Next Article in Journal
Cinnamomum cassia Modulates Key Players of Gut–Liver Axis in Murine Lupus
Previous Article in Journal
Cytotoxicity of Cannabinoids in Combination with Traditional Lymphoma Chemotherapeutic Drugs Against Non-Hodgkin’s Lymphoma
Previous Article in Special Issue
Unlocking the Sugar Code: Implications and Consequences of Glycosylation in Alzheimer’s Disease and Other Tauopathies
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

GalNAc-Transferases in Cancer

by
Shruthi C. Iyer
1,
Dinesh Kumar Srinivasan
2 and
Rajeev Parameswaran
3,*
1
Yong Loo Lin School of Medicine, National University of Singapore, Singapore 117597, Singapore
2
Department of Anatomy, Yong Loo Lin School of Medicine, National University of Singapore, Singapore 117594, Singapore
3
Department of Surgery, National University Hospital, Singapore 119074, Singapore
*
Author to whom correspondence should be addressed.
Biomedicines 2026, 14(1), 5; https://doi.org/10.3390/biomedicines14010005
Submission received: 27 October 2025 / Revised: 7 December 2025 / Accepted: 9 December 2025 / Published: 19 December 2025
(This article belongs to the Special Issue Role of Glycomics in Health and Diseases)

Abstract

Background/Objectives: The polypeptide N-acetylgalactosaminyltransferase (GALNT) family initiates mucin-type O-glycosylation, a post-translational modification that plays a pivotal role in cellular signaling, adhesion, and immune evasion. Dysregulated GALNT expression has been increasingly implicated in carcinogenesis. Methods: We reviewed the literature on the expression, function, and clinical relevance of GALNT isoforms across various cancers, with a focus on their mechanistic roles, biomarker potential, and therapeutic implications. Results: Aberrant GALNT expression is observed in numerous malignancies, including breast, colorectal, gastric, lung, ovarian, and hepatocellular carcinomas. Isoforms such as GALNT1, -T2, -T3, and -T14 contribute to tumorigenesis by modulating the glycosylation of mucins such as Mucin-1 (MUC1), epithelial growth factor receptors (EGFR), and other signaling proteins. These alterations promote cancer cell proliferation, metastasis, epithelial–mesenchymal transition (EMT), and chemoresistance. Deranged GALNT expression is frequently associated with poor prognosis, and certain GALNT genotypes predict treatment response. However, functional redundancy among isoforms poses challenges for selective targeting. Conclusions: Despite their strong potential as modulators of cancer progression, GALNTs face substantial limitations in terms of substrate identification, mechanistic clarity, immune relevance, and therapeutic tractability. Overcoming these challenges requires advanced glycoproteomics, development of isoform-specific tools, and integrated studies across cancer and immunology to fully harness GALNT biology for clinical benefit.

Graphical Abstract

1. Introduction

Protein glycosylation is an essential post-translational modification step in eukaryotes, with many metabolically active substances being glycosylated compounds (glycans). Numerous protein glycosylation types exist in animals, plants, and microorganisms [1]. Mucin-type O-glycosylation is one such type and occurs when glycans are attached to serine (Ser) and threonine (Thr) amino acid residues via O-linked N-acetylgalactosamine (GalNAc) [2]. Biologically, mucin-type O-glycosylation plays a critical role in maintaining epithelial barrier function, modulating immune responses, and regulating protein secretion and turnover. The polypeptide N-acetylgalactosaminyltransferases (GALNTs) thus contribute to proteome complexity and functional diversification, particularly in mucosal tissues where mucins—heavily O-glycosylated glycoproteins—form protective layers [2,3].
Typically, the initiation of O-glycosylation is mediated by only one or two genes encoding glycosyltransferase (GT) enzymes. These genes encode enzymes that attach the first O-glycan to the relevant amino acid residue in the protein. However, the initiation step of mucin-type O-linked glycan synthesis involves a large family of up to 20 homologous genes which encode numerous GALNT enzymes [2]. GALNTs are a type of GT enzyme that catalyze O-glycosylation to form mucin-type O-glycans, and this process is initiated in the Golgi apparatus once most of the protein folding events have been completed [4]. This contrasts with N-glycosylation and other types of O-glycosylation, which is initiated in the endoplasmic reticulum (ER). The larger number of enzymes initiating mucin-type O-glycan synthesis and its occurrence in the Golgi apparatus rather than in the ER make GALNTs unique compared to other types of protein glycosylation enzymes. This is due to their greater potential for differential regulation in cells and tissues [4].
Clinically, aberrant GALNT activity is associated with several pathological conditions. Malfunction of GALNTs is responsible for altered glycosylation and a truncated mucin-type O-glycan structure, which have been associated with increased metastatic potential and poor prognosis in cancer [5,6,7]. In cancer, altered expression of specific GALNTs (e.g., GALNT3, T6, etc.) leads to the synthesis of truncated or atypical O-glycans, such as the Tn and sialyl-Tn antigens, which are hallmarks of tumor progression and metastasis. This review aims to explore the role of GALNTs in cancer, their underlying molecular mechanisms, and their potential as prognostic markers for various cancers.

1.1. Structure of GALNTs

GALNTs, like other Golgi glycosyltransferases, have a type II transmembrane topology. They have a short N-terminal cytoplasmic domain, a single hydrophobic transmembrane domain, and a large luminal/extracellular C-terminus [8]. The short cytoplasmic domain of all GALNTs contains a basic sequence of amino acids that may be involved in interaction with peripheral Golgi membrane protein tethering complexes [8]. The luminal domain consists of a juxtamembrane stem region and a catalytic region. The additional ricin-like lectin domain of GALNTs at the C-terminus makes them unique compared to other eukaryotic GTs [9]. The juxtamembrane stem regions are variable in length across the different isoforms of GALNTs (ranging from around 90 amino acids in GALNT1, -T13, and -T16 to 470 amino acids in GALNT5) [8].
The GALNT catalytic domains are about 230 amino acids long and adopt three main three-dimensional (3D) fold patterns called the GT-A, GT-B, and GT-C, respectively [4,10]. GT-A enzymes are metal-ion-dependent, using a metal ion cofactor (primarily Mn2+) at the active site to catalyze the reaction. The GT-A structural motif, a specific 3D fold common in glycosyltransferases, made up of two tightly interacting β–α–β Rossmann-like folds, contains amino acids that interact with the uracil ring of the Uridine diphosphate N-acetylgalactosamine (UDP-GalNAc) molecule, which is the sugar donor in the enzymatic reaction (as shown in Figure 1) [4,11]. In contrast, GT-B enzymes possess two distinct domains and are metal-ion-independent, thereby catalyzing reactions even without a metal ion cofactor at the active site. More recently, a third folding pattern called GT-C was identified [12]. GALNTs, like other GTs, may be phosphorylated either at their cytosolic or luminal domains, and this provides insight into how GALNT activities could be regulated by phosphorylation. In a similar manner, the GALNT activation (GALA) pathway drives the re-localization of GALNTs from the Golgi to the ER, partly through negative regulation of Extracellular Signal-Regulated Kinase 8 (ERK8). This activation of the GALA pathway leads to the accumulation of truncated O-glycans in cancer cells, which in turn enhances tumor cell migration, invasion, and metastatic behavior [13,14,15,16].

1.2. Types of GALNTs and the Human GALNT Gene Family

GALNTs are members of the GT27 family in the CAZy glycosyltransferase classification (http://www.cazy.org (accessed on 3 August 2024)). In humans, 20 GALNT genes have been identified, most of which encode active polypeptide GALNTs functioning in O-glycosylation [4]. The genes encoding GALNTs, known as GALNT genes, have been highly conserved in sequence throughout evolution and play an essential function in eukaryotic species [4,17]. In 2012, Bennett and colleagues analyzed cladograms (phylogenetic trees based on sequence similarities) and genomic organization (with specific focus on exon-intron boundaries). They grouped the human GALNTs into seven subfamilies (namely subfamilies Ia–g, IIa, and IIb) [4]. Most of the GALNTs show enhanced expression in specific tissues, as shown in Table 1. The differential expression of the GALNT genes across various tissues and developmental stages contributes to the functional diversity of glycosylation patterns observed in human proteins [13].
The GALNT gene family is organized into two main groups based on their structural characteristics and substrate specificity. Group I GALNT genes, such as GALNT1, GALNT2, and GALNT3, are primarily responsible for synthesizing the Tn antigen, a precursor of more complex O-glycans [13]. In contrast, group II gene members, such as GALNT4, GALNT5, and GALNT6, exhibit broader substrate specificity and are involved in generating more extended O-glycan structures (Figure 2) [13].

1.3. Substrate Specificities of GALNTs

The substrate specificities of GALNTs are fundamental in defining the structure and function of mucin-type O-glycans in diverse biological contexts. These enzymes catalyze the initial transfer of GALNT to serine or threonine residues on polypeptide chains, initiating O-glycosylation. Individual GALNT isoforms exhibit distinct preferences for peptide sequences and post-translational modifications, thereby contributing to the complexity and specificity of O-glycan patterns. Certain isoforms preferentially act on peptide motifs enriched in proline or other bulky amino acids, which can modulate both enzymatic activity and glycan structure [18,19]. Importantly, the large family of GALNT isoenzymes acts in a coordinated fashion. Some isoforms initiate O-glycosylation at specific sites, while others act subsequently to glycosylate adjacent residues, thereby promoting glycan extension and density [9,20]. Additionally, some isoforms selectively glycosylate sites already modified by other glycosyltransferases, whereas others display broader substrate tolerance, enabling the modification of a wider range of peptide sequences [19]. Some examples of isoform specific expression affecting substrate selection and glycan function are listed in Table 2 [21,22,23,24,25,26,27,28,29,30,31].

1.4. Expression of GALNTs

The expression of GALNTs is tightly regulated and shifts with cell type, differentiation status, and maturation stage. The expression also varies significantly between normal and cancerous tissues. In noncancerous states, specific isoforms of GALNTs are expressed in a controlled manner, contributing to the synthesis of non-truncated mucin-type O-glycans, usually capped with sialic acid. For example, correct glycosylation mediated by GALNTs is important in the functioning of many peptide hormones, lipid regulation, and bone functions [32,33,34]. In contrast, cancerous tissues exhibit altered expression levels of GALNTs, leading to aberrant glycosylation patterns that can promote tumor progression and metastasis [35]. The mechanisms through which GALNTs play a role in cancer are described in Table 2. The truncated glycoforms seen in cancer states are possibly the result of various mechanisms involving GALNTs, some of which include core-1 β1-3galactosyltransferase-specific chaperone 1 (COSMC) inactivation, through mutations or hypermethylation [36,37,38], relocation of GALNTs to the ER instead of Golgi apparatus [14,39], structural/topological changes in glycosylation enzymes [39,40], and impaired stability of E-cadherins, thereby giving rise to altered cell adhesion molecules [41].

1.5. Role of Golgi Apparatus in GALNT Activity

The Golgi apparatus plays a crucial role in the function and localization of GALNTs. The Golgi is a key organelle of the secretory pathway and consists of stacks of cisternae, which receives the proteins and lipids synthesized in the ER to further process them [42]. The cisternae are interconnected to form Golgi ribbons, which are located near the centromere. The stack is sandwiched at the entrance by the membrane structures called the cis-Golgi network (CGN) and vesicular tubular clusters (VTCs), and at the exit by the trans-Golgi network (TGN) [42,43,44]. The TGN is the major site which sorts proteins to distinct cellular locations.
Inside the Golgi, GALNTs catalyze the addition of GalNAc to the hydroxyl groups of serine and threonine residues on target proteins, initiating O-linked glycosylation to form various glycans (Figure 2). Their precise localization within Golgi sub-compartments is essential for determining where, when, and how extensively O-glycans are added [33,45]. Enzymes involved in glycosylation are generally arranged according to their function: early-acting enzymes are found in the cis-Golgi, while late-acting ones are enriched in the trans-Golgi [17,46,47,48]. However, not all glycosylation enzymes follow a strict cis-to-trans localization. For example, GALNTs involved in the initiation of O-glycosylation—specifically GALNT1, -T2, and -T3—are distributed across different regions of the Golgi [47]. Furthermore, whilst GALNT1 is evenly distributed throughout the stack, GALNT2 localizes mainly to the trans-Golgi, and GALNT3 is enriched in the medial Golgi.
When the GALNTs shift to later Golgi compartments, initiation becomes restricted due to increased protein folding and prior modifications, altering glycan site occupancy and density [49]. The location of GALNTs also determines the timing of competition between initiation and downstream enzymes such as C1GALT1, GCNTs, and sialyltransferases, shaping core structure formation and glycan elongation [33,49,50]. Signal-regulated GALNT relocalization (e.g., via EGFR–ERK) can enhance global initiation, modify receptor glycosylation, and alter cell signaling, a phenomenon frequently observed in cancer, which is discussed in a later section. Thus, GALNT spatial organization governs the fidelity, extent, and biological impact of O-glycan biosynthesis. Mislocalization due to trafficking defects, pH disruption, or Golgi stress contributes to truncated Tn/STn antigens and disease-associated glycosylation abnormalities [49,51,52]. The Tn glycan is extended by the addition of more sugars to produce a mature O-GalNAc glycan, each with one of four core molecules (Figure 3).
The synthesis and movement of the O-glycans are initiated in the cis-Golgi bodies by GALNTs in normal conditions. However, there is an alternative pathway called the GALA pathway, whereby the GALNTs are relocated from the Golgi to the ER, enabling the glycosylation of substrates in the ER, a site where glycosylation typically does not occur [53]. In the GALA system, external stimuli (e.g., growth factors) activate Src, a non-receptor tyrosine kinase, which is essential for initiating the relocation of GALNTs, thereby triggering signaling cascades. Src activation enhances activity of the Coat Protein I (COPI) vesicular transport system, enabling retrograde transport of GALNTs to the ER. This step is GALNT-specific; other Golgi enzymes are not significantly affected. ERK8, a serine/threonine-specific protein kinase, inhibits GALNT relocation. When ERK8 is downregulated or inactivated, GALNTs more easily translocate to the ER. In the ER, they gain access to immature, newly synthesized proteins, allowing for early glycosylation and modification of proteins that would otherwise not be glycosylated. This is an atypical and inducible pathway of O-glycosylation, usually seen in cancer states, which leads to the abundant formation of the Tn antigen [39,54].

2. Role of GALNTs in Cancers

Aberrant expression and mutations in GALNTs can lead to cancer-specific glycosylation patterns, impacting tumor proliferation, induction of EMT, metastasis, immune evasion, and drug resistance [55,56,57]. Many of the functions of receptors such as EGFR, Notch, integrins, etc., rely on proper glycosylation mediated by GALNTs and influence the various cancer signaling pathways. This has been shown in many human cancers, which will be discussed below.

2.1. EGFR-Mediated Pathway

Epithelial growth factor (EGF) and EGF receptor activation are involved in the development and progression of many cancers. Ligand binding of EGFR results in phosphorylation and dimerization, with resultant downstream cascade activation of the various signaling pathways, resulting in increased proliferation, migration, and invasion of cancer cells. Glycosylation of EGFR by GALNTs alters receptor conformation, stability, localization, or internalization [23]. GALNT1 overexpression is associated with increased cell migration and invasion. Using hepatocellular carcinoma (HCC) cell lines, Huang et al. showed that the knockdown of GALNT1 reduced EGF-induced EGFR phosphorylation and enhanced EGFR degradation [23]. Overexpression of GALNT2 suppresses EGF-induced EGFR activation and promotes EGFR degradation. Wu et al. showed that the O-glycosylation status of EGFR was altered by GALNT2 using immunoprecipitation techniques with Vicia Villosa Lectin (VVA-lectin), and specifically, GALNT2 knockdown resulted in higher EGFR phosphorylation and downstream signaling [21]. Similarly, using gastric cancer cells, knockdown of GALNT2 decreased Tn epitope expression on EGFR and increased EGFR/Akt downstream signaling, with a resultant increase in malignant behavior [58]. Glycosylation of EGFR by GALNT5 promotes ligand binding and stability of EGF, leading to the activation of the AKT/ERK axis, promoting malignant behavior in cholangiocarcinoma cell lines; however, the precise O-glycosite residues on EGFR modified by GALNT5 have not been mapped [59].
GALNT3 also acts as a modulator/suppressor of EGFR activity via O-glycosylation of EGFR (or its co-receptors). In pancreatic cancer cell lines, reduced GALNT3 expression was associated with increased phosphorylation of EGFR/ErbB family receptors and increased aggressiveness of poorly differentiated cancer cells [60]. The knockdown of GALNT3 resulted in increased Tn antigen expression on EGFR and human epidermal growth factor receptor 2 (Her2) proteins (i.e., truncated O-glycans) and increased receptor activation. Using pull-down assays with VVA-lectin in chemoresistant ovarian cancer, Li et al. showed GALNT14 to be involved with the EGFR/mTOR pathway that contributed to cisplatin resistance [26]. This study showed that the downregulation of GALNT14 significantly inhibits O-GALNTylation of EGFR and promotes apoptosis and ferroptosis of ovarian cancer cells. Besides direct glycosylation of EGFR, GALNTs may indirectly alter the glycosylation of receptors/co-receptors or integrins that modulate EGFR signaling [55]. The differential effects of GALNTs on EGFR signaling likely reflect site-specific O-glycosylation patterns, with distinct substrate specificities.

2.2. TGF-β Signaling Pathway and Epithelial–Mesenchymal Transition (EMT)

Transforming growth factor beta (TGF-β) plays dual roles in cancer, acting as a tumor suppressor during early stages of cancer, but as a tumor promoter of invasion and metastasis in late stages [61]. TGF-β receptor activation triggers both SMAD-dependent and SMAD-independent pathways that drive EMT, immune evasion, and fibrosis [62]. EMT is characterized by the coordinated loss of epithelial junction proteins, such as E-cadherin, and the acquisition of mesenchymal markers, including vimentin [63]. During EMT, epithelial cells lose their polarity and intercellular adhesion, acquire mesenchymal traits, and become motile and invasive [64].
O-GALNTylation by GALNT enzymes can amplify or dampen TGF-β signaling, reshaping cellular adhesion and migration [56]. GALNT4 acts as a tumor-suppressive modulator of TGF-β signaling, specifically by modifying the TGF-β type II receptor (TβRII) and restraining EMT, migration, and metastasis, as shown by Wu et al. in breast cancer cells [65]. The authors showed that a GALNT moiety attached to Ser31 on TβRII by GALNT4 causes steric hindrance, leading to the decreased dimerization of TβRII. This, in effect, decreases SMAD2/3 phosphorylation, diminishes the transcription of EMT-related genes (e.g., SNAIL, SLUG, ZEB1), and thereby helps to maintain the epithelial phenotype, with decreased invasion and migration. Knockdown of GALNT4 led to the opposite effect with the enhancement of EMT, mesenchymal markers such as SNAIL and N-cadherin, and a decrease in E-cadherin (epithelial marker) levels [65].
GALNT2 normally suppresses TGF-β pathway activation, with reduced expression leading to upregulated TGF-β ligand production, enhanced EMT, and invasion. The expression of GALNT2 can be tumor-suppressive in gastric [66] and hepatocellular carcinoma [23], whereas it can act as a tumor promoter in colorectal cancer [67] and lung cancer [68]. The differential behavior of tumor suppression versus tumor promotion is dependent on substrate activation, such as EGFR, TGF-β, etc. Liu et al., using in vivo studies, showed that knockdown of GALNT2 in SGC-7901 gastric cancer cells increased TGF-β expression, along with MMP-2, leading to increased SMAD, resulting in enhanced EMT and invasive features [66]. GALNT2 also acts independently of TGF-β by modifying O-glycans on AXL (via a proteasome-dependent pathway) to induce migration, invasion, and peritoneal metastasis in colorectal cancer cells [67]. Knockdown of GALNT2 (via siRNA or CRISPR) reduced colorectal cancer cell migration and invasion.
During TGF-β-induced EMT, cancer cells acquire invasive and metastatic traits in parallel with the abnormal regulation of key GALNTs [61]. In breast cancer cell lines, GALNT14 promotes the proliferation, migration, and invasion of breast cancer cells. GALNT14 expression is associated with an increase in the mRNA expression of vimentin, N-cadherin, matrix metalloproteinase-2 (MMP-2), vascular endothelial growth factor (VEGF), TGF-β, and MMP-2 activity, with opposite effects seen in GALNT14-knockdown cells [69]. Induction of oncofetal fibronectin (onfFN) with EMT by TGF-β has been shown in cancer cells, via O-glycan addition in the IIICS domain catalyzed by GALNT3 and/or GALNT6 [70,71]. In human prostate cancer cell lines, silencing GALNT3/GALNT6 reduced onfFN with the blunting of TGF-β-induced EMT (loss of E-cadherin, gain of N-cadherin/vimentin, and increased motility) [70]. Similarly, in breast cancer cell lines, the suppression of fibronectin expression abolished the aberrant proliferative phenotype induced by GALNT6 overexpression, indicating that the GALNT–fibronectin axis is a pivotal driver of cancer progression [29].

2.3. Notch Signaling and GALNTs

Notch signaling is a highly conserved juxtracrine (cell to cell) pathway, whereby a Notch receptor on a cell binds to a ligand on a neighboring cell. This triggers proteolytic cleavages of the receptor and release of the Notch intracellular domain (NICD), which then translocate to the nucleus to regulate gene expression [72,73]. The extracellular domain of Notch can undergo O-glycosylation, which influences ligand binding/activation [74]. This has been shown to be involved with cancer progression, including metastasis [75]. The NOTCH1 receptor can be glycosylated by GALNT11 at the EGF6 and EGF36 sites in the juxtamembrane region close to the S2 cleavage site [76]. Direct evidence linking GALNTs and Notch signaling in cancer is limited in the literature. In lung adenocarcinoma, GALNT2 promotes proliferation and progression of cancer by activation of the Notch/Hes1-PTEN-PI3K/Akt signaling pathway [77]. Inhibition of GALNT2 expression by siRNA in lung cancer cells resulted in decreased levels of Notch1/3 [78]. GALNT11 expression has been shown to be a prognostic marker for Notch-driven chronic lymphoid leukemia subtypes, particularly those with high NOTCH1/2 and JAG1/2 expression [79].

2.4. GALNT and Immune Evasion in Cancers

The role of GALNTs in immune evasion stems from the way glycosylation masks tumor antigens, modulates immune checkpoint ligands, and affects immune cell recruitment and activation. The enzyme GALNT6 plays an important role in immune evasion by its action of MUC1, a heavily glycosylated mucin often overexpressed in epithelial cancers [80,81,82,83]. The heavy mucin O-glycosylation creates a dense “glycan shield” over peptide epitopes, obscuring immune recognition (by antibodies, T cells) or interfering with antigen presentation [84]. In a study using bladder cancer cell lines, Sun et al. showed that GALNT6-driven glycosylation impaired the infiltration, cytotoxic function, and effector activity of CD8+ tumor-infiltrating lymphocytes (TILs) within the tumor microenvironment (TME) [85]. Mechanistically, GALNT6-mediated O-glycosylation upregulates the expression and function of immune checkpoint molecules such as programmed death-ligand 1 (PD-L1), Cytotoxic T-lymphocyte-associated protein 4 (CTLA-4), and Lymphocyte-activation gene 3 (LAG-3), thereby enhancing inhibitory signaling cascades that suppress anti-tumor immunity. This glycosylation-dependent modulation also attenuates immune-related signaling pathways, diminishes CD8+ T cell proliferation and effector function, and promotes the secretion of immunosuppressive cytokines and mediators. Collectively, these alterations foster the development of an immunosuppressive TME, facilitating tumor immune escape and supporting malignant progression.
A study using pancreatic cancer cell lines (PaTu-8988t and KPC) showed a negative association between GALNT6 expression and immune cell infiltration [86]. In this study, knockdown of GALNT6 significantly increased the sensitivity of PDAC cells to cytotoxic T lymphocytes (CTLs) and macrophages and suppressed Programmed Cell Death Ligand 1 (PD-L1) levels in both in vitro and in vivo systems. Mechanistically, GALNT6-mediated O-glycosylation inhibited the translocation of stimulator of interferon genes (STING) and promoted their degradation in a cyclic guanosine monophosphate-adenosine monophosphate (GMP-AMP) synthase (cGAS)-independent manner, leading to reduced expression of type I interferon (IFN-β) and the chemokines C-C motif chemokine ligand 5 (CCL5) and C-X-C motif chemokine ligand 10 (CXCL10) [80,81]. These changes suppressed the recruitment of CD8+ T cells and macrophages, thereby facilitating immune evasion. Furthermore, GALNT6 stabilized PD-L1 by preventing its ubiquitin–proteasome-mediated degradation, contributing to the maintenance of an immunosuppressive tumor microenvironment [82]. The authors showed that the combination of Benzyl-α-GALNT and BMS1166 and the combination of an O-glycosylation inhibitor with PD-1/PD-L1 immune checkpoint blockade synergistically enhanced anti-tumor immune responses and reduced the viability of the tumor cells.
An integrated multi-omics analysis has revealed that post-translational modification (PTM) networks are central drivers of colorectal cancer (CRC) progression and immune evasion [87]. The study identified the widespread dysregulation of post-translational modification (PTM) pathways, with aberrant ubiquitination sustaining Wnt/β-catenin signaling and promoting tumor growth. Notably, GALNT6-mediated glycosylation emerged as a key mechanism of immune escape by stabilizing PD-L1 and preventing CD8+ T cell infiltration [83,84,86]. Single-cell transcriptomics further linked GALNT6 activity to a goblet-cell-like tumor subpopulation associated with immune exclusion. A machine learning-derived 5-gene PTM Activity Signature (including GALNT6, UBE2C, and others) achieved perfect discrimination between CRC and healthy tissues and stratified patients by immune phenotype. Therapeutically, combining O-glycosylation inhibition with PD-1/PD-L1 blockade synergistically enhanced anti-tumor immune responses. These findings establish PTM networks—particularly GALNT6-mediated glycosylation—as pivotal regulators of tumor immune dynamics and underscore their potential as both diagnostic biomarkers and therapeutic targets in colorectal cancer.
The study by Song et al. demonstrated that GALNT14 promotes lung-specific breast cancer metastasis by enhancing tumor cell self-renewal and interactions with the lung microenvironment [88]. Clinically, high GALNT14 expression in primary tumors correlated with poor lung metastasis-free survival. GALNT14 enables breast cancer cells to overcome bone morphogenetic protein (BMP)-mediated inhibition of self-renewal in the lung by glycosylating BMP receptors, supporting stem-like behavior via SOX4 upregulation [88]. Knockdown of c-JUN resulted in the inhibition of the KRAS–PI3K pathway with decreased GALNT14 expression. In vivo data from the study also showed GALNT14′s role in the recruitment of macrophages to the site of metastases, most likely by modulating the secretion of cyto/chemokines. Furthermore, GALNT3 has been shown to exert tumor-suppressive effects by limiting angiogenesis in lung cancer through EMT–dependent pathways, underscoring its broader immunomodulatory role in cancer [89,90].

3. GALNTs as Prognostic Markers in Cancer

Differential expression is associated with prognostic outcomes in various human cancers, mediated via the mechanisms described in Table 3. The gene expression levels of the various GALNTs in cancers compared with normal tissues were extrapolated from GEPIA2 (http://gepia2021.cancer-pku.cn/ (accessed on 13 July 2025)) and are shown in Figure 4 and Figure 5 [91]. GALNTs, therefore, may be used as biomarkers of prognosis and chemotherapeutic resistance. However, redundancy among isoforms complicates targeting due to functional overlap between isoforms and context-specific isoform switching, which makes it difficult for targeted therapies to be developed [92]. Our review emphasizes the fact that differential GALNT expression is associated with prognostic outcomes in various human cancers, mediated via the mechanisms described in Table 4 [93,94,95,96,97,98,99,100,101,102,103,104,105,106]. Nevertheless, further investigations are required to delineate the isoform-specific functional roles of individual GALNT family members, given their partially overlapping yet context-dependent enzymatic activities and substrate specificities. Comprehensive characterization at the transcriptional, post-transcriptional, and post-translational levels will be critical to uncover differential regulatory mechanisms and tissue- or disease-specific functions. Moreover, elucidating the molecular basis of intra-family compensatory dynamics, including potential transcriptional upregulation, altered substrate promiscuity, or subcellular delocalization in response to loss or inhibition of specific isoforms, will be essential for the rational design of targeted therapeutic interventions that can circumvent redundancy and achieve sustained glycosylation modulation in cancers.

4. Conclusions

GALNTs are key regulators of mucin-type O-glycosylation and exert diverse, isoform-specific effects on tumor progression, immune modulation, and metastasis. These enzymes are implicated in cancer progression through effects on proliferation, EMT, metastasis, and microenvironmental interactions. However, their roles remain limited by overlapping isoform functions, poorly defined substrates, and insufficient mechanistic clarity, especially regarding immune modulation. Current models often lack human-specific glycan features, and isoform-selective inhibitors are not yet clinically available. Diagnostic and therapeutic translation is still at an early stage. Future research should focus on mapping GALNT-specific substrates, developing selective inhibitors, characterizing glycan structures, and elucidating their roles in immune evasion. Integration of glycoproteomics with immuno-oncology may enable the creation of novel therapeutic strategies.

Author Contributions

Conceptualization, R.P.; methodology, S.C.I. and R.P.; software, R.P.; validation, S.C.I., D.K.S. and R.P.; formal analysis, S.C.I. and R.P.; investigation, S.C.I.; resources, S.C.I. and R.P.; data curation, S.C.I. and R.P.; writing—original draft preparation, S.C.I.; writing—review and editing, R.P. and D.K.S.; visualization, R.P.; supervision, D.K.S. and R.P.; project administration, R.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
3DThree-dimensional
AKTAk strain transforming
BMPBone morphogenetic protein
C1GALT1Core 1 synthase, glycoprotein-N-acetylgalactosamine 3-beta-galactosyltransferase 1
CCL5C-C motif chemokine ligand 5
CGNCis-Golgi network
c-JunTranscription factor Jun
COPICoat Protein I
COSMCCore-1 β1-3galactosyltransferase-specific chaperone 1
CTLCytotoxic T lymphocytes
CTLA-4Cytotoxic T-lymphocyte-associated protein 4
CRISPRClustered regularly interspaced short palindromic repeats
CXCL10C-X-C motif chemokine ligand 10
EGFEpithelial growth factor
EGFREpithelial growth factor receptors
EMTEpithelial-mesenchymal transition
EREndoplasmic reticulum
ErbBErythroblastic oncogene B
ERK8Extracellular Signal-Regulated Kinase 8
FOXO1Forkhead box O1 transcription factor
FUTFucosyltransferase
GALAGALNT activation
GALNTN-acetylgalactosaminyltransferase
GlcNAcN-Acetylglucosamine
GMP-AMPGuanosine monophosphate-adenosine monophosphate
GTGlycosyltransferase
GT-AGlycosyltransferase fold A
GT-BGlycosyltransferase fold B
GT-CGlycosyltransferase fold C
HCCHepatocellular carcinoma
Her2Human epidermal growth factor receptor 2
KRASKirsten ras oncogene homolog
LAG-3Lymphocyte-activation gene 3
MMPMatrix metalloproteinase-2
mTORmammalian target of rapamycin
MUC1Mucin-1
NICDNotch intracellular domain
onfFNOncofetal fibronectin
PD-L1Programmed Cell Death Ligand 1
PI3KPhosphoinositide 3-kinase
PTENPhosphatase and tensin homolog
PTMPost-translational modification
SerSerine
siRNASmall interfering RNA
SLUGSNAI2 gene
SNAILZinc finger protein SNAI1
ST3GalST3 beta-galactoside alpha-2,3-sialyltransferase 1
STINGStimulator of interferon genes
TGF-βTransforming growth factor beta
ThrThreonine
TnThomsen-nouveau
TGNTrans-Golgi network
TMETumour microenvironment
UBE2CUbiquitin-conjugating enzyme E2C
UDP-GalNAcUridine diphosphate N-acetylgalactosamine
VEGFVascular endothelial growth factor
VTCVesicular tubular clusters
VVAVicia villosa Lectin
ZEB1Zinc finger E-box binding homeobox 1

References

  1. Eichler, J. Protein glycosylation. Curr. Biol. 2019, 29, R229–R231. [Google Scholar] [CrossRef]
  2. Steen, P.V.D.; Rudd, P.M.; Dwek, R.A.; Opdenakker, G. Concepts and principles of O-linked glycosylation. Crit. Rev. Biochem. Mol. Biol. 1998, 33, 151–208. [Google Scholar] [CrossRef]
  3. Hounsell, E.F.; Davies, M.J.; Renouf, D.V. O-linked protein glycosylation structure and function. Glycoconj. J. 1996, 13, 19–26. [Google Scholar] [CrossRef]
  4. 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]
  5. Pinto, D.; Parameswaran, R. Role of Truncated O-GalNAc Glycans in Cancer Progression and Metastasis in Endocrine Cancers. Cancers 2023, 15, 3266. [Google Scholar] [CrossRef] [PubMed]
  6. Munkley, J.; Elliott, D.J. Hallmarks of glycosylation in cancer. Oncotarget 2016, 7, 35478. [Google Scholar] [CrossRef] [PubMed]
  7. Meany, D.L.; Chan, D.W. Aberrant glycosylation associated with enzymes as cancer biomarkers. Clin. Proteom. 2011, 8, 1–14. [Google Scholar] [CrossRef]
  8. Paulson, J.C.; Colley, K.J. Glycosyltransferases: Structure, localization, and control of cell type-specific glycosylation. J. Biol. Chem. 1989, 264, 17615–17618. [Google Scholar] [CrossRef] [PubMed]
  9. Dodd, R.; Drickamer, K. Lectin-like proteins in model organisms: Implications for evolution of carbohydrate-binding activity. Glycobiology 2001, 11, 71R–79R. [Google Scholar] [CrossRef]
  10. Hagen, F.K.; Hazes, B.; Raffo, R.; DeSa, D.; Tabak, L.A. Structure-Function Analysis of the UDP-N-acetyl-d-galactosamine: PolypeptideN-acetylgalactosaminyltransferase: Essential Residues Lie In A Predicted Active Site Cleft Resembling a Lactose Repressor Fold. J. Biol. Chem. 1999, 274, 6797–6803. [Google Scholar] [CrossRef]
  11. Bourne, Y.; Henrissat, B. Glycoside hydrolases and glycosyltransferases: Families and functional modules. Curr. Opin. Struct. Biol. 2001, 11, 593–600. [Google Scholar] [CrossRef]
  12. Varki, A.; Cummings, R.D.; Esko, J.D.; Stanley, P.; Hart, G.W.; Aebi, M.; Darvill, A.G.; Kinoshita, T.; Packer, N.H.; Prestegard, J.H. Essentials of Glycobiology [Internet], 3rd ed.; Cold Spring Harbor Laboratory Press: Cold Spring Harbor, NY, USA, 2015. [Google Scholar]
  13. 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. Essentials of Glycobiology [Internet], 4th ed.; Cold Spring Harbor Laboratory Press: Cold Spring Harbor, NY, USA, 2022. [Google Scholar]
  14. Gill, D.J.; Tham, K.M.; Chia, J.; Wang, S.C.; Steentoft, C.; Clausen, H.; Bard-Chapeau, E.A.; Bard, F.A. Initiation of GalNAc-type O-glycosylation in the endoplasmic reticulum promotes cancer cell invasiveness. Proc. Natl. Acad. Sci. USA 2013, 110, E3152–E3161. [Google Scholar] [CrossRef]
  15. Chia, J.; Wang, S.C.; Wee, S.; Gill, D.J.; Tay, F.; Kannan, S.; Verma, C.S.; Gunaratne, J.; Bard, F.A. Src activates retrograde membrane traffic through phosphorylation of GBF1. Elife 2021, 10, e68678. [Google Scholar] [CrossRef]
  16. Del Giudice, S.; De Luca, V.; Parizadeh, S.; Russo, D.; Luini, A.; Di Martino, R. Endogenous and Exogenous Regulatory Signaling in the Secretory Pathway: Role of Golgi Signaling Molecules in Cancer. Front. Cell Dev. Biol. 2022, 10, 833663. [Google Scholar] [CrossRef]
  17. Hagen, K.G.T.; Fritz, T.A.; Tabak, L.A. All in the family: The UDP-GalNAc:polypeptide N-acetylgalactosaminyltransferases. Glycobiology 2003, 13, 1r–16r. [Google Scholar] [CrossRef]
  18. Mahajan, S.P.; Srinivasan, Y.; Labonte, J.W.; DeLisa, M.P.; Gray, J.J. Structural basis for peptide substrate specificities of glycosyltransferase GalNAc-T2. ACS Catal. 2021, 11, 2977–2991. [Google Scholar] [CrossRef]
  19. Gerken, T.A.; Raman, J.; Fritz, T.A.; Jamison, O. Identification of common and unique peptide substrate preferences for the UDP-GalNAc: Polypeptide α-N-acetylgalactosaminyltransferases T1 and T2 derived from oriented random peptide substrates. J. Biol. Chem. 2006, 281, 32403–32416. [Google Scholar] [CrossRef] [PubMed]
  20. Gerken, T.A.; Jamison, O.; Perrine, C.L.; Collette, J.C.; Moinova, H.; Ravi, L.; Markowitz, S.D.; Shen, W.; Patel, H.; Tabak, L.A. Emerging paradigms for the initiation of mucin-type protein O-glycosylation by the polypeptide GalNAc transferase family of glycosyltransferases. J. Biol. Chem. 2011, 286, 14493–14507. [Google Scholar] [CrossRef]
  21. Wu, Y.M.; Liu, C.H.; Hu, R.H.; Huang, M.J.; Lee, J.J.; Chen, C.H.; Huang, J.; Lai, H.S.; Lee, P.H.; Hsu, W.M.; et al. Mucin glycosylating enzyme GALNT2 regulates the malignant character of hepatocellular carcinoma by modifying the EGF receptor. Cancer Res. 2011, 71, 7270–7279. [Google Scholar] [CrossRef] [PubMed]
  22. Osman, T.E.; Guo, Y.; Li, S. Exploring the combined roles of GALNT1 and GALNT2 in hepatocellular carcinoma malignancy and EGFR modulation. Discov. Oncol. 2025, 16, 337. [Google Scholar] [CrossRef] [PubMed]
  23. Huang, M.-J.; Hu, R.-H.; Chou, C.-H.; Hsu, C.-L.; Liu, Y.-W.; Huang, J.; Hung, J.-S.; Lai, I.-R.; Juan, H.-F.; Yu, S.-L.; et al. Knockdown of GALNT1 suppresses malignant phenotype of hepatocellular carcinoma by suppressing EGFR signaling. Oncotarget 2015, 6, 5650–5665. [Google Scholar] [CrossRef]
  24. Wang, Z.Q.; Bachvarova, M.; Morin, C.; Plante, M.; Gregoire, J.; Renaud, M.C.; Sebastianelli, A.; Bachvarov, D. Role of the polypeptide N-acetylgalactosaminyltransferase 3 in ovarian cancer progression: Possible implications in abnormal mucin O-glycosylation. Oncotarget 2014, 5, 544–560. [Google Scholar] [CrossRef]
  25. Sheta, R.; Woo, C.M.; Roux-Dalvai, F.; Fournier, F.; Bourassa, S.; Droit, A.; Bertozzi, C.R.; Bachvarov, D. A metabolic labeling approach for glycoproteomic analysis reveals altered glycoprotein expression upon GALNT3 knockdown in ovarian cancer cells. J. Proteom. 2016, 145, 91–102. [Google Scholar] [CrossRef]
  26. Li, H.-w.; Liu, M.-b.; Jiang, X.; Song, T.; Feng, S.-x.; Wu, J.-y.; Deng, P.-f.; Wang, X.-y. GALNT14 Regulates Ferroptosis and Apoptosis of Ovarian Cancer Through the EGFR/mTOR Pathway. Future Oncol. 2022, 18, 149–161. [Google Scholar] [CrossRef]
  27. Gao, D.; Zhu, Z.; Yu, J.; Wei, S.; Xing, C. GALNT4-induced O-GalNAc glycosylation of MUC1 accelerates the invasive phenotype of PDAC. Pancreatology 2025, 25, 939–949. [Google Scholar] [CrossRef]
  28. Liu, C.; Li, Z.; Xu, L.; Shi, Y.; Zhang, X.; Shi, S.; Hou, K.; Fan, Y.; Li, C.; Wang, X.; et al. GALNT6 promotes breast cancer metastasis by increasing mucin-type O-glycosylation of α2M. Aging 2020, 12, 11794–11811. [Google Scholar] [CrossRef] [PubMed]
  29. Park, J.H.; Katagiri, T.; Chung, S.; Kijima, K.; Nakamura, Y. Polypeptide N-acetylgalactosaminyltransferase 6 disrupts mammary acinar morphogenesis through O-glycosylation of fibronectin. Neoplasia 2011, 13, 320–326. [Google Scholar] [CrossRef]
  30. Scott, E.; Hodgson, K.; Calle, B.; Turner, H.; Cheung, K.; Bermudez, A.; Marques, F.J.G.; Pye, H.; Yo, E.C.; Islam, K.; et al. Upregulation of GALNT7 in prostate cancer modifies O-glycosylation and promotes tumour growth. Oncogene 2023, 42, 926–937. [Google Scholar] [CrossRef]
  31. Munkley, J.; Vodak, D.; Livermore, K.E.; James, K.; Wilson, B.T.; Knight, B.; McCullagh, P.; McGrath, J.; Crundwell, M.; Harries, L.W.; et al. Glycosylation is an Androgen-Regulated Process Essential for Prostate Cancer Cell Viability. EBioMedicine 2016, 8, 103–116. [Google Scholar] [CrossRef] [PubMed]
  32. Teslovich, T.M.; Musunuru, K.; Smith, A.V.; Edmondson, A.C.; Stylianou, I.M.; Koseki, M.; Pirruccello, J.P.; Ripatti, S.; Chasman, D.I.; Willer, C.J.; et al. Biological, clinical and population relevance of 95 loci for blood lipids. Nature 2010, 466, 707–713. [Google Scholar] [CrossRef] [PubMed]
  33. Schjoldager, K.T.; Narimatsu, Y.; Joshi, H.J.; Clausen, H. Global view of human protein glycosylation pathways and functions. Nat. Rev. Mol. Cell Biol. 2020, 21, 729–749. [Google Scholar] [CrossRef]
  34. Madsen, T.D.; Hansen, L.H.; Hintze, J.; Ye, Z.; Jebari, S.; Andersen, D.B.; Joshi, H.J.; Ju, T.; Goetze, J.P.; Martin, C.; et al. An atlas of O-linked glycosylation on peptide hormones reveals diverse biological roles. Nat. Commun. 2020, 11, 4033. [Google Scholar] [CrossRef] [PubMed]
  35. Häuselmann, I.; Borsig, L. Altered tumor-cell glycosylation promotes metastasis. Front. Oncol. 2014, 4, 28. [Google Scholar] [CrossRef]
  36. 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]
  37. Ju, T.; Aryal, R.P.; Stowell, C.J.; Cummings, R.D. Regulation of protein O-glycosylation by the endoplasmic reticulum-localized molecular chaperone Cosmc. J. Cell Biol. 2008, 182, 531–542. [Google Scholar] [CrossRef]
  38. Ju, T.; Lanneau, G.S.; Gautam, T.; Wang, Y.; Xia, B.; Stowell, S.R.; Willard, M.T.; Wang, W.; Xia, J.Y.; Zuna, R.E.; et al. Human tumor antigens Tn and sialyl Tn arise from mutations in Cosmc. Cancer Res. 2008, 68, 1636–1646. [Google Scholar] [CrossRef] [PubMed]
  39. Gill, D.J.; Chia, J.; Senewiratne, J.; Bard, F. Regulation of O-glycosylation through Golgi-to-ER relocation of initiation enzymes. J. Cell Biol. 2010, 189, 843–858. [Google Scholar] [CrossRef]
  40. Marcos, N.T.; Pinho, S.; Grandela, C.; Cruz, A.; Samyn-Petit, B.; Harduin-Lepers, A.; Almeida, R.; Silva, F.; Morais, V.; Costa, J.; et al. Role of the human ST6GalNAc-I and ST6GalNAc-II in the synthesis of the cancer-associated sialyl-Tn antigen. Cancer Res. 2004, 64, 7050–7057. [Google Scholar] [CrossRef] [PubMed]
  41. Raghu, D.; Mobley, R.J.; Shendy, N.A.M.; Perry, C.H.; Abell, A.N. GALNT3 Maintains the Epithelial State in Trophoblast Stem Cells. Cell Rep. 2019, 26, 3684–3697.e3687. [Google Scholar] [CrossRef]
  42. Klumperman, J. Architecture of the mammalian Golgi. Cold Spring Harb. Perspect. Biol. 2011, 3, a005181. [Google Scholar] [CrossRef] [PubMed]
  43. Ladinsky, M.S.; Mastronarde, D.N.; McIntosh, J.R.; Howell, K.E.; Staehelin, L.A. Golgi structure in three dimensions: Functional insights from the normal rat kidney cell. J. Cell Biol. 1999, 144, 1135–1149. [Google Scholar] [CrossRef]
  44. Szul, T.; Sztul, E. COPII and COPI traffic at the ER-Golgi interface. Physiology 2011, 26, 348–364. [Google Scholar] [CrossRef]
  45. Harada, A.; Kunii, M.; Kurokawa, K.; Sumi, T.; Kanda, S.; Zhang, Y.; Nadanaka, S.; Hirosawa, K.M.; Tokunaga, K.; Tojima, T.; et al. Dynamic movement of the Golgi unit and its glycosylation enzyme zones. Nat. Commun. 2024, 15, 4514–4519. [Google Scholar] [CrossRef]
  46. Roth, J.; Taatjes, D.J.; Lucocq, J.M.; Weinstein, J.; Paulson, J.C. Demonstration of an extensive trans-tubular network continuous with the Golgi apparatus stack that may function in glycosylation. Cell 1985, 43, 287–295. [Google Scholar] [CrossRef] [PubMed]
  47. Röttger, S.; White, J.; Wandall, H.H.; Olivo, J.C.; Stark, A.; Bennett, E.P.; Whitehouse, C.; Berger, E.G.; Clausen, H.; Nilsson, T. Localization of three human polypeptide GalNAc-transferases in HeLa cells suggests initiation of O-linked glycosylation throughout the Golgi apparatus. J. Cell Sci. 1998, 111, 45–60. [Google Scholar] [CrossRef]
  48. Hayes, B.K.; Varki, A. The biosynthesis of oligosaccharides in intact Golgi preparations from rat liver. Analysis of N-linked and O-linked glycans labeled by UDP-[6-3H]N-acetylgalactosamine. J. Biol. Chem. 1993, 268, 16170–16178. [Google Scholar] [CrossRef] [PubMed]
  49. Gill, D.J.; Clausen, H.; Bard, F. Location, location, location: New insights into O-GalNAc protein glycosylation. Trends Cell Biol. 2011, 21, 149–158. [Google Scholar] [CrossRef]
  50. Gerken, T.A.; Owens, C.L.; Pasumarthy, M. Site-specific core 1 O-glycosylation pattern of the porcine submaxillary gland mucin tandem repeat. Evidence for the modulation of glycan length by peptide sequence. J. Biol. Chem. 1998, 273, 26580–26588. [Google Scholar] [CrossRef]
  51. Topaz, O.; Shurman, D.L.; Bergman, R.; Indelman, M.; Ratajczak, P.; Mizrachi, M.; Khamaysi, Z.; Behar, D.; Petronius, D.; Friedman, V.; et al. Mutations in GALNT3, encoding a protein involved in O-linked glycosylation, cause familial tumoral calcinosis. Nat. Genet. 2004, 36, 579–581. [Google Scholar] [CrossRef]
  52. Brockhausen, I. Mucin-type O-glycans in human colon and breast cancer: Glycodynamics and functions. EMBO Rep. 2006, 7, 599–604. [Google Scholar] [CrossRef] [PubMed]
  53. Chia, J.; Tay, F.; Bard, F. The GalNAc-T Activation (GALA) Pathway: Drivers and markers. PLoS ONE 2019, 14, e0214118. [Google Scholar] [CrossRef]
  54. Chia, J.; Tham, K.M.; Gill, D.J.; Bard-Chapeau, E.A.; Bard, F.A. ERK8 is a negative regulator of O-GalNAc glycosylation and cell migration. Elife 2014, 3, e01828. [Google Scholar] [CrossRef] [PubMed]
  55. Hussain, M.R.; Hoessli, D.C.; Fang, M. N-acetylgalactosaminyltransferases in cancer. Oncotarget 2016, 7, 54067–54081. [Google Scholar] [CrossRef] [PubMed]
  56. Beaman, E.-M.; Carter, D.R.F.; Brooks, S.A. GALNTs: Master regulators of metastasis-associated epithelial-mesenchymal transition (EMT)? Glycobiology 2022, 32, 556–579. [Google Scholar] [CrossRef] [PubMed]
  57. Beaman, E.M.; Brooks, S.A. The extended ppGalNAc-T family and their functional involvement in the metastatic cascade. Histol. Histopathol. 2014, 29, 293–304. [Google Scholar] [CrossRef]
  58. Hu, W.T.; Yeh, C.C.; Liu, S.Y.; Huang, M.C.; Lai, I.R. The O-glycosylating enzyme GALNT2 suppresses the malignancy of gastric adenocarcinoma by reducing EGFR activities. Am. J. Cancer Res. 2018, 8, 1739–1751. [Google Scholar]
  59. Detarya, M.; Lert-Itthiporn, W.; Mahalapbutr, P.; Klaewkla, M.; Sorin, S.; Sawanyawisuth, K.; Silsirivanit, A.; Seubwai, W.; Wongkham, C.; Araki, N.; et al. Emerging roles of GALNT5 on promoting EGFR activation in cholangiocarcinoma: A mechanistic insight. Am. J. Cancer Res. 2022, 12, 4140–4159. [Google Scholar]
  60. Chugh, S.; Meza, J.; Sheinin, Y.M.; Ponnusamy, M.P.; Batra, S.K. Loss of N-acetylgalactosaminyltransferase 3 in poorly differentiated pancreatic cancer: Augmented aggressiveness and aberrant ErbB family glycosylation. Br. J. Cancer 2016, 114, 1376–1386. [Google Scholar] [CrossRef]
  61. Zhang, J.; Dijke, P.T.; Wuhrer, M.; Zhang, T. Role of glycosylation in TGF-β signaling and epithelial-to-mesenchymal transition in cancer. Protein. Cell 2020, 12, 89–106. [Google Scholar] [CrossRef]
  62. Aashaq, S.; Batool, A.; Mir, S.A.; Beigh, M.A.; Andrabi, K.I.; Shah, Z.A. TGF-β signaling: A recap of SMAD-independent and SMAD-dependent pathways. J. Cell. Physiol. 2022, 237, 59–85. [Google Scholar] [CrossRef]
  63. Hugo, H.; Ackland, M.L.; Blick, T.; Lawrence, M.G.; Clements, J.A.; Williams, E.D.; Thompson, E.W. Epithelial--mesenchymal and mesenchymal--epithelial transitions in carcinoma progression. J. Cell. Physiol. 2007, 213, 374–383. [Google Scholar] [CrossRef] [PubMed]
  64. Kalluri, R.; Weinberg, R.A. The basics of epithelial-mesenchymal transition. J. Clin. Investig. 2009, 119, 1420–1428. [Google Scholar] [CrossRef]
  65. Wu, Q.; Zhang, C.; Zhang, K.; Chen, Q.; Wu, S.; Huang, H.; Huang, T.; Zhang, N.; Wang, X.; Li, W.; et al. ppGalNAc-T4-catalyzed O-Glycosylation of TGF-β type II receptor regulates breast cancer cells metastasis potential. J. Biol. Chem. 2021, 296, 10019. [Google Scholar] [CrossRef]
  66. Liu, S.Y.; Shun, C.T.; Hung, K.Y.; Juan, H.F.; Hsu, C.L.; Huang, M.C.; Lai, I.R. Mucin glycosylating enzyme GALNT2 suppresses malignancy in gastric adenocarcinoma by reducing MET phosphorylation. Oncotarget 2016, 7, 11251–11262. [Google Scholar] [CrossRef] [PubMed]
  67. Liao, Y.Y.; Chuang, Y.T.; Lin, H.Y.; Lin, N.Y.; Hsu, T.W.; Hsieh, S.C.; Chen, S.T.; Hung, J.S.; Yang, H.J.; Liang, J.T.; et al. GALNT2 promotes invasiveness of colorectal cancer cells partly through AXL. Mol. Oncol. 2023, 17, 119–133. [Google Scholar] [CrossRef]
  68. Hu, Q.; Tian, T.; Leng, Y.; Tang, Y.; Chen, S.; Lv, Y.; Liang, J.; Liu, Y.; Liu, T.; Shen, L.; et al. The O-glycosylating enzyme GALNT2 acts as an oncogenic driver in non-small cell lung cancer. Cell. Mol. Biol. Lett. 2022, 27, 71. [Google Scholar] [CrossRef] [PubMed]
  69. Huanna, T.; Tao, Z.; Xiangfei, W.; Longfei, A.; Yuanyuan, X.; Jianhua, W.; Cuifang, Z.; Manjing, J.; Wenjing, C.; Shaochuan, Q.; et al. GALNT14 mediates tumor invasion and migration in breast cancer cell MCF-7. Mol. Carcinog. 2015, 54, 1159–1171. [Google Scholar] [CrossRef]
  70. Freire-de-Lima, L.; Gelfenbeyn, K.; Ding, Y.; Mandel, U.; Clausen, H.; Handa, K.; Hakomori, S.I. Involvement of O-glycosylation defining oncofetal fibronectin in epithelial-mesenchymal transition process. Proc. Natl. Acad. Sci. USA 2011, 108, 17690–17695. [Google Scholar] [CrossRef]
  71. Ventura, E.; Weller, M.; Macnair, W.; Eschbach, K.; Beisel, C.; Cordazzo, C.; Claassen, M.; Zardi, L.; Burghardt, I. TGF-β induces oncofetal fibronectin that, in turn, modulates TGF-β superfamily signaling in endothelial cells. J. Cell Sci. 2018, 131, jcs209619. [Google Scholar] [CrossRef]
  72. Bolós, V.; Blanco, M.; Medina, V.; Aparicio, G.; Díaz-Prado, S.; Grande, E. Notch signalling in cancer stem cells. Clin. Transl. Oncol. 2009, 11, 11–19. [Google Scholar] [CrossRef]
  73. Miele, L.; Golde, T.; Osborne, B. Notch Signaling in Cancer. Curr. Mol. Med. 2006, 6, 905–918. [Google Scholar] [CrossRef]
  74. Pandey, A.; Niknejad, N.; Jafar-Nejad, H. Multifaceted regulation of Notch signaling by glycosylation. Glycobiology 2021, 31, 8–28. [Google Scholar] [CrossRef]
  75. Sethi, N.; Kang, Y. Notch signalling in cancer progression and bone metastasis. Br. J. Cancer 2011, 105, 1805–1810. [Google Scholar] [CrossRef]
  76. Boskovski, M.T.; Yuan, S.; Pedersen, N.B.; Goth, C.K.; Makova, S.; Clausen, H.; Brueckner, M.; Khokha, M.K. The heterotaxy gene GALNT11 glycosylates Notch to orchestrate cilia type and laterality. Nature 2013, 504, 456–459. [Google Scholar] [CrossRef]
  77. Wang, W.; Sun, R.; Zeng, L.; Chen, Y.; Zhang, N.; Cao, S.; Deng, S.; Meng, X.; Yang, S. GALNT2 promotes cell proliferation, migration, and invasion by activating the Notch/Hes1-PTEN-PI3K/Akt signaling pathway in lung adenocarcinoma. Life Sci. 2021, 276, 119439. [Google Scholar] [CrossRef] [PubMed]
  78. Miao, C.; Sun, R.; Ji, D.; Wu, M.; Fu, Q.; Mei, L.; Wu, Z. Mechanism of the GALNT family proteins in regulating tumorigenesis and development of lung cancer (Review). Mol. Clin. Oncol. 2025, 22, 37. [Google Scholar] [CrossRef] [PubMed]
  79. Libisch, M.G.; Casás, M.; Chiribao, M.; Moreno, P.; Cayota, A.; Osinaga, E.; Oppezzo, P.; Robello, C. GALNT11 as a new molecular marker in chronic lymphocytic leukemia. Gene 2014, 533, 270–279. [Google Scholar] [CrossRef]
  80. Brooks, S.A.; Carter, T.M.; Bennett, E.P.; Clausen, H.; Mandel, U. Immunolocalisation of members of the polypeptide N-acetylgalactosaminyl transferase (ppGalNAc-T) family is consistent with biologically relevant altered cell surface glycosylation in breast cancer. Acta. Histochem. 2007, 109, 273–284. [Google Scholar] [CrossRef]
  81. Liesche, F.; Kölbl, A.C.; Ilmer, M.; Hutter, S.; Jeschke, U.; Andergassen, U. Role of N-acetylgalactosaminyltransferase 6 in early tumorigenesis and formation of metastasis. Mol. Med. Rep. 2016, 13, 4309–4314. [Google Scholar] [CrossRef]
  82. Tarhan, Y.E.; Kato, T.; Jang, M.; Haga, Y.; Ueda, K.; Nakamura, Y.; Park, J.H. Morphological Changes, Cadherin Switching, and Growth Suppression in Pancreatic Cancer by GALNT6 Knockdown. Neoplasia 2016, 18, 265–272. [Google Scholar] [CrossRef] [PubMed]
  83. Gomes, J.; Marcos, N.T.; Berois, N.; Osinaga, E.; Magalhães, A.; Pinto-de-Sousa, J.; Almeida, R.; Gärtner, F.; Reis, C.A. Expression of UDP-N-acetyl-D-galactosamine: Polypeptide N-acetylgalactosaminyltransferase-6 in gastric mucosa, intestinal metaplasia, and gastric carcinoma. J. Histochem. Cytochem. 2009, 57, 79–86. [Google Scholar] [CrossRef] [PubMed]
  84. Zhang, Y.; Sun, L.; Lei, C.; Li, W.; Han, J.; Zhang, J.; Zhang, Y. A Sweet Warning: Mucin-Type O-Glycans in Cancer. Cells 2022, 11, 3666. [Google Scholar] [CrossRef] [PubMed]
  85. Sun, X.; Wu, H.; Tang, L.; Al-Danakh, A.; Jian, Y.; Gong, L.; Li, C.; Yu, X.; Zeng, G.; Chen, Q.; et al. GALNT6 promotes bladder cancer malignancy and immune escape by epithelial-mesenchymal transition and CD8(+) T cells. Cancer Cell Int. 2024, 24, 308. [Google Scholar] [CrossRef] [PubMed]
  86. Chong, S.; Liu, Y.; Bian, Z.; Hu, D.; Guo, S.; Dong, C.; Zeng, J.; Fan, S.; Chen, X. GALNT6 dual regulates innate immunity STING signaling and PD-L1 expression to promote immune evasion in pancreatic ductal adenocarcinoma. Cell. Signal. 2025, 134, 111942. [Google Scholar] [CrossRef]
  87. Yang, G.; Ji, L.; Lv, C.; Zhao, C.; Ma, R.; Li, Y.; Hu, Y.; Pan, L. Integrated multi-omics analysis reveals PTM networks as key regulators of colorectal cancer progression and immune evasion. Discov. Oncol. 2025, 16, 1725. [Google Scholar] [CrossRef]
  88. Song, K.H.; Park, M.S.; Nandu, T.S.; Gadad, S.; Kim, S.C.; Kim, M.Y. GALNT14 promotes lung-specific breast cancer metastasis by modulating self-renewal and interaction with the lung microenvironment. Nat. Commun. 2016, 7, 13796. [Google Scholar] [CrossRef]
  89. Park, M.S.; Yang, A.Y.; Lee, J.E.; Kim, S.K.; Roe, J.-s.; Park, M.-S.; Oh, M.J.; An, H.J.; Kim, M.-Y. GALNT3 suppresses lung cancer by inhibiting myeloid-derived suppressor cell infiltration and angiogenesis in a TNFR and c-MET pathway-dependent manner. Cancer Lett. 2021, 521, 294–307. [Google Scholar] [CrossRef]
  90. Zhang, Q.; Burdette, J.E.; Wang, J.-P. Integrative network analysis of TCGA data for ovarian cancer. BMC Syst. Biol. 2014, 8, 1338. [Google Scholar] [CrossRef]
  91. Tang, Z.; Kang, B.; Li, C.; Chen, T.; Zhang, Z. GEPIA2: An enhanced web server for large-scale expression profiling and interactive analysis. Nucleic Acids Res. 2019, 47, W556–W560. [Google Scholar] [CrossRef]
  92. Schjoldager, K.T.; Joshi, H.J.; Kong, Y.; Goth, C.K.; King, S.L.; Wandall, H.H.; Bennett, E.P.; Vakhrushev, S.Y.; Clausen, H. Deconstruction of O-glycosylation--GalNAc-T isoforms direct distinct subsets of the O-glycoproteome. EMBO Rep. 2015, 16, 1713–1722. [Google Scholar] [CrossRef]
  93. Kimura, R.; Yoshimaru, T.; Matsushita, Y.; Matsuo, T.; Ono, M.; Park, J.H.; Sasa, M.; Miyoshi, Y.; Nakamura, Y.; Katagiri, T. The GALNT6-LGALS3BP axis promotes breast cancer cell growth. Int. J. Oncol. 2020, 56, 581–595. [Google Scholar] [CrossRef]
  94. Fernandez, E.; Ubillos, L.; Elgul, N.; Festari, M.F.; Mazal, D.; Pritsch, O.; Alonso, I.; Osinaga, E.; Berois, N. Polypeptide-GalNAc-Transferase-13 Shows Prognostic Impact in Breast Cancer. Cancers 2021, 13, 5616. [Google Scholar] [CrossRef]
  95. Shibao, K.; Izumi, H.; Nakayama, Y.; Ohta, R.; Nagata, N.; Nomoto, M.; Matsuo, K.; Yamada, Y.; Kitazato, K.; Itoh, H.; et al. Expression of UDP-N-acetyl-alpha-D-galactosamine-polypeptide galNAc N-acetylgalactosaminyl transferase-3 in relation to differentiation and prognosis in patients with colorectal carcinoma. Cancer 2002, 94, 1939–1946. [Google Scholar] [CrossRef]
  96. Yan, X.; Lu, J.; Zou, X.; Zhang, S.; Cui, Y.; Zhou, L.; Liu, F.; Shan, A.; Lu, J.; Zheng, M.; et al. The polypeptide N-acetylgalactosaminyltransferase 4 exhibits stage-dependent expression in colorectal cancer and affects tumorigenesis, invasion and differentiation. FEBS J. 2018, 285, 3041–3055. [Google Scholar] [CrossRef]
  97. Ubillos, L.; Berriel, E.; Mazal, D.; Victoria, S.; Barrios, E.; Osinaga, E.; Berois, N. Polypeptide-GalNAc-T6 expression predicts better overall survival in patients with colon cancer. Oncol. Lett. 2018, 16, 225–234. [Google Scholar] [CrossRef] [PubMed]
  98. Guo, J.M.; Chen, H.L.; Wang, G.M.; Zhang, Y.K.; Narimatsu, H. Expression of UDP-GalNAc:polypeptide N-acetylgalactosaminyltransferase-12 in gastric and colonic cancer cell lines and in human colorectal cancer. Oncology 2004, 67, 271–276. [Google Scholar] [CrossRef] [PubMed]
  99. Onitsuka, K.; Shibao, K.; Nakayama, Y.; Minagawa, N.; Hirata, K.; Izumi, H.; Matsuo, K.; Nagata, N.; Kitazato, K.; Kohno, K.; et al. Prognostic significance of UDP-N-acetyl-alpha-D-galactosamine:polypeptide N-acetylgalactosaminyltransferase-3 (GalNAc-T3) expression in patients with gastric carcinoma. Cancer Sci. 2003, 94, 32–36. [Google Scholar] [CrossRef]
  100. He, H.; Shen, Z.; Zhang, H.; Wang, X.; Tang, Z.; Xu, J.; Sun, Y. Clinical significance of polypeptide N-acetylgalactosaminyl transferase-5 (GalNAc-T5) expression in patients with gastric cancer. Br. J. Cancer 2014, 110, 2021–2029. [Google Scholar] [CrossRef]
  101. Guo, Y.; Shi, J.; Zhang, J.; Li, H.; Liu, B.; Guo, H. Polypeptide N-acetylgalactosaminyltransferase-6 expression in gastric cancer. Onco. Targets Ther. 2017, 10, 3337–3344. [Google Scholar] [CrossRef] [PubMed]
  102. Xu, G.; Wu, Y.L.; Li, N.; Xu, R.; Zhang, J.B.; Ming, H.; Zhang, Y. GALNT10 promotes the proliferation and metastatic ability of gastric cancer and reduces 5-fluorouracil sensitivity by activating HOXD13. Eur. Rev. Med. Pharmacol. Sci. 2020, 24, 11610–11619. [Google Scholar] [CrossRef]
  103. Yamamoto, S.; Nakamori, S.; Tsujie, M.; Takahashi, Y.; Nagano, H.; Dono, K.; Umeshita, K.; Sakon, M.; Tomita, Y.; Hoshida, Y.; et al. Expression of uridine diphosphate N-acetyl-alpha-D-galactosamine: Polypeptide N-acetylgalactosaminyl transferase 3 in adenocarcinoma of the pancreas. Pathobiology 2004, 71, 12–18. [Google Scholar] [CrossRef] [PubMed]
  104. Yan, J.; Gong, H.; Han, S.; Liu, J.; Wu, Z.; Wang, Z.; Wang, T. GALNT5 functions as a suppressor of ferroptosis and a predictor of poor prognosis in pancreatic adenocarcinoma. Am. J. Cancer Res. 2023, 13, 4579–4596. [Google Scholar] [PubMed]
  105. Li, Z.; Yamada, S.; Inenaga, S.; Imamura, T.; Wu, Y.; Wang, K.Y.; Shimajiri, S.; Nakano, R.; Izumi, H.; Kohno, K.; et al. Polypeptide N-acetylgalactosaminyltransferase 6 expression in pancreatic cancer is an independent prognostic factor indicating better overall survival. Br. J. Cancer 2011, 104, 1882–1889. [Google Scholar] [CrossRef]
  106. Mochizuki, Y.; Ito, K.; Izumi, H.; Kohno, K.; Amano, J. Expression of polypeptide N-acetylgalactosaminyl transferase-3 and its association with clinicopathological factors in thyroid carcinomas. Thyroid 2013, 23, 1553–1560. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The domain architecture of GALNT showing linkage between GT-A motif and uracil ring of the UDP-GALNT molecule. The various GALNT enzymes that bind specifically to the alpha and beta domains of the ricin like lectin are highlighted in the green box (as indicated by the arrows). Abbreviations: GALNT—polypeptide N-acetylgalactosaminyltransferase; GT-A—glycosyltransferase-A; UDP—uridine diphosphate.
Figure 1. The domain architecture of GALNT showing linkage between GT-A motif and uracil ring of the UDP-GALNT molecule. The various GALNT enzymes that bind specifically to the alpha and beta domains of the ricin like lectin are highlighted in the green box (as indicated by the arrows). Abbreviations: GALNT—polypeptide N-acetylgalactosaminyltransferase; GT-A—glycosyltransferase-A; UDP—uridine diphosphate.
Biomedicines 14 00005 g001
Figure 2. Differences in action of group I and II GALNTs. Yellow square = GalNac; blue square = GlcNac; yellow circle = galactose. Abbreviations: GALNT—polypeptide N-acetylgalactosaminyltransferase; Tn—Thomsen-nouveau; Ser—serine; Thr—threonine.
Figure 2. Differences in action of group I and II GALNTs. Yellow square = GalNac; blue square = GlcNac; yellow circle = galactose. Abbreviations: GALNT—polypeptide N-acetylgalactosaminyltransferase; Tn—Thomsen-nouveau; Ser—serine; Thr—threonine.
Biomedicines 14 00005 g002
Figure 3. The synthesis and movement of the O-glycans along the compartments of the Golgi apparatus to be expressed on the cell surface. The key to glycans is as follows: yellow square = GalNac; blue square = GlcNac; yellow circle = galactose; purple diamond = sialic acid; red triangle = fucose; blue triangle arrow to shown GALNT acting on the reaction step. The dark bold arrows indicate movement of glycans from Golgi compartment to another. The small arrows in the cis-Golgi show the pathway of UDP-GalNAc to form Tn epitope. Abbreviations: UDP—uridine diphosphate; GALNT—polypeptide N-acetylgalactosaminlytransferase; TGN—trans-Golgi network; C1GALT1—core 1 synthase, glycoprotein-N-acetylgalactosamine 3-beta-galactosyltransferase 1; COSMC—Core-1 β1-3galactosyltransferase-specific chaperone 1; ST3Gal—ST3 beta-galactoside alpha-2,3-sialyltransferase 1; FUT—Fucosyltransferase; GlcNAc—N-Acetylglucosamine.
Figure 3. The synthesis and movement of the O-glycans along the compartments of the Golgi apparatus to be expressed on the cell surface. The key to glycans is as follows: yellow square = GalNac; blue square = GlcNac; yellow circle = galactose; purple diamond = sialic acid; red triangle = fucose; blue triangle arrow to shown GALNT acting on the reaction step. The dark bold arrows indicate movement of glycans from Golgi compartment to another. The small arrows in the cis-Golgi show the pathway of UDP-GalNAc to form Tn epitope. Abbreviations: UDP—uridine diphosphate; GALNT—polypeptide N-acetylgalactosaminlytransferase; TGN—trans-Golgi network; C1GALT1—core 1 synthase, glycoprotein-N-acetylgalactosamine 3-beta-galactosyltransferase 1; COSMC—Core-1 β1-3galactosyltransferase-specific chaperone 1; ST3Gal—ST3 beta-galactoside alpha-2,3-sialyltransferase 1; FUT—Fucosyltransferase; GlcNAc—N-Acetylglucosamine.
Biomedicines 14 00005 g003
Figure 4. (A) Gene expression profile of the various GALNTs in colorectal cancer, showing higher GALNT3/4 and 6 expressions in cancer cells compared to normal cells. (B) Kaplan–Meier survival curve based on GALNT expressions in colorectal cancer showing no significant difference in survival based on GALNT expression. Dotted lines indicate 95% confidence interval. The figures are extrapolated from GEPIA2 (http://gepia2021.cancer-pku.cn/ (accessed on 13 July 2025)) using the TCGA/GTEx expression data. * = p < 0.05.
Figure 4. (A) Gene expression profile of the various GALNTs in colorectal cancer, showing higher GALNT3/4 and 6 expressions in cancer cells compared to normal cells. (B) Kaplan–Meier survival curve based on GALNT expressions in colorectal cancer showing no significant difference in survival based on GALNT expression. Dotted lines indicate 95% confidence interval. The figures are extrapolated from GEPIA2 (http://gepia2021.cancer-pku.cn/ (accessed on 13 July 2025)) using the TCGA/GTEx expression data. * = p < 0.05.
Biomedicines 14 00005 g004
Figure 5. (A) Gene expression profile of the various GALNTs in breast cancer showing higher GALNT3 and 6 expressions in cancer cells compared to normal cells. (B) Kaplan–Meier survival curve based on GALNT expressions in breast cancer showing a slightly improved survival in cancers with low GALNT6 expression. Dotted lines indicate 95% confidence interval. The figures are extrapolated from GEPIA2 (http://gepia2021.cancer-pku.cn/ (accessed on 13 July 2025)) using the TCGA/GTEx expression data. * = p < 0.05.
Figure 5. (A) Gene expression profile of the various GALNTs in breast cancer showing higher GALNT3 and 6 expressions in cancer cells compared to normal cells. (B) Kaplan–Meier survival curve based on GALNT expressions in breast cancer showing a slightly improved survival in cancers with low GALNT6 expression. Dotted lines indicate 95% confidence interval. The figures are extrapolated from GEPIA2 (http://gepia2021.cancer-pku.cn/ (accessed on 13 July 2025)) using the TCGA/GTEx expression data. * = p < 0.05.
Biomedicines 14 00005 g005aBiomedicines 14 00005 g005b
Table 1. Expression of GALNTs in various human tissues (adapted from [4]).
Table 1. Expression of GALNTs in various human tissues (adapted from [4]).
GALNTHuman Tissues
Expressed in
Family SubclassificationChromosome
Locus
T1Widely distributedIa18q12.1
T2Widely distributedIb1q41–q42
T3Bone, spermIc2q24–q31
T4LungIIa12q21.33
T5StomachId2q24.1
T6Stomach, colonIc12q13
T7Sublingual glandIIb4q34.1
T8Colon, testesIe12p13.3
T9BrainIe12q24.33
T10KidneyIIb5q33.2
T11ProstateIf7q36.1
T12ColonIIa9q22.33
T13BrainIa2q24.1
T14KidneyIb2p23.1
T15PlacentaId3p25.1
T16Brain, heartIb14q24.1
T17Brain, ovaryIIb4q34.1
T18Widely distributedIe11p15.3
T19BrainIe7q11.23
T20TestesIf7q36.1
Table 2. Some examples of the various GALNTs and their functional role in human cancers.
Table 2. Some examples of the various GALNTs and their functional role in human cancers.
GALNT IsoformTissue InvolvedSubstrate or Pathway InvolvedFunctional RoleReferences
GALNT1 and -T2Hepatocellular cancerEGFRModifies the activity of EGFR; dysregulation leads to malignant behavior[21,22,23]
GALNT3 and -T14Ovarian cancerMUC1; EGFR/mTORIncreased activity leads to increased MUC1 expression, resulting in increased cell proliferation, cell migration, and invasion[24,25,26]
GALNT4Pancreatic cancerMUC1Induces O-GalNAc glycosylation of MUC1 to enhance MUC1 protein stability[27]
GALNT6Breast cancerMUC1 & MUC4, α2-macroglobulinDrives clustered O-GalNAc addition; activates PI3K/Akt signaling to increase invasion and tumor growth[28,29]
GALNT7Prostate cancerFOXO1, androgen regulatorUpregulation affects immune activity and cell signaling[30,31]
Table 3. The dual role of GALNTs as oncogenic and tumor suppressors in various cancers.
Table 3. The dual role of GALNTs as oncogenic and tumor suppressors in various cancers.
GALNT IsoformTumor RoleMechanismPrimary
Cancer
Source
GALNT1OncogenicGlycosylates Notch to induce
EMT and metastasis
Breast[56]
GALNT2OncogenicActivates Notch/Hes1-PI3K/
Akt axis
Lung[77]
GALNT2SuppressiveMaintains glycan profile;
suppresses invasion
Gastric[58]
GALNT3SuppressiveSuppresses Wnt via TNFR1
and c-MET glycosylation
Lung[89]
GALNT6OncogenicAlters mucin-type
O-glycosylation, causing
Wnt activation
Colorectal[87]
GALNT10SuppressiveCell surface glycoproteome
remodeling, preventing EMT
and migration
Ovarian[90]
GALNT14OncogenicEnhances β-catenin
glycosylation, causing
Wnt activation
Lung[78]
Table 4. Prognostic impact of the various GALNTs on common human cancers.
Table 4. Prognostic impact of the various GALNTs on common human cancers.
Primary
Cancer
GALNT
Expression
Pattern
Prognostic ImpactReferences
Breast (BRCA)High GALNT6, -T13, -T14 in
advanced
disease.
  • GALNT6: ↑ associated with advanced nodal stage and poor prognosis.
  • GALNT13: ↑ associated with tumor progression and metastasis.
  • GALNT14: ↑ linked to lung-specific metastasis and reduced distant metastasis-free survival.
[28,88,93,94]
ColonLow GALNT3, -T4, -T6, -T12 in advanced
disease.
  • GALNT3: ↓ associated with lower survival rate; poor expression in poorly differentiated cancers.
  • GALNT4: ↓ associated with increased metastatic potential and lower survival rate.
  • GALNT6: ↓ associated with shorter survival; lower expression in BRAF-mutated and KRAS wild-type cancers.
  • GALNT12: ↓ associated with shorter survival and metastasis.
[95,96,97,98]
GastricGALNT3, -T5, -T6, -T10 have varying impacts on outcomes
  • GALNT3: ↑ associated with good prognosis.
  • GALNT5: ↓ associated with poorly differentiated gastric adenocarcinomas, advanced stage, lymph node metastasis, and poor prognosis.
  • GALNT6: ↑ associated with correlation with advanced clinical stage, distant metastasis, and poor survival.
  • GALNT10: ↑ associated with shorter survival, lymph node, and distant metastasis.
[99,100,101,102]
PancreaticGALNT3, -T5, -T6 have varying impacts on outcomes
  • GALNT3: ↓ contribute to dedifferentiation and increased aggressiveness during tumor progression.
  • GALNT5: ↑ associated with poor survival.
  • GALNT6: ↑ associated with favorable tumor biology and improved survival.
[103,104,105]
ThyroidHigh GALNT3 expression in poor prognosis.
  • GALNT3: ↑ contribute to dedifferentiation and increased aggressiveness during tumor progression.
[106]
↑ represents increase in GALNT and ↓ represent decrease in GALNT.
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

Iyer, S.C.; Srinivasan, D.K.; Parameswaran, R. GalNAc-Transferases in Cancer. Biomedicines 2026, 14, 5. https://doi.org/10.3390/biomedicines14010005

AMA Style

Iyer SC, Srinivasan DK, Parameswaran R. GalNAc-Transferases in Cancer. Biomedicines. 2026; 14(1):5. https://doi.org/10.3390/biomedicines14010005

Chicago/Turabian Style

Iyer, Shruthi C., Dinesh Kumar Srinivasan, and Rajeev Parameswaran. 2026. "GalNAc-Transferases in Cancer" Biomedicines 14, no. 1: 5. https://doi.org/10.3390/biomedicines14010005

APA Style

Iyer, S. C., Srinivasan, D. K., & Parameswaran, R. (2026). GalNAc-Transferases in Cancer. Biomedicines, 14(1), 5. https://doi.org/10.3390/biomedicines14010005

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