Next Article in Journal
Update in Lung Cancer Molecular Pathology: Technological Advances and Clinical Practice
Next Article in Special Issue
CD44 Intracellular Domain: A Long Tale of a Short Tail
Previous Article in Journal
Clinical Management of Patients with Gastric MALT Lymphoma: A Gastroenterologist’s Point of View
Previous Article in Special Issue
The Landscape of Small Leucine-Rich Proteoglycan Impact on Cancer Pathogenesis with a Focus on Biglycan and Lumican
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Effects of Hyaluronan on Breast Cancer Aggressiveness

Department of Medicine and Surgery, University of Insubria, 21100 Varese, Italy
*
Author to whom correspondence should be addressed.
Cancers 2023, 15(15), 3813; https://doi.org/10.3390/cancers15153813
Submission received: 29 June 2023 / Revised: 21 July 2023 / Accepted: 24 July 2023 / Published: 27 July 2023
(This article belongs to the Special Issue Matrix Effectors and Cancer)

Abstract

:

Simple Summary

Breast cancer is the most common neoplasm in women. Although the primary tumor does not appear in a vital organ, lethality is due to the ability of tumor cells to invade and seed distant organs, causing metastases. Approaches to reduce breast cancer cell aggressiveness target hormone receptors that sustain cell growth and motility. However, other factors contribute to aberrant cell behaviors in cancer cells, and nowadays, the role of the environment surrounding cancer cells is evident. The extracellular matrix polysaccharide hyaluronan is a ubiquitous component of the tumor microenvironment that not only modulates cell growth and movement but also plays a critical role in modulating the inflammatory response. In this review, we discuss the role of hyaluronan in relation to the expression of critical hormone receptors.

Abstract

The expression of the estrogen receptor (ER), progesterone receptor (PR), and human epidermal growth factor receptor 2 (HER2) in breast cancer cells is critical for determining tumor aggressiveness and targeting therapies. The presence of such receptors allows for the use of antagonists that effectively reduce breast cancer growth and dissemination. However, the absence of such receptors in triple-negative breast cancer (TNBC) reduces the possibility of targeted therapy, making these tumors very aggressive with a poor outcome. Cancers are not solely composed of tumor cells, but also include several types of infiltrating cells, such as fibroblasts, macrophages, and other immune cells that have critical functions in regulating cancer cell behaviors. In addition to these cells, the extracellular matrix (ECM) has become an important player in many aspects of breast cancer biology, including cell growth, motility, metabolism, and chemoresistance. Hyaluronan (HA) is a key ECM component that promotes cell proliferation and migration in several malignancies. Notably, HA accumulation in the tumor stroma is a negative prognostic factor in breast cancer. HA metabolism depends on the fine balance between HA synthesis by HA synthases and degradation yielded by hyaluronidases. All the different cell types present in the tumor can release HA in the ECM, and in this review, we will describe the role of HA and HA metabolism in different breast cancer subtypes.

Graphical Abstract

1. Introduction

Breast cancer remains a significant global health concern, as almost 25% of cancer cases among women are breast cancer incidents [1]. Breast cancer is a heterogeneous disease consisting of different entities affecting the same anatomical organ and originating in the same anatomical structure (i.e., the terminal duct-lobular unit). Heterogeneity represents the primary challenge in treating breast cancer [2], and understanding the critical biomarkers, together with their role in carcinogenesis, drug resistance, and their use for diagnosis and therapy, is crucial for effective breast cancer treatment. The traditional histological classification of breast carcinomas, based on the diversity of the morphological features of the tumors, has the main drawback that approximately 80% of all breast cancers will eventually belong to either one of the two major histopathological classes, namely invasive ductal carcinomas or invasive lobular carcinoma. This implies that tumors with very different biological and clinical profiles are grouped, resulting in minimal prognostic and predictive capabilities and modest clinical utility [3]. To compensate for the limited prognostic and predictive power of the histopathological classification of breast carcinomas, Perou et al. introduced a new method of clustering breast tumors depending on their gene expression profiles and the presence or absence of estrogen receptors (ER), progesterone receptors (PR), and human epidermal growth factor receptor 2 (HER2). In such a way, breast carcinomas can be classified into five main molecular subtypes: luminal A (ER+/PR+/HER2 with low or intermediate differentiation), luminal B (ER+/PR/HER2+ with high differentiation grade), basal-like (ER/PR/HER2) also called triple-negative breast cancer (TNBC), HER2-enriched (ER/PR/HER2+), and normal breast-like (overexpressing genes of adipose and other nonepithelial cells) [4,5]. In addition to these five, another class, called Claudin Low, was discovered, characterized by low gene expression of the tight junction proteins claudin 3, 4, 7, and E-cadherin [6].
Almost 70% of breast cancers are ER+ and hormone-dependent, making ER expression an optimal prognostic marker for responsiveness to treatment [7,8]. ER+ breast cancer is generally associated with a lower risk of recurrence, longer overall survival, and eventually a better prognosis than ER breast cancer. This difference is mainly attributed to the availability of hormonal therapies for ER+ breast cancer, including aromatase inhibitors, selective estrogen receptor modulators (SERMs), and selective estrogen receptor degraders (SERDs) [9]. In contrast, lacking ER expression, TNBC patients are more predisposed to adverse outcomes, recurrence, and metastasis than those affected by other breast cancer subtypes [10,11,12]. Their treatment mainly relies on cytotoxic chemotherapy, either before (neoadjuvant) or after surgery (adjuvant), which is inadequate, even though TNBC cases often benefit from chemotherapy to a greater degree than other breast cancer subtypes [13]. Interestingly, ERα silencing in MCF-7 cells enhanced cell proliferation, migration, and invasion, and thus, epithelial-to-mesenchymal transition (EMT), by changing the expression of critical matrix effectors [14]. The lack of ER, PR, and HER2 expression makes the development and use of TNBC’s targeted therapies highly challenging. Apart from anthracyclines (such as doxorubicin) and taxanes (such as paclitaxel), poly (ADP-ribose) polymerase (PARP) inhibitors (such as olaparib) and immunotherapy (such as atezolizumab) have been approved for use in combination with chemotherapy [15,16,17].
The major problem with breast cancer is its tendency to form metastases. Surgical removal of the primary site of the tumor is often insufficient to avoid distant metastases that typically develop in the liver, bones, lungs, and brain. EMT is the process enabling the development of more invasive tumors that, indeed, are characterized by a mesenchymal-like phenotype [18]. During EMT, epithelial tumor cells lose their cell-cell adhesion, dedifferentiate, gain a mesenchymal phenotype, and secrete an abnormal amount of extracellular matrix (ECM) in a process that highly resembles wound healing. Thus, cells acquire both invasive and metastatic capacities, increasing cell motility, invasiveness, and resistance to apoptosis [18,19,20]. These changes allow cells to detach from the primary tumor, invade the surrounding tissues, intravasate into the blood or lymphatic vessels, and ultimately form secondary tumors at distant sites, a process known as metastasis.
The tumor niche plays a pivotal role in breast cancer progression and metastasis [21]. The migratory behavior of cells and EMT itself are driven not only by the expression of a plethora of transcription factors (i.e., the snail/slug family, twist, EF1/ZEB1, SIP1/ZEB2, and E12/E47) [22], but a crucial role is also played by the ECM [23]. Depending on its composition, the chemical and mechanical properties of the ECM can vary, thereby influencing cell behavior. The ECM not only represents a scaffold on which cells attach and grow but also behaves as a sort of cage, trapping cells and limiting their movements. Interestingly, the degradation of ECM makes available growth factors that are normally bound to ECM components, such as proteoglycans (PG), and fragmented ECM molecules can have signaling activity that stimulates cancer cells [24]. Therefore, it is evident that ECM components play a crucial role in tumor biology.
Among the different components of ECM, this review focuses on the role of hyaluronan (HA) in relation to the aggressiveness of breast cancer cells.

2. The Role of HA in Tumor Stroma

HA, a prominent component of the ECM that is almost ubiquitously present in all vertebrates tissues, is an atypical glycosaminoglycan (GAG), as it is not covalently bound to any PG core protein, lacks any chemical modification, and is the only one produced outside the Golgi [25]. Indeed, the enzymes HA synthase 1, 2, and 3 (HAS1, 2, and 3) are located on the cellular membrane and directly extrude the nascent HA molecule outside the cells. HA polymers consist of repeating disaccharide units of D-glucuronic acid (GlcUA) and N-acetyl-D-glucosamine (GlcNAc), linked together by alternating β-1,4 and β-1,3 glycosidic bonds [26].
HA is ubiquitously present in all ECM tissues, showing many important physiological and pathological functions. The best-known role of HA is its structural function; HA acts as a chemical glue, allowing the formation of a stable ECM through the ability to interact with several proteoglycans (such as aggrecan and versican) [27]. Such HA complexes determine tissue mechanical properties, such as flexibility and stiffness. HA binds to several receptors, such as CD44, the HA-mediated motility receptor (HMMR, RHAMM, or CD168), and Lyve-1. It is known that the interaction of HA with these receptors can influence survival, proliferation, and motility through signaling cascades (Figure 1).
HA is synthesized as a high molecular mass polysaccharide with an average size of 4 × 106 Da, and its turnover is assured by a complex family of degrading enzymes, including HYALs, PH-20, CEMIP, and regulatory proteins such as TMEM2 [28,29,30]. HA fragments are known to have distinct cellular functions compared to the full-size molecule; high-molecular-mass HA (HMWHA; >500 kDa) inhibits, whereas low-molecular-mass HA (LMWHA; <120 kDa) promotes angiogenesis, inflammation, and proliferation [31,32,33,34] (Figure 1). However, some exceptions have been observed; specific LMWHA is critical for proper gut development [35]. Notably, LMWHA can be generated by impaired homeostasis, HAS dysregulation, UV light, or oxidative stress [36]. HA accumulation is detected wherever rapid tissue remodeling and repair are present, that is, during embryogenesis, wound healing, inflammation, and even tumorigenesis. Among all the matrix components, HA is one of the most deregulated in human malignancies, and its amount in the tumor stroma affects overall survival and outcome [37,38,39]. High levels of HA are observed in several tumors compared to healthy tissues, such as renal [40], hepatocellular [41], head and neck [42] and lung squamous cell carcinomas [43], lymphoma [44], glioma [45], prostate [46], melanoma [47], breast [48], and ovarian [49] cancers. In most cancer cases, stromal accumulation of HA strongly correlates with an unfavorable prognosis and decreased survival probability [37,49,50]. As previously highlighted [51,52,53,54], obesity and type-2 diabetes are both risk factors for breast cancer. In particular, breast cancer in fatty breasts generally contains high levels of M2-like tumor-associated macrophages, both derived from the general state of inflammation of the tissue and their recruitment [55] and polarization towards the M2-phenotype due to cable-like HA [56].
HA involvement in supporting tumor metastasis has been widely reported; remarkably, other than driving tumor cells into the endothelium, LMWHA also supports endothelial budding and capillary formation in 3D matrices, thus increasing breast tumor angiogenesis and lymphangiogenesis [36,57].
The aberrant accumulation of HA in malignant tissues is mainly due to the deregulated activity of HAS, hyaluronidases, reactive oxygen species (ROS), and hyaladerins (mainly CD44) [58]. Moreover, it must be considered that HA is synthesized not only by cancer cells; stromal cells are also stimulated to produce a high amount of HA under the stimulation of factors released by cancer cells themselves, as well as infiltrating immune cells, cancer-associated fibroblasts, and others [21,49,59,60,61,62]. The first studies demonstrating that stromal cells produce HA upon interaction with tumor cells date back to the 1980s [63,64]. Notably, cancer-associated stromal cells can stimulate epithelial cells to transform into malignant cells owing to the release of growth factors, chemokines, and cytokines, including fibroblast growth factor (FGF), platelet-derived growth factor (PDGF-BB), transforming growth factor β (TGFβ), interleukin-6 (IL-6), IL-8, and chemokine (C-X-C motif) ligand 7 (CXCL7) [21]. Currently, the involvement of HA in tuning the Hanahan and Weinberg hallmarks of cancer is well accepted and has become a very promising target in cancer research [65,66].

3. HA Synthesis and Metabolism

HASs are the main regulators of HA synthesis, possessing all the enzymatic activities to produce the HA polysaccharide chain and extrude it into the ECM. Indeed, they have a single catalytic domain, which transfers two different donor sugars and forms substrate-specific glycosidic linkages, and a transmembrane channel derived from the spatial organization of their membrane-embedded segment, through which the nascent HA chain is secreted into the extracellular space [67]. Recently, the structure of HAS was analyzed using cryoelectron microscopy, shedding light on how HAS selects its substrates, hydrolyzes the first substrate to prime the synthesis reaction, opens an HA-conducting transmembrane channel, ensures alternating substrate polymerization, and coordinates HA inside the channel, confirming the “pendulum” model proposed by Weigel et al. [68,69].
Humans express three different HAS isoforms, derived from three related and evolutionarily highly conserved genes, yet located on three different chromosomes, differing in tissue expression and catalytic activity [70].
An important feature of all HASs is their capability to synthesize HA starting from two UDP-sugar precursors, exploiting the high energy content of the two substrates and thus avoiding the need for ATP [71]. Interestingly, although ATP is not directly involved in HA production, the synthesis of the two UDP-sugar precursors (i.e., UDP-GlcUA and UDP-GlcNAc) is affected not only by the energetic status of the cells but also by other metabolic pathways. The double oxidation of carbon 6 of UDP-glucose to produce UDP-GlcUA leads to the production of two NADH molecules that influence the NAD:NADH ratio, which is critical for the activity of a variety of enzymes, including dehydrogenases and sirtuins [25,26,72]. UDP-GlcUA is an important substrate for glucuronyltransferase detoxifying enzymes [73]; therefore, an increase in UDP-GlcUA contributes to enhanced chemoresistance in aggressive cancers. Together with UDP-GlcUA, UDP-glucose dehydrogenase (UGDH), which is responsible for the conversion of UDP-glucose into UDP-GlcUA, has an important role in drug detoxification. Interestingly, its knockdown in MDA-MB-231 cells contributed to epirubicin chemoresistance, which is associated with increased deposition and catabolism of HA [74] (Figure 1).
UDP-GlcNAc is synthesized by the hexosamine biosynthetic pathway, which integrates the metabolism of sugars, amino acids, fatty acids, and nucleotides. Therefore, in addition to the presence of HASs, the availability of UDP-GlcUA and UDP-GlcNAc is critical for proper HA biosynthesis. Cancer cells show a deep alteration in metabolism, which can greatly modify HA biosynthesis [66]. Indeed, aggressive breast cancer tumors, characterized by deeply disturbed glucose metabolism, have a higher amount of UDP-sugars, which eventually influence the cancer niche by inducing HA production and deposition without affecting the expression levels of any of the HAS [75].
The first stages of tumorigenesis are characterized by a limited oxygen supply, and the hallmark of breast cancer cells is the Warburg effect, which is dominated by an increased flux of glucose through anaerobic glycolysis to generate ATP and lactate [76]. Although experimental data in this early phase of cancer development are scarce, it can be hypothesized that HA synthesis in neoplastic cells is low. Firstly, glucose should be mainly channeled into catabolic processes rather than shunted to anabolism; second, the NADH produced during the oxidation of UDG-glucose to UDP-GlcUA could be difficult to convert to NAD in the absence of oxygen. Interestingly, the elevated HA in cancer stroma could be produced by fibroblasts, as lactate stimulates HA synthesis in this kind of cell [77]. HA is important for breast cancer cells because it ensures efficient lactate efflux [78]. Furthermore, several HA-metabolizing genes, such as CD44 and HYAL1, possess lactate-responsive elements [79], which could favor rapid HA turnover leading to the formation of important building block intermediates (such as ribose and NADPH; see later in Section 4—HA catabolism) needed for rapidly growing cells. In contrast, in the late stages of tumorigenesis, when angiogenesis happens, oxygen availability is no longer limited, and aggressive breast cancer cells can produce an elevated amount of HA, thus favoring several aspects of tumor malignancy, including motility and stemness [80].

3.1. The Role of HASes in Breast Cancer

The presence of three HAS isoenzymes in humans remains unclear, as their exact role in pathophysiology is only partially known. What is known is that they have distinct spatial and temporal expressional and transcriptional patterns and regulators [81,82,83]. Moreover, each HAS can synthesize HA chains of different sizes; HAS1 and -2 catalyze the formation of HMWHA, whereas HAS 3 favors the production of LMWHA [34]. Although all three HASs are found in various tumor cell types and both HAS2 and HAS 3 expression is correlated with malignant transformation, HAS2 is the most efficient HA-synthesizing enzyme [30,84]—its expression is fundamental during embryogenesis, as HAS2-deficient mice result in embryonic lethality [85], while HAS1 and HAS3 deletions have only minor effects on the phenotype [81]. HAS1 expression is generally low in normal breast tissue but is upregulated in breast cancer. Its overexpression enhances the metastatic potential of breast cancer cells [86] and correlates with shorter overall survival, a higher relapse rate, ER negativity, and HER2 positivity [87] (Table 1). Interestingly, its localization seems to be mostly cytoplasmic, where it cooperates in promoting breast cancer cell growth, intratumoral heterogeneity, and a cancer stem cell-like phenotype [88,89].
Several studies have highlighted the predominant transcriptional activity of the HAS2 gene in ER aggressive breast cancer cell lines compared to the limited expression of HAS3 in nonaggressive ER+ cell lines [91,92,93] (Table 1). It has to be highlighted that stromal cells play a pivotal role in producing abnormal amounts of HA, and in these cells, HAS1, and not HAS2, is the strongest isoform affecting tumor relapse and patient prognosis [87]. In breast cancer, both high HAS2 and HA levels have been shown to induce the recruitment of tumor-associated macrophages (TAM) and tumor neovascularization; stromal HA, via macrophage recruitment, remodels the local microenvironment to promote the formation of tumor vasculatures [94,95]. In TNBC (i.e., MDA-MB-231 and Hs578T), HAS2 mRNA levels are significantly higher than those in less aggressive ER+ cell lines (i.e., MCF-7) [96] (Table 1). Notably, HAS2 overexpression has also been correlated with TGFβ-mediate EMT [97].
Furthermore, HAS2 also plays a pivotal role in regulating cell motility and invasion, as its expression levels are higher in bone metastases than in parental MDA-MB-231 cells [93]. HAS2 expression was also found to be elevated in the cancer stem cell population of breast cancer bone metastasis, and its suppression decreased both the incidence and growth of metastatic lesions [98]. In the ER+ breast cancer cell line, overexpression of HAS2 increased the invasive and migratory abilities of cells, together with the downregulation of epithelial markers (e.g., E-cadherin, β-catenin, and ZO-1), upregulation of mesenchymal markers (e.g., N-cadherin and vimentin), and promotion of invadopodia formation [99]. Finally, breast tumor biopsies showed enhanced angiogenesis and the recruitment of inflammatory cells when HAS2 expression was elevated. In support of these findings, HAS2 suppression by interfering RNA or 4-methylumbelliferone (4-MU) administration reduces tumorigenesis [91,100,101].

3.2. HAS Inhibition by 4-MU

Furthermore, 4-MU is a specific inhibitor of HAS activity, as it reduces the availability of UDP-GlcUA [102] and, indirectly, lowers HAS expression [103,104]. Its effectiveness in reducing tumor growth and aggressiveness has been reported in different cancer types [105,106,107,108]. Interestingly, 4-MU has been shown to reduce HA synthesis and HAS2 and CD44 expression, but increases HYAL1 and HYAL2 in breast cancer cell lines, and to a greater extent in ER cells. Moreover, only ER cell lines showed reduced migration, adhesion, and invasion [101]. However, 4-MU treatment is not specific for any particular HAS, and can partially affect the synthesis of other GAG chains [102]. It is reported that 4-MU causes potential detrimental effects, as described for glycocalyx alteration in atherosclerotic animal models [109] and increased tumorigenicity in hepatocellular carcinoma mouse models [110]. Recently, the potential limited systemic efficacy of 4-methylumbelliferyl glucuronide (4-MUG) was tested in mice. Additionally, 4-MUG reduced HA synthesis independently of its conversion into 4-MU without depleting the HA precursor UDP-GlcUA. However, 4-MUG is not specific to any HAS isoform [111].

3.3. HAS2 Regulations

HAS2 is the only enzyme that exhibits fine-tuned activity via post-translational modifications (Figure 1). Adenosine monophosphate-activated protein kinase (AMPK), a cellular energy sensor, induces the phosphorylation of HAS2 threonine 110, blocking its enzymatic activity and eventually connecting HAS2 activity with the energetic status of the cell [112]. AMPK generally activates catabolic processes to restore ATP levels and inhibit ATP-consuming pathways (i.e., anabolism). Since anabolic pathways are preferred to catabolic pathways in tumors, it is not surprising that AMPK is often found to be blocked in malignancies. In breast cancer, AMPK downregulation has been correlated with the loss of ER expression and a poor prognosis [113,114,115,116,117]. Liver kinase B1 (LKB1), an upstream activator of AMPK, is mutated in numerous malignancies, thus blocking the AMPK pathway and favoring HA synthesis [118]. Interestingly, salicylate has been demonstrated to induce AMPK, eventually downregulating HAS2 expression, HA production, and metastatic breast cancer cell proliferation by inducing cell growth arrest [119].
Other kinases, such as extracellular signal-regulated kinases (ERKs) and protein kinase C (PKC), seem to play a role in regulating HAS2 activity [120]. Finally, HAS2 can form dimers or oligomers upon monoubiquitination of the K190 residue, which plays a key role in its activity and dimerization [121].
More than being solely a precursor for HA synthesis, GlcNAc is also the substrate for O-GlcNAc transferase (OGT), which has a central role in the regulation of metabolism in response to nutrient availability. OGT transfers GlcNAc moieties to serine 221 of the HAS2 protein, increasing its stability in the plasma membrane (up to 5 h) and, thus, HA synthesis [122]. Interestingly, O-GlcNAcylation of the NF-κB subunit p65 plays a critical role in the induction of HAS2 transcription, involving HA synthase 2 antisense 1 (HAS2-AS1) long noncoding RNA (lncRNA) as a mediator of chromatin remodeling [122].

3.4. Epigenetic Regulation of HAS2

HAS2-AS1 was first described by Chao and Spicer in 2005 as the natural antisense of the HAS2 sequence, as it is located on chromosome 8q24.13 and is transcribed from the opposite strand of the HAS2 gene locus. In 2011, Michael and colleagues proved for the first time that under the stimulation of interleukin-1 β (IL-1β) and TGFβ, HAS2 mRNA physically interacts with HAS2-AS1 in proximal tubular epithelial cells, finally stabilizing and promoting HAS2 expression [123]. LncRNAs are known to have tissue-specific effects, and HAS2-AS1 is no exception [124]. Its overexpression in osteosarcoma cells reduces HAS2 transcript [125], while in oral squamous cell carcinoma, it induces HAS2 transcription and hypoxia-induced invasiveness [126]. In mouse mammary gland epithelial cells, HAS2-AS1 is known to induce cell transformation via TGFβ [127]. Although it is well accepted that HAS2-AS1 works as a cis epigenetic regulator of HAS2 in aortic smooth muscle cells (AoSMCs) [72,128], this mechanism of action seems to be invalid in breast cancer. Indeed, HAS2-AS1 expression was higher in TNBC cell lines (MDA-MB-231 and Hs 578T) than in ER+ cells (MCF-7 and T-47D), and its overexpression in MDA-MB-231 and Hs 578T cells reduced cell aggressiveness by inducing mesenchymal-to-epithelial transition and reducing invasiveness, motility, and cell viability (Table 1). Notably, in these cells, HAS2-AS1 was involved neither in the regulation of HAS2 nor in HA deposition. Surprisingly, its expression was higher in TNBC cell lines than in ER+ cells, and its expression directly correlates with ER breast cancer patients’ survival [90]. In addition to being solely an epigenetic regulator of the HAS2 gene, HAS2-AS1 can also be found in the cytoplasm [127]. Different studies have reported the intriguing ability of HAS2-AS1 to interact with other RNAs, acting as competing endogenous RNAs (ceRNAs) for miRNAs. An increasing number of studies have reported that HAS2-AS1/miRNA interactions are mainly functional in cancer [129,130].
HA synthesis can also be increased by golgins and, in particular, by the protein c10orf118, which is abundant in the conditioned media of breast cancer cell lines. c10orf118 induces HAS2 expression in dermal fibroblasts, thereby stimulating HA synthesis [131,132].
Some studies have reported the involvement of HAS2 in breast cancer plasticity and [133] chemotherapy drug resistance [134,135]. Notably, HAS2 knockdown in ER+ breast cancer cell lines induces upregulation of Ezrin, downregulation of ER and, thus, antiestrogen resistance [136].

4. HA Catabolism

In mammalian tissues, the length of the HA chain varies significantly [137]. The polydisperse sizing of HA mainly depends on the catalytic abilities of HYALs, which are hydrolases that cleave the β-(1,4) linkage between GlcNAc and GlcUA. There are six known hyaluronidases in humans: HYALs 1-4, HYALP, and PH-20 [138]. HYALs’ degrading activity is not only related to HA; HYAL4, for example, can cleave chondroitin sulfate, with no evidence of HA catabolic activity [139]. Among all HYALs, HYAL1, and HYAL2 are the predominant isoforms that cleave HA in cooperation with CD44. At the plasma membrane, HYAL2 chops HMWHA into small fragments that are soon internalized into the cells via endocytosis and further degraded in the lysosomes by HYAL1. Inside the cells, HA is completely degraded by the coordinated action of HYALs, β-glucuronidase, and hexosaminidase, leading to free GlcUA and GlcNAc [29]. Eventually, GlcUA is converted to xylulose-5-phosphate, which sustains the hexose monophosphate pathway for the synthesis of NADPH and ribose.
Recently, other HA-degrading enzymes have been reported to exhibit hyaluronidase activity, including the HA-binding protein involved in HA depolymerization (HYBID, also known as CEMIP) and TMEM2 [140,141]. In contrast to HYALs, which function at an acidic pH, TMEM2 functions at a physiological pH and, being regulated by SOX4, is known to be associated with metastatic migration and invasion of breast cancer cells [142]. CEMIP degrades HA after clathrin-dependent internalization into endosomes and is known to promote the progression and metastasis of numerous malignancies, including breast cancer [141,143,144,145].
The catabolism of HA results in the generation of bioactive fragments with contrasting size-dependent functions [146,147]. For instance, HMWHA is known to have antiangiogenic, antiproliferative, and immunosuppressive properties. On the other hand, LMWHA induces inflammation and angiogenesis, has immunostimulatory functions, and induces tissue reparative processes, as described in wound healing [137,148]. As proof, naked mole-rats, which possess HA chains longer than 12 million Da, have an unusually long life, approximately ten times longer than other rodents, and an incredible resistance to tumor development and spreading when cancer cells are injected into the dermis [149]. Recently, it was demonstrated that these animals also express very low levels of HYAL1 and HYAL3 [150].

The Role of HA Oligosaccharides in Breast Cancer

Metastatic breast cancer is associated with an increased amount of serum LMWHA, and BT-549 and Hs578T TNBC cell lines have been found to produce a high amount of LMWHA, making it a good prognostic marker [151]. Contrary to what happens in the naked mole rat, tumors typically express high levels of HYALs. These enzymes also contribute to sustaining cancer metabolism and metabolic reprogramming, providing support to the pentose pathway to obtain reducing equivalents and ribose for anabolism and increasing the glycolytic rate to produce ATP [152]. Indeed, matrix remodeling through HYAL-mediated HA digestion increases glycolysis in several tumors, including breast cancer [153]. The combined overexpression of HAS and either HYAL1 [154] or HYAL2 [155] is characteristic of the invasive front of human breast cancer. TNBC cell lines have been shown to express high levels of HYAL2, HAS2, and CD44 [91] (Table 1).
In addition to being enzymatically degraded by HYALs, HA is also susceptible to ROS degradation [156,157]. Significant ROS production is typical in inflamed tissues, and cancer is a major inflammatory disease. Oxidative stress is simultaneously beneficial and detrimental for tumors; if it is essential in initiating and promoting malignant growth, an excess of oxidative stress could lead to the apoptosis of cancer cells. In this case, the abundant HA in the ECM acts as a protective shield against excessive reactive nitrogen species (RNS) and ROS. However, HA cleavage releases an excessive amount of bioactive fragments, promoting inflammatory pathologies, including cancer [156,158]. If HMWHA serves as a scavenger for ROS/RNS, HA fragments, especially those produced by ROS, act as danger signals, inducing the synthesis of cytokines, eventually exacerbating inflammation, and inducing classic tumor-associated pathways such as angiogenesis, as observed in breast cancer [159]. The presence of LMWHA in the serum of breast cancer patients is useful for discriminating between metastatic and nonmetastatic breast cancers [151].

5. Hyaladerins in Breast Cancer

The multiple and even contradictory effects of HA polymers of different sizes can be attributed to the presence of several HA receptors on the cell surface, which can trigger specific and differential signaling cascades. Among these, the most well-known are CD44, RHAMM, lymphatic vessel endothelial receptor 1 (LYVE-1), and the HA receptor for endocytosis (HARE) (Figure 1). It is also known that HA modulates toll-like receptors 2 and 4 (TLR2 and 4); however, HA binding to TLRs remains unclear [160]. Through the engagement of these binding partners, HA perturbs tissue homeostasis by driving cell motility, proliferation, apoptosis, and tissue remodeling.
The most prominent HA receptor is CD44, which is abundant in both inflammatory and cancer cells [160]. CD44 is a single-span transmembrane glycoprotein involved not only in mediating both cell-cell and cell-matrix interactions but is also a key player in transmitting HA signaling in tumor progression [161]. The peculiarity of the HA/CD44 interaction is that the receptor can acquire either a high- or low-affinity HA binding state. In humans, CD44 is encoded by a single gene that contains 20 exons, generating many variants (CD44v) and possessing different levels of glycosylation. The standard isoform of CD44 (CD44s) is ubiquitously expressed in all vertebrates, whereas CD44v is mainly expressed during inflammation and cancer [162]. Even though v6-v10 isoforms are most likely related to procancerous functions [163,164,165], CD44s have also been reported to induce EMT [166]. Notably, the switching between CD44s and CD44v plays a key role in regulating the EMT and plasticity of cancer cells [167]. Different breast cancer subtypes express peculiar CD44v—this diversity in CD44v expression has been ascribed to specific clinical markers (i.e., HER2, ER, PR), suggesting the involvement of CD44v in specific oncogenic signaling pathways [168]. Moreover, the CD44+/CD24 breast cancer cell subpopulation, typically identified as cancer stem cells, is associated with invasive properties and a poor prognosis [169]. In fact, the HA/CD44 interaction promotes cytoskeletal remodeling, thus favoring cell growth, survival, invasion, and metastasis.
The HA/CD44 interaction involves the BX7B HA-binding motif in the N-terminal region of the CD44 receptor. HA interaction with CD44 triggers receptor clustering and interaction with other transmembrane proteins, including receptor tyrosine kinases, serine/threonine kinase receptors, tumor necrosis factor receptor (TNFR)-like receptors, G-protein-coupled receptor CXCR4, Wnt receptor LRP5/6, CD147, c-Met, VEGFR2, PDGF, IGF1R, ErbB family, TGFβ, and ATP-binding cassette transporters [57,167,170,171]. In such a way, the CD44 cytoplasmic domain can be phosphorylated to transduce signals when the ligand binds to the extracellular domain. The signaling cascade triggered by the HA/CD44 interaction mainly stimulates the phosphoinositide 3-kinase/phosphoinositide-dependent kinase-1/protein kinase B (PI3K/PDK1/Akt) pathway, the Ras phosphorylation cascade involving RAF1, mitogen-activated protein kinase (MEK), and ERK1/2, as well as Wnt/b-catenin, NF-κB, and focal adhesion kinases (FAK). Consequently, the HA/CD44 interaction stimulates several oncogenic pathways and microRNA (miRNA) functions related mainly to cell adhesiveness, migration, and infiltration [172].
The CD44 intracellular domain (CD44ICD) can undergo cleavage and then translocate into the nucleus, acting at least in part as a transcriptional regulator [173,174]. In breast cancer, CD44ICD induces the transcription of stemness factors such as Nanog, Sox2, and Oct-4, thereby contributing to breast cancer aggressiveness and tumorigenesis [175].
HAS2 and CD44 are known to be highly expressed in ER breast cancer, promoting tumor aggressiveness. An interesting study highlighted the strict correlation between HAS2 expression, HA/CD44 signaling, and ER-dependent tumor aggressiveness. Indeed, HAS2 overexpression reduces the estrogen dependence of MCF7 cells by inhibiting the transcription of ER-driven genes [176]. This mechanism could at least partially explain the aggressiveness of ER cells, where HAS2 and CD44 expression are higher than in ER+ cells. Moreover, the increase in HA/CD44 interaction promotes EMT by suppressing the anti-EMT effects of estradiol [177].
The second most important HA receptor is RHAMM (also known as CD168), a coiled-coil protein containing the BX7B motif that can be found on cell membranes, in the cytoplasm, nucleus, or extracellular space. High RHAMM expression has been reported in several tumors, including breast, colon, brain, prostate, and endometrial [178,179,180,181]. When binding HA, RHAMM can interact with other receptors, such as the PDGF receptor (PDGFR), TGFβ receptor I (TGFβRI), and CD44 [162]. Consequently, several pathways are induced, including Ras, FAK, ERK1/2, PKC, tyrosine kinase pp60 (c-Src), NF-κB, and PI3K, eventually leading to cell migration, wound healing, tumorigenesis, and EMT [182,183].
A large body of evidence supports the role of TLR2 and -4 in HA signaling. TLRs are membrane receptors typical of the immune system and are mainly related to inflammation during viral infection and tumorigenesis [184,185,186]. Recently, it has been demonstrated that HA interaction with both CD44 and TLR4 sustains colon tumorigenesis, by promoting tumor growth in mice [187]. Similarly, the HA/TLR4 interaction in glioblastoma provokes cell differentiation by activating NF-κB [188].
Another interesting HA-binding protein is TNF-stimulated gene 6 (TSG-6), which yields the formation of cross-links between HA and the heavy chains of the serine protease inhibitor inter-alpha-inhibitor (IαI), thus stabilizing the structural integrity of the ECM. In general, TSG-6 is not constitutively expressed in normal adult tissues, but its expression is upregulated during inflammation, inflammation-like processes (such as ovulation), and cancer progression [189,190]. TSG-6 can bind HA with high affinity, mediating the generation of HA cable-like structures and enhancing HA-CD44 binding, which facilitates leukocyte migration to the site of inflammation. In breast tumors, cancer-associated fibroblasts (CAFs) are responsible for ECM deposition in the tumor niche. Notably, in normal mammary tissues, normal fibroblasts deposit much more HA and TSG-6 than CAFs in breast tumors, where the levels of TSG-6 and, consequently, HA-TSG-6 cross-linked structures are significantly decreased and associated with higher tumor malignancy [191]. Moreover, the interaction between HMWHA and TSG-6 affects the angiogenic behavior of monocytes/macrophages [192].

6. Future Directions

The evidence that HA is a central player in breast cancer, able to fine-tune many aspects of cell aggressivity, is clear and solid and has been confirmed in several experimental models. However, the precise role of each HAS isoenzyme in breast cancer progression remains unknown; therefore, the identification of HAS1, -2, and -3 specific regulators should be a priority to avoid a general alteration of HA homeostasis.
The finding that HA accumulates in aggressive breast cancers would require a thorough investigation regarding the identification of which cells (i.e., neoplastic or cancer-associated) synthesize such HA. For example, it is well known that in pancreatic tumors, cancer cells use HA as an energy source, exploiting the production of HA by stromal cells [193]. Therefore, a deep understanding of the crosstalk among different cell types in the tumor microenvironment is of great importance.
Future approaches to breast cancer could involve engineered immune cells (currently used for hematological malignancies) or molecules capable of starving tumor cells, such as those inhibiting amino acid metabolism (i.e., glutamine antagonists). HA coating around cancer cells can mask critical epitopes for immune cell recognition, and the starving approach could reduce HA synthesis. Therefore, new combined approaches involving HA inhibition should be considered to maximize the efficacy of future therapies.

7. Conclusions

Breast cancer is a highly heterogeneous malignancy with variable morphological, biological, and clinical features, and an efficient response to treatment is challenging. The discovery of the crucial role of hormone receptors in breast cancer cells has revolutionized therapies, allowing targeted approaches. However, it is evident that the tumor microenvironment plays a fundamental role in directing cancer-associated cells, infiltrating immune cells, vascular cells, and fibroblasts toward protumorigenic or antitumorigenic effects. HA is a ubiquitous component of the ECM in cancer and healthy tissues. Although mainly associated with aggressive ER-negative tumors, HA and HA metabolic enzymes play a critical role in breast cancer cell proliferation, metabolism, chemoresistance, and metastasis, and inhibition of HA synthesis with 4-MU could represent a new pharmacological target for the treatment of these cancers within the next few years. The study of new and specific regulators of HA metabolism, such as HAS2-AS1, could represent new strategies for treating this type of cancer in the future. Interestingly, the HA receptor CD44 is highly expressed in the subpopulation of breast cancer cells with stem cell potential, and HA-coated particles could be used as a vehicle to reach such an aggressive population of cells.

Author Contributions

Conceptualization, A.P. (Arianna Parnigoni) and D.V.; writing—original draft preparation, A.P. (Arianna Parnigoni), M.V., P.M., E.K. and D.V.; writing—review and editing, A.P. (Arianna Parnigoni), A.P. (Alberto Passi) and D.V.; supervision, D.V. All authors have read and agreed to the published version of the manuscript.

Funding

FAR2022-University of Insubria to A.P. (Alberto Passi), M.V., E.K. and D.V. Assegno di Ricerca Junior 2022 from the University of Insubria to A.P. (Arianna Parnigoni).

Data Availability Statement

Figures presented in this review were created with BioRender.com, accessed on 29 June 2023, with agreement numbers TC25MV1E8N and WX25MUW9GU to A.P. (Arianna Parnigoni).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sung, H.; Ferlay, J.; Siegel, R.L.; Laversanne, M.; Soerjomataram, I.; Jemal, A.; Bray, F. Global Cancer Statistics 2020: GLOBOCAN Estimates of Incidence and Mortality Worldwide for 36 Cancers in 185 Countries. CA Cancer J. Clin. 2021, 71, 209–249. [Google Scholar] [CrossRef]
  2. Polyak, K. Heterogeneity in Breast Cancer. J. Clin. Investig. 2011, 121, 3786–3788. [Google Scholar] [CrossRef] [Green Version]
  3. Viale, G. The Current State of Breast Cancer Classification. Ann. Oncol. 2012, 23, x207–x210. [Google Scholar] [CrossRef]
  4. Tsang, J.Y.S.S.; Tse, G.M. Molecular Classification of Breast Cancer; Lippincott Williams and Wilkins: Philadelphia, PA, USA, 2020; Volume 27, pp. 27–35. [Google Scholar]
  5. Perou, C.M.; Sørlie, T.; Eisen, M.B.; Van De Rijn, M.; Jeffrey, S.S.; Rees, C.A.; Pollack, J.R.; Ross, D.T.; Johnsen, H.; Akslen, L.A.; et al. Molecular Portraits of Human Breast Tumours. Nature 2000, 406, 747–752. [Google Scholar] [CrossRef] [Green Version]
  6. Fougner, C.; Bergholtz, H.; Norum, J.H.; Sørlie, T. Re-Definition of Claudin-Low as a Breast Cancer Phenotype. Nat. Commun. 2020, 11, 1787. [Google Scholar] [CrossRef] [Green Version]
  7. Osborne, C.K.; Schiff, R.; Fuqua, S.A.W.; Shou, J. Estrogen Receptor: Current Understanding of Its Activation and Modulation I. Clin. Cancer Res. 2001, 7, 4338–4342. [Google Scholar]
  8. Williams, C.; Lin, C.Y. Oestrogen Receptors in Breast Cancer: Basic Mechanisms and Clinical Implications. Ecancermedicalscience 2013, 7, 370. [Google Scholar] [CrossRef]
  9. Mandapati, A.; Lukong, K.E. Triple Negative Breast Cancer: Approved Treatment Options and Their Mechanisms of Action. J. Cancer Res. Clin. Oncol. 2022, 149, 3701–3719. [Google Scholar] [CrossRef]
  10. Dent, R.; Trudeau, M.; Pritchard, K.I.; Hanna, W.M.; Kahn, H.K.; Sawka, C.A.; Lickley, L.A.; Rawlinson, E.; Sun, P.; Narod, S.A. Triple-Negative Breast Cancer: Clinical Features and Patterns of Recurrence. Clin. Cancer Res. 2007, 13, 4429–4434. [Google Scholar] [CrossRef] [Green Version]
  11. Liedtke, C.; Mazouni, C.; Hess, K.R.; André, F.; Tordai, A.; Mejia, J.A.; Symmans, W.F.; Gonzalez-Angulo, A.M.; Hennessy, B.; Green, M.; et al. Response to Neoadjuvant Therapy and Long-Term Survival in Patients with Triple-Negative Breast Cancer. J. Clin. Oncol. 2008, 26, 1275–1281. [Google Scholar] [CrossRef]
  12. Lin, N.U.; Vanderplas, A.; Hughes, M.E.; Theriault, R.L.; Edge, S.B.; Wong, Y.N.; Blayney, D.W.; Niland, J.C.; Winer, E.P.; Weeks, J.C. Clinicopathologic Features, Patterns of Recurrence, and Survival among Women with Triple-Negative Breast Cancer in the National Comprehensive Cancer Network. Cancer 2012, 118, 5463–5472. [Google Scholar] [CrossRef] [Green Version]
  13. Carey, L.A.; Dees, E.C.; Sawyer, L.; Gatti, L.; Moore, D.T.; Collichio, F.; Ollila, D.W.; Sartor, C.I.; Graham, M.L.; Perou, C.M. The Triple Negative Paradox: Primary Tumor Chemosensitivity of Breast Cancer Subtypes. Clin. Cancer Res. 2007, 13, 2329–2334. [Google Scholar] [CrossRef] [Green Version]
  14. Bouris, P.; Skandalis, S.S.; Piperigkou, Z.; Afratis, N.; Karamanou, K.; Aletras, A.J.; Moustakas, A.; Theocharis, A.D.; Karamanos, N.K. Estrogen Receptor Alpha Mediates Epithelial to Mesenchymal Transition, Expression of Specific Matrix Effectors and Functional Properties of Breast Cancer Cells. Matrix Biol. 2015, 43, 42–60. [Google Scholar] [CrossRef]
  15. Pareja, F.; Geyer, F.C.; Marchiò, C.; Burke, K.A.; Weigelt, B.; Reis-Filho, J.S. Triple-Negative Breast Cancer: The Importance of Molecular and Histologic Subtyping, and Recognition of Low-Grade Variants. NPJ Breast Cancer 2016, 2, 16036. [Google Scholar] [CrossRef]
  16. Collignon, J.; Lousberg, L.; Schroeder, H.; Jerusalem, G. Triple-Negative Breast Cancer: Treatment Challenges and Solutions. Breast Cancer 2016, 8, 93–107. [Google Scholar] [CrossRef] [Green Version]
  17. Schmid, P.; Adams, S.; Rugo, H.S.; Schneeweiss, A.; Barrios, C.H.; Iwata, H.; Diéras, V.; Hegg, R.; Im, S.-A.; Shaw Wright, G.; et al. Atezolizumab and Nab-Paclitaxel in Advanced Triple-Negative Breast Cancer. N. Engl. J. Med. 2018, 379, 2108–2121. [Google Scholar] [CrossRef]
  18. Moustakas, A.; Heldin, P. TGFβ and Matrix-Regulated Epithelial to Mesenchymal Transition. Biochim. Biophys. Acta 2014, 1840, 2621–2634. [Google Scholar] [CrossRef] [Green Version]
  19. Bissell, M.J.; Rizki, A.; Mian, I.S. Tissue Architecture: The Ultimate Regulator of Breast Epithelial Function. Curr. Opin. Cell Biol. 2003, 15, 753–762. [Google Scholar] [CrossRef] [Green Version]
  20. Bartolini, B.; Caravà, E.; Caon, I.; Parnigoni, A.; Moretto, P.; Passi, A.; Vigetti, D.; Viola, M.; Karousou, E. Heparan Sulfate in the Tumor Microenvironment. Adv. Exp. Med. Biol. 2020, 1245, 147–161. [Google Scholar] [CrossRef]
  21. Karousou, E.; Parnigoni, A.; Moretto, P.; Passi, A.; Viola, M.; Vigetti, D. Hyaluronan in the Cancer Cells Microenvironment. Cancers 2023, 15, 798. [Google Scholar] [CrossRef]
  22. Ye, X.; Brabletz, T.; Kang, Y.; Longmore, G.D.; Nieto, M.A.; Stanger, B.Z.; Yang, J.; Weinberg, R.A. Upholding a Role for EMT in Breast Cancer Metastasis. Nature 2017, 547, E1. [Google Scholar] [CrossRef]
  23. Marozzi, M.; Parnigoni, A.; Negri, A.; Viola, M.; Vigetti, D.; Passi, A.; Karousou, E.; Rizzi, F. Inflammation, Extracellular Matrix Remodeling, and Proteostasis in Tumor Microenvironment. Int. J. Mol. Sci. 2021, 22, 8102. [Google Scholar] [CrossRef]
  24. Winkler, J.; Abisoye-Ogunniyan, A.; Metcalf, K.J.; Werb, Z. Concepts of Extracellular Matrix Remodelling in Tumour Progression and Metastasis. Nat. Commun. 2020, 11, 5120. [Google Scholar] [CrossRef]
  25. Caon, I.; Parnigoni, A.; Viola, M.; Karousou, E.; Passi, A.; Vigetti, D. Cell Energy Metabolism and Hyaluronan Synthesis. J. Histochem. Cytochem. 2021, 69, 35–47. [Google Scholar] [CrossRef]
  26. Viola, M.; Vigetti, D.; Karousou, E.; D’Angelo, M.L.; Caon, I.; Moretto, P.; De Luca, G.; Passi, A. Biology and Biotechnology of Hyaluronan. Glycoconj. J. 2015, 32, 93–103. [Google Scholar] [CrossRef]
  27. Cowman, M.K.; Turley, E.A. Functional Organization of Extracellular Hyaluronan, CD44, and RHAMM. Proteoglycan Res. 2023, 1, e4. [Google Scholar] [CrossRef]
  28. Riecks, J.; Parnigoni, A.; Győrffy, B.; Kiesel, L.; Passi, A.; Vigetti, D.; Götte, M. The Hyaluronan-Related Genes HAS2, HYAL1-4, PH20 and HYALP1 Are Associated with Prognosis, Cell Viability and Spheroid Formation Capacity in Ovarian Cancer. J. Cancer Res. Clin. Oncol. 2022, 148, 3399–3419. [Google Scholar] [CrossRef]
  29. Csoka, A.B.; Frost, G.I.; Stern, R. The Six Hyaluronidase-like Genes in the Human and Mouse Genomes. Matrix Biol. 2001, 20, 499–508. [Google Scholar] [CrossRef]
  30. Itano, N.; Sawai, T.; Yoshida, M.; Lenas, P.; Yamada, Y.; Imagawa, M.; Shinomura, T.; Hamaguchi, M.; Yoshida, Y.; Ohnuki, Y.; et al. Three Isoforms of Mammalian Hyaluronan Synthases Have Distinct Enzymatic Properties. J. Biol. Chem. 1999, 274, 25085–25092. [Google Scholar] [CrossRef] [Green Version]
  31. Slevin, M.; Kumar, S.; Gaffney, J. Angiogenic Oligosaccharides of Hyaluronan Induce Multiple Signaling Pathways Affecting Vascular Endothelial Cell Mitogenic and Wound Healing Responses. J. Biol. Chem. 2002, 277, 41046–41059. [Google Scholar] [CrossRef] [Green Version]
  32. Hauser-Kawaguchi, A.; Luyt, L.G.; Turley, E. Design of Peptide Mimetics to Block Pro-Inflammatory Functions of HA Fragments. Matrix Biol. 2019, 78–79, 346–356. [Google Scholar] [CrossRef]
  33. Pardue, E.L.; Ibrahim, S.; Ramamurthi, A. Role of Hyaluronan in Angiogenesis and Its Utility to Angiogenic Tissue Engineering. Organogenesis 2008, 4, 203–214. [Google Scholar] [CrossRef] [Green Version]
  34. Tavianatou, A.G.; Piperigkou, Z.; Barbera, C.; Beninatto, R.; Masola, V.; Caon, I.; Onisto, M.; Franchi, M.; Galesso, D.; Karamanos, N.K. Molecular Size-Dependent Specificity of Hyaluronan on Functional Properties, Morphology and Matrix Composition of Mammary Cancer Cells. Matrix Biol. Plus 2019, 3, 100008. [Google Scholar] [CrossRef]
  35. Chaaban, H.; Burge, K.; Eckert, J.; Trammell, M.; Dyer, D.; Keshari, R.S.; Silasi, R.; Regmi, G.; Lupu, C.; Good, M.; et al. Acceleration of Small Intestine Development and Remodeling of the Microbiome Following Hyaluronan 35 Kda Treatment in Neonatal Mice. Nutrients 2021, 13, 2030. [Google Scholar] [CrossRef]
  36. Parnigoni, A.; Viola, M.; Karousou, E.; Rovera, S.; Giaroni, C.; Passi, A.; Vigetti, D. Hyaluronan in Pathophysiology of Vascular Diseases: Specific Roles in Smooth Muscle Cells, Endothelial Cells, and Macrophages. Am. J. Physiol.-Cell Physiol. 2022, 323, C505–C519. [Google Scholar]
  37. Auvinen, P.; Tammi, R.; Parkkinen, J.; Tammi, M.; Ågren, U.; Johansson, R.; Hirvikoski, P.; Eskelinen, M.; Kosma, V.M. Hyaluronan in Peritumoral Stroma and Malignant Cells Associates with Breast Cancer Spreading and Predicts Survival. Am. J. Pathol. 2000, 156, 529–536. [Google Scholar] [CrossRef]
  38. Itano, N.; Kimata, K. Altered Hyaluronan Biosynthesis in Cancer Progression. Semin. Cancer Biol. 2008, 18, 268–274. [Google Scholar] [CrossRef]
  39. Itano, N.; Zhuo, L.; Kimata, K. Impact of the Hyaluronan-Rich Tumor Microenvironment on Cancer Initiation and Progression. Cancer Sci. 2008, 99, 1720–1725. [Google Scholar] [CrossRef]
  40. Jokelainen, O.; Pasonen-Seppänen, S.; Tammi, M.; Mannermaa, A.; Aaltomaa, S.; Sironen, R.; Nykopp, T.K. Cellular Hyaluronan Is Associated with a Poor Prognosis in Renal Cell Carcinoma. Urol. Oncol. 2020, 38, 686.e11–686.e22. [Google Scholar] [CrossRef]
  41. Li, J.H.; Wang, Y.C.; Qin, C.D.; Yao, R.R.; Zhang, R.; Wang, Y.; Xie, X.Y.; Zhang, L.; Wang, Y.H.; Ren, Z.G. Over Expression of Hyaluronan Promotes Progression of HCC via CD44-Mediated Pyruvate Kinase M2 Nuclear Translocation. Am. J. Cancer Res. 2016, 6, 509–521. [Google Scholar]
  42. Wang, S.J.; Bourguignon, L.Y.W. Role of Hyaluronan-Mediated CD44 Signaling in Head and Neck Squamous Cell Carcinoma Progression and Chemoresistance. Am. J. Pathol. 2011, 178, 956–963. [Google Scholar] [CrossRef]
  43. Pirinen, R.T.; Tammi, R.H.; Tammi, M.I.; Pääkkö, P.K.; Parkkinen, J.J.; Ågren, U.M.; Johansson, R.T.; Viren, M.M.T.; Törmänen, U.; Soini, Y.M.J.; et al. Expression of Hyaluronan in Normal and Dysplastic Bronchial Epithelium and in Squamous Cell Carcinoma of the Lung. Int. J. Cancer 1998, 79, 251–255. [Google Scholar] [CrossRef]
  44. Hasselbalch, H.; Hovgaard, D.; Nissen, N.; Junker, P. Serum Hyaluronan Is Increased in Malignant Lymphoma. Am. J. Hematol. 1995, 50, 231–233. [Google Scholar] [CrossRef] [PubMed]
  45. Park, J.B.; Kwak, H.J.; Lee, S.H. Role of Hyaluronan in Glioma Invasion. Cell Adh. Migr. 2008, 2, 202–207. [Google Scholar] [CrossRef] [Green Version]
  46. Tahkola, K.; Ahtiainen, M.; Mecklin, J.P.; Kellokumpu, I.; Laukkarinen, J.; Tammi, M.; Tammi, R.; Väyrynen, J.P.; Böhm, J. Stromal Hyaluronan Accumulation Is Associated with Low Immune Response and Poor Prognosis in Pancreatic Cancer. Sci. Rep. 2021, 11, 12216. [Google Scholar] [CrossRef] [PubMed]
  47. Takabe, P.; Siiskonen, H.; Rönkä, A.; Kainulainen, K.; Pasonen-Seppänen, S. The Impact of Hyaluronan on Tumor Progression in Cutaneous Melanoma. Front. Oncol. 2022, 11, 811434. [Google Scholar] [CrossRef]
  48. Auvinen, P.; Tammi, R.; Kosma, V.M.; Sironen, R.; Soini, Y.; Mannermaa, A.; Tumelius, R.; Uljas, E.; Tammi, M. Increased Hyaluronan Content and Stromal Cell CD44 Associate with HER2 Positivity and Poor Prognosis in Human Breast Cancer. Int. J. Cancer 2013, 132, 531–539. [Google Scholar] [CrossRef]
  49. Anttila, M.A.; Tammi, R.H.; Tammi, M.I.; Syrjänen, K.J.; Saarikoski, S.V.; Kosma, V.M. High Levels of Stromal Hyaluronan Predict Poor Disease Outcome in Epithelial Ovarian Cancer. Cancer Res. 2000, 60, 150–155. [Google Scholar]
  50. Tammi, R.H.; Kultti, A.; Kosma, V.M.; Pirinen, R.; Auvinen, P.; Tammi, M.I. Hyaluronan in Human Tumors: Pathobiological and Prognostic Messages from Cell-Associated and Stromal Hyaluronan. Semin. Cancer Biol. 2008, 18, 288–295. [Google Scholar] [CrossRef]
  51. Lauby-Secretan, B.; Scoccianti, C.; Loomis, D.; Grosse, Y.; Bianchini, F.; Straif, K. Body Fatness and Cancer—Viewpoint of the IARC Working Group. N. Engl. J. Med. 2016, 375, 794–798. [Google Scholar] [CrossRef] [Green Version]
  52. Chan, D.S.M.; Vieira, A.R.; Aune, D.; Bandera, E.V.; Greenwood, D.C.; McTiernan, A.; Navarro Rosenblatt, D.; Thune, I.; Vieira, R.; Norat, T. Body Mass Index and Survival in Women with Breast Cancer-Systematic Literature Review and Meta-Analysis of 82 Follow-up Studies. Ann. Oncol. Off. J. Eur. Soc. Med. Oncol. 2014, 25, 1901–1914. [Google Scholar] [CrossRef]
  53. Tsilidis, K.K.; Kasimis, J.C.; Lopez, D.S.; Ntzani, E.E.; Ioannidis, J.P.A. Type 2 Diabetes and Cancer: Umbrella Review of Meta-Analyses of Observational Studies. BMJ 2015, 350, g7607. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Widschwendter, P.; Friedl, T.W.P.; Schwentner, L.; DeGregorio, N.; Jaeger, B.; Schramm, A.; Bekes, I.; Deniz, M.; Lato, K.; Weissenbacher, T.; et al. The Influence of Obesity on Survival in Early, High-Risk Breast Cancer: Results from the Randomized SUCCESS A Trial. Breast Cancer Res. 2015, 17, 129. [Google Scholar] [CrossRef]
  55. Jokela, T.A.; Kuokkanen, J.; Kärnä, R.; Pasonen-Seppänen, S.; Rilla, K.; Kössi, J.; Laato, M.; Tammi, R.H.; Tammi, M.I. Mannose Reduces Hyaluronan and Leukocytes in Wound Granulation Tissue and Inhibits Migration and Hyaluronan-Dependent Monocyte Binding. Wound Repair Regen. 2013, 21, 247–255. [Google Scholar] [CrossRef] [PubMed]
  56. Kuang, D.M.; Wu, Y.; Chen, N.; Cheng, J.; Zhuang, S.M.; Zheng, L. Tumor-Derived Hyaluronan Induces Formation of Immunosuppressive Macrophages through Transient Early Activation of Monocytes. Blood 2007, 110, 587–595. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Olofsson, B.; Porsch, H.; Heldin, P. Knock-down of CD44 Regulates Endothelial Cell Differentiation via NFκB-Mediated Chemokine Production. PLoS ONE 2014, 9, e90921. [Google Scholar] [CrossRef] [Green Version]
  58. Karousou, E.; Misra, S.; Ghatak, S.; Dobra, K.; Götte, M.; Vigetti, D.; Passi, A.; Karamanos, N.K.; Skandalis, S.S. Roles and Targeting of the HAS/Hyaluronan/CD44 Molecular System in Cancer. Matrix Biol. 2017, 59, 3–22. [Google Scholar] [CrossRef]
  59. Spinelli, F.M.; Vitale, D.L.; Sevic, I.; Alaniz, L. Hyaluronan in the Tumor Microenvironment. Adv. Exp. Med. Biol. 2020, 1245, 67–83. [Google Scholar] [CrossRef]
  60. Donelan, W.; Dominguez-Gutierrez, P.R.; Kusmartsev, S. Deregulated Hyaluronan Metabolism in the Tumor Microenvironment Drives Cancer Inflammation and Tumor-Associated Immune Suppression. Front. Immunol. 2022, 13, 971278. [Google Scholar] [CrossRef]
  61. Liu, Y.; Li, L.; Wang, L.; Lu, L.; Li, Y.; Huang, G.; Song, J. “Two-Faces” of Hyaluronan, a Dynamic Barometer of Disease Progression in Tumor Microenvironment. Discov. Oncol. 2023, 14, 11. [Google Scholar] [CrossRef]
  62. Yuan, Z.; Li, Y.; Zhang, S.; Wang, X.; Dou, H.; Yu, X.; Zhang, Z.; Yang, S.; Xiao, M. Extracellular Matrix Remodeling in Tumor Progression and Immune Escape: From Mechanisms to Treatments. Mol. Cancer 2023, 22, 48. [Google Scholar] [CrossRef]
  63. Knudson, W.; Biswas, C.; Toole, B.P. Interactions between Human Tumor Cells and Fibroblasts Stimulate Hyaluronate Synthesis. Proc. Natl. Acad. Sci. USA 1984, 81, 6767–6771. [Google Scholar] [CrossRef]
  64. Asplund, T.; Versnel, M.A.; Laurent, T.C.; Heldin, P. Human Mesothelioma Cells Produce Factors That Stimulate the Production of Hyaluronan by Mesothelial Cells and Fibroblasts. Cancer Res. 1993, 53, 388–392. [Google Scholar] [PubMed]
  65. Hanahan, D. Hallmarks of Cancer: New Dimensions. Cancer Discov. 2022, 12, 31–46. [Google Scholar] [CrossRef] [PubMed]
  66. Caon, I.; Bartolini, B.; Parnigoni, A.; Caravà, E.; Moretto, P.; Viola, M.; Karousou, E.; Vigetti, D.; Passi, A. Revisiting the Hallmarks of Cancer: The Role of Hyaluronan. Semin. Cancer Biol. 2020, 62, 9–19. [Google Scholar] [CrossRef] [PubMed]
  67. Vigetti, D.; Karousou, E.; Viola, M.; Deleonibus, S.; De Luca, G.; Passi, A. Hyaluronan: Biosynthesis and Signaling. Biochim. Biophys. Acta 2014, 1840, 2452–2459. [Google Scholar] [CrossRef] [Green Version]
  68. Maloney, F.P.; Kuklewicz, J.; Corey, R.A.; Bi, Y.; Ho, R.; Mateusiak, L.; Pardon, E.; Steyaert, J.; Stansfeld, P.J.; Zimmer, J. Structure, Substrate Recognition and Initiation of Hyaluronan Synthase. Nature 2022, 604, 195–201. [Google Scholar] [CrossRef]
  69. Weigel, P.H. Hyaluronan Synthase: The Mechanism of Initiation at the Reducing End and a Pendulum Model for Polysaccharide Translocation to the Cell Exterior. Int. J. Cell Biol. 2015, 2015, 367579. [Google Scholar] [CrossRef] [Green Version]
  70. Spicer, A.P.; Seldin, M.F.; Olsen, A.S.; Brown, N.; Wells, D.E.; Doggett, N.A.; Itano, N.; Kimata, K.; Inazawa, J.; McDonald, J.A. Chromosomal Localization of the Human and Mouse Hyaluronan Synthase Genes. Genomics 1997, 41, 493–497. [Google Scholar] [CrossRef]
  71. Vigetti, D.; Viola, M.; Karousou, E.; De Luca, G.; Passi, A. Metabolic Control of Hyaluronan Synthases. Matrix Biol. 2014, 35, 8–13. [Google Scholar] [CrossRef] [Green Version]
  72. Caon, I.; Bartolini, B.; Moretto, P.; Parnigoni, A.; Caravà, E.; Vitale, D.L.; Alaniz, L.; Viola, M.; Karousou, E.; de Luca, G.; et al. Sirtuin 1 Reduces Hyaluronan Synthase 2 Expression by Inhibiting Nuclear Translocation of NF-ΚB and Expression of the Long-Noncoding RNA HAS2–AS1. J. Biol. Chem. 2020, 295, 3485–3496. [Google Scholar] [CrossRef] [Green Version]
  73. Rowland, A.; Miners, J.O.; Mackenzie, P.I. The UDP-Glucuronosyltransferases: Their Role in Drug Metabolism and Detoxification. Int. J. Biochem. Cell Biol. 2013, 45, 1121–1132. [Google Scholar] [CrossRef] [PubMed]
  74. Vitale, D.L.; Caon, I.; Parnigoni, A.; Sevic, I.; Spinelli, F.M.; Icardi, A.; Passi, A.; Vigetti, D.; Alaniz, L. Initial Identification of UDP-Glucose Dehydrogenase as a Prognostic Marker in Breast Cancer Patients, Which Facilitates Epirubicin Resistance and Regulates Hyaluronan Synthesis in MDA-MB-231 Cells. Biomolecules 2021, 11, 246. [Google Scholar] [CrossRef] [PubMed]
  75. Oikari, S.; Kettunen, T.; Tiainen, S.; Häyrinen, J.; Masarwah, A.; Sudah, M.; Sutela, A.; Vanninen, R.; Tammi, M.; Auvinen, P. UDP-Sugar Accumulation Drives Hyaluronan Synthesis in Breast Cancer. Matrix Biol. 2018, 67, 63–74. [Google Scholar] [CrossRef] [Green Version]
  76. Liberti, M.V.; Locasale, J.W. The Warburg Effect: How Does It Benefit Cancer Cells? Trends Biochem. Sci. 2016, 41, 211. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Stern, R.; Shuster, S.; Neudecker, B.A.; Formby, B. Lactate Stimulates Fibroblast Expression of Hyaluronan and CD44: The Warburg Effect Revisited. Exp. Cell Res. 2002, 276, 24–31. [Google Scholar] [CrossRef]
  78. Slomiany, M.G.; Daniel Grass, G.; Robertson, A.D.; Yang, X.Y.; Maria, B.L.; Beeson, C.; Toole, B.P. Hyaluronan, CD44, and Emmprin Regulate Lactate Efflux and Membrane Localization of Monocarboxylate Transporters in Human Breast Carcinoma Cells. Cancer Res. 2009, 69, 1293–1301. [Google Scholar] [CrossRef] [Green Version]
  79. Formby, B.; Stern, R. Lactate-Sensitive Response Elements in Genes Involved in Hyaluronan Catabolism. Biochem. Biophys. Res. Commun. 2003, 305, 203–208. [Google Scholar] [CrossRef]
  80. Chanmee, T.; Ontong, P.; Izumikawa, T.; Higashide, M.; Mochizuki, N.; Chokchaitaweesuk, C.; Khansai, M.; Nakajima, K.; Kakizaki, I.; Kongtawelert, P.; et al. Hyaluronan Production Regulates Metabolic and Cancer Stem-like Properties of Breast Cancer Cells via Hexosamine Biosynthetic Pathway-Coupled HIF-1 Signaling. J. Biol. Chem. 2016, 291, 24105–24120. [Google Scholar] [CrossRef] [Green Version]
  81. Tammi, R.H.; Passi, A.G.; Rilla, K.; Karousou, E.; Vigetti, D.; Makkonen, K.; Tammi, M.I. Transcriptional and Post-Translational Regulation of Hyaluronan Synthesis. FEBS J. 2011, 278, 1419–1428. [Google Scholar] [CrossRef]
  82. Jacobson, A.; Brinck, J.; Briskin, M.J.; Spicer, A.P.; Heldin, P. Expression of Human Hyaluronan Synthases in Response to External Stimuli. Biochem. J. 2000, 348 Pt 1, 29–35. [Google Scholar] [CrossRef]
  83. Heldin, P.; Laurent, T.C.; Heldin, C.H. Effect of Growth Factors on Hyaluronan Synthesis in Cultured Human Fibroblasts. Biochem. J. 1989, 258, 919–922. [Google Scholar] [CrossRef] [Green Version]
  84. Passi, A.; Vigetti, D.; Buraschi, S.; Iozzo, R.V. Dissecting the Role of Hyaluronan Synthases in the Tumor Microenvironment. FEBS J. 2019, 286, 2937–2949. [Google Scholar] [CrossRef] [Green Version]
  85. Camenisch, T.D.; Spicer, A.P.; Brehm-Gibson, T.; Biesterfeldt, J.; Augustine, M.L.; Calabro, A.; Kubalak, S.; Klewer, S.E.; McDonald, J.A. Disruption of Hyaluronan Synthase-2 Abrogates Normal Cardiac Morphogenesis and Hyaluronan-Mediated Transformation of Epithelium to Mesenchyme. J. Clin. Investig. 2000, 106, 349–360. [Google Scholar] [CrossRef] [Green Version]
  86. Itano, N.; Sawai, T.; Atsumi, F.; Miyaishi, O.; Taniguchi, S.; Kannagi, R.; Hamaguchi, M.; Kimata, K. Selective Expression and Functional Characteristics of Three Mammalian Hyaluronan Synthases in Oncogenic Malignant Transformation. J. Biol. Chem. 2004, 279, 18679–18687. [Google Scholar] [CrossRef] [Green Version]
  87. Auvinen, P.; Rilla, K.; Tumelius, R.; Tammi, M.; Sironen, R.; Soini, Y.; Kosma, V.M.; Mannermaa, A.; Viikari, J.; Tammi, R. Hyaluronan Synthases (HAS1-3) in Stromal and Malignant Cells Correlate with Breast Cancer Grade and Predict Patient Survival. Breast Cancer Res. Treat. 2014, 143, 277–286. [Google Scholar] [CrossRef]
  88. Marotta, L.L.C.; Almendro, V.; Marusyk, A.; Shipitsin, M.; Schemme, J.; Walker, S.R.; Bloushtain-Qimron, N.; Kim, J.J.; Choudhury, S.A.; Maruyama, R.; et al. The JAK2/STAT3 Signaling Pathway Is Required for Growth of CD44+CD24 Stem Cell-like Breast Cancer Cells in Human Tumors. J. Clin. Investig. 2011, 121, 2723–2735. [Google Scholar] [CrossRef]
  89. Shipitsin, M.; Campbell, L.L.; Argani, P.; Weremowicz, S.; Bloushtain-Qimron, N.; Yao, J.; Nikolskaya, T.; Serebryiskaya, T.; Beroukhim, R.; Hu, M.; et al. Molecular Definition of Breast Tumor Heterogeneity. Cancer Cell 2007, 11, 259–273. [Google Scholar] [CrossRef] [Green Version]
  90. Parnigoni, A.; Caon, I.; Teo, W.X.; Hua, S.H.; Moretto, P.; Bartolini, B.; Viola, M.; Karousou, E.; Yip, G.W.; Götte, M.; et al. The Natural Antisense Transcript HAS2-AS1 Regulates Breast Cancer Cells Aggressiveness Independently from Hyaluronan Metabolism. Matrix Biol. 2022, 109, 140–161. [Google Scholar] [CrossRef]
  91. Li, Y.; Li, L.; Brown, T.J.; Heldin, P. Silencing of Hyaluronan Synthase 2 Suppresses the Malignant Phenotype of Invasive Breast Cancer Cells. Int. J. Cancer 2007, 120, 2557–2567. [Google Scholar] [CrossRef]
  92. Udabage, L.; Brownlee, G.R.; Waltham, M.; Blick, T.; Walker, E.C.; Heldin, P.; Nilsson, S.K.; Thompson, E.W.; Brown, T.J. Antisense-Mediated Suppression of Hyaluronan Synthase 2 Inhibits the Tumorigenesis and Progression of Breast Cancer. Cancer Res. 2005, 65, 6139–6150. [Google Scholar] [CrossRef] [Green Version]
  93. Bernert, B.; Porsch, H.; Heldin, P. Hyaluronan Synthase 2 (HAS2) Promotes Breast Cancer Cell Invasion by Suppression of Tissue Metalloproteinase Inhibitor 1 (TIMP-1). J. Biol. Chem. 2011, 286, 42349–42359. [Google Scholar] [CrossRef] [Green Version]
  94. Kobayashi, N.; Miyoshi, S.; Mikami, T.; Koyama, H.; Kitazawa, M.; Takeoka, M.; Sano, K.; Amano, J.; Isogai, Z.; Niida, S.; et al. Hyaluronan Deficiency in Tumor Stroma Impairs Macrophage Trafficking and Tumor Neovascularization. Cancer Res. 2010, 70, 7073–7083. [Google Scholar] [CrossRef] [Green Version]
  95. Koyama, H.; Hibi, T.; Isogai, Z.; Yoneda, M.; Fujimori, M.; Amano, J.; Kawakubo, M.; Kannagi, R.; Kimata, K.; Taniguchi, S.; et al. Hyperproduction of Hyaluronan in Neu-Induced Mammary Tumor Accelerates Angiogenesis through Stromal Cell Recruitment: Possible Involvement of Versican/PG-M. Am. J. Pathol. 2007, 170, 1086–1099. [Google Scholar] [CrossRef] [Green Version]
  96. Heldin, P.; Basu, K.; Olofsson, B.; Porsch, H.; Kozlova, I.; Kahata, K. Deregulation of Hyaluronan Synthesis, Degradation and Binding Promotes Breast Cancer. J. Biochem. 2013, 154, 395–408. [Google Scholar] [CrossRef] [Green Version]
  97. Porsch, H.; Bernert, B.; Mehić, M.; Theocharis, A.D.; Heldin, C.H.; Heldin, P. Efficient TGFβ-Induced Epithelial-Mesenchymal Transition Depends on Hyaluronan Synthase HAS2. Oncogene 2013, 32, 4355–4365. [Google Scholar] [CrossRef]
  98. Okuda, H.; Kobayashi, A.; Xia, B.; Watabe, M.; Pai, S.K.; Hirota, S.; Xing, F.; Liu, W.; Pandey, P.R.; Fukuda, K.; et al. Hyaluronan Synthase HAS2 Promotes Tumor Progression in Bone by Stimulating the Interaction of Breast Cancer Stem-like Cells with Macrophages and Stromal Cells. Cancer Res. 2012, 72, 537–547. [Google Scholar] [CrossRef] [Green Version]
  99. Sheng, Y.; Cao, M.; Liu, Y.; He, Y.; Zhang, G.; Du, Y.; Gao, F.; Yang, C. Hyaluronan Synthase 2 (HAS2) Regulates Cell Phenotype and Invadopodia Formation in Luminal-like Breast Cancer Cells. Mol. Cell. Biochem. 2021, 476, 3383–3391. [Google Scholar] [CrossRef]
  100. Urakawa, H.; Nishida, Y.; Wasa, J.; Arai, E.; Zhuo, L.; Kimata, K.; Kozawa, E.; Futamura, N.; Ishiguro, N. Inhibition of Hyaluronan Synthesis in Breast Cancer Cells by 4-Methylumbelliferone Suppresses Tumorigenicity in Vitro and Metastatic Lesions of Bone in Vivo. Int. J. Cancer 2012, 130, 454–466. [Google Scholar] [CrossRef]
  101. Karalis, T.T.; Heldin, P.; Vynios, D.H.; Neill, T.; Buraschi, S.; Iozzo, R.V.; Karamanos, N.K.; Skandalis, S.S. Tumor-Suppressive Functions of 4-MU on Breast Cancer Cells of Different ER Status: Regulation of Hyaluronan/HAS2/CD44 and Specific Matrix Effectors. Matrix Biol. 2019, 78–79, 118–138. [Google Scholar] [CrossRef]
  102. Kakizaki, I.; Kojima, K.; Takagaki, K.; Endo, M.; Kannagi, R.; Ito, M.; Maruo, Y.; Sato, H.; Yasuda, T.; Mita, S.; et al. A Novel Mechanism for the Inhibition of Hyaluronan Biosynthesis by 4-Methylumbelliferone. J. Biol. Chem. 2004, 279, 33281–33289. [Google Scholar] [CrossRef] [Green Version]
  103. Nagy, N.; Kuipers, H.F.; Frymoyer, A.R.; Ishak, H.D.; Bollyky, J.B.; Wight, T.N.; Bollyky, P.L. 4-Methylumbelliferone Treatment and Hyaluronan Inhibition as a Therapeutic Strategy in Inflammation, Autoimmunity, and Cancer. Front. Immunol. 2015, 6, 123. [Google Scholar] [CrossRef] [Green Version]
  104. Vigetti, D.; Rizzi, M.; Viola, M.; Karousou, E.; Genasetti, A.; Clerici, M.; Bartolini, B.; Hascall, V.C.; De Luca, G.; Passi, A. The Effects of 4-Methylumbelliferone on Hyaluronan Synthesis, MMP2 Activity, Proliferation, and Motility of Human Aortic Smooth Muscle Cells. Glycobiology 2009, 19, 537–546. [Google Scholar] [CrossRef] [Green Version]
  105. Pibuel, M.A.; Poodts, D.; Díaz, M.; Molinari, Y.A.; Franco, P.G.; Hajos, S.E.; Lompardía, S.L. Antitumor Effect of 4MU on Glioblastoma Cells Is Mediated by Senescence Induction and CD44, RHAMM and p-ERK Modulation. Cell Death Discov. 2021, 7, 280. [Google Scholar] [CrossRef]
  106. Rodríguez, M.M.; Onorato, A.; Cantero, M.J.; Domínguez, L.; Bayo, J.; Fiore, E.; García, M.; Atorrasagasti, C.; Canbay, A.; Malvicini, M.; et al. 4-Methylumbelliferone-Mediated Polarization of M1 Macrophages Correlate with Decreased Hepatocellular Carcinoma Aggressiveness in Mice. Sci. Rep. 2021, 11, 6310. [Google Scholar] [CrossRef]
  107. Lokman, N.A.; Price, Z.K.; Hawkins, E.K.; Macpherson, A.M.; Oehler, M.K.; Ricciardelli, C. 4-Methylumbelliferone Inhibits Cancer Stem Cell Activation and Overcomes Chemoresistance in Ovarian Cancer. Cancers 2019, 11, 1187. [Google Scholar] [CrossRef] [Green Version]
  108. Tamura, R.; Yokoyama, Y.; Yoshida, H.; Imaizumi, T.; Mizunuma, H. 4-Methylumbelliferone Inhibits Ovarian Cancer Growth by Suppressing Thymidine Phosphorylase Expression. J. Ovarian Res. 2014, 7, 94. [Google Scholar] [CrossRef]
  109. Nagy, N.; Freudenberger, T.; Melchior-Becker, A.; Röck, K.; Ter Braak, M.; Jastrow, H.; Kinzig, M.; Lucke, S.; Suvorava, T.; Kojda, G.; et al. Inhibition of Hyaluronan Synthesis Accelerates Murine Atherosclerosis: Novel Insights into the Role of Hyaluronan Synthesis. Circulation 2010, 122, 2313–2322. [Google Scholar] [CrossRef] [Green Version]
  110. Mikami, K.; Endo, T.; Sawada, N.; Igarashi, G.; Kimura, M.; Sakuraba, H.; Fukuda, S. Inhibition of Systemic Hyaluronan Synthesis Exacerbates Murine Hepatic Carcinogenesis. Vivo 2018, 32, 273–278. [Google Scholar] [CrossRef] [Green Version]
  111. Nagy, N.; Gurevich, I.; Kuipers, H.F.; Ruppert, S.M.; Marshall, P.L.; Xie, B.J.; Sun, W.; Malkovskiy, A.V.; Rajadas, J.; Grandoch, M.; et al. 4-Methylumbelliferyl Glucuronide Contributes to Hyaluronan Synthesis Inhibition. J. Biol. Chem. 2019, 294, 7864–7877. [Google Scholar] [CrossRef]
  112. Vigetti, D.; Clerici, M.; Deleonibus, S.; Karousou, E.; Viola, M.; Moretto, P.; Heldin, P.; Hascall, V.C.; De Luca, G.; Passi, A. Hyaluronan Synthesis Is Inhibited by Adenosine Monophosphate-Activated Protein Kinase through the Regulation of HAS2 Activity in Human Aortic Smooth Muscle Cells. J. Biol. Chem. 2011, 286, 7917–7924. [Google Scholar] [CrossRef] [Green Version]
  113. Zhao, H.; Orhan, Y.C.; Zha, X.; Esencan, E.; Chatterton, R.T.; Bulun, S.E. AMP-Activated Protein Kinase and Energy Balance in Breast Cancer. Am. J. Transl. Res. 2017, 9, 197–213. [Google Scholar]
  114. Hampsch, R.A.; Wells, J.D.; Traphagen, N.A.; McCleery, C.F.; Fields, J.L.; Shee, K.; Dillon, L.M.; Pooler, D.B.; Lewis, L.D.; Demidenko, E.; et al. AMPK Activation by Metformin Promotes Survival of Dormant ER+ Breast Cancer Cells. Clin. Cancer Res. 2020, 26, 3707–3719. [Google Scholar] [CrossRef]
  115. Vara-Ciruelos, D.; Russell, F.M.; Grahame Hardie, D. The Strange Case of AMPK and Cancer: Dr Jekyll or Mr Hyde? Open Biol. 2019, 9, 190099. [Google Scholar] [CrossRef] [Green Version]
  116. Wei, C.; Zou, H.; Xiao, T.; Liu, X.; Wang, Q.; Cheng, J.; Fu, S.; Peng, J.; Xie, X.; Fu, J. TQFL12, a Novel Synthetic Derivative of TQ, Inhibits Triple-Negative Breast Cancer Metastasis and Invasion through Activating AMPK/ACC Pathway. J. Cell. Mol. Med. 2021, 25, 10101–10110. [Google Scholar] [CrossRef]
  117. Seto-Tetsuo, F.; Arioka, M.; Miura, K.; Inoue, T.; Igawa, K.; Tomooka, K.; Takahashi-Yanaga, F.; Sasaguri, T. DIF-1 Inhibits Growth and Metastasis of Triple-Negative Breast Cancer through AMPK-Mediated Inhibition of the MTORC1-S6K Signaling Pathway. Oncogene 2021, 40, 5579–5589. [Google Scholar] [CrossRef]
  118. Alessi, D.R.; Sakamoto, K.; Bayascas, J.R. LKB1-Dependent Signaling Pathways. Annu. Rev. Biochem. 2006, 75, 137–163. [Google Scholar] [CrossRef]
  119. Karalis, T.T.; Chatzopoulos, A.; Kondyli, A.; Aletras, A.J.; Karamanos, N.K.; Heldin, P.; Skandalis, S.S. Salicylate Suppresses the Oncogenic Hyaluronan Network in Metastatic Breast Cancer Cells. Matrix Biol. Plus 2020, 6–7, 100031. [Google Scholar] [CrossRef]
  120. Suzuki, M.; Asplund, T.; Yamashita, H.; Heldin, C.H.; Heldin, P. Stimulation of Hyaluronan Biosynthesis by Platelet-Derived Growth Factor-BB and Transforming Growth Factor-Β1 Involves Activation of Protein Kinase C. Biochem. J. 1995, 307, 817–821. [Google Scholar] [CrossRef]
  121. Karousou, E.; Kamiryo, M.; Skandalis, S.S.; Ruusala, A.; Asteriou, T.; Passi, A.; Yamashita, H.; Hellman, U.; Heldin, C.H.; Heldin, P. The Activity of Hyaluronan Synthase 2 Is Regulated by Dimerization and Ubiquitination. J. Biol. Chem. 2010, 285, 23647–23654. [Google Scholar] [CrossRef] [Green Version]
  122. Vigetti, D.; Deleonibus, S.; Moretto, P.; Karousou, E.; Viola, M.; Bartolini, B.; Hascall, V.C.; Tammi, M.; De Luca, G.; Passi, A. Role of UDP-N-Acetylglucosamine (GlcNAc) and O-GlcNAcylation of Hyaluronan Synthase 2 in the Control of Chondroitin Sulfate and Hyaluronan Synthesis. J. Biol. Chem. 2012, 287, 35544–35555. [Google Scholar] [CrossRef] [Green Version]
  123. Michael, D.R.; Phillips, A.O.; Krupa, A.; Martin, J.; Redman, J.E.; Altaher, A.; Neville, R.D.; Webber, J.; Kim, M.Y.; Bowen, T. The Human Hyaluronan Synthase 2 (HAS2) Gene and Its Natural Antisense RNA Exhibit Coordinated Expression in the Renal Proximal Tubular Epithelial Cell. J. Biol. Chem. 2011, 286, 19523–19532. [Google Scholar] [CrossRef] [Green Version]
  124. Parnigoni, A.; Caon, I.; Moretto, P.; Viola, M.; Karousou, E.; Passi, A.; Vigetti, D. The Role of the Multifaceted Long Non-Coding RNAs: A Nuclear-Cytosolic Interplay to Regulate Hyaluronan Metabolism. Matrix Biol. Plus 2021, 11, 100060. [Google Scholar] [CrossRef]
  125. Chao, H.; Spicer, A.P. Natural Antisense MRNAs to Hyaluronan Synthase 2 Inhibit Hyaluronan Biosynthesis and Cell Proliferation. J. Biol. Chem. 2005, 280, 27513–27522. [Google Scholar] [CrossRef] [Green Version]
  126. Zhu, G.; Wang, S.; Chen, J.; Wang, Z.; Liang, X.; Wang, X.; Jiang, J.; Lang, J.; Li, L. Long Noncoding RNA HAS2-AS1 Mediates Hypoxia-Induced Invasiveness of Oral Squamous Cell Carcinoma. Mol. Carcinog. 2017, 56, 2210–2222. [Google Scholar] [CrossRef]
  127. Kolliopoulos, C.; Lin, C.Y.; Heldin, C.H.; Moustakas, A.; Heldin, P. Has2 Natural Antisense RNA and Hmga2 Promote Has2 Expression during TGFβ-Induced EMT in Breast Cancer. Matrix Biol. 2019, 80, 29–45. [Google Scholar] [CrossRef]
  128. Vigetti, D.; Deleonibus, S.; Moretto, P.; Bowen, T.; Fischer, J.W.; Grandoch, M.; Oberhuber, A.; Love, D.C.; Hanover, J.A.; Cinquetti, R.; et al. Natural Antisense Transcript for Hyaluronan Synthase 2 (HAS2-AS1) Induces Transcription of HAS2 via Protein O-GlcNAcylation. J. Biol. Chem. 2014, 289, 28816–28826. [Google Scholar] [CrossRef] [Green Version]
  129. Zhang, L.; Wang, H.; Xu, M.; Chen, F.; Li, W.; Hu, H.; Yuan, Q.; Su, Y.; Liu, X.; Wuri, J.; et al. Long Noncoding RNA HAS2-AS1 Promotes Tumor Progression in Glioblastoma via Functioning as a Competing Endogenous RNA. J. Cell. Biochem. 2020, 121, 661–671. [Google Scholar] [CrossRef]
  130. Tong, L.; Wang, Y.; Ao, Y.; Sun, X. CREB1 Induced LncRNA HAS2-AS1 Promotes Epithelial Ovarian Cancer Proliferation and Invasion via the MiR-466/RUNX2 Axis. Biomed. Pharmacother. 2019, 115, 108891. [Google Scholar] [CrossRef]
  131. Hynes, R.O.; Naba, A. Overview of the Matrisome—An Inventory of Extracellular Matrix Constituents and Functions. Cold Spring Harb. Perspect. Biol. 2012, 4, a004903. [Google Scholar] [CrossRef] [Green Version]
  132. Caon, I.; D’Angelo, M.L.; Bartolini, B.; Caravà, E.; Parnigoni, A.; Contino, F.; Cancemi, P.; Moretto, P.; Karamanos, N.K.; Passi, A.; et al. The Secreted Protein C10orf118 Is a New Regulator of Hyaluronan Synthesis Involved in Tumour-Stroma Cross-Talk. Cancers 2021, 13, 1105. [Google Scholar] [CrossRef]
  133. Chanmee, T.; Ontong, P.; Mochizuki, N.; Kongtawelert, P.; Konno, K.; Itano, N. Excessive Hyaluronan Production Promotes Acquisition of Cancer Stem Cell Signatures through the Coordinated Regulation of Twist and the Transforming Growth Factor β (TGF-β)-Snail Signaling Axis. J. Biol. Chem. 2014, 289, 26038–26056. [Google Scholar] [CrossRef] [Green Version]
  134. Yang, C.; Liu, Y.; He, Y.; Du, Y.; Wang, W.; Shi, X.; Gao, F. The Use of HA Oligosaccharide-Loaded Nanoparticles to Breach the Endogenous Hyaluronan Glycocalyx for Breast Cancer Therapy. Biomaterials 2013, 34, 6829–6838. [Google Scholar] [CrossRef]
  135. Yang, C.; Sheng, Y.; Shi, X.; Liu, Y.; He, Y.; Du, Y.; Zhang, G.; Gao, F. CD44/HA Signaling Mediates Acquired Resistance to a PI3Kα Inhibitor. Cell Death Dis. 2020, 11, 831. [Google Scholar] [CrossRef]
  136. Sun, X.; Tang, F.; Guo, Q.; Liu, Y.; He, Y.; Du, Y.; Gao, F.; Zhang, G.; Yang, C. HAS2-Ezrin-ER Axis Plays a Role in Acquired Antiestrogen Resistance of ER-Positive Breast Cancer. Front. Pharmacol. 2022, 13, 1031487. [Google Scholar] [CrossRef]
  137. Tavianatou, A.G.; Caon, I.; Franchi, M.; Piperigkou, Z.; Galesso, D.; Karamanos, N.K. Hyaluronan: Molecular Size-Dependent Signaling and Biological Functions in Inflammation and Cancer. FEBS J. 2019, 286, 2883–2908. [Google Scholar] [CrossRef] [Green Version]
  138. Bohaumilitzky, L.; Huber, A.K.; Stork, E.M.; Wengert, S.; Woelfl, F.; Boehm, H. A Trickster in Disguise: Hyaluronan’s Ambivalent Roles in the Matrix. Front. Oncol. 2017, 7, 269643. [Google Scholar] [CrossRef] [Green Version]
  139. Maciej-Hulme, M.L. New Insights Into Human Hyaluronidase 4/Chondroitin Sulphate Hydrolase. Front. Cell Dev. Biol. 2021, 9, 767924. [Google Scholar] [CrossRef]
  140. Yamamoto, H.; Tobisawa, Y.; Inubushi, T.; Irie, F.; Ohyama, C.; Yamaguchi, Y. A Mammalian Homolog of the Zebrafish Transmembrane Protein 2 (TMEM2) Is the Long-Sought-after Cell-Surface Hyaluronidase. J. Biol. Chem. 2017, 292, 7304–7313. [Google Scholar] [CrossRef] [Green Version]
  141. Yoshida, H.; Nagaoka, A.; Kusaka-Kikushima, A.; Tobiishi, M.; Kawabata, K.; Sayo, T.; Sakai, S.; Sugiyama, Y.; Enomoto, H.; Okada, Y.; et al. KIAA1199, a Deafness Gene of Unknown Function, Is a New Hyaluronan Binding Protein Involved in Hyaluronan Depolymerization. Proc. Natl. Acad. Sci. USA 2013, 110, 5612–5617. [Google Scholar]
  142. Lee, H.; Goodarzi, H.; Tavazoie, S.F.; Alarcón, C.R. TMEM2 Is a SOX4-Regulated Gene That Mediates Metastatic Migration and Invasion in Breast Cancer. Cancer Res. 2016, 76, 4994–5005. [Google Scholar] [CrossRef] [Green Version]
  143. Spataro, S.; Guerra, C.; Cavalli, A.; Sgrignani, J.; Sleeman, J.; Poulain, L.; Boland, A.; Scapozza, L.; Moll, S.; Prunotto, M. CEMIP (HYBID, KIAA1199): Structure, Function and Expression in Health and Disease. FEBS J. 2022. [Google Scholar] [CrossRef]
  144. Dong, X.; Yang, Y.; Yuan, Q.; Hou, J.; Wu, G. High Expression of CEMIP Correlates Poor Prognosis and the Tumur Microenvironment in Breast Cancer as a Promisingly Prognostic Biomarker. Front. Genet. 2021, 12, 768140. [Google Scholar] [CrossRef]
  145. Banach, A.; Jiang, Y.P.; Roth, E.; Kuscu, C.; Cao, J.; Lin, R.Z. CEMIP Upregulates BiP to Promote Breast Cancer Cell Survival in Hypoxia. Oncotarget 2019, 10, 4307–4320. [Google Scholar] [CrossRef] [Green Version]
  146. Stern, R.; Jedrzejas, M.J. Hyaluronidases: Their Genomics, Structures, and Mechanisms of Action. Chem. Rev. 2006, 106, 818–839. [Google Scholar] [CrossRef] [Green Version]
  147. Tolg, C.; Yuan, H.; Flynn, S.M.; Basu, K.; Ma, J.; Tse, K.C.K.; Kowalska, B.; Vulkanesku, D.; Cowman, M.K.; McCarthy, J.B.; et al. Hyaluronan Modulates Growth Factor Induced Mammary Gland Branching in a Size Dependent Manner. Matrix Biol. 2017, 63, 117–132. [Google Scholar] [CrossRef]
  148. Motolese, A.; Vignati, F.; Brambilla, R.; Cerati, M.; Passi, A. Interaction between a Regenerative Matrix and Wound Bed in Nonhealing Ulcers: Results with 16 Cases. BioMed Res. Int. 2013, 2013, 849321. [Google Scholar] [CrossRef]
  149. Tian, X.; Azpurua, J.; Hine, C.; Vaidya, A.; Myakishev-Rempel, M.; Ablaeva, J.; Mao, Z.; Nevo, E.; Gorbunova, V.; Seluanov, A. High-Molecular-Mass Hyaluronan Mediates the Cancer Resistance of the Naked Mole Rat. Nature 2013, 499, 346–349. [Google Scholar] [CrossRef] [Green Version]
  150. del Marmol, D.; Holtze, S.; Kichler, N.; Sahm, A.; Bihin, B.; Bourguignon, V.; Dogné, S.; Szafranski, K.; Hildebrandt, T.B.; Flamion, B. Abundance and Size of Hyaluronan in Naked Mole-Rat Tissues and Plasma. Sci. Rep. 2021, 11, 7951. [Google Scholar] [CrossRef]
  151. Wu, M.; Cao, M.; He, Y.; Liu, Y.; Yang, C.; Du, Y.; Wang, W.; Gao, F. A Novel Role of Low Molecular Weight Hyaluronan in Breast Cancer Metastasis. FASEB J. 2015, 29, 1290–1298. [Google Scholar] [CrossRef]
  152. Boroughs, L.K.; Deberardinis, R.J. Metabolic Pathways Promoting Cancer Cell Survival and Growth. Nat. Cell Biol. 2015, 17, 351–359. [Google Scholar] [CrossRef] [Green Version]
  153. Sullivan, W.J.; Mullen, P.J.; Schmid, E.W.; Flores, A.; Momcilovic, M.; Sharpley, M.S.; Jelinek, D.; Whiteley, A.E.; Maxwell, M.B.; Wilde, B.R.; et al. Extracellular Matrix Remodeling Regulates Glucose Metabolism through TXNIP Destabilization. Cell 2018, 175, 117–132.e21. [Google Scholar] [CrossRef] [Green Version]
  154. Tan, J.X.; Wang, X.Y.; Li, H.Y.; Su, X.L.; Wang, L.; Ran, L.; Zheng, K.; Ren, G.S. HYAL1 Overexpression Is Correlated with the Malignant Behavior of Human Breast Cancer. Int. J. Cancer 2011, 128, 1303–1315. [Google Scholar] [CrossRef]
  155. Udabage, L.; Brownlee, G.R.; Nilsson, S.K.; Brown, T.J. The Over-Expression of HAS2, Hyal-2 and CD44 Is Implicated in the Invasiveness of Breast Cancer. Exp. Cell Res. 2005, 310, 205–217. [Google Scholar] [CrossRef]
  156. Berdiaki, A.; Neagu, M.; Spyridaki, I.; Kuskov, A.; Perez, S.; Nikitovic, D. Hyaluronan and Reactive Oxygen Species Signaling-Novel Cues from the Matrix? Antioxidants 2023, 12, 824. [Google Scholar] [CrossRef]
  157. Cowman, M.K. Hyaluronan and Hyaluronan Fragments. Adv. Carbohydr. Chem. Biochem. 2017, 74, 1–59. [Google Scholar] [CrossRef]
  158. Dogné, S.; Flamion, B. Endothelial Glycocalyx Impairment in Disease: Focus on Hyaluronan Shedding. Am. J. Pathol. 2020, 190, 768–780. [Google Scholar] [CrossRef]
  159. Karihtala, P.; Soini, Y.; Auvinen, P.; Tammi, R.; Tammi, M.; Kosma, V.M. Hyaluronan in Breast Cancer: Correlations with Nitric Oxide Synthases and Tyrosine Nitrosylation. J. Histochem. Cytochem. 2007, 55, 1191–1198. [Google Scholar] [CrossRef]
  160. Jiang, D.; Liang, J.; Noble, P.W. Hyaluronan as an Immune Regulator in Human Diseases. Physiol. Rev. 2011, 91, 221–264. [Google Scholar] [CrossRef] [Green Version]
  161. Morath, I.; Hartmann, T.N.; Orian-Rousseau, V. CD44: More than a Mere Stem Cell Marker. Int. J. Biochem. Cell Biol. 2016, 81, 166–173. [Google Scholar] [CrossRef]
  162. Misra, S.; Hascall, V.C.; Markwald, R.R.; Ghatak, S. Interactions between Hyaluronan and Its Receptors (CD44, RHAMM) Regulate the Activities of Inflammation and Cancer. Front. Immunol. 2015, 6, 137534. [Google Scholar] [CrossRef] [Green Version]
  163. Luo, Z.; Wu, R.R.; Lv, L.; Li, P.; Zhang, L.Y.; Hao, Q.L.; Li, W. Prognostic Value of CD44 Expression in Non-Small Cell Lung Cancer: A Systematic Review. Int. J. Clin. Exp. Pathol. 2014, 7, 3632. [Google Scholar] [PubMed]
  164. Todaro, M.; Gaggianesi, M.; Catalano, V.; Benfante, A.; Iovino, F.; Biffoni, M.; Apuzzo, T.; Sperduti, I.; Volpe, S.; Cocorullo, G.; et al. CD44v6 Is a Marker of Constitutive and Reprogrammed Cancer Stem Cells Driving Colon Cancer Metastasis. Cell Stem Cell 2014, 14, 342–356. [Google Scholar] [CrossRef] [Green Version]
  165. Di Franco, S.; Turdo, A.; Benfante, A.; Colorito, M.L.; Gaggianesi, M.; Apuzzo, T.; Kandimalla, R.; Chinnici, A.; Barcaroli, D.; Mangiapane, L.R.; et al. ΔNp63 Drives Metastasis in Breast Cancer Cells via PI3K/CD44v6 Axis. Oncotarget 2016, 7, 54157–54173. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Cieply, B.; Koontz, C.; Frisch, S.M. CD44S-Hyaluronan Interactions Protect Cells Resulting from EMT against Anoikis. Matrix Biol. 2015, 48, 55–65. [Google Scholar] [CrossRef] [Green Version]
  167. Chen, C.; Zhao, S.; Karnad, A.; Freeman, J.W. The Biology and Role of CD44 in Cancer Progression: Therapeutic Implications. J. Hematol. Oncol. 2018, 11, 64. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Olsson, E.; Honeth, G.; Bendahl, P.O.; Saal, L.H.; Gruvberger-Saal, S.; Ringnér, M.; Vallon-Christersson, J.; Jönsson, G.; Holm, K.; Lövgren, K.; et al. CD44 Isoforms Are Heterogeneously Expressed in Breast Cancer and Correlate with Tumor Subtypes and Cancer Stem Cell Markers. BMC Cancer 2011, 11, 418. [Google Scholar] [CrossRef] [Green Version]
  169. Ibrahim, S.A.; Gadalla, R.; El-Ghonaimy, E.A.; Samir, O.; Mohamed, H.T.; Hassan, H.; Greve, B.; El-Shinawi, M.; Mohamed, M.M.; Götte, M. Syndecan-1 Is a Novel Molecular Marker for Triple Negative Inflammatory Breast Cancer and Modulates the Cancer Stem Cell Phenotype via the IL-6/STAT3, Notch and EGFR Signaling Pathways. Mol. Cancer 2017, 16, 57. [Google Scholar] [CrossRef] [Green Version]
  170. Wang, Z.; Zhao, K.; Hackert, T.; Zöller, M. CD44/CD44v6 a Reliable Companion in Cancer-Initiating Cell Maintenance and Tumor Progression. Front. Cell Dev. Biol. 2018, 6, 397013. [Google Scholar] [CrossRef] [Green Version]
  171. Schmitt, M.; Metzger, M.; Gradl, D.; Davidson, G.; Orian-Rousseau, V. CD44 Functions in Wnt Signaling by Regulating LRP6 Localization and Activation. Cell Death Differ. 2015, 22, 677–689. [Google Scholar] [CrossRef] [Green Version]
  172. Yu, Q.; Stamenkovic, I. Cell Surface-Localized Matrix Metalloproteinase-9 Proteolytically Activates TGF-β and Promotes Tumor Invasion and Angiogenesis. Genes Dev. 2000, 14, 163–176. [Google Scholar] [CrossRef]
  173. Gao, R.; Li, D.; Xun, J.; Zhou, W.; Li, J.; Wang, J.; Liu, C.; Li, X.; Shen, W.; Qiao, H.; et al. CD44ICD Promotes Breast Cancer Stemness via PFKFB4-Mediated Glucose Metabolism. Theranostics 2018, 8, 6248–6262. [Google Scholar] [CrossRef] [PubMed]
  174. Kajita, M.; Itoh, Y.; Chiba, T.; Mori, H.; Okada, A.; Kinoh, H.; Seiki, M. Membrane-Type 1 Matrix Metalloproteinase Cleaves Cd44 and Promotes Cell Migration. J. Cell Biol. 2001, 153, 893–904. [Google Scholar] [CrossRef] [PubMed]
  175. Cho, Y.; Lee, H.-W.; Kang, H.-G.; Kim, H.-Y.; Kim, S.-J.; Chun, K.-H.; Cho, Y.; Lee, H.-W.; Kang, H.-G.; Kim, H.-Y.; et al. Cleaved CD44 Intracellular Domain Supports Activation of Stemness Factors and Promotes Tumorigenesis of Breast Cancer. Oncotarget 2015, 6, 8709–8721. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Vanneste, M.; Hanoux, V.; Bouakka, M.; Bonnamy, P.J. Hyaluronate Synthase-2 Overexpression Alters Estrogen Dependence and Induces Histone Deacetylase Inhibitor-like Effects on ER-Driven Genes in MCF7 Breast Tumor Cells. Mol. Cell. Endocrinol. 2017, 444, 48–58. [Google Scholar] [CrossRef]
  177. Guttilla, I.K.; Adams, B.D.; White, B.A. ERα, MicroRNAs, and the Epithelial–Mesenchymal Transition in Breast Cancer. Trends Endocrinol. Metab. 2012, 23, 73–82. [Google Scholar] [CrossRef]
  178. Korkes, F.; De Castro, M.G.; De Cassio Zequi, S.; Nardi, L.; Del Giglio, A.; De Lima Pompeo, A.C. Hyaluronan-Mediated Motility Receptor (RHAMM) Immunohistochemical Expression and Androgen Deprivation in Normal Peritumoral, Hyperplasic and Neoplastic Prostate Tissue. BJU Int. 2014, 113, 822–829. [Google Scholar] [CrossRef] [Green Version]
  179. Zhang, H.; Ren, L.; Ding, Y.; Li, F.; Chen, X.; Ouyang, Y.; Zhang, Y.; Zhang, D.; Zhang, H.; Ren, L.; et al. Hyaluronan-Mediated Motility Receptor Confers Resistance to Chemotherapy via TGFβ/Smad2-Induced Epithelial-Mesenchymal Transition in Gastric Cancer. FASEB J. 2019, 33, 6365–6377. [Google Scholar] [CrossRef]
  180. Mele, V.; Sokol, L.; Kölzer, V.H.; Pfaff, D.; Muraro, M.G.; Keller, I.; Stefan, Z.; Centeno, I.; Terracciano, L.M.; Dawson, H.; et al. The Hyaluronan-Mediated Motility Receptor RHAMM Promotes Growth, Invasiveness and Dissemination of Colorectal Cancer. Oncotarget 2017, 8, 70617–70629. [Google Scholar] [CrossRef] [Green Version]
  181. Rein, D.T.; Roehrig, K.; Schöndorf, T.; Lazar, A.; Fleisch, M.; Niederacher, D.; Bender, H.G.; Dall, P. Expression of the Hyaluronan Receptor RHAMM in Endometrial Carcinomas Suggests a Role in Tumour Progression and Metastasis. J. Cancer Res. Clin. Oncol. 2003, 129, 161–164. [Google Scholar] [CrossRef]
  182. Kouvidi, K.; Berdiaki, A.; Nikitovic, D.; Katonis, P.; Afratis, N.; Hascall, V.C.; Karamanos, N.K.; Tzanakakis, G.N. Role of Receptor for Hyaluronic Acid-Mediated Motility (RHAMM) in Low Molecular Weight Hyaluronan (LMWHA)- Mediated Fibrosarcoma Cell Adhesion. J. Biol. Chem. 2011, 286, 38509–38520. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Vigetti, D.; Götte, M.; Pavão, M.S.G.; Theocharis, A.D. Cellular Microenvironment in Human Pathologies. BioMed Res. Int. 2013, 2013, 946958. [Google Scholar] [CrossRef] [PubMed]
  184. Hally, K.; Fauteux-Daniel, S.; Hamzeh-Cognasse, H.; Larsen, P.; Cognasse, F. Revisiting Platelets and Toll-Like Receptors (TLRs): At the Interface of Vascular Immunity and Thrombosis. Int. J. Mol. Sci. 2020, 21, 6150. [Google Scholar] [CrossRef]
  185. Di Lorenzo, A.; Bolli, E.; Tarone, L.; Cavallo, F.; Conti, L. Toll-Like Receptor 2 at the Crossroad between Cancer Cells, the Immune System, and the Microbiota. Int. J. Mol. Sci. 2020, 21, 9418. [Google Scholar] [CrossRef]
  186. Zheng, W.; Xu, Q.; Zhang, Y.; Xiaofei, E.; Gao, W.; Zhang, M.; Zhai, W.; Rajkumar, R.S.; Liu, Z. Toll-like Receptor-Mediated Innate Immunity against Herpesviridae Infection: A Current Perspective on Viral Infection Signaling Pathways. Virol. J. 2020, 17, 192. [Google Scholar] [CrossRef] [PubMed]
  187. Makkar, S.; Riehl, T.E.; Chen, B.; Yan, Y.; Alvarado, D.M.; Ciorba, M.A.; Stenson, W.F. Hyaluronic Acid Binding to TLR4 Promotes Proliferation and Blocks Apoptosis in Colon Cancer. Mol. Cancer Ther. 2019, 18, 2446–2456. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Ferrandez, E.; Gutierrez, O.; Segundo, D.S.; Fernandez-Luna, J.L. NFκB Activation in Differentiating Glioblastoma Stem-like Cells Is Promoted by Hyaluronic Acid Signaling through TLR4. Sci. Rep. 2018, 8, 6341. [Google Scholar] [CrossRef]
  189. Mukhopadhyay, D.; Hascall, V.C.; Day, A.J.; Salustri, A.; Fülöp, C. Two Distinct Populations of Tumor Necrosis Factor-Stimulated Gene-6 Protein in the Extracellular Matrix of Expanded Mouse Cumulus Cell–Oocyte Complexes. Arch. Biochem. Biophys. 2001, 394, 173–181. [Google Scholar] [CrossRef]
  190. Rugg, M.S.; Willis, A.C.; Mukhopadhyay, D.; Hascall, V.C.; Fries, E.; Fülöp, C.; Milner, C.M.; Day, A.J. Characterization of Complexes Formed between TSG-6 and Inter-Alpha-Inhibitor That Act as Intermediates in the Covalent Transfer of Heavy Chains onto Hyaluronan. J. Biol. Chem. 2005, 280, 25674–25686. [Google Scholar] [CrossRef] [Green Version]
  191. Zhang, G.; He, Y.; Liu, Y.; Du, Y.; Yang, C.; Gao, F. Reduced Hyaluronan Cross-Linking Induces Breast Cancer Malignancy in a CAF-Dependent Manner. Cell Death Dis. 2021, 12, 586. [Google Scholar] [CrossRef]
  192. Spinelli, F.M.; Vitale, D.L.; Icardi, A.; Caon, I.; Brandone, A.; Giannoni, P.; Saturno, V.; Passi, A.; García, M.; Sevic, I.; et al. Hyaluronan Preconditioning of Monocytes/Macrophages Affects Their Angiogenic Behavior and Regulation of TSG-6 Expression in a Tumor Type-Specific Manner. FEBS J. 2019, 286, 3433–3449. [Google Scholar] [CrossRef] [PubMed]
  193. Kim, P.K.; Halbrook, C.J.; Kerk, S.A.; Radyk, M.; Wisner, S.; Kremer, D.M.; Sajjakulnukit, P.; Andren, A.; Hou, S.W.; Trivedi, A.; et al. Hyaluronic Acid Fuels Pancreatic Cancer Cell Growth. eLife 2021, 10, e62645. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic representation of hyaluronic acid (HA) synthesis, metabolism, degradation, and signaling via hyaladerin interactions. In cancer, HA synthesis is mainly sustained by HA synthase 2 (HAS2), whose activity of is controlled by several growth factors and post-translational modifications, including ubiquitination, O-GlcNAcylation, and phosphorylation, the latter being induced by a decrease in ATP:AMP ratio. Moreover, HAS2 antisense 1 (HAS2-AS1) long noncoding RNA (lncRNA) epigenetically regulates HAS2 transcription. Pharmacological agents, such as 4-methylumbelliferone (4-MU), prevent HA synthesis by downregulating HAS2 activity. Altered glucose metabolism enhances the accumulation of UDP sugars, thereby increasing HA synthesis. At the plasma membrane, HYAL2 chops high-molecular-mass HA (HMWHA) into small fragments that are soon internalized into the tumor cell by endocytosis through coordinated binding with CD44 and further degraded in the lysosomes by HYAL1. The resulting sugars, N-acetyl-D-glucosamine (GlcNAc), and D-glucuronic acid (GlcUA), can be recycled via cell energy metabolism. HA in the tumor microenvironment can interact with different membrane receptors, including CD44 and RHAMM, thus stimulating protumorigenic pathways supporting cell proliferation, migration, invasion, and metastasis. HA also stimulates cell proliferation via inducing a toll-like receptor 4 (TLR-4)-dependent signaling cascade, although HA interaction with TLR receptors has never been demonstrated. Abbreviations: Glc, glucose; G6P, glucose-6-phosphate; F6P, Fructose-6-phosphate; PPP, pentose phosphate pathway.
Figure 1. Schematic representation of hyaluronic acid (HA) synthesis, metabolism, degradation, and signaling via hyaladerin interactions. In cancer, HA synthesis is mainly sustained by HA synthase 2 (HAS2), whose activity of is controlled by several growth factors and post-translational modifications, including ubiquitination, O-GlcNAcylation, and phosphorylation, the latter being induced by a decrease in ATP:AMP ratio. Moreover, HAS2 antisense 1 (HAS2-AS1) long noncoding RNA (lncRNA) epigenetically regulates HAS2 transcription. Pharmacological agents, such as 4-methylumbelliferone (4-MU), prevent HA synthesis by downregulating HAS2 activity. Altered glucose metabolism enhances the accumulation of UDP sugars, thereby increasing HA synthesis. At the plasma membrane, HYAL2 chops high-molecular-mass HA (HMWHA) into small fragments that are soon internalized into the tumor cell by endocytosis through coordinated binding with CD44 and further degraded in the lysosomes by HYAL1. The resulting sugars, N-acetyl-D-glucosamine (GlcNAc), and D-glucuronic acid (GlcUA), can be recycled via cell energy metabolism. HA in the tumor microenvironment can interact with different membrane receptors, including CD44 and RHAMM, thus stimulating protumorigenic pathways supporting cell proliferation, migration, invasion, and metastasis. HA also stimulates cell proliferation via inducing a toll-like receptor 4 (TLR-4)-dependent signaling cascade, although HA interaction with TLR receptors has never been demonstrated. Abbreviations: Glc, glucose; G6P, glucose-6-phosphate; F6P, Fructose-6-phosphate; PPP, pentose phosphate pathway.
Cancers 15 03813 g001
Table 1. Molecular profiles and expression of HA-related gens ER+ vs. ER breast cancer cell lines.
Table 1. Molecular profiles and expression of HA-related gens ER+ vs. ER breast cancer cell lines.
MDA-MB-231Hs 578TSUM149BT-549MCF-7T-47DBT-474
TUMOR 1ACIDCInfDCIDCACIDCIDC
ER 2 ----+++
PR 2----+++
HER2 2------+
HAS1 2±±±±+±±
HAS2 2++++++±±±
HAS3 2+++++±±±
HAS2-AS1 3++++++++++±--
UGDH 2++++++++++++++++
HYAL1 2±+++±+±
HYAL2 2+++++++++
RHAMM 2++++++++++++
CD44 2+++++++++++++++++++
CEMIP 2+++++++++++
TMEM2 2++++++++
1 Information from ATCC, 2 the human protein atlas and 3 [90]. AC, adenocarcinoma; IDC, invasive ductal carcinoma; InfDC, infiltrating ductal carcinoma. +/- data refer to normalized transcript per million (nTPM). -, 0 nTPM; ±, 0–1 nTPM; +, 1.1–50 nTPM; ++, 50.1–100 nTPM; +++, 100.1–200 nTPM; ++++, >200 nTPM.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Parnigoni, A.; Moretto, P.; Viola, M.; Karousou, E.; Passi, A.; Vigetti, D. Effects of Hyaluronan on Breast Cancer Aggressiveness. Cancers 2023, 15, 3813. https://doi.org/10.3390/cancers15153813

AMA Style

Parnigoni A, Moretto P, Viola M, Karousou E, Passi A, Vigetti D. Effects of Hyaluronan on Breast Cancer Aggressiveness. Cancers. 2023; 15(15):3813. https://doi.org/10.3390/cancers15153813

Chicago/Turabian Style

Parnigoni, Arianna, Paola Moretto, Manuela Viola, Evgenia Karousou, Alberto Passi, and Davide Vigetti. 2023. "Effects of Hyaluronan on Breast Cancer Aggressiveness" Cancers 15, no. 15: 3813. https://doi.org/10.3390/cancers15153813

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop