Next Article in Journal
Two Faces of Autophagy in the Struggle against Cancer
Next Article in Special Issue
Poly(ADP-Ribose) Polymerase 1 Promotes Inflammation and Fibrosis in a Mouse Model of Chronic Pancreatitis
Previous Article in Journal
Psoriasis: Pathogenesis, Comorbidities, and Therapy Updated
Previous Article in Special Issue
VTRNA2-1: Genetic Variation, Heritable Methylation and Disease Association
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Epigenetic Regulation of Glycosylation in Cancer and Other Diseases

by
Rossella Indellicato
1,* and
Marco Trinchera
2
1
Department of Health Sciences, University of Milan, 20142 Milan, Italy
2
Department of Medicine and Surgery, University of Insubria, 21100 Varese, Italy
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(6), 2980; https://doi.org/10.3390/ijms22062980
Submission received: 2 March 2021 / Revised: 10 March 2021 / Accepted: 12 March 2021 / Published: 15 March 2021
(This article belongs to the Special Issue Epigenetic Mechanisms and Human Pathology)

Abstract

:
In the last few decades, the newly emerging field of epigenetic regulation of glycosylation acquired more importance because it is unraveling physiological and pathological mechanisms related to glycan functions. Glycosylation is a complex process in which proteins and lipids are modified by the attachment of monosaccharides. The main actors in this kind of modification are the glycoenzymes, which are translated from glycosylation-related genes (or glycogenes). The expression of glycogenes is regulated by transcription factors and epigenetic mechanisms (mainly DNA methylation, histone acetylation and noncoding RNAs). This review focuses only on these last ones, in relation to cancer and other diseases, such as inflammatory bowel disease and IgA1 nephropathy. In fact, it is clear that a deeper knowledge in the fine-tuning of glycogenes is essential for acquiring new insights in the glycan field, especially if this could be useful for finding novel and personalized therapeutics.

1. Introduction

In the vast universe of cell biology, there is a very elaborate mechanism capable of carrying out a myriad of functions: glycosylation of proteins and lipids. It consists of the enzymatic attachment of monosaccharides to lipid or protein molecules [1], giving rise to a class of macromolecules called the glycoconjugates (glycoproteins, proteoglycans, mucins, glycosphingolipids, lipopolysaccharides). Glycoconjugates differ in their glycan (the carbohydrate chain) sequence, length, number and position of branches, and type of connections between sugars [1,2]. The complexity of glycan structures is due to the fact that glycan synthesis is not template-driven, unlike linear molecules such as DNA and proteins [3], and is influenced by many variables, including environmental factors, genetic factors (i.e., single nucleotide polymorphisms), transcription factors, protein transports, altered pH values in subcellular sites (especially in the Golgi apparatus), Golgi organizers, ion channels, oxygen concentration, subcellular localization of enzymes, activated monosaccharide donor substrates, and acceptor substrates availability [3,4,5]. In fact, it is estimated that more than 800 genes are involved in the process of glycosylation [6,7], and, among these, about 500 glycosylation-related genes (or glycogenes) are directly involved in glycan assembly, remodeling and degradation, and account for about 2% of the genome [8,9].
Glycosylation has important implications for numerous processes, including the following.
Physical and structural role: a dense layer of glycans (glycocalyx) covers the surface of all the cells allowing to modulate cell–cell, cell–matrix, and cell–molecule interactions, critical to the development of a complex multicellular organism [10,11]; protein folding [9]: it takes place in the endoplasmic reticulum and ensures protein stability and functioning [10]; transcriptional regulation [9,12]: O-GlcNAcylation participates in the epigenetic regulation of gene expression [12,13,14]; interactions between host and pathogenic microorganism [10,11]: viruses and bacteria bind and get access to host cells through glycan receptors [15,16] and are able to evade the immune system decorating themselves with a layer of host-like glycans (the so-called “molecular mimicry” mechanism) [9,17]. Parasites use another strategy to survive, called “glycan gimmickry”, consisting in targeting host–glycan-binding proteins with their glycans [18].
One of the most important features of glycosylation is that it takes place in a cell- and tissue-specific manner [3,6,19]. For example, it plays a very critical role in the development and function of the nervous system, where characteristic glycan structures regulate axon pathfinding, neurite outgrowth, synaptogenesis, neurotransmission, and other neuronal processes [20]. Moreover, neural-specific glycans are required to carry out high-order brain functions, including learning/memory and the formation of neural networks [6,21].
This review briefly describes the principal epigenetic mechanisms that participate in gene expression and how they are related to glycosylation. A short overview illustrates the normal physiology of epigenetic regulation of glycogenes, and a more extensive section is devoted to cancer and other diseases.

2. Epigenetic Regulation of Glycosylation

Since glycosylation is cell- and tissue-specific, but the DNA template is always the same in every cell of an organism, a master player able to regulate gene expression is required: epigenetics. It was defined as a “stably heritable phenotype resulting from changes in a chromosome without alterations in the DNA sequence” [22]. It is a process that takes place during the differentiation of somatic cells, as well as in response to environmental changes [23]. Epigenetics is one of the reasons why the cells of an organism have a different phenotype, even if they share the same DNA sequence. Epigenetics acts in order to regulate gene expression mainly through the following three mechanisms [24,25,26,27,28].
DNA methylation: it occurs in CpG rich-regions called CpG islands, where CpG dinucleotides tend to cluster. Frequently, methylation of these regions represses gene transcription and expression [29], while unmethylated regions promote gene activation. Basically, methylation requires a methyl group (CH3) covalently attached to the 5-carbon of the cytosine residue (5mC) in the CpG site [26,30,31]. This action is carried out by DNA methyltransferases (DNMTs), and the CH3 group physically interrupts the binding between the proper transcription factor and its recognition sequence. Moreover, gene silencing upon methylation can also occur when methyl-CpG-binding proteins bind to the methylated DNA and recruit co-repressor molecules, such as histone deacetylases, to induce chromatin structure condensation [32].
Histone modifications: the two main modifications that can occur on histones are methylation and acetylation. They alter chromatin structure; in fact, euchromatin (actively transcribed) is characterized by high levels of acetylation and di/trimethylation of H3K4, H3K36 and H3K79 [29,33], while heterochromatin (transcriptionally inactive) is characterized by low levels of acetylation and high levels of H3K9, H3K27 and H4K20 methylation [29]. O-GlcNAcylation is another form of histone modification [12,13,14], and it is the perfect example of how epigenetics and glycosylation are tangled together: epigenetics regulates glycogenes expression, and glycosylation participates in epigenetic regulation.
Noncoding RNAs (ncRNAs): a wide range of RNA molecules belongs to this category, which is included transfer RNA (tRNA), ribosomal RNA (rRNA), microRNA (miRNA), small interfering RNA (siRNA), piwi-interacting RNA (piRNA), small nuclear RNA (snRNA), small nucleolar RNA (snoRNA), and long noncoding RNA (lcnRNA) [34,35,36]. The most studied are miRNAs and lncRNAs. miRNAs are characterized by a short sequence of nucleotides (about 20–30), play a role in gene silencing [31] targeting mRNA 3’UTR regions, and inhibit protein translation or enhance mRNA degradation [26,34]. To date, over 80 glycogenes have been identified as miRNAs targets [37]. LncRNAs are longer than miRNAs (up to more than 100 kilobases) [25] and can function both by repressing or activating gene expression [34], acting as molecular chaperones or scaffolds for various chromatin regulators [31].
Of course, epigenetics is not the only mechanism involved in the transcription of glycogenes. In fact, also transcription factors binding to a gene promoter and enhancer elements are fundamental for this purpose [38]. A prominent example is given by transcription factors hepatocyte nuclear factor 1α (HNF1α) and its downstream target hepatocyte nuclear factor 4α (HNF4α), which were proven by Lauc et al. in the first genome-wide association study (GWAS) of protein glycosylation to regulate the expression of key fucosyltransferases and fucose biosynthesis genes. This finding revealed a new role for HNF1α as a master transcriptional regulator of multiple stages in the fucosylation process [39].

3. Physiological Aspects of Epigenetic Regulation of Glycosylation

Before going deeper into the field of epigenetic regulation associated with pathological glycosylations, it is worth making a brief presentation on how glycogenes are regulated by epigenetics when it comes to normal physiology. Research in this field is just at the beginning, and there are little available data at present. Yet, some prominent studies carried out on the brain elucidated the significance/importance of specific neural glycans since their fine-tuning is pivotal for high-order brain functions (i.e., learning/memory, the formation of the neural network, myelination), and their dysregulation leads to various neurological disorders [40]. The first research group focused on a glycosyltransferase called N-acetylglucosaminyltransferase IX (MGAT5B), that catalyzes the transfer of N-acetylglucosamine (GlcNAc) to the 6-OH position of the mannose residues of GlcNAcβ1,2-Manα on both the α1,3- and α1,6-linked mannose arms in the core structure of N-glycans. It is also responsible for the transfer of GlcNAc in β1,6-linkage to O-mannosyl glycans. The gene encoding this enzyme is MGAT5B, which is exclusively expressed in the brain [41], and it has been proved that it is under the control of neural cell-specific histone modification: active chromatin marks like H3K9ac and H3K4me3 were found in the mouse brain, and repressive chromatin marks like H3K27me3 and H3K9me2 were detected in mouse kidney and liver [42]. The second research group studied two glycosyltransferases involved in lipid glycosylation: B4GALNT1 and ST8SIA1. They are both involved in the biosynthesis of gangliosides, a class of sialic acid-containing glycosphingolipids particularly abundant in the central nervous system. Their peculiarity consists in being ontogenically regulated, and in fact, they are more expressed in the adult brain. Experiments on mice showed that brain gangliosides shift from the simpler ones (GM3 and GD3) in early phases of life to more complex ones during development (GM1, GD1a, GT1a, and GT1b) and that expression of B4galnt1 (prevalently) and St8Sia1, both involved in this shifting, increased, due to histone H3 and H4 acetylation [43,44,45].

4. Epigenetic Regulation of Glycosylation in Cancer

The majority of the studies of epigenetic regulation of glycogenes are about cancer. It is well-established that aberrant glycosylation is one of the hallmarks of tumoral cells [46,47,48] and that these changes are nonrandom: in cancer advancement, only the fittest cells survive, and specific glycan changes are selected for tumor progression [47]. In fact, transcription of a gene tends to be constitutively repressed in cancer, when its epigenetic silencing is advantageous for promoting cancer progression [49]. In particular, incomplete synthesis and neo-synthesis processes are the two principal mechanisms associated with alterations of carbohydrate structures during tumor progression [50]. Incomplete synthesis refers to truncated glycosylation that produces the Tn antigen in mucin-type O-glycans, and neo-synthesis produces abnormal glycosylation patterns such as sialyl Lewis X (sLex) [51,52]. Tn antigen and sLex are typical of lymphocytes and help in their extravasation from the blood, while in cancer, they facilitate metastatic spread [1,53]. Novel glycan structures also have the role of enabling cancer cells to evade the host immune response [15,51,54,55].
All these modifications in glycosylation during the tumoral event are carried out by genetic, epigenetic, metabolic, inflammatory and environmental mechanisms [52], but this review focuses only on epigenetic alterations that affect glycogenes during carcinogenesis. The first studies were based on the methylation status of the promoter region, using demethylating agents such as 5-aza-2-deoxycytidine (5-aza-dC) [56,57,58]. Later on, it was discovered that hypermethylation of a promoter could not be sufficient to maintain gene silencing; in fact, even upon a treatment, only a partial restoration was achieved, and this was due to other epigenetic marks involved such as repressive histone modifications [59,60].
Epigenetic modifications of glycogenes in cancer were extensively reviewed by Dall’Olio and Trinchera [48], and the most recent ones are updated in Table 1, but it is most likely that the number of glycogenes epigenetically regulated in cancer is going to grow in the next years.
Below, we present some prominent examples of epigenetically modified glycogenes involved in tumor progression.

4.1. C1GALT1C1

One of the hallmarks of carcinoma mucins is their incomplete glycosylation. The addition of the first N-acetylgalactosamine (GalNAc) O-linked to serine or threonine of mucin-type glycans leads to the formation of the Tn antigen, which is a well-known cancer-associated structure [48]. On this first GalNAc, a Gal residue could be added by core 1 β1,3-galactosyltransferase (C1GALT1 or T-synthase), which needs the molecular chaperone C1GALT1C1 (encoded by C1GALT1C1 gene) for its functioning. This leads to the formation of the T antigen. At the same time, another enzyme called sialyltransferase ST6GALNAC1 could act on the Tn antigen, adding a residue of α2,6-linked sialic acid, resulting in the formation of the sialyl-Tn (STn) antigen and blocking further chain elongation [52,104]. During carcinogenesis, C1GALT1C1 expression could be downregulated due to genetic mutations [105] or, more interestingly, to an epigenetic modification: hypermethylation of the promoter leads to the silencing of C1GALT1C1 and to the accumulation of the cancer-associated Tn and STn antigens [106,107]. The secreted mucins expressing these antigens often appear in the bloodstream of patients with cancer and are associated with invasion since they potentiate migration of tumor cells through the inhibition of cell–cell contacts [108,109]. Moreover, these carcinoma mucins often decorate the tumor surface, creating clustered sites for antibody attachment, thereby improving their activity as tumor immunogens. In fact, since these glycans infrequently occur in normal tissues, they provoke immune responses in patients, a property that has been exploited for potential immunotherapy [47,108].

4.2. B4GALNT2 (β-1,4-N-acetyl-galactosaminyltransferase 2)

Sda carbohydrate (GalNAcβ1,4[Sialα2,3)Galβ1,4GlcNAc) belongs to the category of the histo-blood group antigens. They were initially found on the erythrocyte surface, but it was soon discovered that this group of antigens is widely distributed in many epithelial tissues (colon, stomach, kidney, oocyte) and secretions (urine, serum, saliva, milk) [110,111,112], and play roles in the regulation of physiological mechanisms. In particular, studies in murine models showed that the Sda antigen is involved in the processes of hemostasis [113,114] and reproduction [115,116]. The last step in the biosynthesis of the Sda antigen is catalyzed by GalNAc transferase B4GALNT2 (also known as Sda synthase), which adds an N-acetylgalactosamine to a terminal α2,3-sialylated galactose residue [117]. Experiments on guinea-pigs [118] and rats [119] proved that the B4GALNT2 gene could be ontogenically regulated; in fact, the enzyme was absent at birth and increased with age. More information on Sda synthase is known as far as concern cancer since the expression and activity of this enzyme are downregulated in gastrointestinal cancer leading to a complete loss of the antigen [112,117,120]. The reason for such differential expression was attributed to the hypermethylation of the B4GALNT2 promoter [120,121], which is embedded in CpG islands. In work by Kawamura et al. [54], the B4GALNT2 gene was found methylated in about one-half of the gastric cancer cases taken under consideration and in the majority of gastric and colon cancer cell lines. They used a demethylating agent, 5-aza-dC, to attempt a recovery of B4GALNT2 transcription, but it worked only partially, inducing a very weak expression of both the glycoenzyme and the Sda antigen. Human colon cancer cells were also treated with the histone acetylase inhibitor butyrate, but neither a slight recovery of the Sda antigen nor that of B4GALNT2 was observed. According to these results, the mechanism of B4GALNT2 downregulation in cancer deserves further investigation [122].

4.3. B3GALT5

B3GALT5 is one of the glycoenzymes involved in the synthesis of type 1 chain carbohydrate antigens, namely the Lewis a (Lea) trisaccharide, the Lewis b (Leb) tetrasaccharide and the sialyl Lewis a (sLea) tetrasaccharide [123,124]. Lea and Leb are involved in various biological contexts, such as microbial adhesion and cancer [125], whereas sLea has been proven to be specifically an E-selectin ligand, favoring the metastatic process and angiogenesis during cancer development [124,126,127]. The peculiarity of B3GALT5 is that its expression is regulated by two promoters: the LTR and native promoters [128].
The LTR promoter, which has retroviral origins and is activated through hepatocyte nuclear factor HNF1α and HNF1β [129,130], is mainly active in the organs of the gastrointestinal tract (such as the colon, stomach, and pancreas). However, HNF1α and HNF1β are not able to modulate transcription, which depends on distal regulatory elements that are active when methylated. In fact, LTR and proximal sequences lack CpG islands, suggesting that methylation-sensitive DNA sequences reside outside the LTR region, presumably distant from the promoter, where they act as potential epigenetic regulators of transcription [130,131].
In the mammary glands, thymus and trachea, as well as in some human cancer cell lines, transcription is mainly driven by a native promoter, which is sensitive to nuclear factor NF-Y [124] and is located nearby two CpG islands [132] epigenetically regulated through methylation [130]. As for the LTR promoter, NF-Y is unable to regulate transcription, which depends on the methylation of the regulatory elements [130,131]. Moreover, histone modification is another mechanism involved in the regulation. High expression of the native transcript is associated with active histone marks (H3K4me3, H3K79me2, H3K9Ac, and H3K9-14Ac), while low levels of the transcript are associated with repressive histone marks (H3K27me2 and H4K20me3) [132].
The differential regulation of B3GALT5 was studied in particular in the pancreas and colon, comparing normal and tumoral tissues [130,131,132]. B3GALT5 is strongly downregulated in colon cancer with respect to the normal mucosa [133,134], and the silencing of the gene is due to the opposite but synergic behavior of the two promoters: hypomethylation of the distant sequences of the LTR promoter and hypermethylation of the native promoter [124]. In the pancreas, both normal and cancer tissues have very low levels of methylation in the native promoter, and the levels of B3GALT5 LTR transcript were similar to those of the native transcript, without difference between normal and tumoral specimens [131,132].

5. Epigenetic Regulation of Glycosylation in Other Diseases

Dysregulation of glycosylation is associated not only with cancer but also with a number of other diseases. The majority of them are caused by genetic mutations such as congenital disorders of glycosylation, diabetes, cardiovascular, immunological, autoimmune (rheumatoid arthritis, Sjögren’s syndrome, systemic lupus erythematosus) and infectious disorders [2,7]. Other diseases are associated with both genetic and epigenetic modifications, such as inflammatory bowel disease (IBD), IgA1 nephropathy (IgAN), and neurodegenerative diseases, briefly reviewed below. Since this is a recent field of research, it is highly probable that the disorders associated with aberrant epigenetic regulation of glycosylation will increase over time, giving a better insight into the disease pathogenesis.

5.1. Inflammatory Bowel Disease

IBD is a chronic inflammatory disorder that affects the gastrointestinal tract and comprises two clinical syndromes: Crohn’s disease (CD) and ulcerative colitis (UC) [135,136]. These diseases have unknown etiology, and there is insufficient information about pathogenesis, but it is believed that a complex interaction of genetic, epigenetic, microbial, environmental and immunological factors are involved [137]. In particular, several studies have evaluated the epigenetic status of IBD patients using candidate gene strategies [138,139,140,141,142] or epigenome-wide association studies [143,144,145,146], trying to elucidate IBD pathogenesis [147]. Cooke and colleagues [143] collected rectal biopsies and identified some glycogenes that have a differential methylation status between patients with CD and UC (inflamed vs. non-inflamed) and healthy controls. In fact, in inflamed UC vs. controls, B3GALT2, GFPT1 and GBGT1 have increased methylation; in inflamed CD vs. controls, GFPT1 and GBGT1 have increased methylation and FUT2 has a decreased methylation; in non-inflamed CD vs. controls, FUT7 and FGF23 have decreased methylation. These altered methylation levels correlated with the development of IBD, contributing to better understand IBD pathogenesis. Another study conducted by Klasíc and colleagues evaluated the methylation status of β-1,4-mannosyl-glycoprotein 4-β-N-acetylglucosaminyltransferase (MGAT3) promoter in CD3+ T cells isolated from the inflamed mucosa of UC patients. They found that the MGAT3 promoter was hypermethylated in UC patients compared with healthy controls. This kind of deregulation might lead to an increase of the proinflammatory properties of IgG through a decrease in galactosylation and sialylation and an increase of bisecting GlcNAc on digalactosylated glycans, thus suggesting a functional role of MGAT3 in IBD pathogenesis [148].

5.2. IgA1 Nephropathy

Several studies led to the conclusion that inhibition of genes involved in glycosylation by miRNAs plays a role in the pathogenesis of IgA1 nephropathy (IgAN), which is characterized by the aggregation of aberrantly glycosylated IgA1 molecules, leading to the synthesis of inflammatory cytokines and glomerulonephritis. The first study conducted by Serino and colleagues brought to the attention the role of miR-148b. It was demonstrated that peripheral blood mononuclear cells (PBMCs) of patients with IgAN show a higher miR-148b expression level compared to healthy controls, and this upregulation leads to a lower C1GALT1 expression. C1GALT1 is involved in the O-glycosylation of the IgA1 heavy chain hinge-region, and without the expression of the gene, hinge-region displays a deficiency of galactose [149]. This group also demonstrated that GALNT2 (UDP-GalNAc: polypeptide N-acetylgalactosaminyltransferase 2) is the target of miRNA let-7b, similarly to C1GALNT1 and miR-148b. GALNT2 initiates the addition of GalNAc to serine or threonine residues of the IgA1 hinge-region. Let-7b was significantly upregulated in IgAN patients, and, as a consequence, GALNT2 levels became lower [150].
Recently, another miRNA was found to be involved in the aberrant glycosylation of IgAN, but not in a direct way. In fact, the direct target of miR-98–5p is CCL3 (C–C motif chemokine ligand 3), which can change the level of Th1 and Th2 cytokines in many diseases. Th1 and Th2 cytokines participate in the pathogenesis of IgAN. In this case, only IL-6 was found to be upregulated in PBMCs of IgAN patients compared to healthy controls. IL-6 reduces the galactosylation of IgA1 by decreasing the expression of C1GALT1 [151].
Another miRNA able to indirectly modify glycosylation in IgAN is miR-374b. The target of this miRNA is C1GALT1C1, which is required for the activity of C1GALT1, and that is downregulated in the B cells isolated from IgAN patients, leading to abnormal glycosylation of IgA1 [152]. Furthermore, miR-320 (upregulated in the renal tissues of IgAN patients) targets C1GALT1C1, which in fact is downregulated in the same patients [67].

5.3. Neurodegenerative Diseases

A critical role of glycosylation is emerging in the field of neuron homeostasis and related neurodegenerative diseases. It is well-known that several glycoconjugates and related processing enzymes, namely glycosyltransferases and glycosidases, are strictly and specifically expressed in the central nervous system, and a set of specific glycosylations, such as ganglioside biosynthesis and GlcNAcylation, are strongly associated with various neurodegenerative disorders [153]. At present, the majority of data arise from genetic defects. Both KO mice of ganglioside glycosyltransferases and congenital disorders of glycosylation affecting ganglioside biosynthesis indicated that ganglioside dysregulation gives rise to neuroinflammation, functional impairment, and in turn neurodegeneration [153,154]. Moreover, non-genetic derangement of glycosyltransferases was associated with Parkinson’s disease (reduced B3GALT4 and ST3GAL2, increased OGT, O-linked GlcNAc transferase), Huntington disease (reduced ST3GAL5, ST3GAL2, ST8SIA3, B4GALNT1), Alzheimer’s disease, and even amyotrophic lateral sclerosis (general ganglioside overexpression) [153,155]. In multiple sclerosis, an autoimmune disease causing inflammation of the central nervous system, glycoproteins are candidate targets of autoreactivity, and glycosyltransferases such as MGAT1, MGAT5 and B4GALT6 are reported as deregulated genes [7,156]. A key role of epigenetic regulation is reported in multiple sclerosis [7] and suggests that such mechanism could be the common trait of some of the other neurodegenerative disorders associated with deranged glycosylation.

6. Concluding Remarks

Epigenetic regulation of glycosylation is an emerging and relatively recent field of research. The analysis of glycogenes expression due to epigenetic mechanisms started with the use of demethylating agents in cancer cell cultures [57,157], and it has become more important over the years. At present, several pathological mechanisms associated with cancer and other diseases are known to be caused by epigenetic dysregulation of glycosylation, as reported in this review. We also reported relevant studies illustrating how epigenetics controls glycosylation under physiological conditions [3,6,44,45,122,158,159]. Altogether these findings help to unravel the roles and functions of glycans, which are candidate targets in the field of personalized medicine through drugs-based inhibitors of their synthesis, glycan antagonists, and glycan-function modulators [52]. In this regard, it is also worth recalling the critical interplays involving glycosylation, epigenetics, and hypoxia since one controls the other. This suggests that drugs affecting glycosylation through epigenetic regulation could be relevant in cancer developing chemo-resistance [26].

Funding

This research was supported by “Aldo Ravelli” Center for Neurotechnology and Experimental Brain Therapeutics (M.T.) and from the University of Insubria (to M.T.). R.I. was supported by the PhD program in Translational Medicine of the University of Milan. The APC was funded by the University of Milan “Biblioteca digitale” (to R.I.).

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

Genes and proteins are named according to the HUGO recommendations.
5-aza-dC5-Aza-2-deoxycytidine
B3GALT5β1,3-Galactosyltransferase isoenzyme 5
B4GALNT2β-1,4-N-Acetylgalactosaminyltransferase 2
C1GALT1Core 1 β1,3-galactosyltransferase
C1GALT1C1C1GALT1-specific chaperone 1
CAFCancer-associated fibroblast
CDCrohn’s disease
FGF23Fibroblast growth factor 23
GalNAcN-Acetylgalactosamine
GBGT1Globoside α-1,3-N-Acetylgalactosaminyltransferase 1
GALNT2UDP-GalNAc: polypeptide N-acetylgalactosaminyltransferase 2
GFPT1Glutamine-fructose-6-phosphate transaminase 1
GlcNAcN-Acetylglucosamine
HNF1αHepatocyte nuclear factor 1α
HNF4αHepatocyte nuclear factor 4α
IBDInflammatory bowel disease
IgANIgA1 nephropathy
MGAT3β-1,4-Mannosyl-glycoprotein 4-β-N-acetylglucosaminyltransferase
OGTO-Linked GlcNAc transferase
PBMCPeripheral blood mononuclear cell
lncRNALong noncoding RNA
miRNAMicroRNA
ncRNANoncoding RNA
piRNAPiwi-interacting RNA
rRNARibosomal RNA
siRNASmall interfering RNA
snRNASmall nuclear RNA
snoRNASmall nucleolar RNA
tRNATransfer RNA
sLexSialyl Lewis X
UCUlcerative colitis

References

  1. Reily, C.; Stewart, T.J.; Renfrow, M.B.; Novak, J. Glycosylation in health and disease. Nat. Rev. Nephrol. 2019, 15, 346–366. [Google Scholar] [CrossRef]
  2. Lauc, G.; Zoldos, V. Protein glycosylation—An evolutionary crossroad between genes and environment. Mol. Biosyst. 2010, 6, 2373–2379. [Google Scholar] [CrossRef] [PubMed]
  3. Lauc, G.; Vojta, A.; Zoldos, V. Epigenetic regulation of glycosylation is the quantum mechanics of biology. Biochim. Biophys. Acta 2014, 1840, 65–70. [Google Scholar] [CrossRef] [PubMed]
  4. Klasic, M.; Kristic, J.; Korac, P.; Horvat, T.; Markulin, D.; Vojta, A.; Reiding, K.R.; Wuhrer, M.; Lauc, G.; Zoldos, V. DNA hypomethylation upregulates expression of the MGAT3 gene in HepG2 cells and leads to changes in N-glycosylation of secreted glycoproteins. Sci. Rep. 2016, 6, 24363. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Nairn, A.V.; York, W.S.; Harris, K.; Hall, E.M.; Pierce, J.M.; Moremen, K.W. Regulation of glycan structures in animal tissues: Transcript profiling of glycan-related genes. J. Biol. Chem. 2008, 283, 17298–17313. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Kizuka, Y.; Kitazume, S.; Okahara, K.; Villagra, A.; Sotomayor, E.M.; Taniguchi, N. Epigenetic regulation of a brain-specific glycosyltransferase N-acetylglucosaminyltransferase-IX (GnT-IX) by specific chromatin modifiers. J. Biol. Chem. 2014, 289, 11253–11261. [Google Scholar] [CrossRef] [Green Version]
  7. Stambuk, T.; Klasic, M.; Zoldos, V.; Lauc, G. N-glycans as functional effectors of genetic and epigenetic disease risk. Mol. Aspects Med. 2020, 100891. [Google Scholar] [CrossRef]
  8. Ng, B.G.; Freeze, H.H. Perspectives on Glycosylation and Its Congenital Disorders. Trends Genet. 2018, 34, 466–476. [Google Scholar] [CrossRef]
  9. Springer, S.A.; Gagneux, P. Glycan evolution in response to collaboration, conflict, and constraint. J. Biol. Chem. 2013, 288, 6904–6911. [Google Scholar] [CrossRef] [Green Version]
  10. Varki, A. Biological roles of glycans. Glycobiology 2017, 27, 3–49. [Google Scholar] [CrossRef] [Green Version]
  11. Varki, A.; Kornfeld, S. Historical Background and Overview. In Essentials of Glycobiology, 3rd. ed.; Varki, A.C.R., Esko, J.D., Eds.; Cold Spring Harbor Laboratory Press: Cold Spring Harbor, NY, USA, 2015; Chapter 1. [Google Scholar] [CrossRef]
  12. Hanover, J.A.; Krause, M.W.; Love, D.C. Bittersweet memories: Linking metabolism to epigenetics through O-GlcNAcylation. Nat. Rev. Mol. Cell Biol. 2012, 13, 312–321. [Google Scholar] [CrossRef]
  13. Leturcq, M.; Lefebvre, T.; Vercoutter-Edouart, A.S. O-GlcNAcylation and chromatin remodeling in mammals: An up-to-date overview. Biochem. Soc. Trans. 2017, 45, 323–338. [Google Scholar] [CrossRef]
  14. Lewis, B.A.; Hanover, J.A. O-GlcNAc and the epigenetic regulation of gene expression. J. Biol. Chem. 2014, 289, 34440–34448. [Google Scholar] [CrossRef] [Green Version]
  15. Lauc, G.; Zoldos, V. Epigenetic regulation of glycosylation could be a mechanism used by complex organisms to compete with microbes on an evolutionary scale. Med. Hypotheses 2009, 73, 510–512. [Google Scholar] [CrossRef] [PubMed]
  16. Varki, A.; Gagneux, P. Biological Functions of Glycans. In Essentials of Glycobiology, 3rd. ed.; Varki, A., Cummings, R.D., Esko, J.D., Stanley, P., Hart, G.W., Aebi, M., Darvill, A.G., Kinoshita, T., Packer, N.H., Eds.; Cold Spring Harbor Press: Cold Spring Harbor, NY, USA, 2015; Chapter 7; pp. 77–88. [Google Scholar] [CrossRef]
  17. Corfield, A.P.; Berry, M. Glycan variation and evolution in the eukaryotes. Trends Biochem. Sci. 2015, 40, 351–359. [Google Scholar] [CrossRef] [PubMed]
  18. Van Die, I.; Cummings, R.D. Glycan gimmickry by parasitic helminths: A strategy for modulating the host immune response? Glycobiology 2010, 20, 2–12. [Google Scholar] [CrossRef] [Green Version]
  19. Moremen, K.W.; Tiemeyer, M.; Nairn, A.V. Vertebrate protein glycosylation: Diversity, synthesis and function. Nat. Rev. Mol. Cell Biol. 2012, 13, 448–462. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Williams, S.E.; Mealer, R.G.; Scolnick, E.M.; Smoller, J.W.; Cummings, R.D. Aberrant glycosylation in schizophrenia: A review of 25 years of post-mortem brain studies. Mol. Psychiatry 2020, 25, 3198–3207. [Google Scholar] [CrossRef] [PubMed]
  21. Weinhold, B.; Seidenfaden, R.; Rockle, I.; Muhlenhoff, M.; Schertzinger, F.; Conzelmann, S.; Marth, J.D.; Gerardy-Schahn, R.; Hildebrandt, H. Genetic ablation of polysialic acid causes severe neurodevelopmental defects rescued by deletion of the neural cell adhesion molecule. J. Biol. Chem. 2005, 280, 42971–42977. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Berger, S.L.; Kouzarides, T.; Shiekhattar, R.; Shilatifard, A. An operational definition of epigenetics. Genes Dev. 2009, 23, 781–783. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Lind, M.I.; Spagopoulou, F. Evolutionary consequences of epigenetic inheritance. Heredity 2018, 121, 205–209. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Bannister, A.J.; Kouzarides, T. Regulation of chromatin by histone modifications. Cell Res. 2011, 21, 381–395. [Google Scholar] [CrossRef]
  25. Cavalli, G.; Heard, E. Advances in epigenetics link genetics to the environment and disease. Nature 2019, 571, 489–499. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Greville, G.; McCann, A.; Rudd, P.M.; Saldova, R. Epigenetic regulation of glycosylation and the impact on chemo-resistance in breast and ovarian cancer. Epigenetics 2016, 11, 845–857. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Tsai, Y.T.; Yu, R.K. Epigenetic activation of mouse ganglioside synthase genes: Implications for neurogenesis. J. Neurochem. 2014, 128, 101–110. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Zoldos, V.; Horvat, T.; Novokmet, M.; Cuenin, C.; Muzinic, A.; Pucic, M.; Huffman, J.E.; Gornik, O.; Polasek, O.; Campbell, H.; et al. Epigenetic silencing of HNF1A associates with changes in the composition of the human plasma N-glycome. Epigenetics 2012, 7, 164–172. [Google Scholar] [CrossRef] [Green Version]
  29. Portela, A.; Esteller, M. Epigenetic modifications and human disease. Nat. Biotechnol. 2010, 28, 1057–1068. [Google Scholar] [CrossRef]
  30. Skvortsova, K.; Iovino, N.; Bogdanovic, O. Functions and mechanisms of epigenetic inheritance in animals. Nat. Rev. Mol. Cell Biol. 2018, 19, 774–790. [Google Scholar] [CrossRef] [Green Version]
  31. Dawson, M.A.; Kouzarides, T. Cancer epigenetics: From mechanism to therapy. Cell 2012, 150, 12–27. [Google Scholar] [CrossRef] [Green Version]
  32. Oda, S.; Fukami, T.; Yokoi, T.; Nakajima, M. Epigenetic regulation is a crucial factor in the repression of UGT1A1 expression in the human kidney. Drug Metab. Dispos. 2013, 41, 1738–1743. [Google Scholar] [CrossRef] [Green Version]
  33. Norouzitallab, P.; Baruah, K.; Vanrompay, D.; Bossier, P. Can epigenetics translate environmental cues into phenotypes? Sci. Total Environ. 2019, 647, 1281–1293. [Google Scholar] [CrossRef]
  34. Hombach, S.; Kretz, M. Non-coding RNAs: Classification, Biology and Functioning. Adv. Exp. Med. Biol. 2016, 937, 3–17. [Google Scholar] [CrossRef] [PubMed]
  35. Huttenhofer, A.; Schattner, P.; Polacek, N. Non-coding RNAs: Hope or hype? Trends Genet. 2005, 21, 289–297. [Google Scholar] [CrossRef]
  36. Wei, J.W.; Huang, K.; Yang, C.; Kang, C.S. Non-coding RNAs as regulators in epigenetics (Review). Oncol. Rep. 2017, 37, 3–9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Thu, C.T.; Mahal, L.K. Sweet Control: MicroRNA Regulation of the Glycome. Biochemistry 2020, 59, 3098–3110. [Google Scholar] [CrossRef] [PubMed]
  38. Neelamegham, S.; Mahal, L.K. Multi-level regulation of cellular glycosylation: From genes to transcript to enzyme to structure. Curr. Opin. Struct. Biol. 2016, 40, 145–152. [Google Scholar] [CrossRef] [Green Version]
  39. Lauc, G.; Essafi, A.; Huffman, J.E.; Hayward, C.; Knezevic, A.; Kattla, J.J.; Polasek, O.; Gornik, O.; Vitart, V.; Abrahams, J.L.; et al. Genomics meets glycomics-the first GWAS study of human N-Glycome identifies HNF1alpha as a master regulator of plasma protein fucosylation. PLoS Genet. 2010, 6, e1001256. [Google Scholar] [CrossRef]
  40. Kizuka, Y.; Nakano, M.; Miura, Y.; Taniguchi, N. Epigenetic regulation of neural N-glycomics. Proteomics 2016, 16, 2854–2863. [Google Scholar] [CrossRef]
  41. Inamori, K.; Endo, T.; Gu, J.; Matsuo, I.; Ito, Y.; Fujii, S.; Iwasaki, H.; Narimatsu, H.; Miyoshi, E.; Honke, K.; et al. N-Acetylglucosaminyltransferase IX acts on the GlcNAc beta 1,2-Man alpha 1-Ser/Thr moiety, forming a 2,6-branched structure in brain O-mannosyl glycan. J. Biol. Chem. 2004, 279, 2337–2340. [Google Scholar] [CrossRef] [Green Version]
  42. Kizuka, Y.; Kitazume, S.; Yoshida, M.; Taniguchi, N. Brain-specific expression of N-acetylglucosaminyltransferase IX (GnT-IX) is regulated by epigenetic histone modifications. J. Biol. Chem. 2011, 286, 31875–31884. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Itokazu, Y.; Tsai, Y.T.; Yu, R.K. Epigenetic regulation of ganglioside expression in neural stem cells and neuronal cells. Glycoconj. J. 2017, 34, 749–756. [Google Scholar] [CrossRef]
  44. Itokazu, Y.; Wang, J.; Yu, R.K. Gangliosides in Nerve Cell Specification. Prog. Mol. Biol. Transl. Sci. 2018, 156, 241–263. [Google Scholar] [CrossRef]
  45. Suzuki, Y.; Yanagisawa, M.; Ariga, T.; Yu, R.K. Histone acetylation-mediated glycosyltransferase gene regulation in mouse brain during development. J. Neurochem. 2011, 116, 874–880. [Google Scholar] [CrossRef] [Green Version]
  46. Tuccillo, F.M.; de Laurentiis, A.; Palmieri, C.; Fiume, G.; Bonelli, P.; Borrelli, A.; Tassone, P.; Scala, I.; Buonaguro, F.M.; Quinto, I.; et al. Aberrant glycosylation as biomarker for cancer: Focus on CD43. BioMed Res. Int. 2014, 2014, 742831. [Google Scholar] [CrossRef] [Green Version]
  47. Varki, A.; Kannagi, R.; Toole, B.; Stanley, P. Glycosylation Changes in Cancer. In Essentials of Glycobiology, 3rd. ed.; Varki, A., Cummings, R.D., Esko, J.D., Stanley, P., Hart, G.W., Aebi, M., Darvill, A.G., Kinoshita, T., Packer, N.H., Eds.; Cold Spring Harbor Press: Cold Spring Harbor, NY, USA, 2015; pp. 597–609, Chapter 47. [Google Scholar] [CrossRef]
  48. Dall’Olio, F.; Trinchera, M. Epigenetic Bases of Aberrant Glycosylation in Cancer. Int. J. Mol. Sci. 2017, 18, 998. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Kannagi, R.; Sakuma, K.; Miyazaki, K.; Lim, K.T.; Yusa, A.; Yin, J.; Izawa, M. Altered expression of glycan genes in cancers induced by epigenetic silencing and tumor hypoxia: Clues in the ongoing search for new tumor markers. Cancer Sci. 2010, 101, 586–593. [Google Scholar] [CrossRef] [PubMed]
  50. Hakomori, S.; Kannagi, R. Glycosphingolipids as tumor-associated and differentiation markers. J. Natl. Cancer Inst. 1983, 71, 21. [Google Scholar]
  51. Kannagi, R.; Yin, J.; Miyazaki, K.; Izawa, M. Current relevance of incomplete synthesis and neo-synthesis for cancer-associated alteration of carbohydrate determinants—Hakomori’s concepts revisited. Biochim. Biophys. Acta 2008, 1780, 525–531. [Google Scholar] [CrossRef]
  52. Pinho, S.S.; Reis, C.A. Glycosylation in cancer: Mechanisms and clinical implications. Nat. Rev. Cancer 2015, 15, 540–555. [Google Scholar] [CrossRef] [PubMed]
  53. Magalhaes, A.; Duarte, H.O.; Reis, C.A. Aberrant Glycosylation in Cancer: A Novel Molecular Mechanism Controlling Metastasis. Cancer Cell 2017, 31, 733–735. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Kawamura, Y.I.; Toyota, M.; Kawashima, R.; Hagiwara, T.; Suzuki, H.; Imai, K.; Shinomura, Y.; Tokino, T.; Kannagi, R.; Dohi, T. DNA hypermethylation contributes to incomplete synthesis of carbohydrate determinants in gastrointestinal cancer. Gastroenterology 2008, 135, 142–151. [Google Scholar] [CrossRef]
  55. Kim, Y.S.; Deng, G. Aberrant expression of carbohydrate antigens in cancer: The role of genetic and epigenetic regulation. Gastroenterology 2008, 135, 305–309. [Google Scholar] [CrossRef]
  56. Kominato, Y.; Hata, Y.; Takizawa, H.; Matsumoto, K.; Yasui, K.; Tsukada, J.; Yamamoto, F. Alternative promoter identified between a hypermethylated upstream region of repetitive elements and a CpG island in human ABO histo-blood group genes. J. Biol. Chem. 2002, 277, 37936–37948. [Google Scholar] [CrossRef] [Green Version]
  57. Kominato, Y.; Hata, Y.; Takizawa, H.; Tsuchiya, T.; Tsukada, J.; Yamamoto, F. Expression of human histo-blood group ABO genes is dependent upon DNA methylation of the promoter region. J. Biol. Chem. 1999, 274, 37240–37250. [Google Scholar] [CrossRef] [Green Version]
  58. Miyazaki, K.; Ohmori, K.; Izawa, M.; Koike, T.; Kumamoto, K.; Furukawa, K.; Ando, T.; Kiso, M.; Yamaji, T.; Hashimoto, Y.; et al. Loss of disialyl Lewis(a), the ligand for lymphocyte inhibitory receptor sialic acid-binding immunoglobulin-like lectin-7 (Siglec-7) associated with increased sialyl Lewis(a) expression on human colon cancers. Cancer Res. 2004, 64, 4498–4505. [Google Scholar] [CrossRef] [Green Version]
  59. Jacinto, F.V.; Ballestar, E.; Esteller, M. Impaired recruitment of the histone methyltransferase DOT1L contributes to the incomplete reactivation of tumor suppressor genes upon DNA demethylation. Oncogene 2009, 28, 4212–4224. [Google Scholar] [CrossRef] [Green Version]
  60. Si, J.; Boumber, Y.A.; Shu, J.; Qin, T.; Ahmed, S.; He, R.; Jelinek, J.; Issa, J.P. Chromatin remodeling is required for gene reactivation after decitabine-mediated DNA hypomethylation. Cancer Res. 2010, 70, 6968–6977. [Google Scholar] [CrossRef] [Green Version]
  61. Fang, T.; Lv, H.; Lv, G.; Li, T.; Wang, C.; Han, Q.; Yu, L.; Su, B.; Guo, L.; Huang, S.; et al. Tumor-derived exosomal miR-1247-3p induces cancer-associated fibroblast activation to foster lung metastasis of liver cancer. Nat. Commun. 2018, 9, 191. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Liu, H.; Chen, D.; Bi, J.; Han, J.; Yang, M.; Dong, W.; Lin, T.; Huang, J. Circular RNA circUBXN7 represses cell growth and invasion by sponging miR-1247-3p to enhance B4GALT3 expression in bladder cancer. Aging 2018, 10, 2606–2623. [Google Scholar] [CrossRef] [PubMed]
  63. Shan, Y.; Ma, J.; Pan, Y.; Hu, J.; Liu, B.; Jia, L. LncRNA SNHG7 sponges miR-216b to promote proliferation and liver metastasis of colorectal cancer through upregulating GALNT1. Cell Death Dis. 2018, 9, 722. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Liu, B.; Pan, S.; Xiao, Y.; Liu, Q.; Xu, J.; Jia, L. LINC01296/miR-26a/GALNT3 axis contributes to colorectal cancer progression by regulating O-glycosylated MUC1 via PI3K/AKT pathway. J. Exp. Clin. Cancer Res. 2018, 37, 316. [Google Scholar] [CrossRef] [Green Version]
  65. Qu, J.J.; Qu, X.Y.; Zhou, D.Z. miR4262 inhibits colon cancer cell proliferation via targeting of GALNT4. Mol. Med. Rep 2017, 16, 3731–3736. [Google Scholar] [CrossRef] [Green Version]
  66. Cao, Q.; Wang, N.; Ren, L.; Tian, J.; Yang, S.; Cheng, H. miR-125a-5p post-transcriptionally suppresses GALNT7 to inhibit proliferation and invasion in cervical cancer cells via the EGFR/PI3K/AKT pathway. Cancer Cell Int. 2020, 20, 117. [Google Scholar] [CrossRef]
  67. Li, C.; Shi, J.; Zhao, Y. MiR-320 promotes B cell proliferation and the production of aberrant glycosylated IgA1 in IgA nephropathy. J. Cell Biochem. 2018, 119, 4607–4614. [Google Scholar] [CrossRef]
  68. Niu, J.T.; Zhang, L.J.; Huang, Y.W.; Li, C.; Jiang, N.; Niu, Y.J. MiR-154 inhibits the growth of laryngeal squamous cell carcinoma by targeting GALNT7. Biochem. Cell Biol. 2018, 96, 752–760. [Google Scholar] [CrossRef]
  69. Wu, H.; Chen, J.; Li, D.; Liu, X.; Li, L.; Wang, K. MicroRNA-30e Functions as a Tumor Suppressor in Cervical Carcinoma Cells through Targeting GALNT7. Transl. Oncol. 2017, 10, 876–885. [Google Scholar] [CrossRef] [PubMed]
  70. Pu, J.; Shen, J.; Zhong, Z.; Yanling, M.; Gao, J. KANK1 regulates paclitaxel resistance in lung adenocarcinoma A549 cells. Artif. Cells Nanomed. Biotechnol. 2020, 48, 639–647. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Banerjee, A.; Mahata, B.; Dhir, A.; Mandal, T.K.; Biswas, K. Elevated histone H3 acetylation and loss of the Sp1-HDAC1 complex de-repress the GM2-synthase gene in renal cell carcinoma. J. Biol. Chem. 2019, 294, 1005–1018. [Google Scholar] [CrossRef] [Green Version]
  72. Sun, X.J.; Wang, M.C.; Zhang, F.H.; Kong, X. An integrated analysis of genome-wide DNA methylation and gene expression data in hepatocellular carcinoma. FEBS Open Biol. 2018, 8, 1093–1103. [Google Scholar] [CrossRef] [PubMed]
  73. Huang, H.; Liu, Y.; Yu, P.; Qu, J.; Guo, Y.; Li, W.; Wang, S.; Zhang, J. MiR-23a transcriptional activated by Runx2 increases metastatic potential of mouse hepatoma cell via directly targeting Mgat3. Sci. Rep. 2018, 8, 7366. [Google Scholar] [CrossRef] [PubMed]
  74. Chai, Y.; Du, Y.; Zhang, S.; Xiao, J.; Luo, Z.; He, F.; Huang, K. MicroRNA-485-5p reduces O-GlcNAcylation of Bmi-1 and inhibits colorectal cancer proliferation. Exp. Cell Res. 2018, 368, 111–118. [Google Scholar] [CrossRef]
  75. Han, D.L.; Wang, L.L.; Zhang, G.F.; Yang, W.F.; Chai, J.; Lin, H.M.; Fu, Z.; Yu, J.M. MiRNA-485-5p, inhibits esophageal cancer cells proliferation and invasion by down-regulating O-linked N-acetylglucosamine transferase. Eur. Rev. Med. Pharmacol. Sci. 2019, 23, 2809–2816. [Google Scholar]
  76. Liu, Y.; Huang, H.; Cao, Y.; Wu, Q.; Li, W.; Zhang, J. Suppression of OGT by microRNA24 reduces FOXA1 stability and prevents breast cancer cells invasion. Biochem. Biophys. Res. Commun. 2017, 487, 755–762. [Google Scholar] [CrossRef] [PubMed]
  77. Liu, Y.; Huang, H.; Liu, M.; Wu, Q.; Li, W.; Zhang, J. MicroRNA-24-1 suppresses mouse hepatoma cell invasion and metastasis via directly targeting O-GlcNAc transferase. Biomed. Pharmacother. 2017, 91, 731–738. [Google Scholar] [CrossRef] [PubMed]
  78. Yu, F.Y.; Zhou, C.Y.; Liu, Y.B.; Wang, B.; Mao, L.; Li, Y. miR-483 is down-regulated in gastric cancer and suppresses cell proliferation, invasion and protein O-GlcNAcylation by targeting OGT. Neoplasma 2018, 65, 406–414. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Wei, Y.; Shao, J.; Wang, Y.; Shen, H.; Yu, S.; Zhang, J.; Yin, L. Hsa-miR-370 inhibited P-selectin-induced cell adhesion in human colon adenocarcinoma cells. Mol. Cell. Biochem. 2019, 450, 159–166. [Google Scholar] [CrossRef] [PubMed]
  80. Han, Y.; Liu, Y.; Fu, X.; Zhang, Q.; Huang, H.; Zhang, C.; Li, W.; Zhang, J. miR-9 inhibits the metastatic ability of hepatocellular carcinoma via targeting beta galactoside alpha-2,6-sialyltransferase 1. J. Physiol. Biochem. 2018, 74, 491–501. [Google Scholar] [CrossRef]
  81. Liu, B.; Liu, Q.; Pan, S.; Huang, Y.; Qi, Y.; Li, S.; Xiao, Y.; Jia, L. The HOTAIR/miR-214/ST6GAL1 crosstalk modulates colorectal cancer procession through mediating sialylated c-Met via JAK2/STAT3 cascade. J. Exp. Clin. Cancer Res. 2019, 38, 455. [Google Scholar] [CrossRef] [Green Version]
  82. Liu, Q.; Ma, H.; Sun, X.; Liu, B.; Xiao, Y.; Pan, S.; Zhou, H.; Dong, W.; Jia, L. The regulatory ZFAS1/miR-150/ST6GAL1 crosstalk modulates sialylation of EGFR via PI3K/Akt pathway in T-cell acute lymphoblastic leukemia. J. Exp. Clin. Cancer Res. 2019, 38, 199. [Google Scholar] [CrossRef]
  83. Liang, L.; Xu, J.; Wang, M.; Xu, G.; Zhang, N.; Wang, G.; Zhao, Y. LncRNA HCP5 promotes follicular thyroid carcinoma progression via miRNAs sponge. Cell Death Dis. 2018, 9, 372. [Google Scholar] [CrossRef]
  84. Jia, L.; Luo, S.; Ren, X.; Li, Y.; Hu, J.; Liu, B.; Zhao, L.; Shan, Y.; Zhou, H. miR-182 and miR-135b Mediate the Tumorigenesis and Invasiveness of Colorectal Cancer Cells via Targeting ST6GALNAC2 and PI3K/AKT Pathway. Dig. Dis. Sci. 2017, 62, 3447–3459. [Google Scholar] [CrossRef]
  85. Haldrup, C.; Pedersen, A.L.; Ogaard, N.; Strand, S.H.; Hoyer, S.; Borre, M.; Orntoft, T.F.; Sorensen, K.D. Biomarker potential of ST6GALNAC3 and ZNF660 promoter hypermethylation in prostate cancer tissue and liquid biopsies. Mol. Oncol. 2018, 12, 545–560. [Google Scholar] [CrossRef] [Green Version]
  86. Verlaat, W.; Snoek, B.C.; Heideman, D.A.M.; Wilting, S.M.; Snijders, P.J.F.; Novianti, P.W.; van Splunter, A.P.; Peeters, C.F.W.; van Trommel, N.E.; Massuger, L.; et al. Identification and Validation of a 3-Gene Methylation Classifier for HPV-Based Cervical Screening on Self-Samples. Clin. Cancer Res. 2018, 24, 3456–3464. [Google Scholar] [CrossRef] [Green Version]
  87. Huang, H.C.; Chao, C.C.; Wu, P.H.; Chung, H.Y.; Lee, H.Y.; Suen, C.S.; Hwang, M.J.; Cai, B.H.; Kannagi, R. Epigenetic silencing of the synthesis of immunosuppressive Siglec ligand glycans by NF-kappaB/EZH2/YY1 axis in early-stage colon cancers. Biochim. Biophys. Acta Gene Regul. Mech. 2019, 1862, 173–183. [Google Scholar] [CrossRef] [PubMed]
  88. Li, W.; Zheng, X.; Ren, L.; Fu, W.; Liu, J.; Xv, J.; Liu, S.; Wang, J.; Du, G. Epigenetic hypomethylation and upregulation of GD3s in triple negative breast cancer. Ann. Transl. Med. 2019, 7, 723. [Google Scholar] [CrossRef] [PubMed]
  89. Qin, Y.; Wu, C.W.; Taylor, W.R.; Sawas, T.; Burger, K.N.; Mahoney, D.W.; Sun, Z.; Yab, T.C.; Lidgard, G.P.; Allawi, H.T.; et al. Discovery, Validation, and Application of Novel Methylated DNA Markers for Detection of Esophageal Cancer in Plasma. Clin. Cancer Res. 2019, 25, 7396–7404. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Shan, Y.; Liu, Y.; Zhao, L.; Liu, B.; Li, Y.; Jia, L. MicroRNA-33a and let-7e inhibit human colorectal cancer progression by targeting ST8SIA1. Int. J. Biochem. Cell Biol. 2017, 90, 48–58. [Google Scholar] [CrossRef]
  91. Ma, W.; Zhao, X.; Liang, L.; Wang, G.; Li, Y.; Miao, X.; Zhao, Y. miR-146a and miR-146b promote proliferation, migration and invasion of follicular thyroid carcinoma via inhibition of ST8SIA4. Oncotarget 2017, 8, 28028–28041. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Wang, F.; Ye, L.J.; Wang, F.J.; Liu, H.F.; Wang, X.L. miR-146a promotes proliferation, invasion, and epithelial-to-mesenchymal transition in oral squamous carcinoma cells. Environ. Toxicol. 2020, 35, 1050–1057. [Google Scholar] [CrossRef]
  93. Wang, Y.; Chen, J.; Chen, X.; Jiang, F.; Sun, Y.; Pan, Y.; Zhang, W.; Zhang, J. MiR-34a suppresses HNSCC growth through modulating cell cycle arrest and senescence. Neoplasma 2017, 64, 543–553. [Google Scholar] [CrossRef] [Green Version]
  94. Li, W.; Li, Y.; Ma, W.; Zhou, J.; Sun, Z.; Yan, X. Long noncoding RNA AC114812.8 promotes the progression of bladder cancer through miR-371b-5p/FUT4 axis. Biomed. Pharmacother. 2020, 121, 109605. [Google Scholar] [CrossRef]
  95. Li, Y.; Sun, Z.; Liu, B.; Shan, Y.; Zhao, L.; Jia, L. Tumor-suppressive miR-26a and miR-26b inhibit cell aggressiveness by regulating FUT4 in colorectal cancer. Cell Death Dis. 2017, 8, e2892. [Google Scholar] [CrossRef] [Green Version]
  96. Liu, B.; Ma, H.; Liu, Q.; Xiao, Y.; Pan, S.; Zhou, H.; Jia, L. MiR-29b/Sp1/FUT4 axis modulates the malignancy of leukemia stem cells by regulating fucosylation via Wnt/beta-catenin pathway in acute myeloid leukemia. J. Exp. Clin. Cancer Res. 2019, 38, 200. [Google Scholar] [CrossRef] [Green Version]
  97. Yuan, X.; Liu, J.; Ye, X. Effect of miR-200c on the proliferation, migration and invasion of breast cancer cells and relevant mechanisms. J. Buon. 2019, 24, 61–67. [Google Scholar]
  98. Zhang, C.; Ge, C. A Simple Competing Endogenous RNA Network Identifies Novel mRNA, miRNA, and lncRNA Markers in Human Cholangiocarcinoma. Biomed Res. Int. 2019, 2019, 3526407. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Zhang, Y.; Zhang, D.; Lv, J.; Wang, S.; Zhang, Q. MiR-125a-5p suppresses bladder cancer progression through targeting FUT4. Biomed. Pharmacother. 2018, 108, 1039–1047. [Google Scholar] [CrossRef] [PubMed]
  100. Zheng, Q.; Cui, X.; Zhang, D.; Yang, Y.; Yan, X.; Liu, M.; Niang, B.; Aziz, F.; Liu, S.; Yan, Q.; et al. miR-200b inhibits proliferation and metastasis of breast cancer by targeting fucosyltransferase IV and alpha1,3-fucosylated glycans. Oncogenesis 2017, 6, e358. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  101. Liang, L.; Gao, C.; Li, Y.; Sun, M.; Xu, J.; Li, H.; Jia, L.; Zhao, Y. miR-125a-3p/FUT5-FUT6 axis mediates colorectal cancer cell proliferation, migration, invasion and pathological angiogenesis via PI3K-Akt pathway. Cell Death Dis. 2017, 8, e2968. [Google Scholar] [CrossRef] [Green Version]
  102. Pan, S.; Liu, Y.; Liu, Q.; Xiao, Y.; Liu, B.; Ren, X.; Qi, X.; Zhou, H.; Zeng, C.; Jia, L. HOTAIR/miR-326/FUT6 axis facilitates colorectal cancer progression through regulating fucosylation of CD44 via PI3K/AKT/mTOR pathway. Biochim. Biophys. Acta Mol. Cell Res. 2019, 1866, 750–760. [Google Scholar] [CrossRef]
  103. Zhang, Z.L.; Jiang, H.Y.; Wang, Y.S.; Shi, M.H. Heparan sulfate D-glucosamine 3-O-sulfotransferase 3B1 is a novel regulator of transforming growth factor-beta-mediated epithelial-to-mesenchymal transition and regulated by miR-218 in nonsmall cell lung cancer. J. Cancer Res. Ther. 2018, 14, 24–29. [Google Scholar] [CrossRef] [PubMed]
  104. Oliveira-Ferrer, L.; Legler, K.; Milde-Langosch, K. Role of protein glycosylation in cancer metastasis. Semin. Cancer Biol. 2017, 44, 141–152. [Google Scholar] [CrossRef]
  105. Schietinger, A.; Philip, M.; Yoshida, B.A.; Azadi, P.; Liu, H.; Meredith, S.C.; Schreiber, H. A mutant chaperone converts a wild-type protein into a tumor-specific antigen. Science 2006, 314, 304–308. [Google Scholar] [CrossRef]
  106. Mi, R.; Song, L.; Wang, Y.; Ding, X.; Zeng, J.; Lehoux, S.; Aryal, R.P.; Wang, J.; Crew, V.K.; van Die, I.; et al. Epigenetic silencing of the chaperone Cosmc in human leukocytes expressing tn antigen. J. Biol. Chem. 2012, 287, 41523–41533. [Google Scholar] [CrossRef] [Green Version]
  107. Xu, F.; Wang, D.; Cui, J.; Li, J.; Jiang, H. Demethylation of the Cosmc Promoter Alleviates the Progression of Breast Cancer Through Downregulation of the Tn and Sialyl-Tn Antigens. Cancer Manag. Res. 2020, 12, 1017–1027. [Google Scholar] [CrossRef] [Green Version]
  108. Fuster, M.M.; Esko, J.D. The sweet and sour of cancer: Glycans as novel therapeutic targets. Nat. Rev. Cancer 2005, 5, 526–542. [Google Scholar] [CrossRef] [PubMed]
  109. Julien, S.; Krzewinski-Recchi, M.A.; Harduin-Lepers, A.; Gouyer, V.; Huet, G.; Le Bourhis, X.; Delannoy, P. Expression of sialyl-Tn antigen in breast cancer cells transfected with the human CMP-Neu5Ac: GalNAc alpha2,6-sialyltransferase (ST6GalNac I) cDNA. Glycoconj. J. 2001, 18, 883–893. [Google Scholar] [CrossRef]
  110. Dall’Olio, F.; Malagolini, N.; Chiricolo, M.; Trinchera, M.; Harduin-Lepers, A. The expanding roles of the Sd(a)/Cad carbohydrate antigen and its cognate glycosyltransferase B4GALNT2. Biochim. Biophys. Acta 2014, 1840, 443–453. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Ravn, V.; Dabelsteen, E. Tissue distribution of histo-blood group antigens. APMIS 2000, 108, 1–28. [Google Scholar] [CrossRef]
  112. Wang, H.R.; Hsieh, C.Y.; Twu, Y.C.; Yu, L.C. Expression of the human Sd(a) beta-1,4-N-acetylgalactosaminyltransferase II gene is dependent on the promoter methylation status. Glycobiology 2008, 18, 104–113. [Google Scholar] [CrossRef] [PubMed]
  113. Ginsburg, D. Identifying novel genetic determinants of hemostatic balance. J. Thromb. Haemost. 2005, 3, 1561–1568. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Mohlke, K.L.; Purkayastha, A.A.; Westrick, R.J.; Smith, P.L.; Petryniak, B.; Lowe, J.B.; Ginsburg, D. Mvwf, a dominant modifier of murine von Willebrand factor, results from altered lineage-specific expression of a glycosyltransferase. Cell 1999, 96, 111–120. [Google Scholar] [CrossRef] [Green Version]
  115. Li, P.T.; Liao, C.J.; Wu, W.G.; Yu, L.C.; Chu, S.T. Progesterone-regulated B4galnt2 expression is a requirement for embryo implantation in mice. Fertil. Steril. 2011, 95, 2404–2409.e3. [Google Scholar] [CrossRef]
  116. Li, P.T.; Liao, C.J.; Yu, L.C.; Wu, W.G.; Chu, S.T. Localization of B4GALNT2 and its role in mouse embryo attachment. Fertil. Steril. 2012, 97, 1206–1212.e3. [Google Scholar] [CrossRef] [PubMed]
  117. Kawamura, Y.I.; Kawashima, R.; Fukunaga, R.; Hirai, K.; Toyama-Sorimachi, N.; Tokuhara, M.; Shimizu, T.; Dohi, T. Introduction of Sd(a) carbohydrate antigen in gastrointestinal cancer cells eliminates selectin ligands and inhibits metastasis. Cancer Res. 2005, 65, 6220–6227. [Google Scholar] [CrossRef] [Green Version]
  118. Dall’Olio, F.; Malagolini, N.; Serafini-Cessi, F. Tissue distribution and age-dependent expression of beta-4-N-acetylgalactosaminyl-transferase in guinea-pig. Biosci. Rep. 1987, 7, 925–932. [Google Scholar] [CrossRef]
  119. Dall’Olio, F.; Malagolini, N.; Di Stefano, G.; Ciambella, M.; Serafini-Cessi, F. Postnatal development of rat colon epithelial cells is associated with changes in the expression of the beta 1,4-N-acetylgalactosaminyltransferase involved in the synthesis of Sda antigen of alpha 2,6-sialyltransferase activity towards N-acetyl-lactosamine. Biochem. J. 1990, 270, 519–524. [Google Scholar] [CrossRef] [PubMed]
  120. Malagolini, N.; Santini, D.; Chiricolo, M.; Dall’Olio, F. Biosynthesis and expression of the Sda and sialyl Lewis x antigens in normal and cancer colon. Glycobiology 2007, 17, 688–697. [Google Scholar] [CrossRef]
  121. Groux-Degroote, S.; Wavelet, C.; Krzewinski-Recchi, M.A.; Portier, L.; Mortuaire, M.; Mihalache, A.; Trinchera, M.; Delannoy, P.; Malagolini, N.; Chiricolo, M.; et al. B4GALNT2 gene expression controls the biosynthesis of Sda and sialyl Lewis X antigens in healthy and cancer human gastrointestinal tract. Int. J. Biochem. Cell Biol. 2014, 53, 442–449. [Google Scholar] [CrossRef]
  122. Zoldos, V.; Grgurevic, S.; Lauc, G. Epigenetic regulation of protein glycosylation. Biomol. Concepts 2010, 1, 253–261. [Google Scholar] [CrossRef] [PubMed]
  123. Indellicato, R.; Zulueta, A.; Caretti, A.; Trinchera, M. Complementary Use of Carbohydrate Antigens Lewis a, Lewis b, and Sialyl-Lewis a (CA19.9 Epitope) in Gastrointestinal Cancers: Biological Rationale Towards A Personalized Clinical Application. Cancers 2020, 12, 1509. [Google Scholar] [CrossRef] [PubMed]
  124. Trinchera, M.; Zulueta, A.; Caretti, A.; Dall’Olio, F. Control of Glycosylation-Related Genes by DNA Methylation: The Intriguing Case of the B3GALT5 Gene and Its Distinct Promoters. Biology 2014, 3, 484–497. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Holgersson, J.; Lofling, J. Glycosyltransferases involved in type 1 chain and Lewis antigen biosynthesis exhibit glycan and core chain specificity. Glycobiology 2006, 16, 584–593. [Google Scholar] [CrossRef] [PubMed]
  126. Chase, S.D.; Magnani, J.L.; Simon, S.I. E-selectin ligands as mechanosensitive receptors on neutrophils in health and disease. Ann. Biomed. Eng. 2012, 40, 849–859. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Trinchera, M.; Aronica, A.; Dall’Olio, F. Selectin Ligands Sialyl-Lewis a and Sialyl-Lewis x in Gastrointestinal Cancers. Biology 2017, 6, 16. [Google Scholar] [CrossRef] [PubMed]
  128. Dunn, C.A.; van de Lagemaat, L.N.; Baillie, G.J.; Mager, D.L. Endogenous retrovirus long terminal repeats as ready-to-use mobile promoters: The case of primate beta3GAL-T5. Gene 2005, 364, 2–12. [Google Scholar] [CrossRef] [PubMed]
  129. Dunn, C.A.; Medstrand, P.; Mager, D.L. An endogenous retroviral long terminal repeat is the dominant promoter for human beta1,3-galactosyltransferase 5 in the colon. Proc. Natl. Acad. Sci. USA 2003, 100, 12841–12846. [Google Scholar] [CrossRef] [Green Version]
  130. Zulueta, A.; Caretti, A.; Signorelli, P.; Dall’olio, F.; Trinchera, M. Transcriptional control of the B3GALT5 gene by a retroviral promoter and methylation of distant regulatory elements. FASEB J. 2014, 28, 946–955. [Google Scholar] [CrossRef] [Green Version]
  131. Aronica, A.; Avagliano, L.; Caretti, A.; Tosi, D.; Bulfamante, G.P.; Trinchera, M. Unexpected distribution of CA19.9 and other type 1 chain Lewis antigens in normal and cancer tissues of colon and pancreas: Importance of the detection method and role of glycosyltransferase regulation. Biochim. Biophys. Acta Gen. Subj. 2017, 1861, 3210–3220. [Google Scholar] [CrossRef]
  132. Caretti, A.; Sirchia, S.M.; Tabano, S.; Zulueta, A.; Dall’Olio, F.; Trinchera, M. DNA methylation and histone modifications modulate the beta1,3 galactosyltransferase beta3Gal-T5 native promoter in cancer cells. Int. J. Biochem. Cell. Biol. 2012, 44, 84–90. [Google Scholar] [CrossRef]
  133. Isshiki, S.; Kudo, T.; Nishihara, S.; Ikehara, Y.; Togayachi, A.; Furuya, A.; Shitara, K.; Kubota, T.; Watanabe, M.; Kitajima, M.; et al. Lewis type 1 antigen synthase (beta3Gal-T5) is transcriptionally regulated by homeoproteins. J. Biol. Chem. 2003, 278, 36611–36620. [Google Scholar] [CrossRef] [Green Version]
  134. Salvini, R.; Bardoni, A.; Valli, M.; Trinchera, M. beta 1,3-Galactosyltransferase beta 3Gal-T5 acts on the GlcNAcbeta 1-->3Galbeta 1-->4GlcNAcbeta 1-->R sugar chains of carcinoembryonic antigen and other N-linked glycoproteins and is down-regulated in colon adenocarcinomas. J. Biol. Chem. 2001, 276, 3564–3573. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Abraham, C.; Cho, J.H. Inflammatory bowel disease. N. Engl. J. Med. 2009, 361, 2066–2078. [Google Scholar] [CrossRef] [PubMed]
  136. Khor, B.; Gardet, A.; Xavier, R.J. Genetics and pathogenesis of inflammatory bowel disease. Nature 2011, 474, 307–317. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Borg-Bartolo, S.P.; Boyapati, R.K.; Satsangi, J.; Kalla, R. Precision medicine in inflammatory bowel disease: Concept, progress and challenges. F1000Research 2020, 9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Gonsky, R.; Deem, R.L.; Targan, S.R. Distinct Methylation of IFNG in the Gut. J. Interferon Cytokine Res. 2009, 29, 407–414. [Google Scholar] [CrossRef]
  139. Lobaton, T.; Azuara, D.; Rodriguez-Moranta, F.; Loayza, C.; Sanjuan, X.; de Oca, J.; Fernandez-Robles, A.; Guardiola, J.; Capella, G. Relationship between methylation and colonic inflammation in inflammatory bowel disease. World J. Gastroenterol. 2014, 20, 10591–10598. [Google Scholar] [CrossRef] [PubMed]
  140. Saito, S.; Kato, J.; Hiraoka, S.; Horii, J.; Suzuki, H.; Higashi, R.; Kaji, E.; Kondo, Y.; Yamamoto, K. DNA methylation of colon mucosa in ulcerative colitis patients: Correlation with inflammatory status. Inflamm. Bowel Dis. 2011, 17, 1955–1965. [Google Scholar] [CrossRef] [Green Version]
  141. Tahara, T.; Shibata, T.; Nakamura, M.; Yamashita, H.; Yoshioka, D.; Okubo, M.; Maruyama, N.; Kamano, T.; Kamiya, Y.; Fujita, H.; et al. Promoter methylation of protease-activated receptor (PAR2) is associated with severe clinical phenotypes of ulcerative colitis (UC). Clin. Exp. Med. 2009, 9, 125–130. [Google Scholar] [CrossRef]
  142. Tahara, T.; Shibata, T.; Nakamura, M.; Yamashita, H.; Yoshioka, D.; Okubo, M.; Maruyama, N.; Kamano, T.; Kamiya, Y.; Nakagawa, Y.; et al. Effect of MDR1 gene promoter methylation in patients with ulcerative colitis. Int. J. Mol. Med. 2009, 23, 521–527. [Google Scholar] [CrossRef] [Green Version]
  143. Cooke, J.; Zhang, H.; Greger, L.; Silva, A.L.; Massey, D.; Dawson, C.; Metz, A.; Ibrahim, A.; Parkes, M. Mucosal genome-wide methylation changes in inflammatory bowel disease. Inflamm. Bowel Dis. 2012, 18, 2128–2137. [Google Scholar] [CrossRef]
  144. Hasler, R.; Feng, Z.; Backdahl, L.; Spehlmann, M.E.; Franke, A.; Teschendorff, A.; Rakyan, V.K.; Down, T.A.; Wilson, G.A.; Feber, A.; et al. A functional methylome map of ulcerative colitis. Genome Res. 2012, 22, 2130–2137. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Karatzas, P.S.; Mantzaris, G.J.; Safioleas, M.; Gazouli, M. DNA methylation profile of genes involved in inflammation and autoimmunity in inflammatory bowel disease. Medicine 2014, 93, e309. [Google Scholar] [CrossRef] [PubMed]
  146. Nimmo, E.R.; Prendergast, J.G.; Aldhous, M.C.; Kennedy, N.A.; Henderson, P.; Drummond, H.E.; Ramsahoye, B.H.; Wilson, D.C.; Semple, C.A.; Satsangi, J. Genome-wide methylation profiling in Crohn’s disease identifies altered epigenetic regulation of key host defense mechanisms including the Th17 pathway. Inflamm. Bowel Dis. 2012, 18, 889–899. [Google Scholar] [CrossRef]
  147. Mateos, B.; Palanca-Ballester, C.; Saez-Gonzalez, E.; Moret, I.; Lopez, A.; Sandoval, J. Epigenetics of Inflammatory Bowel Disease: Unraveling Pathogenic Events. Crohn’s Colitis 360 2019, 1. [Google Scholar] [CrossRef]
  148. Klasic, M.; Markulin, D.; Vojta, A.; Samarzija, I.; Birus, I.; Dobrinic, P.; Ventham, N.T.; Trbojevic-Akmacic, I.; Simurina, M.; Stambuk, J.; et al. Promoter methylation of the MGAT3 and BACH2 genes correlates with the composition of the immunoglobulin G glycome in inflammatory bowel disease. Clin. Epigenet. 2018, 10, 75. [Google Scholar] [CrossRef] [PubMed]
  149. Serino, G.; Sallustio, F.; Cox, S.N.; Pesce, F.; Schena, F.P. Abnormal miR-148b expression promotes aberrant glycosylation of IgA1 in IgA nephropathy. J. Am. Soc. Nephrol. 2012, 23, 814–824. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  150. Serino, G.; Sallustio, F.; Curci, C.; Cox, S.N.; Pesce, F.; De Palma, G.; Schena, F.P. Role of let-7b in the regulation of N-acetylgalactosaminyltransferase 2 in IgA nephropathy. Nephrol. Dial Transplant. 2015, 30, 1132–1139. [Google Scholar] [CrossRef] [Green Version]
  151. Liu, D.; Xia, M.; Liu, Y.; Tan, X.; He, L.; Chen, G.; Liu, H. The upregulation of miR-98-5p affects the glycosylation of IgA1 through cytokines in IgA nephropathy. Int. Immunopharmacol. 2020, 82, 106362. [Google Scholar] [CrossRef] [PubMed]
  152. Hu, S.; Bao, H.; Xu, X.; Zhou, X.; Qin, W.; Zeng, C.; Liu, Z. Increased miR-374b promotes cell proliferation and the production of aberrant glycosylated IgA1 in B cells of IgA nephropathy. FEBS Lett. 2015, 589, 4019–4025. [Google Scholar] [CrossRef] [PubMed]
  153. Moll, T.; Shaw, P.J.; Cooper-Knock, J. Disrupted glycosylation of lipids and proteins is a cause of neurodegeneration. Brain 2020, 143, 1332–1340. [Google Scholar] [CrossRef]
  154. Trinchera, M.; Parini, R.; Indellicato, R.; Domenighini, R.; dall’Olio, F. Diseases of ganglioside biosynthesis: An expanding group of congenital disorders of glycosylation. Mol. Genet. Metab. 2018, 124, 230–237. [Google Scholar] [CrossRef]
  155. Zulueta, A.; Mingione, A.; Signorelli, P.; Caretti, A.; Ghidoni, R.; Trinchera, M. Simple and Complex Sugars in Parkinson’s Disease: A Bittersweet Taste. Mol. Neurobiol. 2020, 57, 2934–2943. [Google Scholar] [CrossRef] [PubMed]
  156. Chao, C.C.; Gutierrez-Vazquez, C.; Rothhammer, V.; Mayo, L.; Wheeler, M.A.; Tjon, E.C.; Zandee, S.E.J.; Blain, M.; de Lima, K.A.; Takenaka, M.C.; et al. Metabolic Control of Astrocyte Pathogenic Activity via cPLA2-MAVS. Cell 2019, 179, 1483–1498. [Google Scholar] [CrossRef] [PubMed]
  157. Iwamoto, S.; Withers, D.A.; Handa, K.; Hakomori, S. Deletion of A-antigen in a human cancer cell line is associated with reduced promoter activity of CBF/NF-Y binding region, and possibly with enhanced DNA methylation of A transferase promoter. Glycoconj. J. 1999, 16, 659–666. [Google Scholar] [CrossRef] [PubMed]
  158. Pink, M.; Ratsch, B.A.; Mardahl, M.; Schroter, M.F.; Engelbert, D.; Triebus, J.; Hamann, A.; Syrbe, U. Identification of two regulatory elements controlling Fucosyltransferase 7 transcription in murine CD4+ T cells. Mol. Immunol. 2014, 62, 1–9. [Google Scholar] [CrossRef] [PubMed]
  159. Syrbe, U.; Jennrich, S.; Schottelius, A.; Richter, A.; Radbruch, A.; Hamann, A. Differential regulation of P-selectin ligand expression in naive versus memory CD4+ T cells: Evidence for epigenetic regulation of involved glycosyltransferase genes. Blood 2004, 104, 3243–3248. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Table 1. List of glycogenes regulated through epigenetics.
Table 1. List of glycogenes regulated through epigenetics.
TargetEpigenetic MechanismEffectTissue/Cells InvolvedReferences
Galactosyltransferases
B4GALT3miR-1247-3p
CircUBXN7/miR-1247-3p axis
Downregulation
Upregulation
CAFs in lung metastasis of liver cancer
Bladder cancer
[61,62]
N-acetyl-galactosaminyl transferases
GALNT1LncRNA SNHG7/miR-216b axisUpregulationColorectal cancer[63]
GALNT3Linc01296/miR-26a axisUpregulationColorectal cancer[64]
GALNT4miR-4262 (downregulated)UpregulationColorectal cancer[65]
GALNT7miR-30e (downregulated)
LncRNA SNHG7/miR-34a axis
miR-154 (downregulated)
miR-125a-5p (downregulated)
Upregulation
Upregulation
Upregulation
Upregulation
Cervical cancer
Colorectal cancer
Laryngeal squamous cell carcinoma
Cervical cancer
[66,67,68,69]
GALNT14HypermethylationDownregulationA549-T cells (paclitaxel-resistant strain of human non-small cell lung cancer)[70]
B4GALNT1Histone acetylation
Hypermethylation
Upregulation
Downregulation
Renal cell carcinoma
Hepatocellular carcinoma
[71,72]
N-acetyl-glucosaminyl transferases
MGAT3miR-23a(upregulated)DownregulationHca-P (mouse) cell line[73]
OGTmiR-24-1 (downregulated)
miR-24 (downregulated)
miR-483 (downregulated)
miR-485-5p (downregulated)
Upregulation
Upregulation
Upregulation
Upregulation
Hca-F (mouse) cell line
High invasive breast cancer cell lines
Gastric cancer
Colorectal cancer and esophageal cancer cell lines
[74,75,76,77,78]
Sialyltransferases
ST3GAL4miR-370 (treatment)DownregulationColo 320 cell line[79]
ST6GAL1miR-9 (downregulated)
LncRNA ZFAS1/miR-150 axis
LncRNA HOTAIR/miR-214 axis
Upregulation
Upregulation
Upregulation
Hepatocellular carcinoma cell lines with high lymphatic metastatic potential
T-cell acute lymphoblastic leukemia
Colorectal cancer
[80,81,82]
ST6GAL2LncRNA HCP5/miR-22-3p, miR-186-5p, miR-216a-5p axisUpregulationFollicular thyroid carcinoma[83]
ST6GALNAC2miR-182 and miR-135bDownregulationColorectal cancer[84]
ST6GALNAC3Promoter hypermethylationDownregulationProstate cancer[85]
ST6GALNAC5Promoter hypermethylationDownregulationCervical cancer[86]
ST6GALNAC6Histone methylation (H3K27me3)DownregulationColon cancer[87]
ST8SIA1miR-33a and let-7e (downregulated)
Promoter hypomethylation
Promoter hypermethylation
Not evaluated
Upregulation
Not evaluated
Colorectal cancer
Triple-negative breast cancer
Esophageal cancer
[88,89,90]
ST8SIA4miR-146a and miR-146b (upregulated)miR-146a (upregulated)Downregulation
Downregulation
Follicular thyroid carcinomaOral squamous carcinoma cell lines[91,92]
Fucosyltransferases
FUT1miR-34a (downregulated)UpregulationHead and neck squamous cell carcinoma[93]
FUT4miR-26a e miR-26b (downregulated)
miR-200b (downregulated)
miR-125a-5p (downregulated)
miR-200c (treatment)
miR-1295b and miR-6715amiR-29b/Sp1 axis
LncRNA AC114812.8/miR-371b-5p axis
Upregulation
Upregulation
Upregulation
Downregulation
Not evaluated
Upregulation
Upregulation
Colorectal cancer
Breast cancer
Bladder cancer cell lines
MCF-7 cell line (breast cancer)
Cholangiocarcinoma
Acute myeloid leukemia
Bladder cancer cell lines
[94,95,96,97,98,99,100]
FUT5miR-125a-3p (downregulated)UpregulationColorectal cancer[101]
FUT6miR-125a-3p (downregulated)
LncRNA HOTAIR/miR-326 axis
Upregulation
Upregulation
Colorectal cancer
Colorectal cancer
[101,102]
Sulfotransferases
HS3ST3B1miR-218 (downregulated)UpregulationNon-small cell lung cancer[103]
Nucleotide donor transporters
DTDSTHistone methylation (H3K27me3)DownregulationColon cancer[87]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Indellicato, R.; Trinchera, M. Epigenetic Regulation of Glycosylation in Cancer and Other Diseases. Int. J. Mol. Sci. 2021, 22, 2980. https://doi.org/10.3390/ijms22062980

AMA Style

Indellicato R, Trinchera M. Epigenetic Regulation of Glycosylation in Cancer and Other Diseases. International Journal of Molecular Sciences. 2021; 22(6):2980. https://doi.org/10.3390/ijms22062980

Chicago/Turabian Style

Indellicato, Rossella, and Marco Trinchera. 2021. "Epigenetic Regulation of Glycosylation in Cancer and Other Diseases" International Journal of Molecular Sciences 22, no. 6: 2980. https://doi.org/10.3390/ijms22062980

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

Article Metrics

Back to TopTop