Next Article in Journal
Characterization and Inhibitory Effects of Essential Oil and Nanoemulsion from Ocotea indecora (Shott) Mez in Aspergillus Species
Previous Article in Journal
Dehydroborylation of Terminal Alkynes Using Lithium Aminoborohydrides
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Altered Glycosylation in Progression and Management of Bladder Cancer

by
Magdalena Wilczak
1,2,
Magdalena Surman
1 and
Małgorzata Przybyło
1,*
1
Department of Glycoconjugate Biochemistry, Faculty of Biology, Institute of Zoology and Biomedical Research, Jagiellonian University, Gronostajowa 9 Street, 30-387 Krakow, Poland
2
Doctoral School of Exact and Natural Sciences, Jagiellonian University, Prof. S. Łojasiewicza 11 Street, 30-348 Krakow, Poland
*
Author to whom correspondence should be addressed.
Submission received: 24 February 2023 / Revised: 5 April 2023 / Accepted: 11 April 2023 / Published: 13 April 2023
(This article belongs to the Section Chemical Biology)

Abstract

:
Bladder cancer (BC) is the 10th most common malignancy worldwide, with an estimated 573,000 new cases and 213,000 deaths in 2020. Available therapeutic approaches are still unable to reduce the incidence of BC metastasis and the high mortality rates of BC patients. Therefore, there is a need to deepen our understanding of the molecular mechanisms underlying BC progression to develop new diagnostic and therapeutic tools. One such mechanism is protein glycosylation. Numerous studies reported changes in glycan biosynthesis during neoplastic transformation, resulting in the appearance of the so-called tumor-associated carbohydrate antigens (TACAs) on the cell surface. TACAs affect a wide range of key biological processes, including tumor cell survival and proliferation, invasion and metastasis, induction of chronic inflammation, angiogenesis, immune evasion, and insensitivity to apoptosis. The purpose of this review is to summarize the current information on how altered glycosylation of bladder cancer cells promotes disease progression and to present the potential use of glycans for diagnostic and therapeutic purposes.

1. Introduction

Bladder cancer (BC) is the 10th most common malignancy worldwide, which accounts for 3% of global cancer diagnoses. The highest incidence rate of BC is observed in men in Southern and Western Europe and in North America [1]. According to GLOBOCAN, the estimated number of new BC cases in 2020 reached 573,000, and there were 213,000 new BC-related deaths [2]. BC is a highly heterogeneous cancer with a spectrum ranging from low-grade non-invasive tumors with good prognosis but high recurrence rate up to high-grade muscle-invasive tumors with a poor prognosis. Histologically, BC is classified into two categories: urothelial, which is diagnosed in more than 90% of cases, and non-urothelial, including squamous cell carcinoma, adenocarcinoma, and small cell carcinoma [3,4]. Based on invasiveness, BC is classified into non-muscle-infiltrating bladder cancer (NMIBC, 70% of cases) and muscle-infiltrating bladder cancer (MIBC, 30% of cases). Different therapeutic strategies for BC are applied based on the BC category. In the case of NMIBC, transurethral resection of bladder tumor (TURBT) is the most common strategy, followed by intravesical chemotherapy or immunotherapy, while in MIBC, cystectomy is used with a combination of chemotherapeutic drugs [5]. Unfortunately, current therapeutic approaches do not reduce high mortality rates in BC due to its frequent recurrence after TURBT and the possibility of redevelopment to more invasive and metastatic tumors [6,7]. Therefore, more attention should be devoted to a better understanding of the mechanisms underlying BC progression to establish new diagnostic and therapeutic tools in the future.
The aim of this review article is to summarize the current information on how altered glycosylation of BC cells promotes disease progression and to present the potential use of glycans for diagnostic and therapeutic purposes. Additionally, recent findings on the glycosylation status of BC-derived extracellular vesicles will be presented.

2. Alteration in Glycosylation Observed in Cancer

Glycosylation is an enzymatic process in which glycan structures are attached to proteins, lipids, and, as recently demonstrated, RNA [8]. Glycan is a general term for any monosaccharide or set of monosaccharides (oligosaccharides or polysaccharides) that are linked via covalent glycosidic bonds to another molecule. Glycosylation leads to the formation of glycoconjugates, which include glycoproteins, proteoglycans, glycosphingolipids, and glycosylated RNA (Figure 1). Additionally, glycans can exist as free oligosaccharides and non-peptide glycosaminoglycans (hyaluronic acid). For a long time, carbohydrates were considered only as a source of energy, a spare material, components of cell walls, membrane structures, mucus, and metabolites. However, ongoing research has shown that glycans play a significant role in many physiological and pathological biological processes [9]. It is also worth noting that the information about glycan synthesis is not completely written in the genome. Glycans are encoded in a complex network of at least several dozen genes that are (besides allelic variants) also affected by epigenetic and environmental factors [10].
The most crucial factors determining glycan formation are the expression of enzymes involved in their biosynthesis (i.e., glycosyltransferases and glycosidases) and the expression of genes encoding them, but also the availability of active monosaccharide donors. Consequently, glycosylation is highly species-, tissue-, and cell-specific. It depends on the current physiological/pathological state of the cell, but in some cases, there are allelic variants determining glycosylation status, e.g., glycan-based blood groups [11]. In the case of glycoproteins, we also deal with microheterogeneity (i.e., the attachment of different glycans to the same glycosylation site), which may contribute to their structural variability and affect their function (in a manner analogous to the effects of protein sequence changes caused by corresponding gene mutations) [12].
Numerous studies have shown that during neoplastic transformation, rapid changes in glycan biosynthesis result in the appearance of the so-called tumor-associated carbohydrate antigens (TACAs) on the cell surface [13]. The emergence of TACAs affects cell-cell and cell-extracellular matrix (ECM) interactions [14], survival and proliferation of cancer cells, tumor invasion, and metastasis, induction of chronic inflammation and angiogenesis, as well as immune evasion and insensitivity to apoptosis [15,16,17]. Cancer-related changes in glycan biosynthesis can result in an underexpression or overexpression of naturally occurring glycans, expression of glycans normally restricted to embryonic tissues, the appearance of incomplete or truncated structures, and, less commonly, the appearance of completely novel structures [15]. The most frequently observed changes in glycosylation in tumors and their functional significance are summarized in Table 1.
Studies on cancer-related changes in glycosylation have focused mainly on glycoproteins. Glycoproteins can be categorized into those containing N-linked glycans (nitrogen-attached), O-linked glycans (oxygen-attached), and less common C-linked glycans (carbon-attached). N-glycosylation occurs only within Asn−X−Ser/Thr amino acid sequence and, in a rare case, within Asn–X–Cys, where X can be any of the amino acids except for Pro. The key step in the N-glycan biosynthesis includes the synthesis of the precursor oligosaccharide, which is transferred en-block onto the newly synthesized peptide and is further processed into one of three types of N-glycans: high-mannose, complex, or hybrid structures (Figure 2) [18,19]. All N-glycans share a common pentasaccharide structure (Man3GlcNAc2).
In the case of O-glycosylation, the glycan is attached to the hydroxyl group of Ser or Thr residues, or to a lesser extent, to the hydroxyl group of hydroxylysine and hydroxyproline. The most common type of O-linked glycans is mucin-type O-glycans, where GalNAc is the first monosaccharide attached to Ser/Thr via α-glycosidic linkage (Figure 1). However, there are several other types of O-glycans found in glycoproteins other than mucin-type O-glycans, including β-linked O-GlcNAc, β-linked O-xylose, α-linked O-mannose, α-linked O-fucose, and also glucose and galactose, which can be both α- and β-linked. The listed monosaccharide residues attached to the above-mentioned amino acid residues by an O-glycosidic bond can then be elongated, which leads to the formation of either short chains with few residues or elongated bi-antennary structures or no elongation at all, as in the case of GlcNAcylation. Mucin-type O-glycans, unlike N-glycans, have fewer structural rules and do not share a common core structure. Subsequent elongation by successive linking of the monosaccharides to the eight types of core structures is followed by capping and sulfonation that leads to the formation of numerous different glycan structures found in mucins, i.e., glycoproteins bearing clusters of GalNAc-based O-glycans. It is worth emphasizing that the urothelium surface of the bladder wall is covered with a gel-like layer of mucins [20]. In the same layer, glycosaminoglycans can be found, which are long, linear, and highly negatively charged polysaccharides. Proteoglycans are formed through the linking of glycosaminoglycans via a special linker to the serine residue within a polypeptide chain (Figure 1). Another type of glycans found in glycoproteins is glycosylphosphatidylinositol (GPI) anchors. GPI are complex glycophospholipids covalently attached to many of the outer proteins of the eukaryote’s cell membrane (Figure 1). The function of GPI is to ensure the stable association of proteins lacking a transmembrane domain with membrane lipid rafts.
Table 1. Changes in glycosylation of cancer cells and their impact on cancer progression.
Table 1. Changes in glycosylation of cancer cells and their impact on cancer progression.
Change in GlycosylationRole in CancerReferences
Increase in β1,6-branched N-glycans due to overexpression of N-acetylglucosaminyltransferase V (gnt-V)Increased rate of metastases in mice[19,21,22,23]
Increase in β1,4-branched tetra-antennary N-glycan due to overexpression of N-acetylglucosaminyltransferase IV (gnt-IV)Enhanced tumor progression by lattice formation via galectin binding to poly-N-acetyllactosamines (lacnac), and the formation of sialyl-Lewis X (slex)[19,23,24]
Increase in N-glycans core fucosylation due to overexpression of α-1,6-fucosyltransferase (FUT8)Promotion of lung cancer and melanoma progression, a critical role in antibody-dependent cellular cytotoxicity (ADCC) and immune evasion, the regulation of TGF-β, EGF, α3β1 integrin, and E-cadherin function[19,23,25]
Increase in bisecting glcnac in N-glycans due to overexpression of
N-acetylglucosaminyltransferase III (gnt-III)
Suppression of tumor progression in melanoma and mouse mammary tumors[23,26]
Presence of Tn and T antigens and their sialylated glycoforms, sialyl-Tn (stn) and sialyl-T (ST), respectivelyInterference with immune cell recognition and blocking or masking of antigenic peptides presentation by major histocompatibility complex (MHC) molecules; enhanced tumorigenic and invasive properties and promotion of immunosuppression[23]
Increase in N-glycan α2,6 sialylation due to β-galactoside α2,6-sialyltransferase 1 (ST6GAL1) up-regulationIncreased integrin-mediated cell motility and protection from apoptosis induced by galectins, death receptor ligands, and chemotherapeutic drugs[27]
A2,8-linked polysialic acids (polysia)Reduced tumor cell anchorage to extracellular matrix components, making it easier for cancer cells to enter the bloodstream and interact with platelets[14]
N-acetylglucosaminylacylation (O-glcnacylation)Cancer risk factor, excessive nutritional intake[28]
O-acetylated gangliosidesProtection from apoptosis[23]
Metabolic incorporation of diet-derived N-glycolylneuraminic acid (Neu5Gc) into human glycansPromotion of tumor growth by enhancing chronic inflammation and angiogenesis[23]
Sialylation of Lewisx and Lewisx/aCrucial in extravasation[14]
Slex and slea epitopes on glycosphingolipidsMetastatic potential in mice and tumor progression, metastatic spread, and poor prognosis of patients[19]
Sialyl-Lewis-related structuresInfluencing tumor progression by interacting with Siglecs that have immunosuppressive functions[23]
Complete loss of glycosylphosphatidylinositol (GPI)-anchored proteinsObserved in some cases of malignant and premalignant states involving the hematopoietic system[23]

3. The Role of β1,6-Branched N-Glycans in Bladder Cancer

The biosynthesis of β1-6 branched N-glycans relies on the activity of N-acetylglucosaminyltransferase V (EC 2.4.1.155, GnT-V, encoded by MGAT5 gene). GnT-V catalyzes the attachment of N-acetylglucosamine (GlcNAc) residue to α1-6-linked mannose in the core structure via β1-6-linkage using UDP-GlcNAc as the donor substrate (Figure 3) [29,30]. This modification takes place in the Golgi apparatus and concerns only particular proteins, including growth factor receptors, cell adhesion molecules, enzymes, and endogenous lectins [14,31,32]. The formation of tri-/tetra antennary β1-6 branched N-glycans can modulate the half-life and stability of extracellular binding proteins, as well as their functional activity.
Furthermore, there is a link between β1,6-branched N-glycans and cancer development [15,32], and the close association between these structures and a poor prognosis of cancer patients has been reported [33,34]. Nevertheless, the clinical implications of the expression of GnT-V and β1,6-branched N-glycans vary depending on the type of cancer. So far, there are few studies related to the role of β1,6-branched N-glycans in BC. In the mid-2000s, swainsonine was used to evaluate the impact of β1,6-branched complex type N-glycans on the adhesion and migration of BC T24 cells [35]. Swainsonine is an indolizidine alkaloid, first isolated from Swainson canescens, which is an inhibitor of Golgi α-mannosidase II [36]. This enzyme catalyzes the removal of mannose residues from the α1-6 arm of GlcNac2Man5GlcNAc, and its inhibition leads to the accumulation of high mannose (Man4GlcNAc2 and Man5GlcNAc2) and hybrid-type N-glycans in glycoproteins. It was shown that after swainsonine treatment, BC T24 cells significantly increased their adhesion to collagen IV, laminin, and fibronectin. Swainsonine treatment also diminished the migration rate in wound healing assay, indicating the role of β1,6-branched complex type N-glycans in this process [35]. Additionally, non-malignant urothelial HCV-29 cells displayed higher adhesiveness and weaker migratory properties than T-24 BC cells [36,37]. It can be partly attributed to the different glycosylation of α3β1 integrin. Only BC cells showed the presence of β1,6-branched N-glycans on both α3 and β1 integrin subunits [38,39], whereas the level of the α3 integrin subunit expression in both cell lines showed no statistically significant differences [40]. Interestingly, in the case of cadherins, the presence of β1,6-branched N-glycans was demonstrated only for T24 cells but not for HCV29 cells [41].
Although in vitro research suggests that β1,6-branched N-glycans promote BC progression, studies using patient tissues showed that GnT-V expression is lower in advanced stages of BC, while higher GnT-V expression correlates with longer disease-free survival [42,43]. Nevertheless, GnT-V activity in BC tissues was shown to be slightly elevated in comparison to non-malignant specimens [44]. Recently, a link between the activity of GnT-V and response to chemotherapeutic gemcitabine in BC was studied [45,46]. Gemcitabine treatment of T24 cells increased the expression of GnT-V and the amount of β1,6-branched N-glycans on human equilibrative nucleoside transporter 1 (hENT1). On the other hand, GnT-V silencing in T24 cells caused a significant decrease in the amount of β1,6-branched N-glycans on hENT1. It led to decreased uptake of gemcitabine by BC cells, inhibition of drug-induced apoptosis (due to reduced expression of active caspases 3, 8, and 9), as well as G2/M cell cycle arrest. Thus, the expression of β,1-6 branched N-glycans, and GnT-V is related to low malignant potential and good prognosis for BC patients.

4. The Role of Bisecting N-Acetylglucosamine in Bladder Cancer

N-acetylglucosaminyltransferase III (EC 2.4.1.144, GnT-III, encoded by MGAT3 gene) catalyzes the transfer of N-acetylglucosamine (GlcNAc) residue from UDP-GlcNAc to core mannose of N-glycan with a β1,4 linkage forming a bisecting GlcNAc structure (Figure 3) [47]. The introduction of a bisecting GlcNAc residue causes a conformational change in the glycan and prevents further processing and the biosynthesis of β1,6-branched N-glycans catalyzed by GnT-V [24]. In the context of a tumor, GnT-III and bisecting GlcNAc are mostly considered suppressors of tumor growth, epithelial-mesenchymal transition, and metastasis [24,48]. To our knowledge, there are only two papers describing the role of bisecting GlcNAc in bladder cancer.
First, it has been shown that there is a link between the increase in bisecting GlcNAc on fibronectin from patients’ urine and the pathological stage and grade of BC [44]. The binding affinity of fibronectin to wheat germ agglutinin (WGA) was, on average, 3.26 times higher for BC samples compared to healthy controls. Additionally, the GnT-III activity was, on average, 34 times higher in BC samples, and its activity positively correlated with the stage and grade of BC. These observations regarding the elevated level of bisecting GlcNAc in BC have recently been confirmed by another study on N-glycosylation of serum immunoglobulins (Igs) in urothelial carcinomas (UC), most of which originate from bladder [49]. The results of that study clearly showed that bisecting GlcNAc on Igs was significantly accumulated in UC patients. Summing up, even though the number of studies on the role of bisecting GlcNAc in BC is limited, they contradict the hypothesis that this modification has tumor-suppressive functions.

5. The Role of O-GlcNAcylation in Bladder Cancer

O-GlcNAcylation is a reversible and dynamic post-translational modification of cytoplasmic, nuclear, mitochondrial, and plasma membrane proteins involving the attachment of a single GlcNAc residue to serine or threonine residues. O-GlcNAcylation modulates the function of proteins involved in fundamental cellular processes such as transcription, translation, signal transduction, metabolism, cytoskeletal functions, and cell division [50,51,52]. The process of O-GlcNAc attachment is catalyzed by O-GlcNAc transferase (EC 2.4.1.255, OGT, encoded by Ogt gene), whereas the removal of the attached GlcNAc residue is catalyzed by O-GlcNAcase (EC 3.2.1.169, OGA, encoded by MGEA5 gene) (Figure 4) [53]. The level of O-GlcNAcylated proteins is regulated by the nutritional state of the cell as well as metabolic, oxidative, and proteotoxic stress. Finally, O-GlcNAcylation directly or indirectly competes with phosphorylation for the same serine/threonine residues. It provides complex cross-communication between the two posttranslational modifications for control of the function of different proteins (Figure 4) [51,54].
Furthermore, O-GlcNAcylation is highly responsive to environmental nutrient concentrations, i.e., glucose and glutamine [55]. Numerous epidemiological studies have shown that obesity and diabetic conditions significantly increase cancer risk and enhance O-GlcNAcylation. [28,56,57,58]. Increased O-GlcNAcylation and changes in OGT and/or OGA expression have been demonstrated in several types of cancer. Moreover, O-GlcNAcylation participates in the regulation of the epigenome and transcription factors activity (MYC, p53, NFκΒ, β-catenin, FOXM1, estrogen receptor (1) as well as in reprogramming of the metabolic network of cancer cells to promote tumor growth. O-GlcNAcylation can also alter cancer cell proliferation by (2) modulating PI3K/Akt activity and participating during the cell cycle at various stages, (3) promoting invasiveness and metastasis by affecting the E-cadherin/β-catenin system, (4) and/or by promoting the expression of metalloproteinases. Finally, O-GlcNAcylation is involved in inducing immune surveillance in the tumor microenvironment and is involved in the activation and differentiation of T lymphocytes and macrophages.
To our knowledge, the first papers on the role of O-GlcNAcylation in BC were published over 20 years ago. It was shown that OGT mRNA was expressed in the urine of 51.7% of BC patients but not in the urine of healthy individuals. Moreover, the OGT mRNA level was significantly higher in the urine of patients with grades II and III of BC in comparison to grade I [59]. Consistent with this observation, other studies demonstrated the tumor-promoting activity of GlcNAcylation in BC [60,61,62]. Evaluation of OGT expression and GlcNAcylation levels by immunohistochemistry, the sequencing data, and Western blotting showed their upregulation in BC tissues and cells vs. normal tissues/cells [60,61,62]. Importantly, OGT levels were higher in MIBC than in NMIBC. On the other hand, OGA expression was found to be lower in cancer tissues [61].
The direct involvement of GlcNAcylation in promoting BC cell proliferation in vitro and tumor xenograft growth in vivo, as well as in inhibiting apoptosis and autophagy, was also proven by OGT knockdown studies [60,61]. In vitro, O-GlcNAcylation was found to promote migration and invasion of BC cells and to induce cell cycle arrest [61]. It was later proven that autophagy in BC cells could be induced by blocking of O-GlcNAcylation of the autophagy suppressor AMP-activated protein kinase (AMPK) [63]. On the other hand, the inhibition of apoptosis in BC cells may result from cyclin-dependent-like kinase 5 (CDK5) modification by O-linked GlcNAc [62]. Summing up, the removal of O-GlcNAc or inhibition of O-GlcNAcylation may help to induce autophagy in BC cells or prevent BC cells from escaping apoptosis.
Furthermore, it has been shown that hyper-O-GlcNAcylation due to nucleolar and spindle-associated protein 1 (NUSAP1) overexpression might promote BC progression [64]. NUSAP1 is a microtubule-binding protein located mainly in the nucleolus, the up-regulation of which enhances cancer cell invasive and migratory potential and is associated with patients’ poor survival. Another study demonstrated that a reduction of OGT expression increases the sensitivity of bladder cancer cells to cisplatin [60]. Moreover, the acquired resistance to gemcitabine and paclitaxel was shown to induce the overexpression of OGT in BC [65]. It suggests the pathological role of OGT in the development of multidrug resistance in BC.

6. The Role of Total Sialylation in Bladder Cancer

Sialylation is an enzymatic process during which sialic acid residues are covalently attached through α2,3, α2,6, and α2,8 binding to the terminal positions of glycan chains on various acceptors, such as glycoproteins, glycolipids, polysaccharides, including polysialic acid chains. Sialic acids (Sia) represent a family of more than 50 derivatives of neuraminic acid, a nine-carbon monosaccharide consisting of a six-membered carbon ring (C1–C6) with a three-carbon C7–C9 glycerol chain attached. The main derivatives of neuraminic acid are N-acetyl-neuraminic acid (Neu5Ac) and N-glycolyl-neuraminic acid (Neu5Gc), which are its N- and O-substituted derivatives, respectively. Neu5Ac and Neu5Gc can be further modified by the addition of O-acetyl, O-methyl, sulfate, O-lactyl, and phosphate groups at various positions. Importantly, NeuGc is selectively expressed in cancer cells, whereas it is absent or present in low amounts in normal tissues because the human CMAH gene is irreversibly mutated. CMP-Sia is a donor of sialic acids in the sialylation reaction that is catalyzed by members of the sialyltransferase family, which in humans consists of 20 subtypes [66,67].
It is known that cancer cells often exhibit elevated levels of sialylated glycoconjugates. This overexpression may result from the deregulation of the enzymatic activity of sialyltransferases or neuraminidases (glycosidic hydrolases that catalyze the removal of sialic acid residues from glycoconjugates) or both [66,68,69]. Abnormal sialylation has been found to affect the adhesive, migratory and invasive properties of tumor cells, epithelial-mesenchymal transition (EMT), metastasis formation, signal transduction, angiogenesis, as well as immune system evasion, avoidance of apoptosis and cell death, and reduced efficacy of chemotherapy and radiotherapy [70,71,72,73,74].
Total sialic acid (TSA) has been widely recognized as a specific tumor marker and potential therapeutic target due to its higher expression on the surface of cancer cells. In BC, the enzyme responsible for the increase of TSA is ST3GAL6, an enzyme that has α-2,3-sialyltransferase activity toward the Gal-β1,4-GlcNAc structure on glycoproteins and glycolipids [75]. ST3GAL6 has been shown to be downregulated by the transcription factor GATA3, which is one of the key factors maintaining luminal differentiation of urothelial cells [76]. On the other hand, overexpression of ST3GAL6 increased the migration and invasive capacity of BC cells and correlated with the poor prognosis of bladder cancer patients [75]. Elevated serum sialic acid levels have also been correlated with the extent of malignancy, and they appear as a promising biomarker of urothelial tumors [77].

7. The Role of Mucin-Type O-Glycans in Bladder Cancer

Another type of glycans that are covalently attached to serine and threonine residues of proteins via O-glycosylic linkage are mucin-type O-glycans. Mucin-type O-glycosylation, which is the dominant type of O-glycosylation, is an evolutionarily conserved post-translational modification in animals. Mucin-type O-glycosylation is initiated in the Golgi apparatus by a family of 20 homologous polypeptide N-acetylgalactosaminyltransferases (EC 2.4.1.41, GalNAc-Ts). They catalyze the first step of biosynthesis by forming GalNAcα1-O Ser/Thr linkage in O-glycoproteins, known as Tn antigen [78,79]. Importantly, GalNAc-T1 is necessary for the self-renewal and tumor-initiating capacity of BC stem cells [80]. The O-GalNAc residues are further processed by the addition of galactose and GlcNAc residues, leading to the synthesis of 8 types of core structures. Core structures are further elongated by the addition of galactose, GlcNAc, fucose, and sialic acid residues (Figure 5). Mucin-type O-glycans are found on cell surface proteins, i.e., mucins and other glycoproteins. Mucins are involved in the protection of epithelial cell surface as well as in the sorting and secretion of glycoproteins, cell adhesion, migration, trafficking of hematopoietic cells, host-pathogen interaction, and immune responses [79,81,82,83,84,85]. Importantly, altered expression of these structures is known to be associated with several diseases, including cancer.
A number of excellent reviews have been published in recent years that describe the role of mucins in cancer, to cite some of them [86,87,88,89,90,91,92,93,94,95,96,97,98,99,100]. Changes in the expression, distribution, and glycosylation of mucins have been well documented in several cancers, providing a rationale for exploring their use as both diagnostic and prognostic markers, as well as targets for the development of cancer therapies. The role of mucins in cancer pathogenesis stems from the effects they exert on numerous signaling pathways, such as NF-kB, ERα, HIF, MAPK, p53, c-Src, Wnt, and JAK-STAT, which results in the reprogramming of cancer-related genes expression profiles. Numerous studies have confirmed that mucins are involved in the regulation of cancer cell proliferation, growth, migratory and invasive properties, their detachment and reattachment during metastasis, self-renewal of cancer stem cells (CSCs), avoidance of apoptosis, escape from immune surveillance, drug metabolism, and induction of drug resistance. Moreover, glycans in mucins serve as ligands for various carbohydrate-binding proteins such as galectins, selectins, and siglectins, thus changes in their structure are responsible for altering the interaction of cancer cells with other cells, e.g., lymphocytes, platelets, and endothelial cells.

7.1. The Role of Mucin Expression in Bladder Cancer

As mentioned above, mucins (MUCs) are the main carriers of mucin-type O-glycans. MUCs are high molecular weight glycosylated proteins present in the epithelial lining of certain tissues and ducts. Above twenty types of mucins have been identified. They have been divided into two major groups, i.e., membrane-bound and secreted. It is well known that abnormal expression and glycosylation of mucins contribute to tumor survival and proliferation in many types of cancers [89]. Regarding BC, MUC1, MUC2, MUC4, MUC6, and MUC7 have been shown to be expressed by urothelial cells [101,102].
It was found that the differences in the distribution of mucins within the epithelium of the urinary tract as well as in blood and urine samples have diagnostic and prognostic potential. MUC-1 expression has been found to increase when primary BC acquires a more aggressive phenotype, and its high expression is correlated with higher tumor grade and worse prognosis [103,104,105,106,107]. MUC2 and MUC6 expression was associated with low-grade tumors and presented an inverse correlation with cancer-specific deaths [105]. On the other hand, MUC4 expression was linked to cancer-specific deaths [104,105]. Finally, MUC7 expression was detected only in samples of invasive transitional cell carcinomas [108,109,110,111].
Another member of the mucin family proteins, MUC16, also known as cancer antigen 125 (CA125), has been associated with shorter disease-free periods and poorer overall survival rates of BC patients [112,113]. MUC16 expression was also associated with the immunosuppressive tumor microenvironment and gemcitabine/cisplatin resistance [112,113]. Furthermore, CD164, another member of this family involved in tumor maintenance and progression, has been associated with poor clinical outcomes of BC patients [114]. Last but not least, Cluster of Differentiation 44 (CD44), the major surface receptor for hyaluronic acid, is another glycoprotein with mucin-type O-linked glycans. CD44 is a multifunctional protein involved in cell adhesion, cell-matrix adhesion, cell migration, leukocyte homing and activation, tumor invasion, and metastasis. CD44 expression has been positively correlated with tumor aggressiveness and the response of bladder cancer to radiation, as well as associated with drug-resistant phenotype and poor prognosis [115,116,117,118]. It has also been proven that CD44 O-glycosylation status regulates cancer aggressiveness [119].
Finally, MUC1 (also known as CA15-3), whose expression correlates with high BC grade, takes part in epithelial surface protection, regulation of receptor kinase signaling, and cell adhesion by interacting with β-catenin and cyclin D1. In cancer, overexpression of MUC1 interferes with cell adhesion, protects tumor cells from immune recognition, and promotes metastasis. Furthermore, MUC1, an autoantigen, is altered during the malignant process and induces immune responses. Recently, MUC1 has been shown to play a key role in the acquisition of drug resistance [99,120] and metabolic reprogramming leading to increased glycolysis, glucose uptake, and lactate production in BC [99]. Interestingly, high expression of MUC1 was associated with a favorable prognosis for patients with BC when the expression of the epidermal growth factor (EGF) receptor HER3 was also high [121,122]. MUC1 is highly O-glycosylated on Ser/Thr residues and carries core 2 mucin-type O-glycans, whose role in bladder cancer is described in Section 7.3.
Summing up, not only changes in the expression of mucins but also in their post-translational modifications are a common feature of many tumors originating from human epithelial cells. Mucin-type O-glycans associated with malignancies are incompletely synthesized O-linked glycans. Their presence, together with the accumulation of precursors, results in the loss of the normal O-glycan antigens.

7.2. The Role of Tn and T Antigens in Bladder Cancer

Tn antigen, T antigen, and their sialic acid counterparts, i.e., the sialyl-Tn (STn) antigen and the (mono-/di-) sialyl-T (ST) antigen (Figure 5), are among the most intensively investigated TACAs in BC. Moreover, in BC tissues, a predominance of sialoglycans over neutral glycoforms of Tn and T antigens has been observed [112]. Tn and STn antigens are not found in normal urothelium because Tn is masked by covalently bounded terminal carbohydrate residues and is, therefore, hardly expressed [123]. The premature stop of Tn antigen elongation by its sialylation is observed in tumors and leads to the synthesis of STn antigen. STn antigen is the most widely studied glycotype in cancer, and its biosynthesis is controlled by the ST6 N-acetylgalactosaminide α-2,6-sialyltransferase 1 (ST6GALNAC-I), which catalyzes the transfer of NeuNAc to the O-6 position of GalNAc residues linked to Ser/Thr. Increased ST6GALNAC-I activity has been associated with the efficacy of treatment with Bacillus Calmette-Guérin (BCG), the gold standard adjuvant immunotherapy for NMIBC after transurethral resection, to which one-third of patients fail to respond. The analysis of BC tissues showed that patients with elevated activity of ST6GALNAC-I, along with high expression of interleukin-6 mRNA, responded to BCG treatment [124]. In addition, STn antigen expression alone or in combination with S6T antigen, i.e., T antigen isotype having α2,6-linked sialic acid residue, has been associated with lower recurrence rates after BCG treatment [125]. Moreover, STn antigen-expressing cancer cells internalized more BCG and showed higher rates of BCG-induced apoptosis compared to STn-negative cancer cells [125].
In BC, the STn antigen has been associated with high-grade tumors, invasion of the muscle layer, and overall poor survival [112,123,125,126,127]. In addition, a strong link between STn antigen expression, tumor dissemination, and metastasis was stated based on high STn antigen expression in circulating tumor cells (CTCs) [128,129]. Moreover, STn antigen has been shown to induce morphological changes in BC cells, modulate their cell-cell and cell-matrix adhesion properties, increase invasive capacity, and enable immune escape [123,125,127,130]. In the immunological context, STn antigen is known to contribute to evasion from immune surveillance. For instance, STn antigen-bearing mucins inhibit the cytotoxicity of natural killer (NK) cells, thereby impairing their function [131]. Glycoproteomic analyses of BC tissues have also revealed that STn antigen was carried by 143 glycoproteins involved in mediating protein binding, cell-cell communication, cell signaling, regulation of metabolic processes, and hydrolase catalytic activities [112]. These observations regarding STn antigen-bearing glycoproteins were in line with previous in vitro studies [127]. Among STn antigen-expressing proteins, several integrins and cadherins, as well as CD44, MUC1, and MUC16, were identified [112,127]. Interestingly, overexpression of STn antigen has been shown to modulate CD44-mediated invasion and regulate oncogenic pathways governed by CD44 [118]. Moreover, there is a positive correlation between STn antigen expression and the higher proliferation index of BC cells based on p53 and Ki-67 biomarkers [132]. Finally, hypoxia, which is typical for advanced-stage bladder tumors [133,134], may promote STn antigen overexpression in BC cells [127]. In contrast, glucose deprivation has led to the expression of more Tn antigens and, to less extent, STn antigens [135]. HOMER3 has been identified as one of the main glycoproteins triggered by hypoxia and glucose deprivation, whose expression in bladder cancer patients has been associated with more aggressive primary tumors and distant metastases [136].
STn antigen expression is clinically important, not only as a diagnostic and prognostic marker in cancer (CA72-4 serological test) but also as a target of therapeutic strategies. STn antigen-based vaccines have had limited success, probably because STn antigen-expressing cancer cells impair dendritic cell (DC) maturation, downregulate the expression of proinflammatory cytokines (IL-12 and TNF-α), and endow DCs with a tolerogenic function [130]. In addition, STn antigen interactions with Siglec-15 present on macrophages have been shown to enhance the secretion of TGF-β, a key regulator of immune tolerance [137]. However, a novel CAR T cell-based treatment strategy targeting STn antigen has recently been developed in mice. The treatment effectively promoted the secretion of proinflammatory cytokines and tumor cell lysis in vitro and in experimental mice [138]. STn antigen has also been selected as a surrogate marker of BC associated with Schistosoma haematobium infection [139].
Furthermore, core 1 β1,3-galactosyltransferase (C1GnT) is the enzyme catalyzing the conversion of Tn antigen to T antigen, also known as core 1 structure. C1GnT has been shown to modulate the expression of target glycoproteins, such as MUC1 in spontaneous gastritis and gastric cancer [140] and MUC16 in pancreatic adenocarcinomas [141]. On the other hand, the loss of C1GnT resulted in Tn antigen enrichment on CD44 in pancreatic cancer [142]. In BC tissues, the expression of C1GnT and T antigens was found to be higher than in noncancerous bladder tissues [143]. In addition, silencing of C1GnT inhibited the migration and proliferation of BC cells by modifying target glycoproteins, including MUC16. It confirmed the pro-metastatic and proliferative function of upregulated C1GnT. Furthermore, C1GnT expression depends on the co-expression of its specific molecular chaperone, i.e., COSMC protein. COSMC, localized in the endoplasmic reticulum, ensures proper folding and prevents aggregation and proteasomal degradation of C1GnT. COSMC can also interact with partly denatured C1GnT and partially restore its activity. Knock out, silencing, or mutations in the COSMC gene disturb the balance of O-GlcNAcylation, and such abnormal levels of O-glycosylation are closely associated with tumor growth and differentiation [144,145].
The T antigen is also the precursor of the ABO blood group determinants. In BC patients, the loss of ABO(H) blood group determinant expression is associated with advanced forms of the disease and an unfavorable prognosis [146]. The overexpression of T antigen and its sialylated form (ST antigen) has also been associated with poor prognosis [147]. T antigen is a substrate for a few sialyltransferases. However, upregulation of ST3Gal-I has been shown to be one of the main mechanisms responsible for T antigen sialylation, and much higher mRNA levels of ST3Gal I were observed in malignant BC [147]. Interestingly, ST3Gal-I overexpression and the consequent replacement of T antigen with ST antigen having α2,3-linked sialic acid have been shown to induce transcriptomic changes in BC cells. These changes included the decreased expression of several genes involved in different mechanisms of DNA repair, the accuracy of chromosomal segregation, increased susceptibility to oxidative damage, and a stronger inflammatory response of macrophages [148]. Thus, ST antigen overexpression mediated by STGAL-I upregulation appears in the initial stages of oncogenic transformation and persists into more advanced stages of BC [112]. Both whole serum mono- and di- sialylated T antigens were increased in BC patients [149]. Finally, glycoproteomic analysis of two BC cell lines (T24 and 5637 cells) has revealed that over 90% of identified proteins potentially expressed ST antigen [135].

7.3. The Role of Core 2 Structures in Bladder Cancer

In addition to sialylation, T antigen can be modified to form core 2 mucin-type O-glycans that provide a scaffold for subsequent poly-N-acetyllactosamine synthesis. The conversion of T antigen to core 2 is catalyzed by core 2 β-1,6-N-acetylglucosaminyltransferase (C2GnT) (Figure 5), the expression of which positively correlates with both BC stage and grade [150]. Moreover, it has been found that BC patients with C2GnT expression had significantly shorter survival rates than patients not expressing C2GnT [151].
Upregulation of C2GnT has been shown to be a mechanism by which BC cells evade NK immunity. NK cells are key antitumor-effective cells that detect and eliminate circulating cancer cells at the early stages of the disease. One of the substrates of C2GnT is MHC class I-related chain A (MICA), a ligand expressed by cancer cells for an NK-activating receptor (NKG2D). Modification of MICA with poly-N-acetyllactosamine reduces its affinity for NKG2D on NK cells, which allows cancer cells to evade NK-mediated cell death [151]. Another mechanism that allows evading NK response is also related to the presence of mucin core 2 with poly-N-acetyllactosamine on MUC1, which enables the binding of MUC1 through to galectin 3 [150]. It interferes with the access of the tumor necrosis factor-related apoptosis-inducing ligand (on NK cells) to the surface of cancer cells, thereby prolonging the life of CTCs. Furthermore, it has been shown that bladder cancer cells expressing C2GnT have been highly susceptible to cytotoxic T lymphocyte (CTL)-mediated response. It is possible due to the presence of O-glycans of mucin core 2 with poly-N-acetyllactosamine on human leukocyte antigen class I (HLA I), which, due to binding with galectin 3, ensures retention of HLA I on the cell-surface [152]. Thus, evasion of CTL-mediated antitumor immunity is associated with the loss of core 2 structures on human leukocyte antigen HLA I as downregulation of C2GnT in metastatic cells has been reported [152].
It has also been found that the total mRNA level of β1,4-galactosyltransferase-1 (β4GalT1), an enzyme that catalyzes the transfer of galactose from UDP-Gal to GlcNAc and forms the poly-N-acetyllactosamine structure, was significantly higher in BC tissues than in normal ones [153]. Moreover, the level of upregulated expression of a long-form of β4GalT1 (β4GalT1-L) was higher than that of short-form β4GalT1 (β4GalT1-S) in BC cells. These forms were also differently localized in cells and played different roles in drug resistance. Chemotherapy upregulated the expression of β4GalT1-S. On the other hand, the expression of β4GalT1-L was downregulated and resulted in the decreased amount of poly-N-acetyllactosamine structure on mouse double minute 2 homolog (MDM2), a key negative regulator of p53 because MDM2 is preferentially glycosylated by β4GalT1-L. Therefore, chemotherapy-induced overexpression of β4GalT1-S may result in the acquisition of drug resistance and cell stemness through the decrease in poly-N-acetyllactosamine on MDM2, followed by upregulation of p53 and its decreased ubiquitination.

8. Fucosylation in Bladder Cancer

L-Fucose (6-deoxy-L-galactose) is another sugar residue found in glycans in all mammalian cells. Fucose is transferred onto the glycan chains from GDP-fucose, and the process is catalyzed by several fucosyltransferases (FUTs) [154]. So far, 13 genes encoding FUTs have been identified in the human genome. The majority of FUTs from the Golgi apparatus can modify N-linked glycans, whereas O-fucosyltransferases are mostly localized in the endoplasmic reticulum [155]. Based on the location of fucose residue within the glycan chain, fucosylation is divided into core fucosylation (α-1,6-linked fucose), whose formation is catalyzed only by FUT8, and terminal fucosylation (α-1,2- or α-1, 3/4 -linked fucose), whose formation is catalyzed by FUT1-FUT4, FUT6, FUT7, and FUT9 [154]. Regarding the biological role of fucosylated glycans, ABO blood antigens are the most common fucose-containing glycans. It has also been demonstrated that fucosylated glycans are crucial for selectin-mediated leukocyte extravasation, lymphocyte homing, and pathogen-host interactions. Fucosylated proteins are also widely involved in signal transduction.
Changes in fucosylation pattern may be due to the altered availability of GDP-fucose, a monosaccharide donor, or abnormal expression of FUTs and/or α-fucosidase, which may affect the proliferation and survival of cancer cells. Overexpression of α-1,6-fucosyltransferase has been found to result in increased attachment of fucose residue to the innermost GlcNAc in N-glycans core structure, and such a change is a major feature of hepatocarcinogenesis in humans [156]. Aberrant fucosylation has also been described in glioblastoma as well as in breast, colorectal, lung, and prostate cancers [157]. BC cells have been shown to possess a high expression of core fucosylated complex type N-glycans and a low expression of terminally fucosylated complex type N-glycans [158]. Importantly, elevated levels of plasma glycoproteins with terminally fucosylated glycans can distinguish BC patients from healthy controls [159].
Data available in the literature have clearly indicated the involvement of fucosylation in the formation of metastasis sites in BC. N-glycan expression profiling in transforming growth factor-beta (TGFβ)-induced EMT using normal ureter epithelial HCV-29 cells showed decreased expression of α-L-fucosidase, which contributed to increased expression of fucosylated N-glycans [160]. Similarly, a comparison of N-glycan profiles of TGFβ-treated vs. control BC cells by MALDI-TOF/TOF-MS revealed increased fucosylation in TGFβ-treated cells. These findings were consistent with the results of the lectin microarray analysis. Fucα1-2Galβ1-4GlcNAc structure, which was recognized by Ulex europaeus agglutinin-I (UEA-I lectin), was approximately two-fold higher in TGFβ-treated cells than in control BC cells [160].
Furthermore, calreticulin (CRT) has been demonstrated to affect cell adhesion and metastasis of BC cells. CRT stabilized mRNA for FUT-1 and thus regulated FUT-1 gene expression. In the same study, CRT indirectly affected β1 integrin-mediated cell adhesion by altering the level of α1,2-linked fucosylation (FUT-1-dependant) of the β1 integrin subunit. The study also indicated that α1,2-fucosylation of the β1 integrin subunit promoted its activation instead of modifying the integrin-binding sites [161]. Another group of cell adhesion molecules involved in BC progression is cadherins. Positive reaction with Aleuria aurantia agglutinin (AAA) has confirmed the presence of fucose residues on N-cadherin from highly invasive T24 BC cell line and their absence in the case of three less invasive cell lines [41]. Since AAA specifically recognizes not only α1,6-linked “core” fucose residue and a Fucα1,2Galβ1,4GlcNAc sequence (blood group H(O) determinant) but also a Galβ1,4(Fucα1,3)GlcNAc sequence (LewisX determinant), N-cadherin from T24 cells might possess both core fucosylated and/or α1,2/3-fucosylated glycans.
Furthermore, FUT7 expression in BC has been shown to correlate positively with the number of tumor-infiltrating lymphocytes, tumor growth and invasiveness, cancer cell migration, and the occurrence of EMT [162]. Moreover, in another study, BC cell lines derived from invasive tumors have been demonstrated to display increased expression of FUT6 and FUT7, whereas cell lines from non-invasive tumors or normal bladder epithelia have been negative for FUT6 and FUT7 expression [163]. These results suggest that elevated levels of FUT6 and FUT7 should be evaluated as potential prognostic biomarkers in BC.
Another potential and specific diagnostic target is the glycoform of α3 integrin subunit (ITGA3) from the urine of BC patients. Detection of fucosylated ITGA3 in the ITGA3-UEA assay (anti-ITGA3 antibody and fucose-specific Ulex europaeus agglutinin (UEA lectin) conjugated on nanoparticles) has shown that this recognized glycovariant significantly discriminate BC patients from benign hyperplasia patients [164]. Such an assay has also been performed using extracellular vesicles (EVs) derived from BC patients’ urine and BC cell-conditioned media [165]. This study showed that a biomarker combination consisting of ITGA3 and fucose is highly expressed on EVs derived from BC cells and BC urine samples and holds promise in the detection of urological pathologies [164]. More information regarding the glycosylation of EVs in BC is provided in Section 10.
Last but not least, α-1-acid glycoprotein (AGP) is a highly heterogeneous protein that can acquire different glycosylation statuses. With the use of capillary zone electrophoresis coupled with tandem mass spectrometry (CZE-MS), it has been shown that samples from BC patients vs. healthy controls display a higher abundance of AGP isoforms containing tri- and tetra-antennary fucosylated N-glycans [166]. This suggests that fucosylated AGP should also be considered as a potential biomarker of BC.
Finally, fucose residues are also present in a very specific group of glycan epitopes called Lewis antigens. The role of fucosylated Lewis antigens in BC will be discussed in detail in Section 9.

9. Lewis Antigens in Bladder Cancer

The Lewis antigen system is a human blood group system consisting of type I and type II Lewis antigens, which are terminally fucosylated glycan epitopes. The difference between type I and type II Lewis antigens is the type of glycosidic bond present in their structure, i.e., Galβ1-3GlcNAc in type I versus Galβ1-4GlcNAc in type II Lewis antigens. All Lewis antigens (H1, H2, Lewisa (Lea), Lewisb (Leb), Lewisx (Lex), and Lewisy (Ley)) contain the same three sugar residues (GlcNAc, Gal, and Fuc). Further elongation with sialic acid creates more complex glycan structures, e.g., sialyl Lewisa (SLea) or sialyl Lewisx (SLex) (Figure 6). All fucosyltransferases involved in the synthesis of Lewis antigens (FUT1–7 and FUT9) possess a unique substrate specificity that contributes to the high complexity and bioavailability of Lewis antigens in naturally occurring glycoconjugates [165]. Overexpression of mentioned fucosyltransferases and Lewis antigens has already been described in many types of cancer, including BC [167].
Neo-expression of the Lex antigen, which is absent in normal urothelium, has been noted in over 85% of urothelial carcinoma regardless of tumor stage and grade [168]. In addition, immunocytological detection of the Lex antigen on exfoliated bladder cells has improved the efficiency of BC detection, particularly of low-grade and low-stage neoplasms [169]. Currently, the Lex antigen is considered a marker of malignant transformation in BC, and its expression has been positively correlated with the stage, grade, and metastatic potential of a tumor [167]. Moreover, the decreased expression of the Lex antigen on human neutrophils has been associated with better prognosis in patients after bladder resection [170].
Furthermore, a recent report on cell lines from various stages of BC has shown that low-grade BC cell lines expressed elevated levels of the fucosylated Lex antigen, while normal bladder epithelial cells lack the expression of this antigen [171]. T24 (grade 3) and TCCSUP (grade 3) cells also lack the Lex antigen expression, whereas J82COT (grade 4) cells express low levels of Lex. These differences in the Lex antigen expression correlate with differences in mRNA expression levels of α1,3/4-fucosyltransferase genes. This suggests that the Lex antigens are promising candidate biomarkers for grading BC. However, the main difference between clinical and in vitro studies should be noted. While in BC patients, the expression of the Lex antigen has been found to increase gradually with cancer progression [167], the cell lines from the most advanced stages have shown a decrease in expression of the Lex antigen in comparison to those cells from earlier stages of the disease [171].
Moreover, Lewis antigen-mediated E-selectin adhesion has been shown to play a role in BC [172]. A titertray assay with immobilized selectins was used to test the adhesion of BC cells isolated from several tumors. The results showed a positive correlation between the number of bound BC cells isolated from several tumors and Lea antigen expression. It suggests that ligands for E-selectin on the surface of BC cells likely bind through α1,4-linked fucose to the penultimate GlcNAc.
Also, increased sialylation of cancer cells often results from the overexpression of the SLea antigen and the SLex antigen, which can be detected as terminal epitopes of N-glycans, O-glycans, and glycolipids [173]. Increased serum levels of the SLea antigen have been associated with higher stage/grade and increased invasiveness of BC. On the other hand, loss or decrease of the SLea antigen expression in tumor tissues has been associated with a higher atypia grade of the tumor. It has also been shown that expression of the sLex antigen is correlated with shorter 5-year and 7-year survival rates [163].
Finally, BC is a frequent health problem in rural areas of Africa and the Middle East, where Schistosoma haematobium is prevalent. It supports a connection between malignant transformation and infection by this parasite, and, importantly, the level of sialylated Lewis antigens is strictly connected to this phenomenon. Infection with S haematobium cause an increased expression of SLea in the majority of benign or pre-malignant BC lesions. It has been shown that S. haematobium was responsible for an increased expression of the SLea antigen in the majority of benign or pre-malignant BC lesions. In addition, S. Haematobium eggs show expression of SLea and SLex antigens that mimic the glycosylation of human leukocytes. This enables evasion from immune surveillance, easier parasite colonization, and cancer development. This consideration may be of critical value for the early identification of infected patients and for facilitating the development of non-invasive diagnostic tools [139].

10. Glycosylation of Extracellular Vesicles in Bladder Cancer

Extracellular vesicles (EVs) are small spherical structures surrounded by a phospholipid bilayer that are released into the intercellular space by almost all cell types [174]. EVs can be found circulating in various body fluids (e.g., blood, cerebrospinal fluid, sperm, saliva, and urine) or in conditioned cell culture media. Based on their size, biogenesis, and molecular composition, EVs are classified into exosomes (30–150 nm in diameter), ectosomes (100–1000 nm), and apoptotic bodies (>1 μm) [175]. The number of released EVs and their molecular composition undergoes constant changes depending on the current state of the parental cell. Bioactive molecules transferred via EVs can then modulate various biological processes in recipient cells, including gene expression, protein biosynthesis, the course of metabolic pathways, and intracellular signaling [176]. Moreover, tumor-derived EVs can participate in angiogenesis, metastasis, multi-drug resistance, apoptosis inhibition, and immunosuppression [177].
The amount of data on BC-derived EV glycosylation is, however, limited. As already mentioned in Section 8, biomarker combinations consisting of ITGA3 and fucose-recognizing UEA lectin using EVs can successfully discriminate BC patients from those with benign lesions and prostate cancer patients [164]. In our recent lectin-based study analyzing the expression of particular glycoepitopes in ectosomes released by normal urothelial HCV-29 and T-24 BC cells, we have observed that the amounts of β1,6-branched tri- and/or tetraantennary complex N-glycans as well α-2,6-linked sialic acids were higher in T-24-derived ectosomes than in HCV-29-derived ectosomes as assessed by immunoblotting [178]. That reflected relationships that were also present in the parental cells. However, flow cytometry analysis has shown that only the levels of ectosomal bisecting GlcNAc and α2,3-linked sialic acids could be considered possible diagnostic BC glycobiomarkers due to their greater amount in T-24-derived ectosomes. Regardless, our study has shown that the glycosylation status of T-24 cells and HCV-24 cells was responsible for the effect that ectosomes released by them had on the viability and motility of the recipient cells. In another study, the antitumor effects of exosomes derived from genetically modified T-24 BC cells expressing glycosyl-phosphatidylinositol-anchored interleukin 2 (GPI-IL-2) have been investigated [179]. Exosomes expressing GPI-IL-2 have been demonstrated to induce the proliferation of T cells and enhance the antigen-specific cytotoxic T-lymphocyte immune response, contrary to exosomes from cells not expressing GPI-IL-2. It suggests that modification of EV cargo with various types of glycans (herein GPI anchor) could be an interesting approach in exosome-based cancer immunotherapy.

11. Proteoglycans in Bladder Cancer

The transitional epithelium of the bladder is lined with a layer of glycocalyx composed of glycosaminoglycans (GAGs) and other molecules. GAGs ensure impermeability in the bladder [180], and the damage to the GAG layers may lead to penetration of urine, toxic agents, and bacteria into the deeper layers of the bladder. GAGs represent a type of linear, large heteropolysaccharides. Based on the chemical structure of their repeating disaccharide units that consist of an amino sugar (D-glucosamine that is N-acetylated or N-sulphated, or N-acetyl-d-galactosamine) and either uronic acid (D-glucuronic acid or L-iduronic acid) or galactose, GAGs are classified into different categories: hyaluronic acid (HA), heparin/heparan sulfate, chondroitin sulfate, dermatan sulfate, and keratan sulfate (Figure 7). One or more GAG chains, except for HA, are covalently linked via a tetrasaccharide linkage to the core protein to form highly polyanionic proteoglycans due to the additional sulfate groups present on most of the units.
By acting directly on cellular receptors or through interactions with growth factors, GAGs can regulate multiple cellular processes and are considered essential in cancer biology. Abnormal forms (with altered chain sulphation patterns, molecular size, composition, etc.) and distribution of GAGs have been reported in various cancers. Their influence on the proliferation, apoptosis, adhesion, migration, invasion, and hematogenous metastasis of cancer cells, as well as the induction of angiogenesis and immunosuppression in the tumor microenvironment, have been well documented [181,182,183,184]. Several potential applications of GAGs in cancer treatment have also been reported [185,186].
GAG content is elevated in BC compared to normal urothelium, and its composition is also affected by tumor stage and grade [187,188]. Hyaluronic acid and dermatan sulfate were found to be decreased, while chondroitin sulfate to be increased in invasive cancers. The decrease in the percentage of heparan sulfate was closely correlated with higher-grade tumors. Urinary GAG excretion was also found to increase in parallel with tumor size, stage, and grade [189] and to help follow-up BC patients after transurethral resection [190]. These early observations were confirmed by more recent studies in which tumor grade and stage directly correlated with GAG levels in tissue samples; however, GAG levels did not predict tumor recurrence [191].
In contrast, urinary levels of HA and its degrading enzyme, hyaluronidase (HAase), have been shown to be elevated in BC patients and were indicated as markers for monitoring tumor recurrence and screening (HA-HAase test) [192,193,194]. HA is the major ligand for CD44, which is a multifunctional mediator of cancer progression involved in promoting proliferation, motility, invasiveness, angiogenesis, and chemoresistance [195,196,197]. Interestingly, CD44 expression has been shown to be significantly associated with tumor aggressiveness in BC [117].
HA is synthesized by hyaluronic acid synthase (HAS), which has three isoforms: HAS1, HAS2, and HAS3. In BC cells and tissues, transcript and protein levels of HAS1 were elevated and correlated with bladder tumor recurrence and response to treatment [198,199]. In in vivo and in vitro models, downregulation of HAS1 expression was shown to reduce growth, invasion, and angiogenesis of BC through induction of apoptosis [200].
Similarly, HAS2 and HAS3 have also been shown to promote tumor growth and metastasis [199,201,202]. In addition, HAS2 expression negatively correlated with amylo-α-1,6-glucosidase-4-α-glucanotransferase (AGL), one of the enzymes catalyzing glycogenolysis, which has been described as a tumor growth suppressor and prognostic marker in human BC [203]. It has been shown that patients with high mRNA expression of HAS2 and low AGL following induction of HA synthesis had poor survival rates [204]. Two major HA-binding cell surface proteins, namely CD44 and Hyaluronan Mediated Motility Receptor (RHAMM), are crucial for the rapid growth of BC driven by the loss of AGL [205]. Loss of either CD44 or RHAMM induces apoptosis in BC cell lines with low AGL. In addition, increased expression of HYAL-1 type HAase has been found in BC tissues and has been indicated as an independent predictor of muscle invasion [199,206]. Finally, a thorough meta-analysis has shown that both HA and HAases can be used as biomarkers in the diagnosis of BC [207].
Also, another GAG, chondroitin sulfate, has been found to inhibit the growth of BC cells [208]. One of the carriers of chondroitin sulfate is decorin, a small leucine-rich class 1 proteoglycan consisting of a core glycoprotein with a molecular weight of 40 kDa and a single GAG chain, chondroitin sulfate or dermatan sulfate, which is bound to the N-terminus of the core protein. Decorin functions as a natural inhibitor of pan-tyrosine kinase and negatively regulates tumor growth [209,210]. In BC tissues, the expression of decorin was downregulated compared to that in normal tissue [211,212]. In various cancers, low expression of decorin correlated with poor patient prognosis, lymph node metastasis, and weak response to treatment [210]. The introduction of decorin into BC cells not expressing this protein upregulated TGF-β1 and MMP2 expression via p21 protein, promoted apoptosis and adhesion, and inhibited bladder cancer cell proliferation and metastasis [212,213]. On the other hand, biglycan, which is another small leucine-rich proteoglycan, has been shown to suppress BC cell proliferation and tumor growth [214]. Biglycan levels have been found to increase in advanced stages of cancer in tissue samples and be associated with increased survival of BC patients.
Interestingly, bladder cancer cells have also been shown to express a distinct form of chondroitin sulfate, namely oncofetal chondroitin sulfate (ofCS), which is normally confined to the placenta [215,216]. The carriers of ofCS are syndecan 1 (SDC1) and chondroitin sulfate proteoglycan 4 (CSPG4), proteoglycans that are highly expressed in BC and thus increase the overall amount of ofCS in the tumor [216]. It has been demonstrated that ofCS modulates tumor cell motility by affecting canonical integrin signaling pathways [217]. The presence of urinary ofCS has been shown to correlate with grade and tumor size [218], as well as with resistance to chemotherapy and poorer survival of BC patients [216].

12. Manipulation of Glycome as a Therapeutic Tool in Bladder Cancer

As already mentioned in Section 10, genetic engineering gives the possibility to manipulate glycosylation profiles of various cells that can be used in cancer therapy. For instance, multiple carcinomas express extracellular N-glycans, the presence of which negatively correlates with chimeric antigen receptor T (CAR T) cell killing. The knockout GnT-V in pancreatic adenocarcinoma revealed that N-glycans protect tumors from CAR T cell killing by interfering with proper immunological synapse formation and cytokine production leading to decreased cytotoxicity [219]. In the same study, treatment with glucose/mannose analog 2-deoxy-d-glucose (2DG) disrupted the N-glycan cover on tumor cells and enhanced CAR T cell activity. Moreover, 2DG treatment interfered with the PD-1–PD-L1 signaling pathway and increased the survival of tumor-infiltrating CAR T cells in vivo [219]. Therefore, manipulation of the glycosylation profile should be considered in the further development of CAR T therapy, also in the treatment of BC.
Furthermore, core 1 synthase glycoprotein-N-acetylgalactosamine 3-β-galactosyltransferase 1 (C1GALT1) is the key enzyme in the conversion of Tn antigen to T antigen, and its upregulation is observed in BC. Silencing of C1GALT1 suppressed the migratory ability and proliferation of BC YTS-1 cells in vitro and in vivo in YTS-1 inoculated mice [143]. Moreover, miR-1-3p, typically downregulated in urothelial carcinoma, inhibits the transcription of C1GALT1. Artificially induced miR-1-3p overexpression in YTS-1 cells has resulted in their decreased migration and proliferation. It suggests that targeting glycosylation-related enzymes with miRNAs should be considered when designing novel therapies for BC [143].
Another study showed that also manipulation of O-GlcNAc levels might help in disease containment. Melatonin treatment was shown to decrease O-GlcNAc levels which are typically elevated in BC. Melatonin decreased the amount of O-GlcNAc by reducing glucose uptake and availability of GlcNAc donor, i.e., UDP-GlcNAc. This treatment inhibited the proliferation and migration of BC cells through the inhibition of cyclin-dependent-like kinase 5 (CDK5) expression and O-GlcNAcylation of CDK5 [62].
Finally, specific glycan modification may be used to improve the drug delivery process to BC tumors. Intravesical injection of chemotherapeutic drugs such as epirubicin (EPI) is routinely used after transurethral tumor resection. Unfortunately, EPI lacks tumor selectivity and often leads to damage of normal bladder urothelium and other side effects [100]. Because mannose is the most aberrantly expressed glycan on the surface of BC cells, the lectin-drug conjugate has been created by linking concanavalin A (ConA) (mannose-specific lectin) with EPI. This conjugate showed selective cytotoxicity to cancer cells without attacking normal bladder tissues [100].

13. Conclusions and Future Perspectives

In this review, we have discussed the role of aberrant glycosylation in BC development and progression. Presented alterations facilitate metastasis formation, immunosuppression, drug and apoptosis resistance, and angiogenesis during tumor progression. A wide range of alterations in glycosylation are observed in the advanced stages of BC. Therefore, it is crucial to compare full glycosylation profiles of early- and advanced-stages of BC to determine how it changes during disease progression.
Nevertheless, the structural complexity of glycans, heterogeneity in glycosylation sites, and the role of epigenetic and environmental factors in glycan structures make a complete characterization of BC glycome a challenge. However, the development of high-performance liquid chromatography (HPLC) and mass spectrometry (MS) enables detailed structural analysis of glycoconjugates. These advanced techniques can be used to analyze not only samples from cell cultures but also tissue and urine samples from BC patients. However, it should be noted that lectins and antibodies are still commonly used to analyze glycosylation, considering the costs of more sophisticated analytical techniques. Additionally, the creation of a BC-related glycosylation database would be useful to analyze the observed alternations of glycan structures, mutations in glycan-related genes, and the role of glycoconjugates in cancer progression.
Recently, several large-scale cancer-related glycoproteomic studies on intact glycopeptides emerged. Analysis of intact glycopeptides instead of isolated glycans allows the identification of glycosylated proteins, including those expressing TACAs. Moreover, such studies can precisely determine glycosylation sites within proteins/peptides. It also creates diagnostic and/or therapeutic opportunities, for instance, involving genetic engineering. Elimination or introduction of particular glycosylation sites could make it possible to regulate the function of chosen carcinogenesis-related proteins by altering their glycosylation status. It has already been shown for ovarian cancer that integrated glycoproteomic analysis of tumor tissues enables subtyping (clustering) and outcome prediction for high-grade tumors [220,221]. Nevertheless, glycoproteomic strategies still battle some technical limitations. Glycopeptide signals are more difficult to recognize than signals from naked peptides due to their microheterogeneity. Moreover, working in the data-dependent acquisition (DDA) is limited to the MS2 level of tandem MS1/MS2 analysis. It is impossible to identify all glycoforms present in the sample using only MS1 scans due to the high peak intensities. However, recently, efficient software tools have been developed. One of them is pGlycoQuant, which applies a deep learning model that reduces missing values by 19–89% [222].
Targeting aberrant glycosylation and using it as a diagnostic and therapeutic tool in BC is strongly supported by studies presented in this review. For instance, the use of glycosyltransferase inhibitors or glycosidases in BC treatment could limit the side effects of classical chemotherapy, especially affecting the kidneys and liver. Moreover, BC cell lines derived from invasive tumors displayed increased expression of FUT6 and FUT7. It gives an opportunity to target not only glycosylation-related enzymes in the therapy of BC but also in diagnosis. Diagnosis of BC still relies on invasive methods such as cystoscopy and TURBT. Therefore, focusing on TACAs differential expression may lead to the establishment of new diagnostic tools with higher specificity for early detection, prognosis, and grading BC.
In summary, glycosylation status plays a vital role in BC progression (Figure 8). In recent years we observed incredible progress in understanding the effects caused by altered glycosylation. Numerous studies have focused on glycans as promising biomarkers and therapeutic tools in BC, highlighting the importance of glycobiology in cancer-related studies.

Author Contributions

Conceptualization, M.P. and M.W.; writing—original draft preparation, M.P., M.W. and M.S; writing—review and editing, M.P., M.W. and M.S.; funding acquisition, M.P. All authors have read and agreed to the published version of the manuscript.

Funding

The writing of this manuscript was supported by a grant from Jagiellonian University. Poland (N18/DBS/000007).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

β4GalT1β1:4-galactosyltransferase-1
2DG2-deoxy-d-glucose
ADCCantibody-dependent cellular cytotoxicity
AGPα-1-Acid glycoprotein
BCBladder cancer
BCGBacillus Calmette-Guérin
C1GnTCore 1 β1,3-galactosyltransferase
CAR TChimeric antigen receptor T
CD44Cluster of Differentiation 44
CDK5Cyclin-dependent-like kinase 5
CRTCalreticulin
CSCsCancer stem cells
CTCsCirculating tumor cells
CTLsCytotoxic T-lymphocytes
DCsDendritic cells
ECMExtracellular matrix
EGFEpidermal growth factor
EMTEpithelial-mesenchymal transition
EPIEpirubicin
EVsExtracellular vesicles
FUTsFucosyltransferases
GlcNAcN-acetylglucosamine
GnT IIIN-acetylglucosaminyltransferase III
GnT IVN-acetylglucosaminyltransferase IV
GnT VN-acetylglucosaminyltransferase V
GPIGlycosylphosphatidylinositol
GPI-IL-2Glycosyl-phosphatidylinositol-anchored interleukin 2
HLAHuman leukocyte antigen
IgsImmunoglobulins
ITGA3α3 integrin subunit
LeaLewisa
LebLewisb
LexLewisx
LeyLewisy
MIBCMuscle-infiltrating bladder cancer
Neu5AcN-acetyl-neuraminic acid
Neu5GcN-glycolyl-neuraminic acid
NKNatural killer
NMIBCNon-muscle-infiltrating bladder cancer
OGTO-GlcNAc transferase
PHALPhaseolus vulgaris leucoagglutinin
SiaSialic acids
SLexsialyl-Lewis X
ST antigenSialilated T antigen
ST6GALNAC-IST6 N-acetylgalactosaminide α-2,6-sialyltransferase 1
TACATumor-associated carbohydrate antigens
TGFβTransforming growth factor β
TURBTTransurethral resection of bladder tumor
UCUrothelial carcinomas
UEAUlex europaeus agglutinin

References

  1. Saginala, K.; Barsouk, A.; Aluru, J.S.; Rawla, P.; Padala, S.A.; Barsouk, A. Epidemiology of bladder cancer. Med. Sci. 2020, 8, 15. [Google Scholar] [CrossRef] [Green Version]
  2. 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, 7, 209–249. [Google Scholar] [CrossRef]
  3. Chalasani, V.; Chin, J.L.; Izawa, J.I. Histologic variants of urothelial bladder cancer and nonurothelial histology in bladder cancer. Can. Urol. Assoc. J. 2009, 3 (Suppl. 4), S193–S198. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Yousef, P.G.; Gabril, M.Y. An update on the molecular pathology of urinary bladder tumors. Pathol. Res. Pract. 2017, 214, 1–6. [Google Scholar] [CrossRef]
  5. Nadal, R.; Bellmunt, J. Management of metastatic bladder cancer. Cancer Treat. Rev. 2019, 76, 10–21. [Google Scholar] [CrossRef]
  6. Sternberg, C.N.; Bellmunt, J.; Sonpavde, G.; Siefker-Radtke, A.O.; Stadler, W.M.; Bajorin, D.F.; Dreicer, R.; George, D.J.; Milowsky, M.I.; Theodorescu, D.; et al. ICUD-EAU International Consultation on Bladder Cancer 2012: Chemotherapy for urothelial carcinoma-neoadjuvant and adjuvant settings. Eur. Urol. 2013, 63, 58–66. [Google Scholar] [CrossRef]
  7. Kim, Y.J.; Kim, W.J. Can we use methylation markers as diagnostic and prognostic indicators for bladder cancer? Investig. Clin. Urol. 2016, 57 (Suppl. 1), S77–S88. [Google Scholar] [CrossRef] [PubMed]
  8. Flynn, R.A.; Pedram, K.; Malaker, S.A.; Batista, P.J.; Smith, B.A.H.; Johnson, A.G.; George, B.M.; Majzoub, K.; Villalta, P.W.; Carette, J.E.; et al. Small RNAs are modified with N-glycans and displayed on the surface of living cells. Cells 2021, 184, 3109–3124.e22. [Google Scholar] [CrossRef] [PubMed]
  9. Varki, A. Biological roles of glycans. Glycobiology 2017, 27, 3–49. [Google Scholar] [CrossRef] [Green Version]
  10. Stambuk, T.; Klasic, M.; Zoldos, V.; Lauc, G. N-glycans as functional effectors of genetic and epigenetic disease risk. Mol. Asp. Med. 2021, 79, 100891. [Google Scholar] [CrossRef]
  11. Lauc, G.; Krištić, J.; Zoldoš, V. Glycans—The third revolution in evolution. Front. Genet. 2014, 5, 145. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Kristic, J.; Zaytseva, O.O.; Ram, R.; Nguyen, Q.; Novokmet, M.; Vuckovic, F.; Vilaj, M.; Trbojevic-Akmacic, I.; Pezer, M.; Davern, K.M.; et al. Profiling and genetic control of the murine immunoglobulin glycome. Nat. Chem. Biol. 2018, 14, 516–524. [Google Scholar] [CrossRef]
  13. Reis, C.A.; Osorio, H.; Silva, L.; Gomes, C.; David, L. Alterations in glycosylation as biomarkers for cancer detection. J. Clin. Pathol. 2010, 63, 322–329. [Google Scholar] [CrossRef] [PubMed]
  14. Hoja-Łukowicz, D.; Przybyło, M.; Duda, M.; Pocheć, E.; Bubka, M. On the trail of the glycan codes stored in cancer-related cell adhesion proteins. Biochim. Biophys. Acta Gen. Subj. 2017, 1861, 3237–3257. [Google Scholar] [CrossRef] [PubMed]
  15. Pinho, S.S.; Reis, C.A. Glycosylation in cancer: Mechanisms and clinical implications. Nat. Rev. Cancer 2015, 15, 540–555. [Google Scholar] [CrossRef]
  16. RodrÍguez, E.; Schetters, S.; van Kooyk, Y. The tumour glyco-code as a novel immune checkpoint for immunotherapy. Nat. Rev. Immunol. 2018, 18, 204–211. [Google Scholar] [CrossRef]
  17. Mereiter, S.; Balmaña, M.; Campos, D.; Gomes, J.; Reis, C.A. Glycosylation in the era of cancer-targeted therapy: Where are we heading? Cancer Cell 2019, 36, 6–16. [Google Scholar] [CrossRef]
  18. Drickamer, K.; Taylor, M.E. Introduction to Glycobiology, 2nd ed.; Oxford University Press: New York, NY, USA, 2006. [Google Scholar]
  19. Stanley, P.; Moremen, K.W.; Lewis, N.E.; Taniguchi, N.; Aebi, M. N-glycans. In Essentials of Glycobiology, 4th ed.; Varki, A., Cummings, R.D., Esko, J.D., Stanley, P., Hart, G.W., Aebi, M., Mohnen, D., Kinoshita, T., Packer, N.H., Prestegard, J.H., et al., Eds.; Cold Spring Harbor (NY): Cold Spring Harbor Laboratory Press: Long Island, NY, USA, 2022; Chapter 9. [Google Scholar] [CrossRef]
  20. Zoqlam, R.; Lazauskaite, S.; Glickman, S.; Zaitseva, L.; Ilie, P.C.; Qi, S. Emerging molecular mechanisms and genetic targets for developing novel therapeutic strategies for treating bladder diseases. Eur. J. Pharm. Sci. 2022, 173, 106167. [Google Scholar] [CrossRef]
  21. Miyoshi, E.; Terao, M.; Kamada, Y. Physiological roles of N-acetylglucosaminyltransferase V (GnT-V) in mice. BMB Rep. 2012, 45, 554–559. [Google Scholar] [CrossRef] [Green Version]
  22. Nagae, M.; Kizuka, Y.; Mihara, E.; Kitago, Y.; Hanashima, S.; Ito, Y.; Takagi, J.; Taniguchi, N.; Yamaguchi, Y. Structure and mechanism of cancer-associated N-acetylglucosaminyltransferase-V. Nat. Commun. 2018, 9, 3380. [Google Scholar] [CrossRef] [Green Version]
  23. Bellis, S.L.; Reis, C.A.; Varki, A.; Kannagi, R.; Stanley, P. Glycosylation Changes in Cancer. In Essentials of Glycobiology, 4th ed.; Varki, A., Cummings, R.D., Esko, J.D., Stanley, P., Hart, G.W., Aebi, M., Mohnen, D., Kinoshita, T., Packer, N.H., Prestegard, J.H., et al., Eds.; Cold Spring Harbor (NY): Cold Spring Harbor Laboratory Press: Long Island, NY, USA, 2022; Chapter 47. [Google Scholar] [CrossRef]
  24. Kizuka, Y.; Taniguchi, N. Enzymes for N-glycan branching and their genetic and nongenetic regulation in cancer. Biomolecules 2016, 6, 25. [Google Scholar] [CrossRef] [Green Version]
  25. Bastian, K.; Scott, E.; Elliott, D.J.; Munkley, J. FUT8 alpha-(1,6)-fucosyltransferase in cancer. Int. J. Mol. Sci. 2021, 22, 455. [Google Scholar] [CrossRef] [PubMed]
  26. Link-Lenczowski, P.; Bubka, M.; Balog, C.I.A.; Koeleman, C.A.M.; Butters, T.D.; Wuhrer, M.; Lityńska, A. The glycomic effect of N-acetylglucosaminyltransferase III overexpression in metastatic melanoma cells. GnT-III modifies highly branched N-glycans. Glycoconj. J. 2018, 35, 217–231. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Garnham, R.; Scott, E.; Livermore, K.E.; Munkley, J. ST6GAL1: A key player in cancer. Oncol. Lett. 2019, 20, 983–989. [Google Scholar] [CrossRef] [Green Version]
  28. Fardini, Y.; Dehennaut, V.; Lefebvre, T.; Issad, T. O-GlcNAcylation: A new cancer hallmark? Front. Endocrinol. 2013, 4, 99. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Hirata, T.; Nagae, M.; Osuka, R.F.; Mishra, S.K.; Yamada, M.; Kizuka, Y. Recognition of glycan and protein substrates by N-acetylglucosaminyltransferase-V. Biochim. Biophys. Acta Gen. Subj. 2020, 1864, 129726. [Google Scholar] [CrossRef]
  30. Osuka, R.F.; Hirata, T.; Nagae, M.; Nakano, M.; Shibata, H.; Okamoto, R.; Kizuka, Y. N-acetylglucosaminyltransferase-V requires a specific noncatalytic luminal domain for its activity toward glycoprotein substrates. J. Biol. Chem. 2022, 298, 101666. [Google Scholar] [CrossRef]
  31. Demetriou, M.; Granovsky, M.; Quaggin, S.; Dennis, J.W. Negative regulation of T-cell activation and autoimmunity by Mgat5 N-glycosylation. Nature 2001, 409, 733–739. [Google Scholar] [CrossRef]
  32. Taniguchi, N.; Kizuka, Y. Glycans and cancer: Role of N-Glycans in cancer biomarker, progression and metastasis, and therapeutics. Adv. Cancer Res. 2015, 126, 11–51. [Google Scholar] [CrossRef]
  33. Zhao, Y.; Sato, Y.; Isaji, T.; Fukuda, T.; Matsumoto, A.; Miyoshi, E.; Gu, J.; Taniguchi, N. Branched N-glycans regulate the biological functions of integrins and cadherins. FEBS J. 2008, 275, 1939–1948. [Google Scholar] [CrossRef]
  34. Zhao, Y.Y.; Takahashi, M.; Gu, J.G.; Miyoshi, E.; Matsumoto, A.; Kitazume, S.; Taniguchi, N. Functional roles of N-glycans in cell signaling and cell adhesion in cancer. Cancer Sci. 2008, 99, 1304–1310. [Google Scholar] [CrossRef]
  35. Przybyło, M.; Lityńska, A.; Pocheć, E. Different adhesion and migration properties of human HCV29 non-malignant urothelial and T24 bladder cancer cells: Role of glycosylation. Biochimie 2005, 87, 133–142. [Google Scholar] [CrossRef] [PubMed]
  36. Olden, K.; Breton, P.; Grzegorzewski, K.; Yasuda, Y.; Gause, B.L.; Oredipe, O.A.; Newton, S.A.; White, S.L. The potential importance of swainsonine in therapy of cancer and immunology. Pharmacol. Ther. 1991, 50, 285–290. [Google Scholar] [CrossRef]
  37. Lityńska, A.; Przybyło, M.; Pocheć, E.; Laidler, P. Adhesion properties of human bladder cell lines with extracellular matrix components: The role of integrins and glycosylation. Acta Biochim. Pol. 2002, 49, 643–650. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Pocheć, E.; Lityńska, A.; Bubka, M.; Amoresano, A.; Casbarra, A. Characterization of the oligosaccharide component of α3β1 integrin from human bladder carcinoma cell line T24 and its role in adhesion and migration. Eur. J. Cell Biol. 2006, 85, 47–57. [Google Scholar] [CrossRef] [PubMed]
  39. Lityńska, A.; Przybyło, M.; Ksiazek, D.; Laidler, P. Differences of α3β1 integrin glycans from different human bladder cell lines. Acta Biochim. Pol. 2000, 47, 427–434. [Google Scholar] [CrossRef] [Green Version]
  40. Laidler, P.; Gil, D.; Pituch-Noworolska, A.; Ciołczyk, D.; Ksiazek, D.; Przybyło, M.; Lityńska, A. Expression of β1-integrins and N-cadherin in bladder cancer and melanoma cell lines. Acta Biochim. Pol. 2000, 47, 1159–1170. [Google Scholar] [CrossRef]
  41. Przybyło, M.; Hoja-Lukowicz, D.; Litynska, A.; Laidler, P. Different glycosylation of cadherins from human bladder non-malignant and cancer cell lines. Cancer Cell Int. 2002, 2, 6. [Google Scholar] [CrossRef] [PubMed]
  42. Ishimura, H.; Takahashi, T.; Nakagawa, H.; Nishimura, S.I.; Arai, Y.; Horikawa, Y.; Habuchi, T.; Miyoshi, E.; Kyan, A.; Hagisawa, S.; et al. N-acetylglucosaminyltransferase V and β1-6 branching N-linked oligosaccharides are associated with good prognosis of patients with bladder cancer. Clin. Cancer Res. 2006, 12, 2506–2511. [Google Scholar] [CrossRef] [Green Version]
  43. Takahashi, T.; Hagisawa, S.; Yoshikawa, K.; Tezuka, F.; Kaku, M.; Ohyama, C. Predictive value of N-acetylglucosaminyltransferase-V for superficial bladder cancer recurrence. J. Urol. 2006, 175, 90–93. [Google Scholar] [CrossRef] [PubMed]
  44. Guo, J.M.; Zhang, X.Y.; Chen, H.L.; Wang, G.M.; Zhang, Y.K. Structural alterations of sugar chains in urine fibronectin from bladder cancer patients and its enzymatic mechanism. J. Cancer Res. Clin. Oncol. 2001, 127, 512–519. [Google Scholar] [CrossRef] [PubMed]
  45. Karavana, S.Y.; Şenyiğit, Z.A.; Çalışkan, C.; Sevin, G.; Özdemir, D.I.; Erzurumlu, Y.; Şen, S.; Baloğlu, E. Gemcitabine hydrochloride microspheres used for intravesical treatment of superficial bladder cancer: A comprehensive in vitro/ex vivo/in vivo evaluation. Drug Des. Dev. Ther. 2018, 12, 1959–1975. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Tang, Y.; Cong, X.; Wang, S.; Fang, S.; Dong, X.; Yuan, Y.; Fan, J. GnT-V promotes chemosensitivity to gemcitabine in bladder cancer cells through β1,6 GlcNAc branch modification of human equilibrative nucleoside transporter 1. Biochem. Biophys. Res. Commun. 2018, 503, 3142–3148. [Google Scholar] [CrossRef] [PubMed]
  47. Stanley, P. Biological consequences of overexpressing or eliminating N-acetylglucosaminyltransferase-TIII in the mouse. Biochim. Biophys. Acta Gen. Subj. 2002, 1573, 363–368. [Google Scholar] [CrossRef] [PubMed]
  48. Takahashi, M.; Hasegawa, Y.; Maeda, K.; Kitano, M.; Taniguchi, N. Role of glycosyltransferases in carcinogenesis; growth factor signaling and EMT/MET programs. Glycoconj. J. 2022, 39, 167–176. [Google Scholar] [CrossRef]
  49. Tanaka, T.; Yoneyama, T.; Noro, D.; Imanishi, K.; Kojima, Y.; Hatakeyama, S.; Tobisawa, Y.; Mori, K.; Yamamoto, H.; Imai, A.; et al. Aberrant N-glycosylation profile of serum immunoglobulins is a diagnostic biomarker of urothelial carcinomas. Int. J. Mol. Sci. 2017, 18, 2632. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Yang, X.; Qian, K. Proteins O-GlcNAcylation: Emerging mechanisms and functions. Nat. Rev. Mol. Cell Biol. 2017, 18, 452–465. [Google Scholar] [CrossRef] [Green Version]
  51. Hart, G.W. Nutrient regulation of signaling and transcription. J. Biol. Chem. 2019, 294, 2211–2231. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Mueller, T.; Ouyang, X.; Johnson, M.S.; Qian, W.J.; Chatham, J.C.V.; Darley-Usmar, V.; Zhang, J. New insights into the biology of protein o-glcnacylation: Approaches and observations. Front. Aging 2021, 1, 620382. [Google Scholar] [CrossRef]
  53. Elbatrawy, A.A.; Kim, E.J.; Nam, G. O-GlcNAcase: Emerging Mechanism, Substrate Recognition and Small-Molecule Inhibitors. Chem. Med. Chem. 2020, 14, 1244–1257. [Google Scholar] [CrossRef]
  54. Hart, G.W.; Slawson, C.; Ramirez-Correa, G.; Lagerlof, O. Cross talk between O-GlcNAcylation and phosphorylation: Roles in signaling, transcription, and chronic disease. Annu. Rev. Biochem. 2011, 80, 825–858. [Google Scholar] [CrossRef] [Green Version]
  55. Bond, M.R.; Hanover, J.A. A little sugar goes a long way: The cell biology of O-GlcNAc. J. Cell Biol. 2015, 208, 869–880. [Google Scholar] [CrossRef] [Green Version]
  56. Nie, H.; Yi, W. O-GlcNAcylation, a sweet link to the pathology of diseases. J. Zhejiang Univ. Sci. B 2019, 20, 437–448. [Google Scholar] [CrossRef] [PubMed]
  57. Lee, J.B.; Pyo, K.H.; Kim, H.R. Role and Function of O-GlcNAcylation in Cancer. Cancers 2021, 13, 5365. [Google Scholar] [CrossRef] [PubMed]
  58. Lu, Q.; Zhang, X.; Liang, T.; Bai, X. O-GlcNAcylation: An important post-translational modification and a potential therapeutic target for cancer therapy. Mol. Med. 2022, 28, 115. [Google Scholar] [CrossRef] [PubMed]
  59. Rozanski, W.; Krzeslak, A.; Forma, E.; Brys, M.; Blewniewski, M.; Wozniak, P.; Lipinski, M. Prediction of bladder cancer based on urinary content of MGEA5 and OGT mRNA level. Clin. Lab. 2012, 58, 579–583. [Google Scholar] [PubMed]
  60. Wang, L.; Chen, S.; Zhang, Z.; Zhang, J.; Mao, S.; Zheng, J.; Xuan, Y.; Liu, M.; Cai, K.; Zhang, W.; et al. Suppressed OGT expression inhibits cell proliferation while inducing cell apoptosis in bladder cancer. BMC Cancer 2018, 18, 1141. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  61. Jin, L.; Lu, M.H.; Dai, G.C.; Yao, Q.; Xiang, H.; Wang, L.X.; Xue, B.X.; Liu, X. O-GlcNAcylation promotes malignant phenotypes of bladder cancer cells. Neoplasma 2020, 67, 880–888. [Google Scholar] [CrossRef] [Green Version]
  62. Wu, J.; Tan, Z.; Li, H.; Lin, M.; Jiang, Y.; Liang, L.; Ma, O.; Gou, J.; Ning, L.; Li, X.; et al. Melatonin reduces proliferation and promotes apoptosis of bladder cancer cells by suppressing O-GlcNAcylation of cyclin-dependent-like kinase 5. J. Pineal Res. 2021, 71, e12765. [Google Scholar] [CrossRef]
  63. Jin, L.; Yuan, F.; Dai, G.; Yao, Q.; Xiang, H.; Wang, L.; Xue, B.; Shan, Y.; Liu, X. Blockage of O-linked GlcNAcylation induces AMPK-dependent autophagy in bladder cancer cells. Cell. Mol. Biol. Lett. 2020, 5, 17. [Google Scholar] [CrossRef]
  64. Chen, Y.; Liu, J.; Zhang, W.; Kadier, A.; Wang, R.; Zhang, H.; Yao, X. O-GlcNAcylation enhances NUSAP1 stability and promotes bladder cancer aggressiveness. Onco. Targets Ther. 2021, 14, 445–454. [Google Scholar] [CrossRef]
  65. Lee, H.W.; Kang, M.J.; Kwon, Y.J.; Nansa, S.A.; Jung, E.H.; Kim, S.H.; Lee, S.J.; Jeong, K.C.; Kim, Y.; Cheong, H.; et al. Targeted inhibition of O-linked β-N-acetylglucosamine transferase as a promising therapeutic strategy to restore chemosensitivity and attenuate aggressive tumor traits in chemoresistant urothelial carcinoma of the bladder. Biomedicines 2021, 10, 1162. [Google Scholar] [CrossRef]
  66. Harduin-Lepers, A.; Krzewinski-Recchi, M.A.; Colomb, F.; Foulquier, F.; Groux-Degroote, S.; Delannoy, P. Sialyltransferases functions in cancers. Front. Biosci. 2012, 4, 499–515. [Google Scholar] [CrossRef] [PubMed]
  67. Li, Y.; Chen, X. Sialic acid metabolism and sialyltransferases: Natural functions and applications. Appl. Microbiol. Biotechnol. 2012, 94, 887–905. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Miyagi, T.; Takahashi, K.; Hata, K.; Shiozaki, K.; Yamaguchi, K. Sialidase significance for cancer progression. Glycoconj. J. 2012, 29, 567–577. [Google Scholar] [CrossRef] [PubMed]
  69. Vajaria, B.N.; Patel, K.R.; Begum, R.; Patel, P.S. Sialylation: An avenue to target cancer cells. Pathol. Oncol. Res. 2016, 22, 443–447. [Google Scholar] [CrossRef]
  70. Kolasińska, E.; Przybyło, M.; Janik, M.; Lityńska, A. Towards understanding the role of sialylation in melanoma progression. Acta Biochim. Pol. 2016, 63, 533–541. [Google Scholar] [CrossRef]
  71. Bordron, A.; Bagacean, C.; Mohr, A.; Tempescul, A.; Bendaoud, B.; Deshayes, S.; Dalbies, F.; Buors, C.; Saad, H.; Berthou, C.; et al. Resistance to complement activation, cell membrane hypersialylation and relapses in chronic lymphocytic leukemia patients treated with rituximab and chemotherapy. Oncotarget 2018, 9, 31590–31605. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Li, F.; Ding, J. Sialylation is involved in cell fate decision during development, reprogramming and cancer progression. Protein Cell 2019, 10, 550–565. [Google Scholar] [CrossRef] [PubMed]
  73. Dobie, C.; Skropeta, D. Insights into the role of sialylation in cancer progression and metastasis. Br. J. Cancer 2021, 124, 76–90. [Google Scholar] [CrossRef]
  74. Xu, C.; Wang, S.; Wu, Y.; Sun, X.; Yang, D.S. Recent advance in understanding the roles of sialyltransferases in tumor angiogenesis and metastasis. Glycoconj. J. 2021, 38, 119–127. [Google Scholar] [CrossRef]
  75. Dalangood, S.; Zhu, Z.; Ma, Z.; Li, J.; Zeng, Q.; Yan, Y.; Shen, B.; Yan, J.; Huang, R. Identification of glycogene-type and validation of ST3GAL6 as a biomarker predicts clinical outcome and cancer cell invasion in urinary bladder cancer. Theranostics 2020, 10, 10078–10091. [Google Scholar] [CrossRef] [PubMed]
  76. Chan, K.S.; Volkmer, J.P.; Weissman, I. Cancer stem cells in bladder cancer: A revisited and evolving concept. Curr. Opin. Urol. 2010, 20, 393–397. [Google Scholar] [CrossRef] [Green Version]
  77. Chen, D.; Li, D.; Cui, Z.; Zhang, C.; Zhang, Z.; Yan, L. Evaluation of the value of preoperative sialic acid levels in diagnosis and localization of urothelial tumors. J. Cancer 2021, 12, 5066–5075. [Google Scholar] [CrossRef] [PubMed]
  78. Bennett, E.P.; Mandel, U.; Clausen, H.; Gerken, T.A.; Fritz, T.A.; Tabak, L.A. Control of mucin-type O-glycosylation: A classification of the polypeptide GalNAc-transferase gene family. Glycobiology 2012, 22, 736–756. [Google Scholar] [CrossRef] [Green Version]
  79. Dhanisha, S.S.; Guruvayoorappan, C.; Drishya, S.; Abeesh, P. Mucins: Structural diversity, biosynthesis, its role in pathogenesis and as possible therapeutic targets. Crit. Rev. Oncol. Hematol. 2018, 122, 98–122. [Google Scholar] [CrossRef] [PubMed]
  80. Li, C.; Du, Y.; Yang, Z.; He, L.; Wang, Y.; Hao, L.; Ding, M.; Yan, R.; Wang, J.; Fan, Z. GALNT1- mediated glycosylation and activation of Sonic Hedgehog Signaling maintains the self-renewal and tumor-initiating capacity of bladder cancer stem cells. Cancer Res. 2016, 76, 1273–1283. [Google Scholar] [CrossRef] [Green Version]
  81. Bafna, S.; Kaur, S.; Batra, S.K. Membrane-bound mucins: The mechanistic basis for alterations in the growth and survival of cancer cells. Oncogene 2010, 29, 2893–2904. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Tran, D.T.; Ten Hagen, K.G. Mucin-type O-glycosylation during development. J. Biol. Chem. 2013, 288, 6921–6929. [Google Scholar] [CrossRef] [Green Version]
  83. Corfield, A.P. Mucins: A biologically relevant glycan barrier in mucosal protection. Biochim. Biophys. Acta 2015, 1850, 236–252. [Google Scholar] [CrossRef]
  84. Melhem, H.; Regan-Komito, D.; Niess, J.H. Mucins dynamics in physiological and pathological conditions. Int. J. Mol. Sci. 2021, 22, 13642. [Google Scholar] [CrossRef]
  85. Sheng, Y.H.; Hasnain, S.Z. Mucus and mucins: The underappreciated host defence system. Front. Cell. Infect. Microbiol. 2022, 12, 856962. [Google Scholar] [CrossRef]
  86. Jonckheere, N.; Skrypek, N.; Van Seuningen, I. Mucins and tumor resistance to chemotherapeutic drugs. Biochim. Biophys. Acta 2014, 1846, 142–151. [Google Scholar] [CrossRef] [Green Version]
  87. Chugh, S.; Gnanapragassam, V.S.; Jain, M.; Rachagani, S.; Ponnusamy, M.P.; Batra, S.K. Pathobiological implications of mucin glycans in cancer: Sweet poison and novel targets. Biochim. Biophys. Acta 2015, 1856, 211–225. [Google Scholar] [CrossRef] [Green Version]
  88. Hanson, R.L.; Hollingsworth, M.A. Functional consequences of differential O-glycosylation of MUC1, MUC4, and MUC16 (downstream effects on signaling). Biomolecules 2016, 6, 34. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. van Putten, J.P.M.; Strijbis, K. Transmembrane mucins: Signaling receptors at the intersection of inflammation and cancer. J. Innate Immun. 2017, 9, 281–299. [Google Scholar] [CrossRef] [PubMed]
  90. Aithal, A.; Rauth, S.; Kshirsagar, P.; Shah, A.; Lakshmanan, I.; Junker, W.M.; Jain, M.; Ponnusamy, M.P.; Batra, S.K. MUC16 as a novel target for cancer therapy. Expert Opin. Ther. Targets 2018, 22, 675–686. [Google Scholar] [CrossRef]
  91. Bhatia, R.; Gautam, S.K.; Cannon, A.; Thompson, C.; Hall, B.R.; Aithal, A.; Banerjee, K.; Jain, M.; Solheim, J.C.; Kumar, S.; et al. Cancer-associated mucins: Role in immune modulation and metastasis. Cancer Metastasis Rev. 2019, 38, 223–236. [Google Scholar] [CrossRef] [PubMed]
  92. Gupta, R.; Leon, F.; Rauth, S.; Batra, S.K.; Ponnusamy, M.P. A systematic review on the implications of O-linked glycan branching and truncating enzymes on cancer progression and metastasis. Cells 2020, 9, 446. [Google Scholar] [CrossRef] [Green Version]
  93. Ganguly, K.; Rauth, S.; Marimuthu, S.; Kumar, S.; Batra, S.K. Unraveling mucin domains in cancer and metastasis: When protectors become predators. Cancer Metastasis Rev. 2020, 39, 647–659. [Google Scholar] [CrossRef]
  94. Brockhausen, I.; Melamed, J. Mucins as anti-cancer targets: Perspectives of the glycobiologist. Glycoconj. J. 2021, 38, 459–474. [Google Scholar] [CrossRef] [PubMed]
  95. Chen, W.; Zhang, Z.; Zhang, S.; Zhu, P.; Ko, J.K.; Yung, K.K. MUC1: Structure, function, and clinical application in epithelial cancers. Int. J. Mol. Sci. 2021, 22, 6567. [Google Scholar] [CrossRef] [PubMed]
  96. Marimuthu, S.; Rauth, S.; Ganguly, K.; Zhang, C.; Lakshmanan, I.; Batra, S.K.; Ponnusamy, M.P. Mucins reprogram stemness, metabolism and promote chemoresistance during cancer progression. Cancer Metastasis Rev. 2021, 40, 575–588. [Google Scholar] [CrossRef] [PubMed]
  97. Ratan, C.; Cicily, K.D.D.; Nair, B.; Nath, L.R. MUC glycoproteins: Potential biomarkers and molecular targets for cancer therapy. Curr. Cancer Drug Targets 2021, 21, 132–152. [Google Scholar] [CrossRef]
  98. Wi, D.H.; Cha, J.H.; Jung, Y.S. Mucin in cancer: A stealth cloak for cancer cells. BMB Rep. 2021, 54, 344–355. [Google Scholar] [CrossRef] [PubMed]
  99. Qing, L.; Li, O.; Yang, Y.; Xu, W.; Dong, Z. A prognosis marker MUC1 correlates with metabolism and drug resistance in bladder cancer: A bioinformatics research. BMC Urol. 2022, 22, 114. [Google Scholar] [CrossRef]
  100. Zhang, W.; Li, P.; Ding, L.; Yan, C. Mannose-targeting Concanavalin A-Epirubicin Conjugate for Targeted Intravesical Chemotherapy of Bladder Cancer. Chem. Asian J. 2022, 17, e202200342. [Google Scholar] [CrossRef]
  101. N’Dow, J.; Pearson, J.P.; Bennett, M.K.; Neal, D.E.; Robson, C.N. Mucin gene expression in human urothelium and in intestinal segments transposed into the urinary tract. J. Urol. 2000, 164, 1398–1404. [Google Scholar] [CrossRef]
  102. Lau, S.K.; Weiss, L.M.; Chu, P.G. Differential expression of MUC1, MUC2, and MUC5AC in carcinomas of various sites: An immunohistochemical study. Am. J. Clin. Pathol. 2004, 122, 61–69. [Google Scholar] [CrossRef]
  103. Garbar, C.; Mascaux, C. Expression of MUC1 (Ma695) in noninvasive papillary urothelial neoplasm according to the 2004 World Health Organization classification of the noninvasive urothelial neoplasm. An immunologic tool for the pathologist? Anal. Quant. Cytol. Histol. 2011, 33, 277–282. [Google Scholar]
  104. Kaur, S.; Momi, N.; Chakraborty, S.; Wagner, D.G.; Horn, A.J.; Lele, S.M.; Theodorescu, D.; Batra, S.K. Altered expression of transmembrane mucins, MUC1 and MUC4, in bladder cancer: Pathological implications in diagnosis. PLoS ONE 2014, 9, e92742. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Stojnev, S.; Ristic-Petrovic, A.; Jankovic Velickovic, L.; Krstic, M.; Bogdanovic, D.; Khanh, D.T.; Ristic, A.; Conic, I.; Stefanovic, V. Prognostic significance of mucin expression in urothelial bladder cancer. Int. J. Clin. Exp. Pathol. 2014, 7, 4945–4958. [Google Scholar] [PubMed]
  106. Kaymaz, E.; Ozer, E.; Unverdi, H.; Hucumenglu, S. Evaluation of MUC1 and P53 expressions in noninvasive papillary urothelial neoplasms of bladder, their relationship with tumor grade and role in the differential diagnosis. Indian J. Pathol. Microbiol. 2017, 60, 510–514. [Google Scholar] [CrossRef] [PubMed]
  107. Kobayashi, G.; Hayashi, T.; Sentani, K.; Takemoto, K.; Sekino, Y.; Uraoka, N.; Hanamoto, M.; Nose, H.; Teishima, J.; Arihiro, K.; et al. Clinicopathological significance of the overexpression of MUC1 in upper tract urothelial carcinoma and possible application as a diagnostic marker. Pathol. Int. 2022, 72, 606–616. [Google Scholar] [CrossRef] [PubMed]
  108. Retz, M.; Lehmann, J.; Amann, E.; Wullich, B.; Röder, C.; Stöckle, M. Mucin 7 and cytokeratin 20 as new diagnostic urinary markers for bladder tumor. J. Urol. 2003, 169, 86–89. [Google Scholar] [CrossRef]
  109. Okegawa, T.; Kinjo, M.; Horie, S.; Nutahara, K.; Higashihara, E. Detection of mucin 7 gene expression in exfoliated cells in urine from patients with bladder tumor. Urology 2003, 62, 182–186. [Google Scholar] [CrossRef]
  110. Kinjo, M.; Okegawa, T.; Horie, S.; Nutahara, K.; Higashihara, E. Detection of circulating MUC7-positive cells by reverse transcription-polymerase chain reaction in bladder cancer patients. Int. J. Urol. 2004, 11, 38–43. [Google Scholar] [CrossRef]
  111. Pu, X.Y.; Wang, Z.P.; Chen, Y.R.; Wang, X.H.; Wu, Y.L.; Wang, H.P. The value of combined use of survivin, cytokeratin 20 and mucin 7 mRNA for bladder cancer detection in voided urine. J. Cancer Res. Clin. Oncol. 2008, 134, 659–665. [Google Scholar] [CrossRef]
  112. Cotton, S.; Azevedo, R.; Gaiteiro, C.; Ferreira, D.; Lima, L.; Peixoto, A.; Fernandes, E.; Neves, M.; Neves, D.; Amaro, T.; et al. Targeted O-glycoproteomics explored increased sialylation and identified MUC16 as a poor prognosis biomarker in advanced-stage bladder tumours. Mol. Oncol. 2017, 11, 895–912. [Google Scholar] [CrossRef] [Green Version]
  113. Yamashita, T.; Higashi, M.; Sugiyama, H.; Morozumi, M.; Momose, S.; Tamaru, J.I. Cancer antigen 125 expression enhances the gemcitabine/cisplatin-resistant tumor microenvironment in bladder cancer. Am. J. Pathol. 2022, 193, 350–361. [Google Scholar] [CrossRef]
  114. Zhang, X.G.; Zhang, T.; Li, C.Y.; Zhang, M.H.; Chen, F.M. CD164 promotes tumor progression and predicts the poor prognosis of bladder cancer. Cancer Med. 2018, 7, 3763–3772. [Google Scholar] [CrossRef] [PubMed]
  115. Kobayashi, K.; Matsumoto, H.; Matsuyama, H.; Fujii, N.; Inoue, R.; Yamamoto, Y.; Nagao, K. Clinical significance of CD44 variant 9 expression as a prognostic indicator in bladder cancer. Oncol. Rep. 2016, 36, 2852–2860. [Google Scholar] [CrossRef] [Green Version]
  116. Wu, C.T.; Lin, W.Y.; Chang, Y.H.; Chen, W.C.; Chen, M.F. Impact of CD44 expression on radiation response for bladder cancer. J. Cancer 2017, 8, 1137–1144. [Google Scholar] [CrossRef] [Green Version]
  117. Wu, C.T.; Lin, W.Y.; Chen, W.C.; Chen, M.F. Predictive value of CD44 in muscle-invasive bladder cancer and its relationship with IL-6 signaling. Ann. Surg. Oncol. 2018, 25, 3518–3526. [Google Scholar] [CrossRef]
  118. Gaiteiro, C.; Soares, J.; Relvas-Santos, M.; Peixoto, A.; Ferreira, D.; Paulo, P.; Brandão, A.; Fernandes, E.; Azevedo, R.; Palmeira, C.; et al. Glycoproteogenomics characterizes the CD44 splicing code associated with bladder cancer invasion. Theranostics 2022, 12, 3150–3177. [Google Scholar] [CrossRef]
  119. Liao, C.; Wang, Q.; An, J.; Chen, J.; Li, X.; Long, Q.; Xiao, L.; Guan, X.; Liu, J. CD44 glycosylation as a therapeutic target in oncology. Front. Oncol. 2022, 12, 883831. [Google Scholar] [CrossRef]
  120. Shigeta, K.; Hasegawa, M.; Kikuchi, E.; Yasumizu, Y.; Kosaka, T.; Mizuno, R.; Mikami, S.; Miyajima, A.; Kufe, D.; Oya, M. Role of the MUC1-C oncoprotein in the acquisition of cisplatin resistance by urothelial carcinoma. Cancer Sci. 2020, 111, 3639–3652. [Google Scholar] [CrossRef]
  121. Nielsen, T.O.; Borre, M.; Nexo, E.; Sorensen, B.S. Co-expression of HER3 and MUC1 is associated with a favourable prognosis in patients with bladder cancer. BJU Int. 2015, 115, 163–165. [Google Scholar] [CrossRef] [PubMed]
  122. Memon, A.A.; Sorensen, B.S.; Melgard, P.; Fokdal, L.; Thykjaer, T.; Nexo, E. Expression of HER3, HER4 and their ligand heregulin-4 is associated with better survival in bladder cancer patients. Br. J. Cancer 2004, 91, 2034–2041. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Ferreira, J.A.; Videira, P.A.; Lima, L.; Pereira, S.; Silva, M.; Carrascal, M.; Severino, P.F.; Fernandes, E.; Almeida, A.; Costa, C.; et al. Overexpression of tumour-associated carbohydrate antigen sialyl-Tn in advanced bladder tumours. Mol. Oncol. 2013, 7, 719–731. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Severino, P.F.; Silva, M.; Carrascal, M.; Malagolini, N.; Chiricolo, M.; Venturi, G.; Astolfi, A.; Catera, M.; Videira, P.A.; Dall’Olio, F. Expression of sialylTn sugar antigen in bladder cancer cells affects response to Bacillus Calmette Guérin (BCG) and to oxidative damage. Oncotarget 2017, 8, 54506–54517. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Lima, L.; Severino, P.F.; Silva, M.; Miranda, A.; Tavares, A.; Pereira, S.; Fernandes, E.; Cruz, R.; Amaro, T.; Reis, C.A.; et al. Response of high-risk of recurrence/progression bladder tumours expressing sialyl-Tn and sialyl-6-T to BCG immunotherapy. Br. J. Cancer 2013, 109, 2106–2114. [Google Scholar] [CrossRef]
  126. Costa, C.; Pereira, S.; Lima, L.; Peixoto, A.; Fernandes, E.; Neves, D.; Neves, M.; Gaiteiro, C.; Tavares, A.; Gil da Costa, R.M.; et al. Abnormal protein glycosylation and activated PI3K/Akt/mTOR pathway: Role in bladder cancer prognosis and targeted therapeutics. PLoS ONE 2015, 10, e0141253. [Google Scholar] [CrossRef]
  127. Peixoto, A.; Fernandes, E.; Gaiteiro, C.; Lima, L.; Azevedo, R.; Soares, J.; Cotton, S.; Parreira, B.; Neves, M.; Amaro, T.; et al. Hypoxia enhances the malignant nature of bladder cancer cells and concomitantly antagonizes protein O-glycosylation extension. Oncotarget 2016, 7, 63138–63157. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Lima, L.; Neves, M.; Oliveira, M.I.; Dieguez, L.; Freitas, R.; Azevedo, R.; Gaiteiro, C.; Soares, J.; Ferreira, D.; Peixoto, A.; et al. Sialyl-Tn identifies muscle-invasive bladder cancer basal and luminal subtypes facing decreased survival, being expressed by circulating tumor cells and metastases. Urol. Oncol. 2017, 35, 675.e1–675.e8. [Google Scholar] [CrossRef] [PubMed]
  129. Neves, M.; Azevedo, R.; Lima, L.; Oliveira, M.I.; Peixoto, A.; Ferreira, D.; Soares, J.; Fernandes, E.; Gaiteiro, C.; Palmeira, C.; et al. Exploring sialyl-Tn expression in microfluidic-isolated circulating tumour cells: A novel biomarker and an analytical tool for precision oncology applications. N. Biothchnol. 2019, 49, 77–87. [Google Scholar] [CrossRef]
  130. Carrascal, M.A.; Severino, P.F.; Guadalupe Cabral, M.; Silva, M.; Ferreira, J.A.; Calais, F.; Quinto, H.; Pen, C.; Ligeiro, D.; Santos, L.L.; et al. Sialyl Tn-expressing bladder cancer cells induce a tolerogenic phenotype in innate and adaptive immune cells. Mol. Oncol. 2014, 8, 753–765. [Google Scholar] [CrossRef]
  131. Ogata, S.; Maimonis, P.J.; Itzkowitz, S.H. Mucins bearing the cancer-associated sialosyl-Tn antigen mediate inhibition of natural killer cell cytotoxicity. Cancer Res. 1992, 52, 4741–4746. [Google Scholar]
  132. Bernardo, C.; Costa, C.; Amaro, T.; Goncalves, M.; Lopes, P.; Freitas, R.; Gartner, F.; Amado, F.; Ferreira, J.A.; Santos, L. Patient-derived sialyl-Tn-positive invasive bladder cancer xenografts in nude mice: An exploratory model study. Anticancer Res. 2014, 34, 735–744. [Google Scholar]
  133. Theodoropoulos, V.E.; Lazaris, A.; Sofras, F.; Gerzelis, I.; Tsoukala, V.; Ghikonti, I.; Manikas, K.; Kastriotis, I. Hypoxiainducible factor 1α expression correlates with angiogenesis and unfavorable prognosis in bladder cancer. Eur. Urol. 2004, 46, 200–208. [Google Scholar] [CrossRef]
  134. Chai, C.Y.; Chen, W.T.; Hung, W.C.; Kang, W.Y.; Huang, Y.C.; Su, Y.C.; Yang, C.H. Hypoxia-inducible factor-1α expression correlates with focal macrophage infiltration, angiogenesis and unfavourable prognosis in urothelial carcinoma. J. Clin. Pathol. 2008, 61, 658–664. [Google Scholar] [CrossRef] [PubMed]
  135. Peixoto, A.; Freitas, R.; Ferreira, D.; Relvas-Santos, M.; Paulo, P.; Cardoso, M.; Soares, J.; Gaiteiro, C.; Palmeira, C.; Teixeira, F.; et al. Metabolomics, transcriptomics and functional glycomics reveals bladder cancer cells plasticity and enhanced aggressiveness facing hypoxia and glucose deprivation. BioRxiv, 2021; preprint. [Google Scholar] [CrossRef]
  136. Peixoto, A.; Ferreira, D.; Azevedo, R.; Freitas, R.; Fernandes, E.; Relvas-Santos, M.; Gaiteiro, C.; Soares, J.; Cotton, S.; Teixeira, B.; et al. Glycoproteomics identifies HOMER3 as a potentially targetable biomarker triggered by hypoxia and glucose deprivation in bladder cancer. J. Exp. Clin. Cancer Res. 2021, 40, 191. [Google Scholar] [CrossRef] [PubMed]
  137. Takamiya, R.; Ohtsubo, K.; Takamatsu, S.; Taniguchi, N.; Angata, T. The interaction between Siglec-15 and tumor-associated sialyl-Tn antigen enhances TGF-β secretion from monocytes/macrophages through the DAP12-Syk pathway. Glycobiology 2013, 23, 178–187. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Loureiro, L.R.; Feldmann, A.; Bergmann, R.; Koristka, S.; Berndt, N.; Máthé, D.; Hegedüs, N.; Szigeti, K.; Videira, P.A.; Bachmann, M.; et al. Extended half-life target module for sustainable UniCAR T-cell treatment of STn-expressing cancers. J. Exp. Clin. Cancer Res. 2020, 39, 77. [Google Scholar] [CrossRef] [PubMed]
  139. Santos, J.; Fernandes, E.; Ferreira, J.A.; Lima, L.; Tavares, A.; Peixoto, A.; Parreira, B.; Correia da Costa, J.M.; Brindley, P.J.; Lopes, C.; et al. P53 and cancer-associated sialylated glycans are surrogate markers of cancerization of the bladder associated with Schistosoma haematobium infection. PLoS Negl. Trop. Dis. 2014, 8, e3329. [Google Scholar] [CrossRef] [Green Version]
  140. Liu, F.; Fu, J.; Bergstrom, K.; Shan, X.; McDaniel, J.M.; McGee, S.; Bai, X.; Chen, W.; Xia, L. Core 1-derived mucin-type O-glycosylation protects against spontaneous gastritis and gastric cancer. J. Exp. Med. 2020, 217, e20182325. [Google Scholar] [CrossRef] [Green Version]
  141. Chugh, S.; Barkeer, S.; Rachagani, S.; Nimmakayala, R.K.; Perumal, N.; Pothuraju, R.; Atri, P.; Mahapatra, S.; Thapa, I.; Talmon, G.A.; et al. Disruption of C1galt1 gene promotes development and metastasis of pancreatic adenocarcinomas in mice. Gastroenterology 2018, 155, 1608–1624. [Google Scholar] [CrossRef]
  142. Leon, F.; Seshacharyulu, P.; Nimmakayala, R.K.; Chugh, S.; Karmakar, S.; Nallasamy, P.; Vengoji, R.; Rachagani, S.; Cox, J.L.; Mallya, K.; et al. Reduction in O-glycome induces diferentially glycosylated CD44 to promote stemness and metastasis in pancreatic cancer. Oncogene 2022, 41, 57–71. [Google Scholar] [CrossRef]
  143. Tan, Z.; Jiang, Y.; Liang, L.; Wu, J.; Cao, L.; Zhou, X.; Song, Z.; Ye, Z.; Zhao, Z.; Feng, H.; et al. Dysregulation and prometastatic function of glycosyltransferase C1GALT1 modulated by cHP1BP3/ miR-1-3p axis in bladder cancer. J. Exp. Clin. Cancer Res. 2022, 41, 228. [Google Scholar] [CrossRef]
  144. Xiang, T.; Qiao, M.; Xie, J.; Li, Z.; Xie, H. Emerging roles of the unique molecular chaperone Cosmc in the regulation of health and disease. Biomolecules 2022, 12, 1732. [Google Scholar] [CrossRef]
  145. Wang, Y.; Ju, T.; Ding, X.; Xia, B.; Wang, W.; Xia, L.; He, M.; Cummings, R.D. Cosmc is an essential chaperone for correct protein O-glycosylation. Proc. Natl. Acad. Sci. USA 2010, 107, 9228–9233. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Klatte, T.; Xylinas, E.; Rieken, M.; Kluth, L.A.; Rouprêt, M.; Pycha, A.; Fajkovic, H.; Seitz, C.; Karakiewicz, P.I.; Lotan, Y.; et al. Impact of ABO blood type on outcomes in patients with primary nonmuscle invasive bladder cancer. J. Urol. 2014, 191, 1238–1243. [Google Scholar] [CrossRef] [PubMed]
  147. Videira, P.A.; Correia, M.; Malagolini, N.; Crespo, H.C.; Ligeiro, D.; Calais, F.M.; Trindade, H.; Dall’Olio, F. ST3Gal.I sialyltransferase relevance in bladder cancer tissues and cell lines. BMC Cancer 2009, 9, 357. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Severino, P.F.; Silva, M.; Carrascal, M.; Malagolini, N.; Chiricolo, M.; Venturi, G.; Barbaro Forleo, R.; Astolfi, A.; Catera, M.; Videira, P.A.; et al. Oxidative damage and response to Bacillus Calmette-Guérin in bladder cancer cells expressing sialyltransferase ST3GAL1. BMC Cancer 2018, 18, 198. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Takeuchi, M.; Amano, M.; Kitamura, H.; Tsukamoto, T.; Masumori, N.; Hirose, K.; Ohashi, T.; Nishimura, S.I. N- and O-glycome analysis of serum and urine from bladder cancer patients using a high-throughput glycoblotting method. J. Glycom. Lipidom. 2013, 3, 1000108. [Google Scholar] [CrossRef]
  150. Suzuki, Y.; Sutoh, M.; Hatakeyama, S.; Mori, K.; Yamamoto, H.; Koie, T.; Saitoh, H.; Yamaya, K.; Funyu, T.; Habuchi, T.; et al. MUC1 carrying core 2 O-glycans functions as a molecular shield against NK cell attack, promoting bladder tumor metastasis. Int. J. Oncol. 2012, 40, 1831–1838. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  151. Tsuboi, S.; Sutoh, M.; Hatakeyama, S.; Hiraoka, N.; Habuchi, T.; Horikawa, Y.; Hashimoto, Y.; Yoneyama, T.; Mori, K.; Koie, T.; et al. A novel strategy for evasion of NK cell immunity by tumours expressing core2 O-glycans. EMBO J. 2011, 30, 3173–3185. [Google Scholar] [CrossRef] [Green Version]
  152. Sutoh Yoneyama, M.; Tobisawa, Y.; Hatakeyama, S.; Sato, M.; Tone, K.; Tatara, Y.; Kakizaki, I.; Funyu, T.; Fukuda, M.; Hoshi, S.; et al. A mechanism for evasion of CTL immunity by altered O-glycosylation of HLA class I. J. Biochem. 2017, 161, 479–492. [Google Scholar] [CrossRef] [Green Version]
  153. Li, H.; Yang, F.; Chang, K.; Yu, X.; Guan, F.; Li, X. The synergistic function of long and short forms of β4GalT1 in p53-mediated drug resistance in bladder cancer cells. Biochim. Biophys. Acta. Mol. Cell Res. 2023, 1870, 119409. [Google Scholar] [CrossRef]
  154. Jia, L.; Zhang, J.; Ma, T.; Guo, Y.; Yu, Y.; Cui, J. The Function of Fucosylation in Progression of Lung Cancer. Front. Oncol. 2018, 8, 565. [Google Scholar] [CrossRef] [PubMed]
  155. Luo, Y.; Haltiwanger, R.S. O-fucosylation of notch occurs in the endoplasmic reticulum. J. Biol. Chem. 2005, 280, 11289–11294. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Aoyagi, Y.; Saitoh, A.; Suzuki, Y.; Igarashi, K.; Oguro, M.; Yokota, T.; Mori, S.; Suda, T.; Isemura, M.; Asakura, H. Fucosylation index of alpha-fetoprotein, a possible aid in the early recognition of hepatocellular carcinoma in patients with cirrhosis. Hepatology 1993, 17, 50–52. [Google Scholar] [CrossRef] [PubMed]
  157. Shan, M.; Yang, D.; Dou, H.; Zhang, L. Fucosylation in cancer biology and its clinical applications. Prog. Mol. Biol. Transl. Sci. 2019, 162, 93–119. [Google Scholar] [CrossRef] [PubMed]
  158. Yang, G.; Tan, Z.; Lu, W.; Guo, J.; Yu, H.; Yu, J.; Sun, C.; Qi, X.; Li, Z.; Guan, F. Quantitative glycome analysis of N-glycan patterns in bladder cancer vs normal bladder cells using an integrated strategy. J. Proteome Res. 2015, 14, 639–653. [Google Scholar] [CrossRef]
  159. Ferdosi, S.; Ho, T.H.; Castle, E.P.; Stanton, M.L.; Borges, C.R. Behavior of blood plasma glycan features in bladder cancer. PLoS ONE 2018, 13, e0201208. [Google Scholar] [CrossRef]
  160. Guo, J.; Xiang, L.; Zengqi, T.; Wei, L.; Ganglong, Y. Alteration of N-glycans and Expression of Their Related Glycogenes in the Epithelial-Mesenchymal Transition of HCV29 Bladder Epithelial Cells. Molecules 2014, 19, 20073–20090. [Google Scholar] [CrossRef] [Green Version]
  161. Lu, Y.C.; Chen, C.N.; Chu, C.Y.; Lu, J.; Wang, B.J.; Chen, C.H.; Huang, M.C.; Lin, T.H.; Pan, C.C.; Chen, S.S.; et al. Calreticulin activates β1 integrin via fucosylation by fucosyltransferase 1 in J82 human bladder cancer cells. Biochem. J. 2014, 460, 69–78. [Google Scholar] [CrossRef]
  162. Liu, M.; Zheng, Q.; Chen, S.; Liu, J.; Li, S. Fut7 promotes the epithelial–mesenchymal transition and immune infiltration in bladder urothelial carcinoma. J. Inflamm. Res. 2021, 14, 1069–1084. [Google Scholar] [CrossRef]
  163. Numahata, K.; Satoh, M.; Handa, K.; Saito, S.; Ohyama, C.; Ito, A.; Takahashi, T.; Hoshi, S.; Orikasa, S.; Hakomori, S.I. Sialosyl-Le(x) expression defines invasive and metastatic properties of bladder carcinoma. Cancer 2002, 94, 673–685. [Google Scholar] [CrossRef]
  164. Islam, M.K.; Syed, P.; Dhondt, B.; Gidwani, K.; Pettersson, K.; Lamminmäki, U.; Leivo, J. Detection of bladder cancer with aberrantly fucosylated ITGA3. Anal. Biochem. 2021, 628, 114283. [Google Scholar] [CrossRef] [PubMed]
  165. Islam, K. Extracellular Vesicle Glycosylations as Novel Biomarkers of Urological Cancers: Nanoparticle-Aided Glycovariant Assay to Detect Vesicles for the Early Detection of Cancer. Ph.D. Thesis, Turun Yliopisto, Turku, Finland, 2022. [Google Scholar]
  166. Ongay, S.; Martín-Álvarez, P.J.; Neusüss, C.; de Frutos, M. Statistical evaluation of CZE-UV and CZE-ESI-MS data of intact α-1-acid glycoprotein isoforms for their use as potential biomarkers in bladder cancer. Electrophoresis 2010, 31, 3314–3325. [Google Scholar] [CrossRef] [Green Version]
  167. Blanas, A.; Sahasrabudhe, N.M.; Rodríguez, E.; van Kooyk, Y.; van Vliet, S.J. Fucosylated antigens in cancer: An alliance toward tumor progression, metastasis, and resistance to chemotherapy. Front. Oncol. 2018, 8, 39. [Google Scholar] [CrossRef] [PubMed]
  168. Sheinfeld, J.; Reuter, V.E.; Sarkis, A.S.; Cordon-Cardo, C. Blood group antigens in normal and neoplastic urothelium. J. Cell. Biochem. 1992, 50, 50–55. [Google Scholar] [CrossRef]
  169. Sheinfeld, J.; Reuter, V.E.; Melamed, M.R.; Fair, W.R.; Morse, M.; Sogani, P.C.; Herr, H.W.; Whitmore, W.F.; Cordon-Cardo, C. Enhanced bladder cancer detection with the Lewis X antigen as a marker of neoplastic transformation. J. Urol. 1990, 143, 285–288. [Google Scholar] [CrossRef]
  170. Pal, S.K.; Pham, A.; Vuong, W.; Liu, X.; Lin, Y.; Ruel, N.; Yuh, B.E.; Chan, K.; Wilson, T.; Lerner, S.P.; et al. Prognostic significance of neutrophilic infiltration in benign lymph nodes in patients with muscle-invasive bladder cancer. Eur. Urol. Focus 2017, 3, 130–135. [Google Scholar] [CrossRef] [PubMed]
  171. Ezeabikwa, B.; Nandini, M.; Aristotelis, A.; Stuart, M.H.; Yasuyuki, M.; Miguel, M.C.; Sylvain, L.; Msano, M.; Ali, I.; Jamie, H.-M.; et al. Major differences in glycosylation and fucosyltransferase expression in low-grade versus high-grade bladder cancer cell lines. Glycobiology 2021, 31, 1444–1463. [Google Scholar] [CrossRef] [PubMed]
  172. Skorstengaard, K.; Vestergaard, E.M.; Langkilde, N.C.; Christensen, L.L.; Wolf, H.; Orntoft, T.F. Lewis antigen mediated adhesion of freshly removed human bladder tumors to E-selectin. J. Urol. 1999, 161, 1316–1323. [Google Scholar] [CrossRef]
  173. Ferreira, J.A.; Magalhaes, A.; Gomes, J.; Peixoto, A.; Gaiteiro, C.; Fernandes, E.; Santos, L.L.; Reis, C.A. Protein glycosylation in gastric and colorectal cancers: Toward cancer detection and targeted therapeutics. Cancer Lett. 2017, 387, 32–45. [Google Scholar] [CrossRef]
  174. Subedi, P.; Schneider, M.; Philipp, J.; Azimzadeh, O.; Metzger, F.; Moertl, S. Comparison of methods to isolate proteins from extracellular vesicles for mass spectrometry-based proteomic analyses. Anal. Biochem. 2019, 584, 113390. [Google Scholar] [CrossRef]
  175. Zaborowski, M.P.; Balaj, L.; Breakefield, X.O.; Lai, C.P.-K. Extracellular vesicles: Composition, biological relevance, and methods of study. Bioscience 2015, 65, 783–797. [Google Scholar] [CrossRef] [Green Version]
  176. Bebelman, M.P.; Smit, M.J.; Pegtel, D.M.; Baglio, S.R. Biogenesis and function of extracellular vesicles in cancer. Pharmacol. Ther. 2018, 188, 1–11. [Google Scholar] [CrossRef] [PubMed]
  177. Chang, W.H.; Cerione, R.A.; Antonyak, M.A. Extracellular vesicles and their roles in cancer progression. Methods Mol. Biol. 2021, 2174, 143–170. [Google Scholar] [CrossRef] [PubMed]
  178. Surman, M.; Wilczak, M.; Przybyło, M. Lectin-based study reveals the presence of disease-relevant glycoepitopes in bladder cancer cells and ectosomes. Int. J. Mol. Sci. 2022, 23, 14368. [Google Scholar] [CrossRef] [PubMed]
  179. Zhang, J.; Zhang, Y.; Luo, C.; Xia, Y.; Chen, H.; Wu, X. Glycosyl-phosphatidylinositol-anchored interleukin-2 expressed on tumor-derived exosomes induces antitumor immune response in vitro. Tumori 2010, 96, 452–459. [Google Scholar] [CrossRef] [PubMed]
  180. Klingler, C.H. Glycosoaminoglycans: How much do we know about their role in the bladder. Urol. J. 2016, 83 (Suppl. 1), S11–S14. [Google Scholar] [CrossRef]
  181. Singleton, P.A. Hyaluronan regulation of endothelial barrier function in cancer. Adv. Cancer Res. 2014, 123, 191–209. [Google Scholar] [CrossRef] [Green Version]
  182. Morla, S. Glycosaminoglycans and glycosaminoglycan mimetics in cancer and inflammation. Int. J. Mol. Sci. 2019, 20, 1963. [Google Scholar] [CrossRef] [Green Version]
  183. Tcheocharis, A.D.; Karamanos, N.K. Proteoglycans remodeling in cancer: Underlying molecular mechanism. Matrix Biol. 2019, 75–76, 220–259. [Google Scholar] [CrossRef]
  184. Wei, J.; Hu, M.; Huang, K.; Lin, S.; Du, H. Roles of proteoglycans and glycosaminoglycans in cancer development and progression. Int. J. Mol. Sci. 2020, 21, 5983. [Google Scholar] [CrossRef]
  185. Belting, M. Glycosaminoglycans in cancer treatment. Thromb. Res. 2014, 133 (Suppl. 2), S95–S101. [Google Scholar] [CrossRef] [PubMed]
  186. Wieboldt, R.; Läubli, H. Glycosaminoglycans in cancer therapy. Am. J. Physiol. Cell Physiol. 2022, 322, C1187–C1200. [Google Scholar] [CrossRef]
  187. Rübben, H.; Friedrichs, R.; Stuhlsatz, H.; Lutzeyer, W. Glycosaminoglycans in urothelial carcinomas. Urol. Res. 1983, 11, 163–166. [Google Scholar] [CrossRef] [PubMed]
  188. De Klerk, D.P. The glycosaminoglycans of human bladder cancers of varying grade and stage. J. Urol. 1985, 134, 978–981. [Google Scholar] [CrossRef] [PubMed]
  189. Atahan, O.; Kayigil, O.; Hizel, N.; Yavuz, O.; Metin, A. Urinary glycosaminoglycan excretion in bladder carcinoma. Scand. J. Urol. Nephrol. 1996, 30, 173–177. [Google Scholar] [CrossRef] [PubMed]
  190. Konukoğlu, D.; Akçay, T.; Erözenci, A. The importance urinary glycosaminoglycan as a marker for superficial bladder tumors. Cancer Biochem. Res. 1995, 15, 91–95. [Google Scholar]
  191. Bojanić, N.; Nale, D.; Mićić, S.; Lalić, N.; Vuksanović, A.; Tulić, C. Glycosaminoglycans in the urinary bladder mucosa, tumor tissue and mucosal tissue around tumor. Vojnosanit. Pregl. 2012, 69, 147–150. [Google Scholar] [CrossRef]
  192. Lokeshwar, V.B.; Obek, C.; Pham, H.T.; Wei, D.; Young, M.J.; Duncan, R.C.; Soloway, M.S.; Bloch, N.L. Urinary hyaluronic acid and hyaluronidase: Markers for bladder cancer detection and evaluation of grade. J. Urol. 2000, 163, 348–356. [Google Scholar] [CrossRef]
  193. Lokeshwar, V.B.; Schroeder, G.L.; Selzer, M.G.; Hautmann, S.H.; Posey, T.; Duncan, R.C.; Watson, R.; Rose, L.; Markowitz, S.; Soloway, M.S. Bladder tumor markers for monitoring recurrence and screening comparison of hyaluronic acid-hyaluronidase and BTA-Stat tests. Cancer 2002, 95, 61–72. [Google Scholar] [CrossRef]
  194. Schroeder, G.L.; Lorenzo-Gomez, M.F.; Hautmann, S.H.; Friedrich, M.G.; Ekici, S.; Huland, H.; Lokeshwar, V. A side by side comparison of cytology and biomarkers for bladder cancer detection. J. Urol. 2004, 172, 1123–1126. [Google Scholar] [CrossRef] [PubMed]
  195. Chen, L.; Fu, C.; Zhang, Q.; He, C.; Zhang, F.; Wei, Q. The role of CD44 in pathological angiogenesis. FASEB J. 2020, 34, 13125–13139. [Google Scholar] [CrossRef]
  196. Yaghobi, Z.; Movassaghpour, A.; Talebi, M.; Abdoli Shadbad, M.; Hajiasgharzadeh, K.; Pourvahdani, S.; Baradaran, B. The role of CD44 in cancer chemoresistance: A concise review. Eur. J. Pharmacol. 2021, 903, 174147. [Google Scholar] [CrossRef] [PubMed]
  197. Guo, Q.; Yang, C.; Gao, F. The state of CD44 activation in cancer progression and therapeutic targeting. FEBS J. 2022, 289, 7970–7986. [Google Scholar] [CrossRef] [PubMed]
  198. Golshani, R.; Hautmann, S.H.; Estrella, V.; Cohen, B.L.; Kyle, C.C.; Manoharan, M.; Jorda, M.; Soloway, M.S.; Lokeshwar, V.B. HAS1 expression in bladder cancer and its relation to urinary HA test. Int. J. Cancer 2007, 120, 1712–1720. [Google Scholar] [CrossRef]
  199. Kramer, M.W.; Escudero, D.O.; Lokeshwar, S.D.; Golshani, R.; Ekwenna, O.O.; Acosta, K.; Mersegurger, A.S.; Soloway, M.S.; Lokeshwar, V.B. Association of hyaluronic acid family members (HAS1, HAS2, and HYAL-1) with bladder cancer diagnosis and prognosis. Cancer 2011, 117, 1197–1209. [Google Scholar] [CrossRef] [Green Version]
  200. Golshani, R.; Lopez, L.; Estrella, V.; Kramer, M.; Iida, N.; Lokeshwar, V.B. Hyaluronic acid synthase-1 expression regulates bladder cancer growth, invasion, and angiogenesis through CD44. Cancer Res. 2008, 68, 483–491. [Google Scholar] [CrossRef] [Green Version]
  201. 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] [PubMed]
  202. Bharadwaj, A.G.; Kovar, J.L.; Loughman, E.; Elowsky, C.; Oakley, G.G.; Simpson, M.A. Spontaneous metastasis of prostate cancer is promoted by excess hyaluronan synthesis and processing. Am. J. Pathol. 2009, 174, 1027–1036. [Google Scholar] [CrossRef] [Green Version]
  203. Guin, S.; Pollard, C.; Ru, Y.; Ritterson Lew, C.; Duex, J.E.; Dancik, G.; Owens, C.; Spencer, A.; Knight, S.; Holemon, H.; et al. Role in tumor growth of a glycogen debranching enzyme lost in glycogen storage disease. J. Natl. Cancer Inst. 2014, 106, dju062. [Google Scholar] [CrossRef] [Green Version]
  204. Guin, S.; Ru, Y.; Agarwal, N.; Lew, C.R.; Owens, C.; Comi, G.P.; Theodorescu, D. Loss of glycogen debranching enzyme AGL drives bladder tumor growth via induction of hyaluronic acid synthesis. Clin. Cancer Res. 2016, 22, 1274–1283. [Google Scholar] [CrossRef] [Green Version]
  205. Oldenburg, D.; Ru, Y.; Weinhaus, B.; Cash, S.; Theodorescu, D.; Guin, S. CD44 and RHAMM are essential for rapid growth of bladder cancer driven by loss of Glycogen Debranching Enzyme (AGL). BMC Cancer 2016, 16, 713. [Google Scholar] [CrossRef]
  206. Kramer, M.W.; Golshani, R.; Merseburger, A.S.; Kramer, M.W.; Golshani, R.; Merseburger, A.S.; Knapp, J.; Garcia, A.; Hennenlotter, J.; Duncan, R.C.; et al. HYAL-1 Hyaluronidase: A potential prognostic indicator for progression to muscle invasion and recurrence in bladder cancer. Eur. Urol. 2009, 57, 86–93. [Google Scholar] [CrossRef] [Green Version]
  207. Liang, Z.; Zhang, Q.; Wang, C.H.; Shi, F.; Cao, H.; Yu, Y.; Zhang, M.; Liu, X. Hyaluronic acid/ hyaluronidase as biomarkers for bladder cancer: A diagnostic meta-analysis. Neoplasma 2017, 64, 901–908. [Google Scholar] [CrossRef]
  208. Ferro, M.; Giuberti, G.; Zappavigna, S.; Perdonà, S.; Facchini, G.; Sperlongano, P.; Porto, S.; Di Lorenzo, G.; Buonerba, C.; Abbruzzese, A.; et al. Chondroitin sulphate enhances the antitumor activity of gemcitabine and mitomycin-C in bladder cancer cells with different mechanisms. Oncol. Rep. 2012, 27, 409–415. [Google Scholar] [CrossRef]
  209. Morrione, A.; Neill, T.; Iozzo, R.V. Dichotomy of decorin activity on the insulin-like growth factor-I system. FEBS J. 2013, 280, 2138–2149. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  210. Appunni, S.; Anand, V.; Khandelwal, M.; Gupta, N.; Rubens, M.; Sharma, A. Small leucine rich proteoglycans (decorin, biglycan and lumican) in cancer. Clin. Chim. Acta 2019, 491, 1–7. [Google Scholar] [CrossRef] [PubMed]
  211. Sainio, A.; Nyman, M.; Lund, R.; Vuorikoski, S.; Boström, P.; Laato, M.; Boström, P.J.; Järveläinen, H. Lack of decorin expression by human bladder cancer cells offers new tools in the therapy of urothelial malignancies. PLoS ONE 2013, 8, e76190. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  212. Chen, H.; Wang, Z.; Yang, N.; Zhang, J.; Liang, Z. Decorin inhibits proliferation and metastasis in human bladder cancer cells by upregulating P21. Medicine 2022, 101, e29760. [Google Scholar] [CrossRef]
  213. El Behi, M.; Krumeich, S.; Lodillinsky, C.; Kamoun, A.; Tibaldi, L.; Sugano, G.; De Reynies, A.; Chapeaublanc, E.; Laplanche, A.; Lebret, T.; et al. An essential role for decorin in bladder cancer invasiveness. EMBO Mol. Med. 2013, 5, 1835–1851. [Google Scholar] [CrossRef] [PubMed]
  214. Niedworok, C.; Röck, K.; Kretschmer, I.; Freudenberger, T.; Nagy, N.; Szarvas, T.; vom Dorp, F.; Reis, H.; Rübben, H.; Fischer, J.W. Inhibitory role of the small leucine-rich proteoglycan biglycan in bladder cancer. PLoS ONE 2013, 8, e80084. [Google Scholar] [CrossRef] [Green Version]
  215. Salanti, A.; Clausen, T.M.; Agerbæk, M.Ø.; Al Nakouzi, N.; Dahlbäck, M.; Oo, H.Z.; Lee, S.; Gustavsson, T.; Rich, J.R.; Hedberg, B.J.; et al. Targeting human cancer by a glycosaminoglycan binding malaria protein. Cancer Cell 2015, 28, 500–514. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  216. Seiler, R.; Oo, H.Z.; Tortora, D.; Clausen, T.M.; Wang, C.K.; Kumar, G.; Pereira, M.A.; Ørum-Madsen, M.S.; Agerbæk, M.Ø.; Gustavsson, T.; et al. An oncofetal glycosaminoglycan modification provides therapeutic access to cisplatin-resistant bladder cancer. Eur. Urol. 2017, 72, 142–150. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Clausen, T.M.; Pereira, M.A.; Al Nakouzi, N.; Oo, H.Z.; Agerbæk, M.Ø.; Lee, S.; Ørum-Madsen, M.S.; Kristensen, A.R.; El-Naggar, A.; Grandgenett, P.M.; et al. Oncofetal chondroitin sulfate glycosaminoglycans are key players in integrin signaling and tumor cell motility. Mol. Cancer Res. 2016, 14, 1288–1299. [Google Scholar] [CrossRef] [Green Version]
  218. Clausen, T.M.; Kumar, G.; Ibsen, E.K.; Ørum-Madsen, M.S.; Hurtado-Coll, A.; Gustavsson, T.; Agerbæk, M.Ø.; Gatto, F.; Todenhöfer, T.; Basso, U.; et al. A simple method for detecting oncofetal chondroitin sulfate glycosaminoglycans in bladder cancer urine. Cell Death Discov. 2020, 6, 65. [Google Scholar] [CrossRef]
  219. Greco, B.; Malacarne, V.; De Girardi, F.; Scotti, G.M.; Manfredi, F.; Angelino, E.; Sirini, C.; Camisa, B.; Falcone, L.; Moresco, M.A.; et al. Disrupting N-glycan expression on tumor cells boosts chimeric antigen receptor T cell efficacy against solid malignancies. Sci. Transl. Med. 2022, 14, eabg3072. [Google Scholar] [CrossRef]
  220. Kong, S.; Gong, P.; Zeng, W.F.; Jiang, B.; Hou, X.; Zhang, Y.; Zhao, H.; Liu, M.; Yan, G.; Zhou, X.; et al. pGlycoQuant with a deep residual network for quantitative glycoproteomics at intact glycopeptide level. Nat. Commun. 2022, 13, 7539. [Google Scholar] [CrossRef] [PubMed]
  221. Hu, Y.; Pan, J.; Shah, P.; Ao, M.; Thomas, S.N.; Liu, Y.; Chen, L.; Schnaubelt, M.; Clark, D.J.; Rodriguez, H.; et al. Integrated proteomic and glycoproteomic characterization of human high-grade serous ovarian carcinoma. Cell Rep. 2020, 33, 108276. [Google Scholar] [CrossRef] [PubMed]
  222. Pan, J.; Hu, Y.; Sun, S.; Chen, L.; Schnaubelt, M.; Clark, D.; Ao, M.; Zhang, Z.; Chan, D.; Qian, J.; et al. Glycoproteomics-based signatures for tumor subtyping and clinical outcome prediction of high-grade serous ovarian cancer. Nat. Commun. 2020, 11, 6139. [Google Scholar] [CrossRef]
Figure 1. Schematic presentation of classes of glycoconjugates. Glycosylation leads to the creation of the so-called glycoconjugates, which include glycoproteins, proteoglycans, glycosphingolipids, and glycosylated RNA. In addition, glycans not covalently bound to macromolecules are present.
Figure 1. Schematic presentation of classes of glycoconjugates. Glycosylation leads to the creation of the so-called glycoconjugates, which include glycoproteins, proteoglycans, glycosphingolipids, and glycosylated RNA. In addition, glycans not covalently bound to macromolecules are present.
Molecules 28 03436 g001
Figure 2. Schematic presentation of N-glycans types.
Figure 2. Schematic presentation of N-glycans types.
Molecules 28 03436 g002
Figure 3. GnT-V catalyzes the formation of β1,6 branched complex type N-glycans. GnT-III catalyzes the attachment of so-called bisecting GlcNAc to complex and hybrid (not shown) type N-glycans. Glycans possessing bisecting GlcNAc are not substrates for GnT-V, which leads to the inhibition of the biosynthesis of β1,6 branched N-glycans.
Figure 3. GnT-V catalyzes the formation of β1,6 branched complex type N-glycans. GnT-III catalyzes the attachment of so-called bisecting GlcNAc to complex and hybrid (not shown) type N-glycans. Glycans possessing bisecting GlcNAc are not substrates for GnT-V, which leads to the inhibition of the biosynthesis of β1,6 branched N-glycans.
Molecules 28 03436 g003
Figure 4. The crosstalk between O-GlcNAcylation and phosphorylation of serine and threonine residues, and the antagonistic relation between OGT-kinase and OGA-phosphatase.
Figure 4. The crosstalk between O-GlcNAcylation and phosphorylation of serine and threonine residues, and the antagonistic relation between OGT-kinase and OGA-phosphatase.
Molecules 28 03436 g004
Figure 5. Mucin-type O-glycosylation is initiated in the Golgi apparatus by polypeptide N-acetylgalactosaminyltransferases (GalNAc-Ts, twenty isoforms), which initiates the action of numerous glycosyltransferases catalyzing the attachment of galactose, GlcNAc, fucose, and sialic acid residues that result in the extension of GalNAc into numerous different structures. Mucin-type O-glycosylation leads to the synthesis of eight different core structures. (A) Core structures of mucin-type O-glycans. (B) Biosynthesis of O-glycan core 1–4 structures, their sialylated forms, and enzymes taking part in these processes. During the neoplastic transformation, shortened O-glycan chains are formed. The most common forms are the Tn and T antigens and their sialylated forms.
Figure 5. Mucin-type O-glycosylation is initiated in the Golgi apparatus by polypeptide N-acetylgalactosaminyltransferases (GalNAc-Ts, twenty isoforms), which initiates the action of numerous glycosyltransferases catalyzing the attachment of galactose, GlcNAc, fucose, and sialic acid residues that result in the extension of GalNAc into numerous different structures. Mucin-type O-glycosylation leads to the synthesis of eight different core structures. (A) Core structures of mucin-type O-glycans. (B) Biosynthesis of O-glycan core 1–4 structures, their sialylated forms, and enzymes taking part in these processes. During the neoplastic transformation, shortened O-glycan chains are formed. The most common forms are the Tn and T antigens and their sialylated forms.
Molecules 28 03436 g005
Figure 6. Types of Lewis (Le) antigens and their sialylated forms. Le antigens are terminal fucosylated carbohydrates present on the cell surface, glycoproteins, or glycolipids. Lewis antigens are covalently attached to the protein by binding to asparagine (Asn) or serine and threonine (Ser/Thr) residues. The presence of Lewis antigens is dependent on the expression of fucosyltransferase (FUT).
Figure 6. Types of Lewis (Le) antigens and their sialylated forms. Le antigens are terminal fucosylated carbohydrates present on the cell surface, glycoproteins, or glycolipids. Lewis antigens are covalently attached to the protein by binding to asparagine (Asn) or serine and threonine (Ser/Thr) residues. The presence of Lewis antigens is dependent on the expression of fucosyltransferase (FUT).
Molecules 28 03436 g006
Figure 7. Disacharide composition of GAGs. Heparan sulfate, chondroitin sulfate, keratan sulfate, and dermatan sulfate have a common tetrasaccharide linkage region consisting of xylose (Xyl), two galactose (Gal) residues, and single glucuronic acid (GlcA). Hyaluronic acid consists of alternately appearing N-acetyloglucosamine (GlcNAc) and iduronic acid (IdoA). The figure presents chemical structures of GAGs (A) and their schematic visualization (B).
Figure 7. Disacharide composition of GAGs. Heparan sulfate, chondroitin sulfate, keratan sulfate, and dermatan sulfate have a common tetrasaccharide linkage region consisting of xylose (Xyl), two galactose (Gal) residues, and single glucuronic acid (GlcA). Hyaluronic acid consists of alternately appearing N-acetyloglucosamine (GlcNAc) and iduronic acid (IdoA). The figure presents chemical structures of GAGs (A) and their schematic visualization (B).
Molecules 28 03436 g007
Figure 8. Summarized role of altered glycosylation in bladder cancer progression.
Figure 8. Summarized role of altered glycosylation in bladder cancer progression.
Molecules 28 03436 g008
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

Wilczak, M.; Surman, M.; Przybyło, M. Altered Glycosylation in Progression and Management of Bladder Cancer. Molecules 2023, 28, 3436. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules28083436

AMA Style

Wilczak M, Surman M, Przybyło M. Altered Glycosylation in Progression and Management of Bladder Cancer. Molecules. 2023; 28(8):3436. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules28083436

Chicago/Turabian Style

Wilczak, Magdalena, Magdalena Surman, and Małgorzata Przybyło. 2023. "Altered Glycosylation in Progression and Management of Bladder Cancer" Molecules 28, no. 8: 3436. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules28083436

Article Metrics

Back to TopTop