messersmith and T. G. Park, Bioinspired surface immobilization of hyaluronic acid on monodisperse magnetite nanocrystals for targeted cancer imaging, Adv. Mater. Deerfield, 2008, 20(21), 4154–4157 Search PubMed.
  69. L. Shiftan, T. Israely, M. Cohen, V. Frydman, H. Dafni, R. Stern and M. Neeman, Magnetic resonance imaging visualization of hyaluronidase in ovarian carcinoma, Cancer Res., 2005, 65(22), 10316–23 CrossRef CAS.
  70. J. Kim, K. S. Kim, G. Jiang, H. Kang, S. Kim, B. S. Kim, M. H. Park and S. K. Hahn, In vivo real-time bioimaging of hyaluronic acid derivatives using quantum dots, Biopolymers, 2008, 89(12), 1144–53 CrossRef CAS.
  71. K. S. Kim, W. Hur, S. J. Park, S.W. Hong, J. E. Choi, E. J. Goh, S. K. Yoon and S. K. Hahn, Bioimaging for targeted delivery of hyaluronic acid derivatives to the livers in cirrhotic mice using quantum dots, ACS Nano, 2010, 4(6), 3005–14 CrossRef CAS.
  72. S. H. Bhang, N. Won, T. J. Lee, H. Jin, J. Nam, J. Park, H. Chung, H. S. Park, Y. E. Sung, S. K. Hahn, B. S. Kim and S. Kim, Hyaluronic acid-quantum dot conjugates for in vivo lymphatic vessel imaging, ACS Nano, 2009, 3(6), 1389–98 CrossRef CAS.
  73. E. N. Harris, S. V. Kyosseva, J. A. Weigel and P. H. Weigel, Expression, processing, and glycosaminoglycan binding activity of the recombinant human 315-kDa hyaluronic acid receptor for endocytosis (HARE), J. Biol. Chem., 2007, 282(5), 2785–97 CAS.
  74. S. Gustafson, The influence of sulfated polysaccharides on the circulating levels of hyaluronan, Glycobiology, 1997, 7(8), 1209–14 CrossRef CAS.
  75. (a) B. Zhou, J. A. Weigel, L. Fauss and P. H. Weigel, Identification of the hyaluronan receptor for endocytosis (HARE), J. Biol. Chem., 2000, 275(48), 37733–41 CrossRef CAS; (b) S. Sugahara, S. Okuno, T. Yano, H. Hamana and K. Inoue, Characteristics of tissue distribution of various polysaccharides as drug carriers: influences of molecular weight and anionic charge on tumor targeting, Biol. Pharm. Bull., 2001, 24(5), 535–43 CrossRef CAS; (c) K. Y. Choi, H. Chung, K. H. Min, H. Y. Yoon, K. Kim, J. H. Park, I. C. Kwon and S. Y. Jeong, Self-assembled hyaluronic acid nanoparticles for active tumor targeting, Biomaterials, 2010, 31(1), 106–14 CrossRef CAS; (d) V. Orian-Rousseau, CD44, a therapeutic target for metastasising tumours, Eur. J. Cancer, 2010, 46(7), 1271–7 CrossRef CAS.
  76. J. Greiner, M. Schmitt, L. Li, K. Giannopoulos, K. Bosch, A. Schmitt, K. Dohner, R. F. Schlenk, J. R. Pollack, H. Dohner and L. Bullinger, Expression of tumor-associated antigens in acute myeloid leukemia: Implications for specific immunotherapeutic approaches, Blood, 2006, 108(13), 4109–17 CrossRef CAS.
  77. C. A. Maxwell, J. J. Keats, A. R. Belch, L. M. Pilarski and T. Reiman, Receptor for hyaluronan-mediated motility correlates with centrosome abnormalities in multiple myeloma and maintains mitotic integrity, Cancer Res., 2005, 65(3), 850–60 CAS.
  78. (a) K. Giannopoulos, A. Dmoszynska, M. Kowal, J. Rolinski, E. Gostick, D. A. Price, J. Greiner, M. Rojewski, S. Stilgenbauer, H. Dohner and M. Schmitt, Peptide vaccination elicits leukemia-associated antigen-specific cytotoxic CD8+ T-cell responses in patients with chronic lymphocytic leukemia, Leukemia, 2010, 24(4), 798–805 CrossRef CAS; (b) M. Schmitt, A. Schmitt, M. T. Rojewski, J. Chen, K. Giannopoulos, F. Fei, Y. Yu, M. Gotz, M. Heyduk, G. Ritter, D. E. Speiser, S. Gnjatic, P. Guillaume, M. Ringhoffer, R. F. Schlenk, P. Liebisch, D. Bunjes, H. Shiku, H. Dohner and J. Greiner, RHAMM-R3 peptide vaccination in patients with acute myeloid leukemia, myelodysplastic syndrome, and multiple myeloma elicits immunologic and clinical responses, Blood, 2008, 111(3), 1357–65 CAS.

Footnote

Published as part of an Integrative Biology themed issue in honour of Mina J. Bissell: Guest Editor Mary Helen Barcellos-Hoff.

This journal is © The Royal Society of Chemistry 2011