Next Article in Journal
Cellular and Immunological Analysis of 2D2/Th Hybrid Mice Prone to Experimental Autoimmune Encephalomyelitis in Comparison with 2D2 and Th Lines
Previous Article in Journal
Metabolomic and Transcriptomic Analyses Reveal the Potential Mechanisms of Dynamic Ovarian Development in Goats during Sexual Maturation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

O-GlcNAcylation: Crosstalk between Hemostasis, Inflammation, and Cancer

by
Itzel Patricia Vásquez Martínez
1,†,
Eduardo Pérez-Campos
2,†,
Laura Pérez-Campos Mayoral
1,
Holanda Isabel Cruz Luis
1,
María del Socorro Pina Canseco
1,
Edgar Zenteno
3,
Irma Leticia Bazán Salinas
1,
Margarito Martínez Cruz
2,
Eduardo Pérez-Campos Mayoral
1 and
María Teresa Hernández-Huerta
4,*
1
UNAM-UABJO Faculty of Medicine Research Center, Faculty of Medicine and Surgery, Autonomous University “Benito Juarez” of Oaxaca, Oaxaca 68020, Mexico
2
National Institute of Technology of Mexico, Technological Institute of Oaxaca, Oaxaca 68033, Mexico
3
Department of Biochemistry, Faculty of Medicine, National Autonomous University of Mexico, Mexico City 04510, Mexico
4
National Council of Humanities, Sciences and Technologies (CONAHCYT), Faculty of Medicine and Surgery, Autonomous University “Benito Juarez” of Oaxaca, Oaxaca 68120, Mexico
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2024, 25(18), 9896; https://doi.org/10.3390/ijms25189896
Submission received: 2 August 2024 / Revised: 3 September 2024 / Accepted: 10 September 2024 / Published: 13 September 2024
(This article belongs to the Topic Cancer Cell Metabolism (2nd Edition))

Abstract

:
O-linked β-N-acetylglucosamine (O-GlcNAc, O-GlcNAcylation) is a post-translational modification of serine/threonine residues of proteins. Alterations in O-GlcNAcylation have been implicated in several types of cancer, regulation of tumor progression, inflammation, and thrombosis through its interaction with signaling pathways. We aim to explore the relationship between O-GlcNAcylation and hemostasis, inflammation, and cancer, which could serve as potential prognostic tools or clinical predictions for cancer patients’ healthcare and as an approach to combat cancer. We found that cancer is characterized by high glucose demand and consumption, a chronic inflammatory state, a state of hypercoagulability, and platelet hyperaggregability that favors thrombosis; the latter is a major cause of death in these patients. Furthermore, we review transcription factors and pathways associated with O-GlcNAcylation, thrombosis, inflammation, and cancer, such as the PI3K/Akt/c-Myc pathway, the nuclear factor kappa B pathway, and the PI3K/AKT/mTOR pathway. We also review infectious agents associated with cancer and chronic inflammation and potential inhibitors of cancer cell development. We conclude that it is necessary to approach both the diagnosis and treatment of cancer as a network in which multiple signaling pathways are integrated, and to search for a combination of potential drugs that regulate this signaling network.

1. Introduction

Post-translational protein modifications (PTM) allow cells to respond to cellular or environmental signals. O-linked β-d-N-acetylglucosamine (O-GlcNAc, O-GlcNAcylation, or O-acetylglucosaminylation) is a highly dynamic and reversible PTM process [1,2], through the addition of N-acetylglucosamine (GlcNAc) to the hydroxyl group of serine or threonine residues into nuclear, cytoplasmic, and mitochondrial proteins [3]. The enzymes responsible for O-GlcNAcylation are O-GlcNAc transferase (OGT) and O-GlcNAcase (OGA), which are responsible for adding and removing GlcNAc to proteins, respectively. O-GlcNAcylation is the endpoint of the hexosamine biosynthetic pathway (HBP) and terminates with uridine diphosphate N-acetylglucosamine (UDP-GlcNAc), which is the substrate of OGT [4,5]. UDP-GlcNAc biosynthesis is regulated by almost all metabolic pathways in the cell [2] (Figure 1).
O-GlcNAcylation participates in many cellular processes [6] such as signal transduction [7], transcription [8], cell cycle control [9,10], epigenetic control of gene expressions in response to stress [11], and nutrient imbalance [12]. It is also present in almost all cellular compartments; and in humans, it is more abundant in the pancreas, liver, brain, skeletal muscle, adipose tissue, and other organs and tissues [13]. For this reason, O-GlcNAcylation alterations have been related to metabolic diseases [14,15], cardiovascular diseases [16], neurodegenerative diseases [17,18], autoimmune diseases, and cancer [19] (Figure 2).
Furthermore, O-GlcNAcylation alters signaling pathways, particularly in cancer, by promoting tumor growth and immune escape. Among the signaling pathways that are modified are the HBP, which is associated with excess glucose and cancer [20], and the O-GlcNAc/c-Myc pathway, which is involved in the regulation of megakaryopoiesis and thrombopoiesis and, therefore, in hemostasis [21]. Platelets can modulate tumor-associated inflammation and stimulate metastasis and venous thrombosis. Tumor cell-induced platelet aggregation has been demonstrated in several cell lines. In addition, other important molecules in the signaling of thrombo-inflammation and cancer are thrombin, tumor-expressed ADP, podoplanin (PDPN), tissue factor (TF), and ADAM-13, among many others [22]. The most studied pathways associated with O-GlcNAcylation are thrombosis, inflammation and cancer in the PI3K/Akt/c-Myc pathway [23], the nuclear factor-kappa B (NF-κB) signaling pathway [24,25], and PI3K/AKT/mTOR pathway [26,27] (Figure 3, modified of https://www.kegg.jp/pathway/map=map05200&keyword=cancer (accessed on 2 September 2024)). In recent years, the role of O-GlcNacylation has been demonstrated in cancer cell proliferation, angiogenesis, and metastasis, as well as in the state of cancer-associated inflammation, and drug resistance has been demonstrated. In this review, we aim to explore the relationship between O-GlcNAcylation and hemostasis, inflammation, and cancer, which could serve as potential prognostic tools or clinical predictions for cancer patients’ healthcare and as an approach to combat cancer.

2. Cancer, Hemostasis, and O-GlcNAcylation

Cancer-associated thrombosis is a major cause of morbidity and mortality for cancer patients [28]. Likewise, tumors can induce platelet activation, aggregation, release of platelet-derived macrovesicles into circulation, and promote thrombocytosis (Figure 2). There are factors associated with cancer and thrombosis that have been recently reviewed, such as leukocytes (eosinophils, monocytes, and neutrophils), TF, thrombocytosis, and its leukocyte-related indices [29], PDPN, plasminogen activator inhibitor-1, the intrinsic coagulation pathway, von Willebrand factor [30], and miRNAs (e.g., miR-126) [31]. In this way, increased platelet count and platelet-associated clinical laboratory indexes could be used as predictive biomarkers of tumors [29]. In cancer cells, O-GlcNAc and its enzyme OGT are highly elevated, thus coupling the alteration in nutrient status to signaling activities, contributing to reprogramming cellular metabolism and cancer progression [32]. Blood O-GlcNAc levels have been described as increasing in response to inflammation and stress tissue [33].
Cancer cells reprogram their energy metabolism and signaling networks to promote growth, survival, proliferation, and long-term maintenance. Cancer progression is characterized by an altered metabolic state and even different metabolic phenotypes in the cells of the same cancer. Under normal conditions, cells process glucose under aerobic conditions and promote glycolysis under anaerobic conditions; here, the final product of glycolysis depends on the level of available oxygen (Pasteur effect). However, even in high oxygen, cancer cells prefer glycolysis instead of oxidative phosphorylation, called the Warburg effect [26]. This metabolic reprogramming allows cancer cells to absorb a large part of the nutrients from their environment (glucose and glutamine) producing low amounts of adenosine triphosphate (ATP) as a by-product and secreting large amounts of carbon and nitrogen in the form of lactate and ammonia into the extracellular environment so that they can be used by tumor cells as fuel [34,35]. In addition, cancer and proliferating cells adjust their energy metabolism by increasing glucose concentration when there is a decrease in oxidative phosphorylation via the Crabtree effect [36]. In summary, cancer cells are characterized by a high demand and consumption of glucose.
Excess glucose in cancer cells primarily enters glycolysis and increases flow to branching glucose pathways such as HBP [37]. HBP integrates pathways of glucose, amino acid, fatty acid, and nucleotide metabolism to promote the synthesis of UDP-GlcNAc, the end product of HBP and substrate for O-GlcNAcylation [38]. Activation of HBP promotes the proliferation of lung cancer cells, and the inhibition of enzymes related to HBP such as glutamine fructose-6-phosphate amidotransferase 1 (GFAT1) reduces programmed death ligand 1 (PD-L1) levels and, therefore, the progression of lung cancer. O-GlcNAc suppresses the degradation of PD-L1 in cancer cells, which is responsible for immune evasion in cancer [39]; high expression of PD-L1 in tumor cells has been correlated with poor prognosis in cancer patients, such as non-small cell lung cancer (NSCLC) [40]. While studies show that PD-L1 does not appear to be associated with the incidence of venous thromboembolism in NSCLC [41], others have found that the presence of PD-L1 in NSCLC may be associated with an increased risk of thromboembolism [42].
Oncogenes, tumor suppressors, and proteins involved in tumor biology are also regulated by O-GlcNAc [43] such as phosphoglycerate kinase 1 (PGK1), enolase 1 (ENO1), glucose-6-phosphate dehydrogenase (G6PD) c-Myc, p53, Ras, AMPK, and NF-κB, among others. PGK1 regulates the expression of urokinase-type plasminogen activator receptor (uPAR) mRNA associated with metastasis in gastric cancer and breast cancer [44,45]. O-GlcNAcylation activates PGK1 activity to enhance lactate production and simultaneously induces PGK1 translocation into mitochondria to inhibit the pyruvate dehydrogenase (PDH) complex and, thus, reduce oxidative phosphorylation [46]. Excessive lactate production by tumor and stromal cells is associated with increased aggressiveness, due to extracellular acidification, which also induces invasion and metastasis, inhibition of the antitumor immune response, and resistance to therapy [47]. Increased blood lactate levels impair the coagulation system [48].
ENO1 acts as a plasminogen receptor that is converted into plasmin and induces fibrinolysis; in cancer cells, it enhances their ability to invade through the stroma, mainly by activating collagenases and degrading fibrin [49]. ENO1 is associated with a better prognosis only in early-stage breast cancer, which may be related to the different effects of ENO1 on immune infiltration [50]. ENO1 promotes glycolysis, cell proliferation, and migration through activation of the phosphatidylinositol 3-kinase (PI3K)/Akt and adenosine monophosphate-activated protein kinase (AMPK)/mTOR pathways [51,52,53,54], and c-Myc pathways [55].
G6PD (EC 1.1.1.49) is responsible of the production of nicotinamide adenine dinucleotide phosphate (NADPH) for the prevention of cellular damage caused by reactive oxygen species. In response to increased oxidative stress, a reduced ability to induce the innate immune response has been found in G6PD-deficient cells [56]. O-GlcNAcylation of G6PD facilitates lung cancer while phosphorylation of G6PD by polo-like kinase 1 (Plk1) affects cell cycle progression and cell proliferation of multiple cancers [57]. Furthermore, primary G6PD deficiency can cause non-immune hemolytic anemia, whereas severe deficiency presents with intravascular hemolysis and, therefore, renal failure [58]. Other studies suggest that G6PD deficiency is associated with increased release of the proinflammatory cytokine IL-8, and decreased release of the anti-inflammatory cytokine IL-10 [59]. This is important because IL-8-induced O-GlcNAc modification through glucose transporter 3 (GLUT3) and glucosamine fructose-6-phosphate aminotransferase (GFAT) has been reported to regulate cancer stem cell-like properties in colon and lung cancer cells [60].
c-Myc is a helix–loop–helix leucine transcription factor involved in many cellular processes, including cell potency, apoptosis, and differentiation [61]. It regulates genes involved in glycolysis, purine/pyrimidine metabolism, and lipids [62], as well as genes involved in glutamine metabolism, mitochondria biosynthesis, cell cycle control, and HBP genes. In cancer cells, c-Myc is often elevated, which promotes energy production and biomolecule synthesis [63] by regulating genes for anabolic enzymes such as aspartate transcarbamylase and dihydroorotase (CAD), serine hydroxymethyl transferase (SHMT), fatty acid synthase (FAS), and ornithine decarboxylase (ODC). Elevated c-Myc expression in cancer cells has been reported to increase the need for glutamine [64]. Increased O-GlcNAcylation is observed in cancers, such as breast, pancreatic, liver, lung, colon [65], and leukemia [6]. O-GlcNAcylation participates in initiation, progression, and metastasis through various pathways, such as O-GlcNAcylation of G6PD in the pentose phosphate pathway (PPP) [66], O-GlcNAcylation of NF-κB and p53 [10,67] and c-Myc. Increased O-GlcNAcylation of c-Myc promotes the proliferation of pre-B cells as well as some B-cell cancers, such as chronic lymphocytic leukemia [6] and B acute lymphocytic leukemia [68]. Furthermore, the glycolytic pathway is modulated by the PI3K/Akt/c-Myc pathway in pre-B acute lymphoblastic leukemia (pre-ALL) [65]. Interconnections of c-Myc through the O-GlcNAc/c-Myc axis confer on it a regulatory function of O-GlcNAcylation in the processes of megakaryopoiesis and thrombopoiesis [69]. In the case of colon cells, their proliferation is related to platelets, which induce the positive regulation of the c-Myc oncoprotein. This upregulation of platelets can be reduced with aspirin and is correlated with the negative regulation of COX-2 and c-Myc expression [70].
The PI3K/Akt/mTOR signaling pathway is a key mechanism involved in the growth and control of glucose metabolism in cells [71]. The PI3K/Akt pathway regulates the uptake and utilization of glucose [72], its activation increases the expression of glucose transporters on the cell surface, while the activation of hexokinase (HK) and the phosphofructokinase-2-dependent allosteric activation of phosphofructokinase-1 (PFK1) phosphorylate glucose to promote glycolysis [73]. Furthermore, activation of the PI3K/Akt/mTOR pathway enhances carbohydrate, lipid, and protein biosynthesis. PI3K and Akt promote carbon flux from glucose into mitochondrial-dependent biosynthetic pathways such as fatty acid, cholesterol, and isoprenoid synthesis [74], while mTOR regulates cell growth from various amino acid precursors from transamination of mitochondrial intermediates (Figure 4).
The PI3K/Akt/mTOR signaling pathway has been described as one of the most activated pathways in various types of cancer [75], as a result of mutations in the phosphatidylinositol-4,5-bisphosphate 3-kinase catalytic subunit alpha (PIK3CA) or by the loss of phosphatidylinositol-3,4,5-trisphosphate 3-phosphatase (PTEN). Activation of this pathway in human cancers is caused by somatic alterations at specific sites in the pathway and by activation by receptor tyrosine kinases (RTKs) [76]. Somatic mutations in PIK3CA are common in a variety of tumor types including breast, colon, endometrial, and glioblastoma [77,78].
In prostate cancer, common pathways between HBP and PI3K/AKT/mTOR are observed. These pathways promote the expression and levels of androgen receptor, and this increase is mediated, in part, by the overexpression of OGT and stability of the pro-oncogene Myc, which benefits proliferation and metastasis [79]. Patients with prostate cancer have an increased risk of developing thromboembolic disease, which is increased by hormonal treatment [80]. On the other hand, the Akt kinase of the PI3K/AKT/mTOR pathway is associated with platelet activation [81], which, together with mutations in PI3K enzymes, may explain both metastasis and thrombosis in prostate cancer [11].
In the Src family, some members have been linked to O-GlcNAcylation and cancer, such as p53 [82] and p53/56lyn [83]. Mutations in p53 are the most common genetic change in cancer [84]. p53 is a transcription factor and tumor suppressor known for its involvement in various signaling pathways such as cell cycle arrest, cell apoptosis, autophagy, metabolism, response to DNA damage, and apoptosis; it plays an important role in the regulation of glycolysis and oxidative phosphorylation [85,86]. p53 can also arrest the cell cycle to activate DNA repair pathways; if the damage is extensive and cannot be repaired, it will induce the transcription of proteins involved in cell death by apoptosis [87]. In addition, p53 decreases the glycolytic rate through various mechanisms, represses the transcription of glucose transporters (GLUT) [88], and the translocation of GLUT1 to the plasma membrane [89] to suppress glucose uptake. It downregulates the levels of enzymes such as HK [90] and protein phosphoglycerate mutase 1 (PGAM1) that inhibit glycolysis [91], Figure 5. p53 inhibits the PPP to suppress glucose consumption through its binding to G6PD [92], and it also negatively regulates PI3K/Akt signaling through PTEN transcriptional induction [93]. However, in cancer cells, p53 is suppressed, resulting in loss of control of its functions, promoting glycolysis [94].
p53/56lyn is involved in polymorphonuclear leukocyte (PMN) [95], platelets signaling [96], and in TGF-beta 1-induced apoptosis in M-07e leukemic cells [97]. TNF-alpha-mediated stimulation of human PMN adherent to fibrinogen involves p53/56lyn [92] and could be related to neutrophil defects and inflammation [98]. Furthermore, p53/56lyn, along with other Src family kinases, is involved in signaling through the collagen receptor GPVI on the platelet [93].
O-GlcNAcylation has been considered as a “nutritional sensor”, since it exerts effects on the regulation of cell signaling, and transcription in cancer and inflammatory cells in response to nutrients and stress in the tumor microenvironment (TME) [30]. The signaling pathways transform environmental signals into intracellular events, such as immune cell activation and inflammation [99]. In addition, it interacts with immune evasion mechanisms and signaling pathways involved in thromboregulation and thrombosis in cancer-associated [100], e.g., some studies indicate that clotting protease thrombin, FVIIa, and FXa contribute to cancer immune evasion via unique mechanisms such as TF/FVIIa/PAR2 signaling [101].

3. O-GlcNAcylation and Inflammatory State of Cancer

The relationship between inflammation and cancer was first discovered in 1893 by Rudolf Virchow, who observed leukocytes infiltrating tumor tissue, suggesting that cancer may arise from chronic inflammation [102]. Inflammation is defined as the body’s response to tissue damage caused by physical injury, ischemic injury, infection, exposure to toxins, or other types of traumas. The inflammatory response triggers cellular changes and immune responses that result in the repair of damaged tissue and cell proliferation at the site of the injured tissue [103].
O-GlcNAcylation promotes precancerous inflammation and acts as a key orchestrator at the intersection of intrinsic and extrinsic inflammation through O-GlcNAcylation of key transcription factors and functional proteins in the activation of inflammatory cells to trigger cancer-associated inflammation in the TME [29].
Cancer-extrinsic inflammation is characterized by long-term chronic inflammatory conditions that predispose a person to cancer [104]. Risk factors related to this type of inflammation include bacterial and viral infections, obesity, autoimmune diseases, smoking, and excessive alcohol consumption. Approximately 20% of all cancers are related to chronic infection, chronic inflammation, or autoimmunity, in the same tissue or organ (Table 1) by intervening in various signal pathways with promotion (proliferation, metastasis, stemness, or drug resistance) or inhibitory effects (apoptosis) [105].
Cancer-intrinsic inflammation develops in most cancers, where cancer cells recruit immune cells and secrete inflammatory mediators to remodel the TME and, thus, promote cancer progression. This type of inflammation is caused by genetic and/or epigenetic mutations (oncogenes) [100]. Therefore, both extrinsic and intrinsic inflammation trigger the activation of transcription factors and signaling pathways to regulate the inflammatory response through soluble mediators (cytokines, chemokines) and other cellular components [35].
During cancer inflammation, several signal transduction pathways are deregulated to stimulate malignant transformation. Transcription factors associated with O-GlcNacylation and cancer (Table 2) such as NF-κB, the family of activators of transcription (STAT) STAT1/STAT3, and hypoxia-inducible factor (HIF) modulate the inflammatory response through inflammatory mediators (cytokines, chemokines) and immune cell infiltration promoting tumorigenesis [35].
NF-κB is a family of transcription factors including RelA/p65, RelB, c-Rel/Rel, p105/p50, and p100/p52 [126], and plays an important role in inflammatory, immune, and anti-apoptotic responses and hemostasis [127,128,129]. NF-κB is implicated in atherosclerosis and its pathological complication in atherothrombotic diseases due to its transcriptional role in maintaining pro-survival and pro-inflammatory states in vascular and blood cells [125]; this has a role in platelet survival, priming, activation, and aggregation. O-GlcNAcylation of NF-κB is involved in hyperglycemia-induced NF-κB activation and is required for lymphocyte activation [130]. NF-κB O-GlcNAcylation by increased uptake of glucose and glutamine by cancer cells regulates cancer cell proliferation, survival, and metastasis and acts as a link between inflammation and cancer [10,131].
Together with NF-κB, the Janus activator of transcription and signal transduction/kinase (JAK–STAT) signaling pathway is involved in the regulation of cytokine-dependent inflammation and immunity in carcinogenesis [132]. The binding of various ligands to cell surface receptors causes the receptor to activate the JAK family, which phosphorylate tyrosine residues on the receptor and recruits STAT family proteins. The STAT family comprises seven members: STAT1, STAT2, STAT3, STAT4, STAT5A, STAT5B, and STAT6 [133]. Some studies indicate that O-GlcNAcylation may regulate cancer-related inflammation independently or by modulating the production of HIF proteins [134].
Table 2. Transcription factors associated with O-GlcNacylation and cancer.
Table 2. Transcription factors associated with O-GlcNacylation and cancer.
Transcription FactorsO-GlcNAcylation SiteEffectsType of CancerReferences
NF-κB p65Thr322
Thr352
↑ IL-6, TNF-α. Better invasion and metastasis capacity.
↑ CXCR4 expression promotes metastasis.
Cervix[135,136]
NF-κB c-RelSer350↑ IL2, IFN-γ, GM-CSF.
↑ c-Rel
Pancreatic[137,138]
TAB1Ser395Regulator ↑NF-κB
↑ IL-6 and TNF-α.
↑ TAB1
Non-small cell
lung carcinoma
[139,140]
TAB3Ser408↑ Metastasis, NF-κBBreast[141]
STAT3Thr717↑ Migration and invasion by ↑ IL-6/STAT3 signaling.Lung[142,143]
STAT5AThr92↑ Oncogenic transcription, myeloid transformation. Leukemic[144]
RIPK3Thr467↓ Inflammation, inflammation-associated necroptosis.
↑ mRNA RIPk3
Hepatocellular, cervix[145,146,147]
Abbreviations: NF-κB: nuclear factor κB; STAT: activator of transcription; TAB: activated binding adapter kinase 1 (TAK1) proteins; CXCR4: C-X-C motif chemokine receptor 4; TNF-α: tumor necrosis factor α; IFN-γ: interferon γ; GM-CSF: granulocyte–macrophage colony-stimulating factor; IL: interleukine; mRNA: messenger ribonucleic acid; ↑ increase; ↓ decreases.

4. O-GlcNAcylation, Inflammation, Hypercoagulation, and Cancer

In 1856, Rudolf Virchow postulated a triad of conditions that lead to thrombosis: endothelial injury, circulatory stasis, and abnormalities in the components of blood coagulation (hypercoagulable state, hypercoagulation) [148]. A hypercoagulable state occurs when the activation of the hemostatic mechanism of the plasma exceeds its physiological anticoagulant capacity, resulting in a predominance of prothrombotic activities [149].
The cancer-induced hypercoagulable state contributes to thrombosis, which is one of the main causes of death in cancer [150]. In cancer patients, hypercoagulability has been associated with an increased risk of venous thromboembolism, as well as with proliferation, tumor progression, and metastatic spread [151,152].
Different pathophysiological pathways have been identified regarding the interaction between cancer and the different components of the hemostatic system. Overexpression of procoagulant molecules such as TF has been found in various types of cancer. Likewise, plasminogen activator inhibitor 1 (PAI-1) is overexpressed and, as the main inhibitor of fibrin degradation, contributes to the procoagulant state of the TME. The expression of substances with effects on hemostasis has been reported in different types of cancer, some of which are mentioned in Table 3.
O-GlcNAc regulates the expression of PAI-1, fibronectin, and transforming growth factor-β by high glucose [153]. Another mechanism associated with facilitating immune evasion in breast cancer is that mediated by TF-FVIIa through the induction of PD-L1 [154]. O-GlcNAcylation also facilitates immune evasion through inhibition of lysosomal degradation of PD-L1 [39].
Table 3. Hemostatic abnormalities in cancer.
Table 3. Hemostatic abnormalities in cancer.
Type of CancerType of PatientsAbnormalities in HemostasisModifications O-GlcNAcReferences
BreastPatients with metastatic breast cancer↑ D-dimer, ↑ fibrinogen, ↑ prothrombin fragment 1 + 2. There is a correlation between hyperactive/activated platelets and tumor progression and thrombus formation.O-GlcNAc modification regulates MTA1 transcriptional activity during breast cancer cell genotoxic adaptation.[125,148,155,156,157,158]
ProstatePatients with prostate cancer tissues and enhances malignancy of prostate cancer cells↑ VWF
A higher incidence for venous thromboembolisms in prostate cancer exists. Platelets synthesize testosterone in patients with prostate cancer, which may explain relapses in castration-resistant prostate cancer. Insulin resistance in prostate cancer patients predisposes them to acute ischemic heart disease through dermcidin, a protein that can induce abnormal platelet aggregation.
Inhibiting the formation of the E-cadherin/catenin/cytoskeleton complex may underly the O-GlcNAc-induced prostate cancer progression.[146,159,160,161,162,163]
OvaryPatients with ovarian cancer without treatment↑TF
Thrombophilia due to a factor V and prothrombin mutation.
↑ D-dimer, hs-CRP, and IL-6 compared to the control group. Tissue factor pathway inhibitor (TFPI)-2 has been implicated in the suppression of epithelial ovarian cancer, particularly clear cell carcinoma (CCC). It negatively regulates plasmin. TFPI-2 is secreted by CCC, platelets, and vascular endothelial cells.
O-GlcNAcylation increases the motility of ovarian cancer cells via the RhoA/ROCK/MLC signaling pathway.[164,165,166,167,168]
CervixPatients with adenosquamous carcinoma, adenocarcinoma, or squamous cell carcinoma↑ TF. Pretreatment thrombocytosis predicts pelvic lymph node metastasis and larger tumor size. It is also associated with markers of subclinical inflammation such as the neutrophil to lymphocyte ratio (NLR) and the platelet to lymphocyte ratio (PLR).Increased O-GlcNAcylation promotes IGF-1 receptor/phosphatidyl inositol-3 kinase/Akt pathway.
HPV E6 upregulates OGT, increases O-GlcNAc, stabilizes c-MYC via O-GlcNAc, and enhances HPV oncogene activities.
[169,170,171,172,173]
Colorectal1. Patients with worse prognosis.
2. Patients with colorectal cancer diagnosed without treatment.
3. Recently diagnosed patients
1. ↑ VWF in patients with worse prognosis.
↑ D-dimer and ↑ fibrinogen.
2. ↑ P-Selectin, CRP, IL-6 compared to the control group, ↑ in the group with metastasis compared to the group without metastasis.
3. ↑ IL-6, IL-1b, and TNF-α compared to the control group. ↑ IL-6, CRP, and fibrinogen are more advanced stages. Platelet infiltration in tumors, particularly in CRC, is associated with a poor prognosis, as observed in the analysis of postoperative survival.
O-GlcNAcylation, which is negatively regulated by microRNA-101, likely promotes CRC metastasis by enhancing EZH2 protein stability and function.[174,175,176,177,178]
PancreasPatients with pancreatic adenocarcinomas are at increased risk for hypercoagulability↑ TF in pancreatic neoplasms
↑ TF ↑ VEGF in carcinomas is related to a higher rate of venous thromboembolism (26.3%)
Preoperative hypercoagulability can be identified with rotational thromboelastometry and is associated with lymphovascular/perineural invasion and advanced-staged disease in cancer.
Glycosylation of Sox2 by OGT can affect its transcriptional activity and thereby regulate self-renewal in cancer.[179,180,181]
GastricPatients in early, advanced stages or metastasis↑ IL-6 in patients compared to the control group, ↑ in early-stage patients, ↑ number of platelets in the locally advanced and metastatic groups
↑ CRP in cancer patients, ↑ platelets, MPV, and LPLT or PLT in the metastasis group; in addition, MPV ↑ is associated with active inflammation.
Hyper-O-GlcNAcylation significantly promotes GC cells proliferation by modulating cell cycle related proteins and ERK 1/2 signaling.
O-GlcNAcylation on RTN2 was pivotal for its oncogenic functions in gastric cancer.
[182,183,184,185]
LungPatients with early lung cancer (stage I and II) treated with surgery↓ D-dimer, MLR, NLR, and PLR are associated with significantly better overall survival.Hyper-O-GlcNAcylation induces cisplatin resistance via regulation of p53 and c-Myc in human lung carcinoma.[62,186]
HepatocellularPatients with early lung cancer (stage I and II) treated with surgery↑ D-dimer in AFP-negative patients.
↑ Prealbumin was correlated with tumor size.
HCC-related cerebral infarction patients are at high risk of hypercoagulability
OGT mediated by O-GlcNAcylation stabilized RAB10, thus accelerating HCC progression.
OGT plays an oncogenic role in NAFLD-associated HCC by regulating palmitic acid and inducing ER stress, activating oncogenic JNK/c-Jun/AP-1 and NF-κB cascades.
[187,188,189,190]
Leukemia and lymphomaConsecutive patients with acute leukemia and lymphoma↑ Fibrinogen, D-dimer, prothrombin < fragment 1 + 2, FVIII, and VWF in most patients. Fibrinogen increased if IL-6 was high, and a correlation was observed between IL-6 and D-dimer.OGT is significantly upregulated in AML tissues compared with normal tissues. The high level of OGT expression is significantly related to poor overall survival in AML. Inhibition of OGT can inhibit AML cell proliferation and promote AML cell apoptosis.[191]
Abbreviations: VWF, von Willebrand factor; TF, tissue factor; VEGF, vascular endothelial growth factor; AFP, α-Fetoprotein; MTA1, chromatin modifier metastasis-associated protein 1; hs-CRP, high-sensitivity C-reactive protein; CRP, C-reactive protein; MLR, monocyte to lymphocyte ratio; NLR, neutrophil to lymphocyte ratio; PLR, platelet to lymphocyte ratio; IL, interleukine; OGT, O-GlcNAc transferase. RhoA, Ras homolog family member A; ROCK, Rho-associated protein kinase; MLC, myosin light chain; HCC, hepatocellular carcinoma; LPLT, percentage of large platelets; IGF-1, insulin-like growth factor I; HPV E6, high-risk human papillomavirus oncogene, CRC, colorectal cancer; EZH2, histone methyltransferase enhancer of zeste homolog; Sox2, SRY-box transcription factor 2; ERK, extracellular signal-regulated kinase; ER, endoplasmic reticulum; RTN2, reticulon 2; RAB10, Ras-related protein; NAFLD, non-alcoholic fatty liver; JNK, c-Jun N-terminal kinase; AP-1, activator protein-1; NF-κB, nuclear factor-kappaB; AML, acute myeloid leukemia; ↑ increase; ↓ decreases.
In addition to coagulation or fibrinolysis abnormalities, cancer patients exhibit increased platelet responses. All of these, along with platelet alterations, are high-risk factors for thrombosis in cancer patients, as cancer cells can activate platelets and stimulate aggregation through different mechanisms [7,30], Figure 6; tumor cell-induced platelet aggregation (TCIPA) has been correlated with increased metastatic potential [192]. Tumor cells induce platelet aggregation by secreting thrombin, adenosine diphosphate (ADP) [193], and TF. Thrombin activates coagulation factors V, VIII, and XIII and protease-activated receptors (PARs) [194]. ADP activates platelets via the P2Y1 and P2Y12 receptors, causing the platelets to release more ADP from their dense granules and activate more platelets [195]. TF is the main activator of the coagulation cascade when it interacts with factor VIIa [196,197]. Platelets activated by cancer cells can regulate hematopoietic and immune cell migration to the tumor site, which may also contribute to cancer metastasis and progression by stimulating deep venous thrombosis [7].
On the other hand, platelets release proinflammatory cytokines that recruit and activate leukocytes [198] and growth factors that induce tumor growth and angiogenesis [199]; they also release P-selectin, which favors the formation of neutrophil extracellular traps (NETs) [200], while the release of Toll-like receptor 4 (TLR4) triggers NETosis in activated neutrophils [201]. NETs are structures derived from the chromatin and granular content of neutrophils [202], and they play a key role in defense by trapping and killing microorganisms [196]. Some studies suggest that they may be involved in tumor progression, metastasis, and cancer-associated thrombosis. NETs are procoagulant factors because they promote fibrin deposition, recruit red blood cells, and improve platelet activation [203]; in cancer patients, they mask cancer [151].

5. Applications of O-GlcNAcylation in Cancer

The metabolism of cancer cells is influenced by the availability of nutrients in the TME. This is detected by a mechanism that senses nutrients, such as O-GlcNAcylation, mainly by the hexosamine biosynthetic pathway [32]. The regulation of glycolysis and lipid metabolism by O-GlcNAc and OGT has been described. For example, in liver cancer, the depletion of PGK1 dramatically inhibited cancer cell glycolysis, proliferation, and tumorigenesis [204]. In colon cancer, the blocking of O-GlcNAcylation of PGK1 decreases cell proliferation and suppresses glycolysis [46]. The blocking PKM2 O-GlcNAcylation attenuated tumor growth in breast cancer and cervical cancer [205]. In breast cancer, the reduced O-GlcNAcylation leads to an increase in phosphorylated AMPK pathway and decreases cell proliferation via SREBP-1 regulation, and OGT suppression reduces levels of apoptotic marker cleaved PARP and decreases cell proliferation [206]. Other reports have shown that altered levels of O-GlcNAc and OGT are associated with the promotion of resistance to anti-cancer therapeutic agents [207]. Utilizing this knowledge could help in devising strategies to enhance the effectiveness of cancer treatments (Table 4). Furthermore, creating powerful and specific inhibitors of OGT or OGA may hold promise in treating diseases marked by abnormal O-GlcNAcylation [208].

6. Conclusions

The state of thrombosis or hypercoagulability associated with cancer is rarely considered, but it can lead to the patient´s death. Therefore, it is very important to consider that the increase in O-GlcNAc is involved in the reprogramming of metabolism in cancer cells as well as in the alteration of various pathways related to inflammation and hemostasis. For this reason, a better understanding of the O-GlcNacylation mechanisms could be considered as a potential for improved prognostic or clinical prediction tools in the health care of cancer patients, as well as helping facilitate the generation of therapeutic strategies that allow the prevention of complications. In conclusion, it is necessary to approach the diagnosis and treatment of cancer as a network that integrates multiple signaling pathways such as PI3K/Akt/c-Myc, the nuclear factor kappa B, and PI3K/AKT/mTOR, and to explore potential drug combinations that could regulate this signaling network.

Author Contributions

Conceptualization, M.T.H.-H.; investigation, I.P.V.M., E.P.-C., L.P.-C.M. and M.T.H.-H.; writing—original draft preparation I.P.V.M., E.P.-C., L.P.-C.M., H.I.C.L., M.d.S.P.C. and M.T.H.-H.; writing—review and editing, I.P.V.M., E.P.-C. and M.T.H.-H.; supervision, E.Z., I.L.B.S., M.M.C., E.P.-C.M. and M.T.H.-H.; project administration, M.T.H.-H. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no funding.

Acknowledgments

Thanks to the National Council of Humanities, Sciences and Technologies (CONAHCYT) for the scholarship granted to I.P.V.M. and I.L.B.S. The authors would like to thank Eli Cruz Parada for their support throughout the work.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Trinca, G.M.; Hagan, C.R. O-GlcNAcylation in Women’s Cancers: Breast, Endometrial and Ovarian. J. Bioenerg. Biomembr. 2018, 50, 199–204. [Google Scholar] [CrossRef] [PubMed]
  2. Hart, G.W.; Slawson, C.; Ramirez-Correa, G.; Lagerlof, O. Cross Talk Between O-GlcNAcylation and Phosphorylation: Roles in Signaling, Transcription, and Chronic Disease. Annu. Rev. Biochem. 2011, 80, 825–858. [Google Scholar] [CrossRef] [PubMed]
  3. Chang, Y.H.; Weng, C.L.; Lin, K.I. O-GlcNAcylation and its role in the immune system. J. Biomed. Sci. 2020, 27, 57. [Google Scholar] [CrossRef] [PubMed]
  4. Dorfman, A.; Roseman, S.; Moses, F.E.; Ludowieg, J.; Mayeda, M. The biosynthesis of hyaluronic acid by group A Streptococcus. II. Origin of the N-acetylglucosamine moiety. J. Biol. Chem. 1955, 212, 583–591. [Google Scholar] [CrossRef]
  5. Ghosh, S.; Blumenthal, H.J.; Davidson, E.; Roseman, S. Glucosamine metabolism. V. Enzymatic synthesis of glucosamine 6-phosphate. J. Biol. Chem. 1960, 235, 1265–1273. [Google Scholar] [CrossRef]
  6. Slawson, C.; Hart, G.W. O-GlcNAc signalling: Implications for cancer cell biology. Nat. Rev. Cancer 2011, 11, 678–684. [Google Scholar] [CrossRef]
  7. Yang, X.; Ongusaha, P.P.; Miles, P.D.; Havstad, J.C.; Zhang, F.; So, W.V.; Kudlow, J.E.; Michell, R.H.; Olefsky, J.M.; Field, S.J.; et al. Phosphoinositide signalling links O-GlcNAc transferase to insulin resistance. Nature 2008, 451, 964–969. [Google Scholar] [CrossRef]
  8. Chou, T.Y.; Hart, G.W.; Dang, C.V. c-Myc is glycosylated at threonine 58, a known phosphorylation site and a mutational hot spot in lymphomas. J. Biol. Chem. 1995, 270, 18961–18965. [Google Scholar] [CrossRef]
  9. Slawson, C.; Housley, M.P.; Hart, G.W. O-GlcNAc cycling: How a single sugar post-translational modification is changing the way we think about signaling networks. J. Cell. Biochem. 2006, 97, 71–83. [Google Scholar] [CrossRef]
  10. Sakabe, K.; Wang, Z.; Hart, G.W. β-N-acetylglucosamine (O-GlcNAc) is part of the histone code. Proc. Natl. Acad. Sci. USA 2010, 107, 19915–19920. [Google Scholar] [CrossRef]
  11. Zachara, N.E.; O’Donnell, N.; Cheung, W.D.; Mercer, J.J.; Marth, J.D.; Hart, G.W. Dynamic O-GlcNAc modification of nucleocyto-plasmic proteins in response to stress. A survival response of mammalian cells. J. Biol. Chem. 2004, 279, 30133–30142. [Google Scholar] [CrossRef]
  12. Wells, L.; Vosseller, K.; Hart, G.W. Glycosylation of nucleocytoplasmic proteins: Signal transduction and O-GlcNAc. Science 2001, 291, 2376–2378. [Google Scholar] [CrossRef] [PubMed]
  13. Morino, K.; Maegawa, H. Role of O-linked N-acetylglucosamine in the homeostasis of metabolic organs, and its potential links with diabetes and its complications. J. Diabetes Investig. 2021, 12, 130–136. [Google Scholar] [CrossRef] [PubMed]
  14. Lagerlöf, O.; Slocomb, J.E.; Hong, I.; Aponte, Y.; Blackshaw, S.; Hart, G.W.; Huganir, R.L. The nutrient sensor OGT in PVN neurons regulates feeding. Science 2016, 351, 1293–1296. [Google Scholar] [CrossRef] [PubMed]
  15. Vosseller, K.; Wells, L.; Lane, M.D.; Hart, G.W. Elevated nucleocytoplasmic glycosylation by O-GlcNAc results in insulin re-sistance associated with defects in Akt activation in 3T3-L1 adipocytes. Proc. Natl. Acad. Sci. USA 2002, 99, 5313–5318. [Google Scholar] [CrossRef]
  16. Laczy, B.; Hill, B.G.; Wang, K.; Paterson, A.J.; White, C.R.; Xing, D.; Chen, Y.F.; Darley-Usmar, V.; Oparil, S.; Chatham, J.C. Protein O-GlcNAcylation: A new signaling paradigm for the cardiovascular system. Am. J. Physiol. Heart Circ. Physiol. 2009, 296, H13–H28. [Google Scholar] [CrossRef]
  17. Shan, X.; Vocadlo, D.J.; Krieger, C. Reduced protein O-glycosylation in the nervous system of the mutant SOD1 transgenic mouse model of amyotrophic lateral sclerosis. Neurosci. Lett. 2012, 516, 296–301. [Google Scholar] [CrossRef]
  18. Zhu, Y.; Shan, X.; Yuzwa, S.A.; Vocadlo, D.J. The Emerging Link between O-GlcNAc and Alzheimer Disease. J. Biol. Chem. 2014, 289, 34472–34481. [Google Scholar] [CrossRef]
  19. Ma, Z.; Vosseller, K. Cancer Metabolism and Elevated O-GlcNAc in Oncogenic Signaling. J. Biol. Chem. 2014, 289, 34457–34465. [Google Scholar] [CrossRef]
  20. Zou, Y.; Liu, Z.; Liu, W.; Liu, Z. Current knowledge and potential intervention of hexosamine biosynthesis pathway in lung can-cer. World J. Surg. Oncol. 2023, 21, 334. [Google Scholar] [CrossRef]
  21. Spaner, D.E. O-GlcNAcylation in Chronic Lymphocytic Leukemia and Other Blood Cancers. Front. Immunol. 2021, 12, 772304. [Google Scholar] [CrossRef] [PubMed]
  22. Palacios-Acedo, A.L.; Mège, D.; Crescence, L.; Dignat-George, F.; Dubois, C.; Panicot-Dubois, L. Platelets, Thrombo-Inflammation, and Cancer: Collaborating with the Enemy. Front. Immunol. 2019, 10, 1805. [Google Scholar] [CrossRef] [PubMed]
  23. Nakajima, H.; Murakami, K. O-GlcNAcylation: Implications in normal and malignant hematopoiesis. Exp. Hematol. 2021, 101–102, 16–24. [Google Scholar] [CrossRef] [PubMed]
  24. Wang, Z.; Fang, C.; Yao, M.; Wu, D.; Chen, M.; Guo, T.; Mo, J. Research progress of NF-κB signaling pathway and thrombosis. Front. Immunol. 2023, 14, 1257988. [Google Scholar] [CrossRef] [PubMed]
  25. Liu, A.R.; Ramakrishnan, P. Regulation of Nuclear Factor-kappaB Function by O-GlcNAcylation in Inflammation and Cancer. Front. Cell Dev. Biol. 2021, 9, 751761. [Google Scholar] [CrossRef]
  26. Thibault, B.; Ramos-Delgado, F.; Guillermet-Guibert, J. Targeting Class I-II-III PI3Ks in Cancer Therapy: Recent Advances in Tumor Biology and Preclinical Research. Cancers 2023, 15, 784. [Google Scholar] [CrossRef]
  27. Very, N.; Vercoutter-Edouart, A.S.; Lefebvre, T.; Hardivillé, S.; El Yazidi-Belkoura, I. Cross-Dysregulation of O-GlcNAcylation and PI3K/AKT/mTOR Axis in Human Chronic Diseases. Front. Endocrinol. 2018, 9, 602. [Google Scholar] [CrossRef]
  28. Abu Zaanona, M.I.; Mantha, S. Cancer-Associated Thrombosis. [Updated 17 July 2023]. In StatPearls [Internet]; StatPearls Publishing: Treasure Island, FL, USA, 2024. Available online: https://www.ncbi.nlm.nih.gov/books/NBK562222/ (accessed on 10 June 2024).
  29. Ding, S.; Dong, X.; Song, X. Tumor educated platelet: The novel BioSource for cancer detection. Cancer Cell Int. 2023, 23, 91. [Google Scholar] [CrossRef]
  30. Tatsumi, K. The pathogenesis of cancer-associated thrombosis. Int. J. Hematol. 2024, 119, 495–504. [Google Scholar] [CrossRef]
  31. Tavares, V.; Neto, B.V.; Marques, I.S.; Assis, J.; Pereira, D.; Medeiros, R. Cancer-associated thrombosis: What about microRNAs targeting the tissue factor coagulation pathway? Biochimica et biophysica acta. Rev. Cancer 2024, 1879, 189053. [Google Scholar] [CrossRef]
  32. Le Minh, G.; Esquea, E.M.; Young, R.G.; Huang, J.; Reginato, M.J. On a sugar high: Role of O-GlcNAcylation in cancer. JBC 2023, 299, 105344. [Google Scholar] [CrossRef] [PubMed]
  33. Ferreira, M.; Persello, A.; Souab, F.; Gaillard, C.; Denis, M.; Blangy-Letheule, A.; Erraud, A.; Maillard, A.; Dupas, T.; Bigot-Corbel, E.; et al. O-GlcNAcylation blood levels are increased in response to stress induced by cardiopulmonary bypass. Arch. Cardiovasc. 2021, 13, 217. [Google Scholar] [CrossRef]
  34. Boroughs, L.K.; DeBerardinis, R.J. Metabolic pathways promoting cancer cell survival and growth. Nat. Cell Biol. 2015, 17, 351–359. [Google Scholar] [CrossRef] [PubMed]
  35. Ferrer, C.M.; Sodi, V.L.; Reginato, M.J. O-GlcNAcylation in Cancer Biology: Linking Metabolism and Signaling. J. Mol. Biol. 2016, 428, 3282–3294. [Google Scholar] [CrossRef]
  36. Suchorolski, M.T.; Paulson, T.G.; Sanchez, C.A.; Hockenbery, D.; Reid, B.J. Warburg and Crabtree Effects in Premalignant Barrett’s Esophagus Cell Lines with Active Mitochondria. PLoS ONE 2013, 8, e56884. [Google Scholar] [CrossRef]
  37. Ouyang, M.; Yu, C.; Deng, X.; Zhang, Y.; Zhang, X.; Duan, F. O-GlcNAcylation and Its Role in Cancer-Associated Inflammation. Front. Immunol. 2022, 13, 861559. [Google Scholar] [CrossRef]
  38. Slawson, C.; Copeland, R.J.; Hart, G.W. O-GlcNAc Signaling: A Metabolic Link between Diabetes and Cancer? Trends Biochem. Sci. 2010, 35, 547–555. [Google Scholar] [CrossRef]
  39. Zhu, Q.; Wang, H.; Chai, S.; Xu, L.; Lin, B.; Yi, W.; Wu, L. O-GlcNAcylation promotes tumor immune evasion by inhibiting PD-L1 lysosomal degradation. Proc. Natl. Acad. Sci. USA 2023, 120, e2216796120. [Google Scholar] [CrossRef]
  40. Tian, X.; Li, Y.; Huang, Q.; Zeng, H.; Wei, Q.; Tian, P. High PD-L1 Expression Correlates with an Immunosuppressive Tumour Immune Microenvironment and Worse Prognosis in ALK-Rearranged Non-Small Cell Lung Cancer. Biomolecules 2023, 13, 991. [Google Scholar] [CrossRef]
  41. Aguiar De Azevedo, L.; Orione, C.; Tromeur, C.; Couturaud, F.; Descourt, R.; Geier, M. Incidence of venous thromboembolism and association with PD-L1 expression in advanced non-small cell lung cancer patients treated with first-line chemo-immunotherapy. Front. Oncol. 2024, 13, 1221106. [Google Scholar] [CrossRef]
  42. Söyler, Y.; Akın, K.P.; Kavurgacı, S.; Akyürek, N.; Demirağ, F.; Yılmaz, Ü. Could PD-L1 positivity be associated with venous thrombosis in patients with non-small cell lung cancer? J. Thromb. Thrombolysis 2023, 55, 382–391. [Google Scholar] [CrossRef] [PubMed]
  43. de Queiroz, R.M.; Carvalho, E.; Dias, W.B. O-GlcNAcylation: The Sweet Side of the Cancer. Front. Oncol. 2014, 4, 132. [Google Scholar] [CrossRef] [PubMed]
  44. Kita, Y.; Fukagawa, T.; Mimori, K.; Kosaka, Y.; Ishikawa, K.; Aikou, T.; Natsugoe, S.; Sasako, M.; Mori, M. Expression of uPAR mRNA in peripheral blood is a favourite marker for metastasis in gastric cancer cases. Br. J. Cancer 2009, 100, 153–159. [Google Scholar] [CrossRef] [PubMed]
  45. Kotzsch, M.; Farthmann, J.; Meye, A.; Fuessel, S.; Baretton, G.; Tjan-Heijnen, V.C.; Schmitt, M.; Luther, T.; Sweep, F.C.; Magdolen, V.; et al. Prognostic relevance of uPAR-del4/5 and TIMP-3 mRNA expression levels in breast cancer. Eur. J. Cancer 2005, 41, 2760–2768. [Google Scholar] [CrossRef]
  46. Nie, H.; Ju, H.; Fan, J.; Shi, X.; Cheng, Y.; Cang, X.; Zheng, Z.; Duan, X.; Yi, W. O-GlcNAcylation of PGK1 coordinates glycolysis and TCA cycle to promote tumor growth. Nat. Commun. 2020, 11, 36. [Google Scholar] [CrossRef]
  47. de la Cruz-López, K.G.; Castro-Muñoz, L.J.; Reyes-Hernández, D.O.; García-Carrancá, A.; Manzo-Merino, J. Lactate in the Regulation of Tumor Microenvironment and Therapeutic Approaches. Front. Oncol. 2019, 9, 1143. [Google Scholar] [CrossRef]
  48. Engstrom, M.; Schott, U.; Nordstrom, C.H.; Romner, B.; Reinstrup, P. Increased lactate levels impair the coagulation system—A potential contributing factor to progressive hemorrhage after traumatic brain injury. J. Neurosurg. Anesthesiol. 2006, 18, 200–204. [Google Scholar] [CrossRef]
  49. Schofield, L.; Lincz, L.F.; Skelding, K.A. Unlikely role of glycolytic enzyme α-enolase in cancer metastasis and its potential as a prognostic biomarker. J. Cancer Metastasis Treat. 2020, 6, 10. [Google Scholar] [CrossRef]
  50. Shi, Y.Y.; Chen, X.L.; Chen, Q.X.; Yang, Y.Z.; Zhou, M.; Ren, Y.X.; Tang, L.Y.; Ren, Z.F. Association of Enolase-1 with Prognosis and Immune Infiltration in Breast Cancer by Clinical Stage. J. Inflamm. Res. 2023, 16, 493–503. [Google Scholar] [CrossRef]
  51. Qiao, G.; Wu, A.; Chen, X.; Tian, Y.; Lin, X. Enolase 1, a Moonlighting Protein, as a Potential Target for Cancer Treatment. Int. J. Biol. Sci. 2021, 17, 3981–3992. [Google Scholar] [CrossRef]
  52. Huang, C.K.; Sun, Y.; Lv, L.; Ping, Y. ENO1 and Cancer. Mol. Ther. Oncolytics 2022, 24, 288–298. [Google Scholar] [CrossRef] [PubMed]
  53. Wu, X.; Ding, C.; Liu, Y.; Dong, K.; Zhang, H. B7-H3 promotes proliferation and migration of lung cancer cells by modulating PI3K/AKT pathway via ENO1 activity. Transl. Cancer Res. 2024, 13, 833–846. [Google Scholar] [CrossRef] [PubMed]
  54. Qian, X.; Xu, W.; Xu, J.; Shi, Q.; Li, J.; Weng, Y.; Jiang, Z.; Feng, L.; Wang, X.; Zhou, J.; et al. Enolase 1 stimulates glycolysis to promote chemoresistance in gastric cancer. Oncotarget 2017, 8, 47691–47708. [Google Scholar] [CrossRef] [PubMed]
  55. Liu, H.; Wang, X.; Shen, P.; Ni, Y.; Han, X. The basic functions of phosphoglycerate kinase 1 and its roles in cancer and other diseases. Eur. J. Pharmacol. 2022, 920, 174835. [Google Scholar] [CrossRef] [PubMed]
  56. Yen, W.C.; Wu, Y.H.; Wu, C.C.; Lin, H.R.; Stern, A.; Chen, S.H.; Shu, J.C.; Tsun-Yee Chiu, D. Impaired inflammasome activation and bacterial clearance in G6PD deficiency due to defective NOX/p38 MAPK/AP-1 redox signaling. Redox Biol. 2020, 28, 101363. [Google Scholar] [CrossRef]
  57. Yang, H.C.; Wu, Y.H.; Yen, W.C.; Liu, H.Y.; Hwang, T.L.; Stern, A.; Chiu, D.T. The Redox Role of G6PD in Cell Growth, Cell Death, and Cancer. Cells 2019, 8, 1055. [Google Scholar] [CrossRef]
  58. Dore, M.P.; Parodi, G.; Portoghese, M.; Pes, G.M. The Controversial Role of Glucose-6-Phosphate Dehydrogenase Deficiency on Cardiovascular Disease: A Narrative Review. Oxid. Med. Cell Longev. 2021, 2021, 5529256. [Google Scholar] [CrossRef]
  59. Errigo, A.; Bitti, A.; Galistu, F.; Salis, R.; Pes, G.M.; Dore, M.P. Relationship between Glucose-6-Phosphate Dehydrogenase Defi-ciency, X-Chromosome Inactivation and Inflammatory Markers. Antioxidants 2023, 12, 334. [Google Scholar] [CrossRef]
  60. Shimizu, M.; Tanaka, N. IL-8-induced O-GlcNAc modification via GLUT3 and GFAT regulates cancer stem cell-like properties in colon and lung cancer cells. Oncogene 2019, 38, 1520–1533. [Google Scholar] [CrossRef]
  61. Meyer, N.; Penn, L.Z. Reflecting on 25 years with MYC. Nat Rev Cancer 2008, 8, 976–990. [Google Scholar] [CrossRef]
  62. Dang, C.V. Re-thinking the Warburg Effect with Myc Micro-managing Glutamine Metabolism. Cancer Res. 2010, 70, 859–862. [Google Scholar] [CrossRef]
  63. Yeung, S.J.; Pan, J.; Lee, M.H. Roles of p53, MYC and HIF-1 in regulating glycolysis—The seventh hallmark of cancer. Cell Mol. Life Sci. 2008, 65, 3981–3999. [Google Scholar] [CrossRef]
  64. Wise, D.R.; DeBerardinis, R.J.; Mancuso, A.; Sayed, N.; Zhang, X.Y.; Pfeiffer, H.K.; Nissim, I.; Daikhin, E.; Yudkoff, M.; McMahon, S.B.; et al. Myc regulates a transcriptional program that stimulates mitochondrial glutaminolysis and leads to glu-tamine addiction. Proc. Natl. Acad. Sci. USA 2008, 105, 18782–18787. [Google Scholar] [CrossRef] [PubMed]
  65. Luanpitpong, S.; Angsutararux, P.; Samart, P.; Chanthra, N.; Chanvorachote, P.; Issaragrisil, S. Hyper-O-GlcNAcylation induces cisplatin resistance via regulation of p53 and c-Myc in human lung carcinoma. Sci. Rep. 2017, 7, 10607. [Google Scholar] [CrossRef] [PubMed]
  66. Rao, X.; Duan, X.; Mao, W.; Li, X.; Li, Z.; Li, Q.; Zheng, Z.; Xu, H.; Chen, M.; Wang, P.G.; et al. O-GlcNAcylation of G6PD promotes the pentose phosphate pathway and tumor growth. Nat. Commun. 2015, 6, 8468. [Google Scholar] [CrossRef] [PubMed]
  67. Motolani, A.; Martin, M.; Wang, B.; Jiang, G.; Alipourgivi, F.; Huang, X.; Safa, A.; Liu, Y.; Lu, T. Critical Role of Novel O-GlcNAcylation of S550 and S551 on the p65 Subunit of NF-κB in Pancreatic Cancer. Cancers 2023, 15, 4742. [Google Scholar] [CrossRef]
  68. Zhang, B.; Zhou, P.; Li, X.; Shi, Q.; Li, D.; Ju, X. Bitterness in sugar: O-GlcNAcylation aggravates pre-B acute lymphocytic leukemia through glycolysis via the PI3K/Akt/c-Myc pathway. Am. J. Cancer Res. 2017, 7, 1337–1349. [Google Scholar]
  69. Luanpitpong, S.; Poohadsuan, J.; Klaihmon, P.; Kang, X.; Tangkiettrakul, K.; Issaragrisil, S. Metabolic sensor O-GlcNAcylation regulates megakaryopoiesis and thrombopoiesis through c-Myc stabilization and integrin perturbation. Stem Cells 2021, 39, 787–802. [Google Scholar] [CrossRef]
  70. Mitrugno, A.; Sylman, J.L.; Ngo, A.T.; Pang, J.; Sears, R.C.; Williams, C.D.; McCarty, O.J. Aspirin therapy reduces the ability of platelets to promote colon and pancreatic cancer cell proliferation: Implications for the oncoprotein c-MYC. Am. J. Physiol. Cell Physiol. 2017, 312, C176–C189. [Google Scholar] [CrossRef]
  71. Jóźwiak, P.; Forma, E.; Bryś, M.; Krześlak, A. O-GlcNAcylation and Metabolic Reprograming in Cancer. Front. Endocrinol. 2014, 5, 145. [Google Scholar] [CrossRef]
  72. Vander Heiden, M.G.; Cantley, L.C.; Thompson, C.B. Understanding the Warburg Effect: The Metabolic Requirements of Cell Proliferation. Science 2009, 324, 1029–1033. [Google Scholar] [CrossRef] [PubMed]
  73. Ward, P.S.; Thompson, C.B. Metabolic Reprogramming: A Cancer Hallmark Even Warburg Did Not Anticipate. Cancer Cell 2012, 21, 297–308. [Google Scholar] [CrossRef] [PubMed]
  74. Wakil, S.J.; Porter, J.W.; Gibson, D.M. Studies on the mechanism of fatty acid synthesis. I. Preparation and purification of an enzymes system for reconstruction of fatty acid synthesis. Biochim. Biophys. Acta 1957, 24, 453–461. [Google Scholar] [CrossRef] [PubMed]
  75. Sun, L.; Lv, S.; Song, T. O-GlcNAcylation links oncogenic signals and cancer epigenetics. Discov. Oncol. 2021, 12, 54. [Google Scholar] [CrossRef] [PubMed]
  76. Khan, K.H.; Yap, T.A.; Yan, L.; Cunningham, D. Targeting the PI3K-AKT-mTOR signaling network in cancer. Chin. J. Cancer 2013, 32, 253–265. [Google Scholar] [CrossRef]
  77. Samuels, Y.; Wang, Z.; Bardelli, A.; Silliman, N.; Ptak, J.; Szabo, S.; Yan, H.; Gazdar, A.; Powell, S.M.; Riggins, G.J.; et al. High frequency of mutations of the PIK3CA gene in human can-cers. Science 2004, 304, 554. [Google Scholar] [CrossRef]
  78. Yuan, T.; Cantley, L. PI3K pathway alterations in cancer: Variations on a theme. Oncogene 2008, 27, 5497–5510. [Google Scholar] [CrossRef]
  79. Makwana, V.; Rudrawar, S.; Anoopkumar-Dukie, S. Signalling transduction of O-GlcNAcylation and PI3K/AKT/mTOR-axis in prostate cancer. Biochim. Biophys. Acta Mol. Basis Dis. 2021, 1867, 166129. [Google Scholar] [CrossRef]
  80. Van Hemelrijck, M.; Adolfsson, J.; Garmo, H.; Bill-Axelson, A.; Bratt, O.; Ingelsson, E.; Lambe, M.; Stattin, P.; Holmberg, L. Risk of thromboembolic diseases in men with prostate cancer: Results from the population-based PCBaSe Sweden. Lancet Oncol. 2010, 11, 450–458. [Google Scholar] [CrossRef]
  81. Woulfe, D.S. Akt signaling in platelets and thrombosis. Expert. Rev. Hematol. 2010, 3, 81–91. [Google Scholar] [CrossRef]
  82. Lu, Q.; Zhang, X.; Liang, T.; Bai, X. O-GlcNAcylation: An important post-translational modification and a potential therapeutic target for cancer therapy. Mol. Med. 2022, 28, 115. [Google Scholar] [CrossRef] [PubMed]
  83. Wu, J.L.; Chiang, M.F.; Hsu, P.H.; Tsai, D.Y.; Hung, K.H.; Wang, Y.H.; Angata, T.; Lin, K.I. O-GlcNAcylation is required for B cell homeostasis and antibody responses. Nat. Commun. 2017, 8, 1854. [Google Scholar] [CrossRef] [PubMed]
  84. Chuaire-Noack, L.; Sánchez-Corredor, M.C.; Ramírez-Clavijo, S. p53 and its role in the ovarian surface epithelium. A review. Investig. Clín. 2008, 49, 561–593. [Google Scholar]
  85. Vousden, K.H.; Ryan, K.M. p53 and metabolism. Nat. Rev. Cancer 2009, 9, 691–700. [Google Scholar] [CrossRef]
  86. Wang, H.; Guo, M.; Wei, H.; Chen, Y. Targeting p53 pathways: Mechanisms, structures, and advances in therapy. Signal Transduct. Target. Ther. 2023, 8, 92. [Google Scholar] [CrossRef]
  87. Borrero, L.J.H.; El-Deiry, W.S. Tumor Suppressor p53: Biology, Signaling Pathways, and Therapeutic Targeting. Biochim. Biophys. Acta Rev. Cancer 2021, 1876, 188556. [Google Scholar] [CrossRef]
  88. Schwartzenberg-Bar-Yoseph, F.; Armoni, M.; Karnieli, E. The tumor suppressor p53 down-regulates glucose transporters GLUT1 and GLUT4 gene expression. Cancer Res. 2004, 64, 2627–2633. [Google Scholar] [CrossRef]
  89. Zhang, C.; Liu, J.; Wu, R.; Liang, Y.; Lin, M.; Liu, J.; Chan, C.S.; Hu, W.; Feng, Z. Tumor suppressor p53 negatively regulates glycoly-sis stimulated by hypoxia through its target RRAD. Oncotarget 2014, 5, 5535–5546. [Google Scholar] [CrossRef]
  90. Wang, L.; Xiong, H.; Wu, F.; Zhang, Y.; Wang, J.; Zhao, L.; Guo, X.; Chang, L.J.; Zhang, Y.; You, M.J.; et al. Hexokinase 2-Mediated Warburg Effect Is Required for PTEN and p53-Deficiency Driven Prostate Cancer Growth. Cell Rep. 2014, 8, 1461–1474. [Google Scholar] [CrossRef]
  91. Kondoh, H.; Lleonart, M.E.; Gil, J.; Wang, J.; Degan, P.; Peters, G.; Martinez, D.; Carnero, A.; Beach, D. Glycolytic enzymes can modulate cellular life span. Cancer Res. 2005, 65, 177–185. [Google Scholar] [CrossRef]
  92. Jiang, P.; Du, W.; Wang, X.; Mancuso, A.; Gao, X.; Wu, M.; Yang, X. p53 regulates biosynthesis through direct inactivation of glucose-6-phosphate dehydrogenase. Nat. Cell Biol. 2011, 13, 310–316. [Google Scholar] [CrossRef] [PubMed]
  93. Stambolic, V.; MacPherson, D.; Sas, D.; Lin, Y.; Snow, B.; Jang, Y.; Benchimol, S.; Mak, T.W. Regulation of PTEN transcription by p53. Mol. Cell 2001, 8, 317–325. [Google Scholar] [CrossRef]
  94. Soga, T. Cancer metabolism: Key players in metabolic reprogramming. Cancer Sci. 2013, 104, 275–281. [Google Scholar] [CrossRef] [PubMed]
  95. Yan, S.R.; Novak, M.J. Src-family kinase-p53/ Lyn p56 plays an important role in TNF-alpha-stimulated production of O2- by human neutrophils adherent to fibrinogen. Inflammation 1999, 23, 167–178. [Google Scholar] [CrossRef] [PubMed]
  96. Briddon, S.J.; Watson, S.P. Evidence for the involvement of p59fyn and p53/56lyn in collagen receptor signalling in human platelets. Biochem. J. 1999, 338 Pt 1, 203–209. [Google Scholar] [CrossRef] [PubMed]
  97. Zhang, X.; Hu, M.; Lan, Y.; Yu, M.; Chen, B.D. Cleavage of Bcl-2 Protein by Activated Caspase-3 Is Associated with Inactivation of Lyn(p53/56) Kinase Activity in Human M-07e Leukemic Cells during Apoptosis. Zhongguo Shi Yan Xue Ye Xue Za Zhi 2000, 8, 166–175. [Google Scholar]
  98. Yan, S.R.; Byers, D.M.; Bortolussi, R. Role of protein tyrosine kinase p53/56lyn in diminished lipopolysaccharide priming of formylmethionylleucyl- phenylalanine-induced superoxide production in human newborn neutrophils. Infect. Immun. 2004, 72, 6455–6462. [Google Scholar] [CrossRef]
  99. Li, Y.; Xie, M.; Men, L.; Du, J. O-GlcNAcylation in immunity and inflammation: An intricate system (Review). Int. J. Mol. Med. 2019, 44, 363–374. [Google Scholar] [CrossRef]
  100. Metelli, A.; Wu, B.X.; Riesenberg, B.; Guglietta, S.; Huck, J.D.; Mills, C.; Li, A.; Rachidi, S.; Krieg, C.; Rubinstein, M.P.; et al. Thrombin con tributes to cancer immune eva-sion via proteolysis of platelet-bound GARP to activate LTGF-β. Sci. Transl. Med. 2020, 12, eaay4860. [Google Scholar] [CrossRef]
  101. Paul, S.; Mukherjee, T.; Das, K. Coagulation Protease-Driven Cancer Immune Evasion: Potential Targets for Cancer Immunotherapy. Cancers 2024, 16, 1568. [Google Scholar] [CrossRef]
  102. Balkwill, F.; Mantovani, A. Inflammation and cancer: Back to Virchow? Lancet 2001, 357, 539–545. [Google Scholar] [CrossRef]
  103. Singh, N.; Baby, D.; Rajguru, J.P.; Patil, P.B.; Thakkannavar, S.S.; Pujari, V.B. Inflammation and Cancer. Ann. Afr. Med. 2019, 18, 121–126. [Google Scholar] [CrossRef] [PubMed]
  104. Del Prete, A.; Allavena, P.; Santoro, G.; Fumarulo, R.; Corsi, M.M.; Mantovani, A. Molecular pathways in cancer-related inflam-mation. Biochem. Med. 2011, 21, 264–275. [Google Scholar] [CrossRef] [PubMed]
  105. Grivennikov, S.I.; Greten, F.R.; Karin, M. Immunity, Inflammation, and Cancer. Cell 2010, 140, 883–899. [Google Scholar] [CrossRef] [PubMed]
  106. Fardini, Y.; Dehennaut, V.; Lefebvre, T.; Issad, T. O-GlcNAcylation: A New Cancer Hallmark? Front. Endocrinol. 2013, 4, 99. [Google Scholar] [CrossRef] [PubMed]
  107. Salvatori, S.; Marafini, I.; Laudisi, F.; Monteleone, G.; Stolfi, C. Helicobacter pylori and Gastric Cancer: Pathogenetic Mechanisms. Int. J. Mol. Sci. 2023, 24, 2895. [Google Scholar] [CrossRef]
  108. Parsonnet, J.; Hansen, S.; Rodriguez, L.; Gelb, A.B.; Warnke, R.A.; Jellum, E.; Orentreich, N.; Vogelman, J.H.; Friedman, G.D. Helicobacter pylori Infection and Gastric Lymphoma. N. Engl. J. Med. 1994, 330, 1267–1271. [Google Scholar] [CrossRef]
  109. Laoruangroj, C.; Habermann, T.M.; Wang, Y.; King, R.L.; Lester, S.C.; Thompson, C.A.; Witzig, T.E. Should All Patients with Stage IE Gastric Mucosa-Associated Lymphoid Tissue Lymphoma Receive Antibiotic Eradication Therapy for Helicobacter pylori? JCO Oncol. Pract. 2024, 20, 1103–1108. [Google Scholar] [CrossRef]
  110. Neuveut, C.; Wei, Y.; Buendia, M.A. Mechanisms of HBV-related hepatocarcinogenesis. J. Hepatol. 2010, 52, 594–604. [Google Scholar] [CrossRef]
  111. Yang, G.; Wan, P.; Zhang, Y.; Tan, Q.; Qudus, M.S.; Yue, Z.; Luo, W.; Zhang, W.; Ouyang, J.; Li, Y.; et al. Innate Immunity, Inflam-mation, and Intervention in HBV Infection. Viruses 2022, 14, 2275. [Google Scholar] [CrossRef]
  112. Jiang, Y.; Han, Q.; Zhao, H.; Zhang, J. The Mechanisms of HBV-Induced Hepatocellular Carcinoma. J. Hepatocell. Carcinoma 2021, 8, 435–450. [Google Scholar] [CrossRef] [PubMed]
  113. Münger, K.; Baldwin, A.; Edwards, K.M.; Hayakawa, H.; Nguyen, C.L.; Owens, M.; Grace, M.; Huh, K. Mechanisms of human papillomavirus-induced oncogenesis. J. Virol. 2004, 78, 11451–11460. [Google Scholar] [CrossRef] [PubMed]
  114. Narisawa-Saito, M.; Kiyono, T. Basic mechanisms of high-risk human papillomavirus-induced carcinogenesis: Roles of E6 and E7 proteins. Cancer Sci. 2007, 98, 1505–1511. [Google Scholar] [CrossRef] [PubMed]
  115. García-Quiroz, J.; Vázquez-Almazán, B.; García-Becerra, R.; Díaz, L.; Avila, E. The Interaction of Human Papillomavirus Infection and Prostaglandin E2 Signaling in Carcinogenesis: A Focus on Cervical Cancer Therapeutics. Cells 2022, 11, 2528. [Google Scholar] [CrossRef] [PubMed]
  116. Kakegawa, T.; Bae, Y.; Ito, T.; Uchida, K.; Sekine, M.; Nakajima, Y.; Furukawa, A.; Suzuki, Y.; Kumagai, J.; Akashi, T.; et al. Frequency of Propionibacterium acnes Infection in Prostate Glands with Negative Biopsy Results Is an Independent Risk Factor for Prostate Cancer in Patients with Increased Serum PSA Titers. PLoS ONE 2017, 12, e0169984. [Google Scholar] [CrossRef]
  117. Li, Q.; Wu, W.; Gong, D.; Shang, R.; Wang, J.; Yu, H. Propionibacterium acnes overabundance in gastric cancer promote M2 polarization of macrophages via a TLR4/PI3K/Akt signaling. Gastric Cancer 2021, 24, 1242–1253. [Google Scholar] [CrossRef]
  118. Radej, S.; Szewc, M.; Maciejewski, R. Prostate Infiltration by Treg and Th17 Cells as an Immune Response to Propionibacterium acnes Infection in the Course of Benign Prostatic Hyperplasia and Prostate Cancer. Int. J. Mol. Sci. 2022, 23, 8849. [Google Scholar] [CrossRef]
  119. Efared, B.; Bako, A.B.A.; Idrissa, B.; Alhousseini, D.; Boureima, H.S.; Sodé, H.C.; Nouhou, H. Urinary bladder Schistosoma haematobium-related squamous cell carcinoma: A report of two fatal cases and literature review. Trop. Dis. Travel. Med. Vaccines 2022, 8, 3. [Google Scholar] [CrossRef]
  120. Pennington, L.F.; Alouffi, A.; Mbanefo, E.C.; Ray, D.; Heery, D.M.; Jardetzky, T.S.; Hsieh, M.H.; Falcone, F.H. H-IPSE Is a Pathogen-Secreted Host Nucleus-Infiltrating Protein (Infiltrin) Expressed Exclusively by the Schistosoma haematobium Egg Stage. Infect. Immun. 2017, 85, e00301-17. [Google Scholar] [CrossRef]
  121. Ishida, K.; Hsieh, M.H. Understanding Urogenital Schistosomiasis-Related Bladder Cancer: An Update. Front. Med. 2018, 5, 223. [Google Scholar] [CrossRef]
  122. Jalloh, M.; Cassell, A.; Diallo, T.; Gaye, O.; Ndoye, M.; Mbodji, M.M.; Mahamat, M.A.; Diallo, A.; Dial, C.; Labou, I.; et al. Is Schistosomiasis a Risk Factor for Bladder Cancer? Evidence-Based Facts. J. Trop. Med. 2020, 2020, 8270810. [Google Scholar] [CrossRef] [PubMed]
  123. Mathlouthi, N.E.H.; Kriaa, A.; Keskes, L.A.; Rhimi, M.; Gdoura, R. Virulence Factors in Colorectal Cancer Metagenomes and Association of Microbial Siderophores with Advanced Stages. Microorganisms 2022, 10, 2365. [Google Scholar] [CrossRef]
  124. Galardini, M.; Clermont, O.; Baron, A.; Busby, B.; Dion, S.; Schubert, S.; Beltrao, P.; Denamur, E. Major role of iron uptake systems in the intrinsic extra-intestinal virulence of the genus Escherichia revealed by a genome-wide association study. PLoS Genet. 2020, 16, e1009065. [Google Scholar] [CrossRef] [PubMed]
  125. Veziant, J.; Gagnière, J.; Jouberton, E.; Bonnin, V.; Sauvanet, P.; Pezet, D.; Barnich, N.; Miot-Noirault, E.; Bonnet, M. Association of colorectal cancer with pathogenic Escherichia coli: Focus on mechanisms using optical imaging. World J. Clin. Oncol. 2016, 7, 293–301. [Google Scholar] [CrossRef] [PubMed]
  126. Ma, Z.; Chalkley, R.J.; Vosseller, K. Hyper-O-GlcNAcylation activates nuclear factor κ-light-chain-enhancer of activated B cells (NF-κB) signaling through interplay with phosphorylation and acetylation. J. Biol. Chem. 2017, 292, 9150–9163. [Google Scholar] [CrossRef]
  127. Hayden, M.S.; Ghosh, S. Signaling to NF-κB. Genes. Dev. 2004, 18, 2195–2224. [Google Scholar] [CrossRef]
  128. Li, Q.; Verma, I.M. NF-κB regulation in the immune system. Nat. Rev. Immunol. 2002, 2, 725–734. [Google Scholar] [CrossRef]
  129. Kojok, K.; El-Kadiry, A.E.-H.; Merhi, Y. Role of NF-κB in Platelet Function. Int. J. Mol. Sci. 2019, 20, 4185. [Google Scholar] [CrossRef]
  130. Golks, A.; Tran, T.T.; Goetschy, J.F.; Guerini, D. Requirement for O-linked N-acetylglucosaminyltransferase in lymphocytes activation. EMBO J. 2007, 26, 4368–4379. [Google Scholar] [CrossRef]
  131. Karin, M.; Cao, Y.; Greten, F.R.; Li, Z.W. NF-κB in cancer: From innocent bystander to major culprit. Nat. Rev. Cancer 2002, 2, 301–310. [Google Scholar] [CrossRef]
  132. Yu, H.; Pardoll, D.; Jove, R. STATs in cancer inflammation and immunity: A leading role for STAT3. Nat. Rev. Cancer 2009, 9, 798–809. [Google Scholar] [CrossRef]
  133. Hu, X.; Li, J.; Fu, M.; Zhao, X.; Wang, W. The JAK/STAT signaling pathway: From bench to clinic. Signal Transduct. Target. Ther. 2021, 6, 402. [Google Scholar] [CrossRef] [PubMed]
  134. Triner, D.; Shah, Y.M. Hypoxia-inducible factors: A central link between inflammation and cancer. J. Clin. Investig. 2016, 126, 3689–3698. [Google Scholar] [CrossRef] [PubMed]
  135. Yang, W.H.; Park, S.Y.; Nam, H.W.; Kim, D.H.; Kang, J.G.; Kang, E.S.; Kim, Y.S.; Lee, H.C.; Kim, K.S.; Cho, J.W. NFκB activation is associated with its O-GlcNAcylation state under hyperglycemic conditions. Proc. Natl. Acad. Sci. USA 2008, 105, 17345–17350. [Google Scholar] [CrossRef]
  136. Ali, A.; Kim, S.H.; Kim, M.J.; Choi, M.Y.; Kang, S.S.; Cho, G.J.; Kim, Y.S.; Choi, J.Y.; Choi, W.S. O-GlcNAcylation of NF-κB Pro-motes Lung Metastasis of Cervical Cancer Cells via Upregulation of CXCR4 Expression. Mol. Cells 2017, 40, 476–484. [Google Scholar] [CrossRef] [PubMed]
  137. Ramakrishnan, P.; Clark, P.M.; Mason, D.E.; Peters, E.C.; Hsieh-Wilson, L.C.; Baltimore, D. Activation of the Transcriptional Function of the NF-κB Protein c-Rel by O-GlcNAc Glycosylation. Sci. Signal 2013, 6, ra75. [Google Scholar] [CrossRef]
  138. Gieling, R.G.; Elsharkawy, A.M.; Caamaño, J.H.; Cowie, D.E.; Wright, M.C.; Ebrahimkhani, M.R.; Burt, A.D.; Mann, J.; Ray-chaudhuri, P.; Liou, H.C.; et al. The c-Rel subunit of nuclear factor-kappaB regulates murine liver inflamma-tion, wound-healing, and hepatocyte proliferation. Hepatology 2010, 51, 922–931. [Google Scholar] [CrossRef]
  139. Pathak, S.; Borodkin, V.S.; Albarbarawi, O.; Campbell, D.G.; Ibrahim, A.; van Aalten, D.M. O-GlcNAcylation of TAB1 modulates TAK1-mediated cytokine release. EMBO J. 2012, 31, 1394–1404. [Google Scholar] [CrossRef]
  140. Shuang, T.; Wang, M.; Zhou, Y.; Shi, C.; Wang, D. NF-κB1, c-Rel, and ELK1 inhibit miR-134 expression leading to TAB1 upregu-lation in paclitaxel-resistant human ovarian cancer. Oncotarget 2017, 8, 24853–24868. [Google Scholar] [CrossRef]
  141. Tao, T.; He, Z.; Shao, Z.; Lu, H. TAB3 O-GlcNAcylation promotes metastasis of triple negative breast cancer. Oncotarget 2016, 7, 22807–22818. [Google Scholar] [CrossRef]
  142. de Jesus, T.; Shukla, S.; Ramakrishnan, P. Too Sweet to Resist: Control of Immune Cell Function by O-GlcNAcylation. Cell Immunol. 2018, 333, 85–92. [Google Scholar] [CrossRef] [PubMed]
  143. Ge, X.; Peng, X.; Li, M.; Ji, F.; Chen, J.; Zhang, D. OGT regulated O-GlcNacylation promotes migration and invasion by activating IL-6/STAT3 signaling in NSCLC cells. Pathol. Res. Pract. 2021, 225, 153580. [Google Scholar] [CrossRef] [PubMed]
  144. Freund, P.; Kerenyi, M.A.; Hager, M.; Wagner, T.; Wingelhofer, B.; Pham, H.T.T.; Elabd, M.; Han, X.; Valent, P.; Gouilleux, F.; et al. O-GlcNAcylation of STAT5 controls tyrosine phosphorylation and oncogenic tran-scription in STAT5-dependent malignancies. Leukemia 2017, 31, 2132–2142. [Google Scholar] [CrossRef]
  145. Li, X.; Gong, W.; Wang, H.; Li, T.; Attri, K.S.; Lewis, R.E.; Kalil, A.C.; Bhinderwala, F.; Powers, R.; Yin, G.; et al. O-GlcNAc Transferase Suppresses In-flammation and Necroptosis by Targeting Receptor-Interacting Serine/Threonine-Protein Kinase. Immunity 2019, 50, 576–590. [Google Scholar] [CrossRef] [PubMed]
  146. López, A.A.; Gao, Y.; Hou, X.; Lauvau, G.; Yates, J.R.; Wu, P. Profiling of Protein O-GlcNAcylation in Murine CD8+ Effector- and Memory-like T Cells. ACS Chem. Biol. 2017, 12, 3031–3038. [Google Scholar] [CrossRef]
  147. Seo, J.; Kim, H.; Won Cho, J. Role of O-GlcNAcylation of RIP3 Kinase in Necroptotic Cell Death. Korean Society of Glucose Science Conference. 2016, pp. 44–46. Available online: https://www.earticle.net/Article/A294699 (accessed on 16 July 2024).
  148. Blann, A.D.; Lip, G.Y. Venous thromboembolism. BMJ 2006, 332, 215–219. [Google Scholar] [CrossRef]
  149. Páramo, J.A. Diagnóstico de hipercoagulabilidad. Rev. Clínica Esp. 2001, 201, 30–32. [Google Scholar] [CrossRef]
  150. Fainchtein, K.; Tera, Y.; Kearn, N.; Noureldin, A.; Othman, M. Hypercoagulability and Thrombosis Risk in Prostate Cancer: The Role of Thromboelastography. Semin. Thromb. Hemost. 2023, 49, 111–118. [Google Scholar] [CrossRef]
  151. Falanga, A.; Marchetti, M.; Russo, L. Hemostatic Biomarkers and Cancer Prognosis: Where Do We Stand? Semin. Thromb. Hemost. 2021, 47, 962–971. [Google Scholar] [CrossRef]
  152. Giaccherini, C.; Marchetti, M.; Masci, G.; Verzeroli, C.; Russo, L.; Celio, L.; Sarmiento, R.; Gamba, S.; Tartari, C.J.; Diani, E.; et al. Thrombotic biomarkers for risk prediction of malignant disease recurrence in patients with early stage breast cancer. Haematologica 2020, 105, 1704–1711. [Google Scholar] [CrossRef]
  153. Goldberg, H.; Whiteside, C.; Fantus, I.G. O-linked β-N-acetylglucosamine supports p38 MAPK activation by high glucose in glomerular mesangial cells. Am. J. Physiol. Endocrinol. Metab. 2011, 301, E713–E726. [Google Scholar] [CrossRef] [PubMed]
  154. Paul, S.; Das, K.; Ghosh, A.; Chatterjee, A.; Bhoumick, A.; Basu, A.; Sen, P. Coagulation factor VIIa enhances programmed death-ligand 1 expression and its stability in breast cancer cells to promote breast cancer immune evasion. J. Thromb. Haemost. 2023, 21, 3522–3538. [Google Scholar] [CrossRef] [PubMed]
  155. Li, B.; Lu, Z.; Yang, Z.; Zhang, X.; Wang, M.; Chu, T.; Wang, P.; Qi, F.; Anderson, G.J.; Jiang, E.; et al. Monitoring circulating platelet activity to predict cancer-associated thrombosis. Cell Rep. Methods 2023, 3, 100513. [Google Scholar] [CrossRef] [PubMed]
  156. Xie, X.; Wu, Q.; Zhang, K.; Liu, Y.; Zhang, N.; Chen, Q.; Wang, L.; Li, W.; Zhang, J.; Liu, Y. O-GlcNAc modification regulates MTA1 transcriptional activity during breast cancer cell genotoxic adaptation. Biochim. Biophys. Acta Gen. Subj. 2021, 1865, 129930. [Google Scholar] [CrossRef] [PubMed]
  157. Dirix, L.Y.; Salgado, R.; Weytjens, R.; Colpaert, C.; Benoy, I.; Huget, P.; van Dam, P.; Prové, A.; Lemmens, J.; Vermeulen, P. Plasma fibrin D-dimer levels correlate with tumour volume, progression rate and survival in patients with metastatic breast cancer. Br. J. Cancer 2002, 86, 389–395. [Google Scholar] [CrossRef]
  158. Akella, N.M.; Le, M.G.; Ciraku, L.; Mukherjee, A.; Bacigalupa, Z.A.; Mukhopadhyay, D.; Sodi, V.L.; Reginato, M.J. O-GlcNAc Transferase Regulates Cancer Stem-like Potential of Breast Cancer Cells. Mol. Cancer Res. 2020, 18, 585–598. [Google Scholar] [CrossRef]
  159. Ray, U.; Bank, S.; Jayawardana, M.W.; Bhowmik, J.; Redwig, F.; Jana, P.; Bhattacharya, S.; Manna, E.; De, S.K.; Maiti, S.; et al. Insulin resistance in prostate cancer patients and predisposing them to acute ischemic heart disease. Biosci. Rep. 2019, 39, BSR20182313. [Google Scholar] [CrossRef]
  160. Zaslavsky, A.B.; Gloeckner-Kalousek, A.; Adams, M.; Putluri, N.; Venghatakrishnan, H.; Li, H.; Morgan, T.M.; Feng, F.Y.; Tewari, M.; Sreekumar, A.; et al. Platelet-Synthesized Testosterone in Men with Prostate Cancer Induces Androgen Receptor Signaling. Neoplasia 2015, 17, 490–496. [Google Scholar] [CrossRef]
  161. Gu, Y.; Gao, J.; Han, C.; Zhang, X.; Liu, H.; Ma, L.; Sun, X.; Yu, W. O-GlcNAcylation is increased in prostate cancer tissues and enhances malignancy of prostate cancer cells. Mol. Med. Rep. 2014, 10, 897–904. [Google Scholar] [CrossRef]
  162. Gil-Bazo, I.; Catalán, V.; Páramo, J.; Quero, C.; Escrivá de Romaní, S.; Pérez-Ochoa, A.; Arbea, L.; Navarro, V.; De la Cámara, J.; Garrán, C.; et al. Von Willebrand factor as an intermediate between hemostasis and angiogenesis of tumor origin. Rev. Med. Univ. Navarra 2003, 47, 22–28. [Google Scholar]
  163. Kamigaito, T.; Okaneya, T.; Kawakubo, M.; Shimojo, H.; Niahizawa, O.; Nakayama, J. Overexpression of O-GlcNAc by prostate cancer cells is significantly associated with poor prognosis of patients. Prostate Cancer Prostatic Dis. 2014, 17, 18–22. [Google Scholar] [CrossRef] [PubMed]
  164. Kobayashi, H.; Imanaka, S. Toward an understanding of tissue factor pathway inhibitor-2 as a novel serodiagnostic marker for clear cell carcinoma of the ovary. J. Obstet. Gynaecol. Res. 2021, 47, 2978–2989. [Google Scholar] [CrossRef] [PubMed]
  165. Li, H.; Sun, L.; Chen, L.; Kang, Z.; Hao, G.; Bai, F. Effects of adiponectin, plasma D-dimer, inflammation and tumor markers on clinical characteristics and prognosis of patients with ovarian cancer. J. Med. Biochem. 2022, 41, 71–78. [Google Scholar] [CrossRef] [PubMed]
  166. Niu, Y.; Xia, Y.; Wang, J.; Shi, X. O-GlcNAcylation promotes migration and invasion in human ovarian cancer cells via the RhoA/ROCK/MLC pathway. Mol. Med. Rep. 2017, 15, 2083–2089. [Google Scholar] [CrossRef]
  167. Cocco, E.; Varughese, J.; Buza, N.; Bellone, S.; Lin, K.Y.; Bellone, M.; Todeschini, P.; Silasi, D.A.; Azodi, M.; Schwartz, P.E.; et al. Tissue factor expression in ovarian cancer: Implications for immunotherapy with hI-con1, a factor VII-IgGF(c) chimeric protein targeting tissue factor. Clin. Exp. Metastasis 2011, 28, 689–700. [Google Scholar] [CrossRef]
  168. Finsterer, J.; Kuntscher, D.; Brunner, S.; Krugluger, W. Pseudotumor cerebri from sinus venous thrombosis, associated with polycystic ovary syndrome and hereditary hypercoagulability. Gynecol. Endocrinol. 2007, 23, 179–182. [Google Scholar] [CrossRef]
  169. Cheng, J.; Zeng, Z.; Ye, Q.; Zhang, Y.; Yan, R.; Liang, C.; Wang, J.; Li, M.; Yi, M. The association of pretreatment thrombocytosis with prognosis and clinicopathological significance in cervical cancer: A systematic review and meta-analysis. Oncotarget 2017, 8, 24327–24336. [Google Scholar] [CrossRef]
  170. Tas, M.; Yavuz, A.; Ak, M.; Ozcelik, B. Neutrophil-to-Lymphocyte Ratio and Platelet-to-Lymphocyte Ratio in Discriminating Precancerous Pathologies from Cervical Cancer. J. Oncol. 2019, 2019, 2476082. [Google Scholar] [CrossRef]
  171. Zeng, Q.; Zhao, R.X.; Chen, J.; Li, Y.; Li, X.D.; Liu, X.L.; Zhang, W.M.; Quan, C.S.; Wang, Y.S.; Zhai, Y.X.; et al. O-linked GlcNAcylation elevated by HPV E6 mediates viral onco-genesis. Proc. Natl. Acad. Sci. USA 2016, 113, 9333–9338. [Google Scholar] [CrossRef]
  172. Zhao, X.; Cheng, C.; Gou, J.; Yi, T.; Qian, Y.; Du, X.; Zhao, X. Expression of tissue factor in human cervical carcinoma tissue. Exp. Ther. Med. 2018, 16, 4075–4081. [Google Scholar] [CrossRef]
  173. Jiménez-Castillo, V.; Illescas-Barbosa, D.; Zenteno, E.; Avila-Curiel, B.X.; Castañeda-Patlán, M.C.; Robles-Flores, M.; Montante, D.; Pérez-Campos, E.; Torres-Rivera, A.; Bouaboud, A.; et al. Increased O-GlcNAcylation promotes IGF-1 receptor/PhosphatidyI Inositol-3 kinase/Akt pathway in cervical cancer cells. Sci. Rep. 2022, 12, 4464. [Google Scholar] [CrossRef] [PubMed]
  174. Miao, Y.; Xu, Z.; Feng, W.; Zheng, M.; Xu, Z.; Gao, H.; Li, W.; Zhang, Y.; Zong, Y.; Lu, A.; et al. Platelet infiltration predicts survival in postsurgical colorectal cancer patients. Int. J. Cancer 2022, 150, 509–520. [Google Scholar] [CrossRef] [PubMed]
  175. Dymicka-Piekarska, V.; Matowicka-Karna, J.; Gryko, M.; Kemona-Chętnik, I.; Kemona, H. Relationship between soluble P-selectin and inflammatory factors (interleukin-6 and C-reactive protein) in colorectal cancer. Thromb. Res. 2007, 120, 585–590. [Google Scholar] [CrossRef] [PubMed]
  176. Florescu, D.; Boldeanu, M.; Șerban, R.; Florescu, L.; Serbanescu, M.; Ionescu, M.; Streba, L.; Constantin, C.; Constantin, C. Correla-tion of the Pro-Inflammatory Cytokines IL-1β, IL-6, and TNF-α, Inflammatory Markers, and Tumor Markers with the Diag-nosis and Prognosis of Colorectal Cancer. Life Basel Switz. 2023, 13, 2261. [Google Scholar] [CrossRef]
  177. Jiang, M.; Xu, B.; Li, X.; Shang, Y.; Chu, Y.; Wang, W.; Chen, D.; Wu, N.; Hu, S.; Zhang, S.; et al. O-GlcNAcylation promotes colorectal cancer metastasis via the miR-101-O-GlcNAc/EZH2 regulatory feedback cir-cuit. Oncogene 2019, 38, 301–316. [Google Scholar] [CrossRef]
  178. Wang, W.S.; Lin, J.K.; Lin, T.C.; Chiou, T.J.; Liu, J.H.; Yen, C.C.; Chen, P.M. Plasma von Willebrand factor level as a prognostic indicator of patients with metastatic colorectal carcinoma. World J. Gastroenterol. 2005, 11, 2166–2170. [Google Scholar] [CrossRef]
  179. Sharma, N.S.; Gupta, V.K.; Dauer, P.; Kesh, K.; Hadad, R.; Giri, B.; Chandra, A.; Dudeja, V.; Slawson, C.; Banerjee, S.; et al. O-GlcNAc modification of Sox2 regulates self-renewal in pancreatic cancer by promoting its stability. Theranostics 2019, 9, 3410–3424. [Google Scholar] [CrossRef]
  180. Khorana, A.A.; Ahrendt, S.A.; Ryan, C.K.; Francis, C.W.; Hruban, R.H.; Hu, Y.C.; Hostetter, G.; Harvey, J.; Taubman, M.B. Tissue factor expression, angiogenesis, and thrombosis in pancreatic cancer. Clin. Cancer Res. 2007, 13, 2870–2875. [Google Scholar] [CrossRef]
  181. Thorson, C.M.; Van Haren, R.M.; Ryan, M.L.; Curia, E.; Sleeman, D.; Levi, J.U.; Livingstone, A.S.; Proctor, K.G. Pre-existing hypercoagulability in patients undergoing potentially curative cancer resection. Surgery 2014, 155, 134–144. [Google Scholar] [CrossRef]
  182. Detopoulou, P.; Panoutsopoulos, G.I.; Mantoglou, M.; Michailidis, P.; Pantazi, I.; Papadopoulos, S.; Rojas Gil, A.P. Relation of Mean Platelet Volume (MPV) with Cancer: A Systematic Review with a Focus on Disease Outcome on Twelve Types of Cancer. Curr. Oncol. 2023, 30, 3391–3420. [Google Scholar] [CrossRef]
  183. Matowicka-Karna, J.; Kamocki, Z.; Polińska, B.; Osada, J.; Kemona, H. Platelets and inflammatory markers in patients with gastric cancer. Clin. Dev. Immunol. 2013, 2013, 401623. [Google Scholar] [CrossRef] [PubMed]
  184. Jiang, M.; Qiu, Z.; Zhang, S.; Fan, X.; Cai, X.; Xu, B.; Li, X.; Zhou, J.; Zhang, X.; Chu, Y.; et al. Elevated O-GlcNAcylation promotes gastric cancer cells proliferation by modulating cell cycle related proteins and ERK 1/2 signaling. Oncotarget 2016, 7, 61390–61402. [Google Scholar] [CrossRef] [PubMed]
  185. Wang, G.; Xu, Z.; Sun, J.; Liu, B.; Ruan, Y.; Gu, J.; Song, S. O-GlcNAcylation enhances Reticulon 2 protein stability and its promo-tive effects on gastric cancer progression. Cell Signal 2023, 108, 110718. [Google Scholar] [CrossRef] [PubMed]
  186. Wang, J.; Li, H.; Xu, R.; Lu, T.; Zhao, J.; Zhang, P.; Qu, L.; Zhang, S.; Guo, J.; Zhang, L. The MLR, NLR, PLR and D-dimer are associated with clinical outcome in lung cancer patients treated with surgery. BMC Pulm. Med. 2022, 22, 104. [Google Scholar] [CrossRef]
  187. Lv, Z.; Ma, G.; Zhong, Z.; Xie, X.; Li, B.; Long, D. O-GlcNAcylation of RAB10 promotes hepatocellular carcinoma progression. Carcinogenesis 2023, 44, 785–794. [Google Scholar] [CrossRef]
  188. Xu, W.; Zhang, X.; Wu, J.L.; Fu, L.; Liu, K.; Liu, D.; Chen, G.G.; Lai, P.B.; Wong, N.; Yu, J. O-GlcNAc transferase promotes fatty liver-associated liver cancer through inducing palmitic acid and activating endoplasmic reticulum stress. J. Hepatol. 2017, 67, 310–320. [Google Scholar] [CrossRef]
  189. Jing, W.; Peng, R.; Zhu, M.; Lv, S.; Jiang, S.; Ma, J.; Ming, L. Differential Expression and Diagnostic Significance of Pre-Albumin, Fibrinogen Combined with D-Dimer in AFP-Negative Hepatocellular Carcinoma. Pathol. Oncol. Res. 2020, 26, 1669–1676. [Google Scholar] [CrossRef]
  190. Cen, G.; Song, Y.; Chen, S.; Liu, L.; Wang, J.; Zhang, J.; Li, J.; Li, G.; Li, H.; Liang, H.; et al. The investigation on the hypercoag-ulability of hepatocellular carcinoma-related cerebral infarction with thromboelastography. Brain Behav. 2023, 13, e2961. [Google Scholar] [CrossRef]
  191. Mohren, M.; Jentsch-Ullrich, K.; Koenigsmann, M.; Kropf, S.; Schalk, E.; Lutze, G. High coagulation factor VIII and von Wil-lebrand factor in patients with lymphoma and leukemia. Int. J. Hematol. 2016, 103, 189–195. [Google Scholar] [CrossRef]
  192. Xu, X.R.; Zhang, D.; Oswald, B.E.; Carrim, N.; Wang, X.; Hou, Y.; Zhang, Q.; Lavalle, C.; McKeown, T.; Marshall, A.H.; et al. Platelets are versatile cells: New discoveries in hemostasis, thrombosis, immune responses, tumor metastasis and beyond. Crit. Rev. Clin. Lab. Sci. 2016, 53, 409–430. [Google Scholar] [CrossRef]
  193. Zucchella, M.; Dezza, L.; Pacchiarini, L.; Meloni, F.; Tacconi, F.; Bonomi, E.; Grignani, G.; Notario, A. Human tumor cells cul-tured «in vitro» activate platelet function by producing ADP or thrombin. Haematologica 1989, 74, 541–545. [Google Scholar] [PubMed]
  194. De Candia, E. Mechanisms of platelet activation by thrombin: A short history. Thromb. Res. 2012, 129, 250–256. [Google Scholar] [CrossRef] [PubMed]
  195. Menter, D.G.; Tucker, S.C.; Kopetz, S.; Sood, A.K.; Crissman, J.D.; Honn, K.V. Platelets and cancer: A casual or causal relation-ship: Revisited. Cancer Metastasis Rev. 2014, 33, 231–269. [Google Scholar] [CrossRef] [PubMed]
  196. Holinstat, M. Normal platelet function. Cancer Metastasis Rev. 2017, 36, 195–198. [Google Scholar] [CrossRef]
  197. Han, X.; Guo, B.; Li, Y.; Zhu, B. Tissue factor in tumor microenvironment: A systematic review. J. Hematol. Oncol. 2014, 7, 54. [Google Scholar] [CrossRef]
  198. Olsson, A.K.; Cedervall, J. The pro-inflammatory role of platelets in cancer. Platelets 2018, 29, 569–573. [Google Scholar] [CrossRef]
  199. Wojtukiewicz, M.Z.; Sierko, E.; Hempel, D.; Tucker, S.C.; Honn, K.V. Platelets and cancer angiogenesis nexus. Cancer Metastasis Rev. 2017, 36, 249–262. [Google Scholar] [CrossRef]
  200. Cedervall, J.; Hamidi, A.; Olsson, A.K. Platelets, NETs and cancer. Thromb. Res. 2018, 164 (Suppl. S1), S148–S152. [Google Scholar] [CrossRef]
  201. Cools-Lartigue, J.; Spicer, J.; Najmeh, S.; Ferri, L. Neutrophil extracellular traps in cancer progression. Cell Mol. Life Sci. 2014, 71, 4179–4194. [Google Scholar] [CrossRef]
  202. Brinkmann, V.; Reichard, U.; Goosmann, C.; Fauler, B.; Uhlemann, Y.; Weiss, D.S.; Weinrauch, Y.; Sychlinsky, A. Neutrophil extracellular traps kill bacteria. Science 2004, 303, 1532–1535. [Google Scholar] [CrossRef]
  203. Fuchs, T.A.; Brill, A.; Duerschmied, D.; Schatzberg, D.; Monestier, M.; Myers, D.D.; Wrobleski, S.K.; Wakefield, T.W.; Hartwing, J.H.; Wagner, D.D. Extracellular DNA traps promote thrombosis. Proc. Natl. Acad. Sci. USA 2010, 107, 15880–15885. [Google Scholar] [CrossRef]
  204. Hu, H.; Zhu, W.; Qin, J.; Chen, M.; Gong, L.; Li, L.; Liu, X.; Tao, Y.; Yin, H.; Zhou, H.; et al. Acetylation of PGK1 promotes liver cancer cell proliferation and tumorigenesis. Hepatology 2017, 65, 515–528. [Google Scholar] [CrossRef] [PubMed]
  205. Wang, Y.; Liu, J.; Jin, X.; Zhang, D.; Li, D.; Hao, F.; Feng, Y.; Gu, S.; Meng, F.; Tian, M.; et al. O-GlcNAcylation destabilizes the active tetrameric PKM2 to promote the Warburg effect. Proc. Nat. Acad. Sci. USA 2017, 114, 13732–13737. [Google Scholar] [CrossRef] [PubMed]
  206. Sodi, V.L.; Bacigalupa, Z.A.; Ferrer, C.M.; Lee, J.V.; Gocal, W.A.; Mukhopadhyay, D.; Wellen, K.E.; Ivan, M.; Reginato, M.J. Nutrient sensor O-GlcNAc transferase controls cancer lipid metabolism via SREBP-1 regulation. Oncogene 2018, 37, 924–934. [Google Scholar] [CrossRef] [PubMed]
  207. Liu, Y.; Cao, Y.; Pan, X.; Shi, M.; Wu, Q.; Huang, T.; Jiang, H.; Li, W.; Zhang, J. O-GlcNAc elevation through activation of the hexosamine biosynthetic pathway enhances cancer cell chemoresistance. Cell Death Dis. 2018, 9, 485. [Google Scholar] [CrossRef]
  208. Quiang, A.; Slawson, C.; Fields, P.E. The Role of O-GlcNAcylation in Immune Cell Activation. Front. Endocrinol. 2021, 12, 596617. [Google Scholar] [CrossRef]
  209. Barkovskaya, A.; Seip, K.; Prasmickaite, L.; Mills, I.G.; Moestue, S.A.; Itkonen, H.M. Inhibition of O-GlcNAc transferase activates tumor-suppressor gene expression in tamoxifen-resistant breast cancer cells. Sci. Rep. 2020, 10, 16992. [Google Scholar] [CrossRef]
  210. Barkovskaya, A.; Seip, K.; Hilmarsdottir, B.; Maelandsmo, G.M.; Moestue, S.A.; Itkonen, H.M. O-GlcNAc Transferase Inhibition Differentially Affects Breast Cancer Subtypes. Sci. Rep. 2019, 9, 5670. [Google Scholar] [CrossRef]
  211. Lee, S.J.; Lee, D.E.; Choi, S.Y.; Kwon, O.S. OSMI-1 Enhances TRAIL-Induced Apoptosis through ER Stress and NF-κB Signaling in Colon Cancer Cells. Int. J. Mol. Sci. 2021, 22, 11073. [Google Scholar] [CrossRef]
  212. Feinberg, D.; Ramakrishnan, P.; Wong, D.P.; Asthana, A.; Parameswaran, R. Inhibition of O-GlcNAcylation Decreases the Cytotoxic Function of Natural Killer Cells. Front. Immunol. 2022, 13, 841299. [Google Scholar] [CrossRef]
  213. Zhang, J.; Yang, P.; Liu, D.; Gao, M.; Wang, J.; Yu, T.; Zhang, X.; Liu, Y. Inhibiting Hyper-O-GlcNAcylation of c-Myc accelerate diabetic wound healing by alleviating keratinocyte dysfunction. Burn. Trauma 2021, 9, tkab031. [Google Scholar] [CrossRef] [PubMed]
  214. Gross, B.J.; Kraybill, B.C.; Walker, S. Discovery of O-GlcNAc transferase inhibitors. J. Am. Chem. Soc. 2005, 127, 14588–14589. [Google Scholar] [CrossRef] [PubMed]
  215. Itkonen, H.M.; Gorad, S.S.; Duveau, D.Y.; Martin, S.E.; Barkovskaya, A.; Bathen, T.F.; Moestue, S.A.; Mills, I.G. Inhibition of O-GlcNAc transferase activity reprograms prostate cancer cell metabolism. Oncotarget 2016, 7, 12464–12476. [Google Scholar] [CrossRef] [PubMed]
  216. Liu, Y.; Ren, Y.; Cao, Y.; Huang, H.; Wu, Q.; Li, W.; Wu, S.; Zhang, J. Discovery of a Low Toxicity O-GlcNAc Transferase (OGT) Inhibitor by Structure-based Virtual Screening of Natural Products. Sci. Rep. 2017, 7, 12334. [Google Scholar] [CrossRef] [PubMed]
  217. Luanpitpong, S.; Tangkiettrakul, K.; Kang, X.; Srisook, P.; Poohadsuan, J.; Samart, P.; Klaihmon, P.; Janan, M.; Lorthongpanich, C.; Laowtammathron, C.; et al. OGT and OGA gene-edited human induced pluripotent stem cells for dissecting the functional roles of O-GlcNAcylation in hematopoiesis. Front. Cell Dev. Biol. 2024, 12, 1361943. [Google Scholar] [CrossRef]
  218. Olvera, A.; Martinez, J.P.; Casadellà, M.; Llano, A.; Rosás, M.; Mothe, B.; Ruiz-Riol, M.; Arsequell, G.; Valencia, G.; Noguera-Julian, M.; et al. Benzyl-2-Acetamido-2-Deoxy-α-d-Galactopyranoside Increases Human Immunodeficiency Virus Replication and Viral Outgrowth Efficacy In Vitro. Front. Immunol. 2018, 8, 2010. [Google Scholar] [CrossRef]
  219. Pantaleon, M.; Tan, H.Y.; Kafer, G.R.; Kaye, P.L. Toxic effects of hyperglycemia are mediated by the hexosamine signaling pathway and o-linked glycosylation in early mouse embryos. Biol. Reprod. 2010, 82, 751–758. [Google Scholar] [CrossRef]
  220. Zhang, D.; Qi, Y.; Inuzuka, H.; Liu, J.; Wei, W. O-GlcNAcylation in tumorigenesis and its implications for cancer therapy. J. Biol. Chem. 2024; Advance online publication. [Google Scholar] [CrossRef]
  221. Toleman, C.; Paterson, A.J.; Shin, R.; Kudlow, J.E. Streptozotocin inhibits O-GlcNAcase via the production of a transition state analog. Biochem. Biophys. Res. Commun. 2006, 340, 526–534. [Google Scholar] [CrossRef]
  222. Konrad, R.J.; Mikolaenko, I.; Tolar, J.F.; Liu, K.; Kudlow, J.E. The potential mechanism of the diabetogenic action of streptozotocin: Inhibition of pancreatic beta-cell O-GlcNAc-selective N-acetyl-beta-D-glucosaminidase. Biochem. J. 2001, 356 Pt 1, 31–41. [Google Scholar] [CrossRef]
  223. Nakai, K.; Umehara, M.; Minamida, A.; Yamauchi-Sawada, H.; Sunahara, Y.; Matoba, Y.; Okuno-Ozeki, N.; Nakamura, I.; Nakata, T.; Yagi-Tomita, A.; et al. Streptozotocin induces renal proximal tubular injury through p53 signaling activation. Sci. Rep. 2023, 13, 8705. [Google Scholar] [CrossRef] [PubMed]
  224. Korish, A.A.; Abdel Gader, A.G.M.; Alhaider, A.A. Comparison of the hypoglycemic and antithrombotic (anticoagulant) actions of whole bovine and camel milk in streptozotocin-induced diabetes mellitus in rats. J. Dairy Sci. 2020, 103, 30–41. [Google Scholar] [CrossRef] [PubMed]
  225. Lee, T.N.; Alborn, W.E.; Knierman, M.D.; Konrad, R.J. Alloxan is an inhibitor of O-GlcNAc-selective N-acetyl-beta-D-glucosaminidase. Biochem. Biophys. Res. Commun. 2006, 350, 1038–1043. [Google Scholar] [CrossRef] [PubMed]
  226. Stefanini, M. Influence of alloxan diabetes on coagulability of the blood. Proceedings of the Society for Experimental Biology and Medicine. Soc. Exp. Biol. Med. 1948, 69, 309–311. [Google Scholar] [CrossRef]
  227. Aivazian, A.I.; Petrishchev, N.N. Svertyvanie krovi i fibrinoliz u krys s alloksanovym diabetom [Blood coagulation and fibrinolysis in rats with alloxan diabetes]. Patol. Fiziol. I Eksperimental’naia Ter. 1973, 17, 83–85. [Google Scholar]
  228. Ding, N.; Ping, L.; Shi, Y.; Feng, L.; Zheng, X.; Song, Y.; Zhu, J. Thiamet-G-mediated inhibition of O-GlcNAcase sensitizes human leukemia cells to microtubule-stabilizing agent paclitaxel. Biochem. Biophys. Res. Commun. 2014, 453, 392–397. [Google Scholar] [CrossRef]
Figure 1. Modulating O-GlcNAcylation and some inhibitors. Abbreviations: HK, hexokinase; GPI, glucose-6-phosphate isomerase; GFAT, glutamine fructose-6-phosphate amidotransferase; GNPNAT/Emeg32, glucosamine–phosphate N-acetyltransferase; CoA, coenzyme A; GlcNAc, N-acetylglucosamine; PGM3, phosphoglucomutase 3; AGM1, phosphoacetylglucosamine mutase 1; UTP, uridine triphosphate; UDP, uridine diphosphate; UAP1, UDP-N-acetylglucosamine pyrophosphorylase 1; UDP-GlcNAc, UDP-N-acetylglucosamine; AGX1, UDP-N-acetylgalactosamine pyrophosphorylase; OGT, O-GlcNAc transferase; OGA, O-GlcNAcase; Ser, serine; Thr, threonine.
Figure 1. Modulating O-GlcNAcylation and some inhibitors. Abbreviations: HK, hexokinase; GPI, glucose-6-phosphate isomerase; GFAT, glutamine fructose-6-phosphate amidotransferase; GNPNAT/Emeg32, glucosamine–phosphate N-acetyltransferase; CoA, coenzyme A; GlcNAc, N-acetylglucosamine; PGM3, phosphoglucomutase 3; AGM1, phosphoacetylglucosamine mutase 1; UTP, uridine triphosphate; UDP, uridine diphosphate; UAP1, UDP-N-acetylglucosamine pyrophosphorylase 1; UDP-GlcNAc, UDP-N-acetylglucosamine; AGX1, UDP-N-acetylgalactosamine pyrophosphorylase; OGT, O-GlcNAc transferase; OGA, O-GlcNAcase; Ser, serine; Thr, threonine.
Ijms 25 09896 g001
Figure 2. O-GlcNAcylation of proteins affects biological and pathological homeostasis.
Figure 2. O-GlcNAcylation of proteins affects biological and pathological homeostasis.
Ijms 25 09896 g002
Figure 3. Pathways associated with O-GlcNAcylation, thrombosis, inflammation, and cancer. Abbreviations: TGFα, transforming growth factor alpha; EGF, epidermal growth factor; PDGF, platelet-derived growth factor subunit A; IGF, insulin-like growth factor; KITLG, KIT ligand; FLT3LG, fms-related tyrosine kinase 3 ligand; HGF, hepatocyte growth factor; FGF, fibroblast growth factor 1; EGFR, epidermal growth factor receptor; ERBB2, receptor tyrosine protein kinase erbB-2; PDGFR, platelet-derived growth factor receptor alpha; IGFR, insulin-like growth factor 1 receptor; C-KIT, proto-oncogene tyrosine protein kinase Kit; FLT3, fms-related tyrosine kinase 3; MET, proto-oncogene tyrosine protein kinase Met; FGFR, fibroblast growth factor receptor 1; Grb2, growth factor receptor-bound protein 2; Sos, son of sevenless; Ras, GTPase Ras; Raf, A-Raf proto-oncogene serine/threonine protein kinase; MEK or MAP2K1, mitogen-activated protein kinase kinase 1; ERK, mitogen-activated protein kinase 1/3; c-Myc, Myc proto-oncogene protein; c-Fos, proto-oncogene protein c-fos; c-Jun, transcription factor AP-1; VEGF, vascular endothelial growth factor; MMPs, matrix metalloproteinases (interstitial collagenase); IL-8, interleukin 8; CyclinD1, G1/S-specific cyclin-D1; CDK4, cyclin-dependent kinase 4; PLCγ, phosphatidylinositol phospholipase C, gamma-1; DAG, diacylglycerol; IP3, inositol 1,4,5-trisphosphate; Ca2+, calcium 2+; PKC, classical protein kinase C; CAM, calmodulin; CAMK, calcium/calmodulin-dependent protein kinase (CaM kinase) II; Elk, ETS domain-containing protein Elk-1; PTEN, phosphatidylinositol-3,4,5-trisphosphate 3-phosphatase and dual-specificity protein phosphatase PTEN; PI3K, phosphatidylinositol-4,5-bisphosphate 3-kinase catalytic subunit alpha/beta/delta; Akt, RAC serine/threonine protein kinase; mTOR, serine/threonine protein kinase mTOR; S6K, ribosomal protein S6 kinase beta; IKK, inhibitor of nuclear factor kappa-B kinase; Iκβα, NF-kappa-B inhibitor alpha; NFκβ, nuclear factor NF-kappa-B p105 subunit; COX-2, prostaglandin endoperoxide synthase 2; iNOS, nitric oxide synthase, inducible.
Figure 3. Pathways associated with O-GlcNAcylation, thrombosis, inflammation, and cancer. Abbreviations: TGFα, transforming growth factor alpha; EGF, epidermal growth factor; PDGF, platelet-derived growth factor subunit A; IGF, insulin-like growth factor; KITLG, KIT ligand; FLT3LG, fms-related tyrosine kinase 3 ligand; HGF, hepatocyte growth factor; FGF, fibroblast growth factor 1; EGFR, epidermal growth factor receptor; ERBB2, receptor tyrosine protein kinase erbB-2; PDGFR, platelet-derived growth factor receptor alpha; IGFR, insulin-like growth factor 1 receptor; C-KIT, proto-oncogene tyrosine protein kinase Kit; FLT3, fms-related tyrosine kinase 3; MET, proto-oncogene tyrosine protein kinase Met; FGFR, fibroblast growth factor receptor 1; Grb2, growth factor receptor-bound protein 2; Sos, son of sevenless; Ras, GTPase Ras; Raf, A-Raf proto-oncogene serine/threonine protein kinase; MEK or MAP2K1, mitogen-activated protein kinase kinase 1; ERK, mitogen-activated protein kinase 1/3; c-Myc, Myc proto-oncogene protein; c-Fos, proto-oncogene protein c-fos; c-Jun, transcription factor AP-1; VEGF, vascular endothelial growth factor; MMPs, matrix metalloproteinases (interstitial collagenase); IL-8, interleukin 8; CyclinD1, G1/S-specific cyclin-D1; CDK4, cyclin-dependent kinase 4; PLCγ, phosphatidylinositol phospholipase C, gamma-1; DAG, diacylglycerol; IP3, inositol 1,4,5-trisphosphate; Ca2+, calcium 2+; PKC, classical protein kinase C; CAM, calmodulin; CAMK, calcium/calmodulin-dependent protein kinase (CaM kinase) II; Elk, ETS domain-containing protein Elk-1; PTEN, phosphatidylinositol-3,4,5-trisphosphate 3-phosphatase and dual-specificity protein phosphatase PTEN; PI3K, phosphatidylinositol-4,5-bisphosphate 3-kinase catalytic subunit alpha/beta/delta; Akt, RAC serine/threonine protein kinase; mTOR, serine/threonine protein kinase mTOR; S6K, ribosomal protein S6 kinase beta; IKK, inhibitor of nuclear factor kappa-B kinase; Iκβα, NF-kappa-B inhibitor alpha; NFκβ, nuclear factor NF-kappa-B p105 subunit; COX-2, prostaglandin endoperoxide synthase 2; iNOS, nitric oxide synthase, inducible.
Ijms 25 09896 g003
Figure 4. PI3K/Akt/mTOR signaling pathway by O-GlcNAcylation. Abbreviations: IRS1-2, insulin receptor substrates 1 and 2; PI3K, phosphatidylinositol 3-kinase; PTEN, phosphatidylinosi-tol-3,4,5-trisphosphate 3-phosphatase; PIP2, phosphatidylinositol bisphosphate; PIP3, phosphatidylinositol triphosphate; PDK1, 3-phosphoinositide-dependent protein kinase 1; Akt, Ser/Thr kinase protein kinase B; mTORC2, mechanistic target of rapamycin complex 1; AMPK, adenosine monophosphate-activated protein kinase; TSC2, tuberous sclerosis complex 2; mTORC1, mechanistic target of rapamycin complex 1; 4EBP1, eukaryotic translation initiation factor 4E-binding protein; S6K1, ribosomal protein S6 kinase 1; HIF-1α, hypoxia-inducible factor 1 alpha subunit; SREBP1, sterol regulatory element binding protein 1; FASN, fatty acid synthase; HSP90A, heat shock protein 90 alpha.
Figure 4. PI3K/Akt/mTOR signaling pathway by O-GlcNAcylation. Abbreviations: IRS1-2, insulin receptor substrates 1 and 2; PI3K, phosphatidylinositol 3-kinase; PTEN, phosphatidylinosi-tol-3,4,5-trisphosphate 3-phosphatase; PIP2, phosphatidylinositol bisphosphate; PIP3, phosphatidylinositol triphosphate; PDK1, 3-phosphoinositide-dependent protein kinase 1; Akt, Ser/Thr kinase protein kinase B; mTORC2, mechanistic target of rapamycin complex 1; AMPK, adenosine monophosphate-activated protein kinase; TSC2, tuberous sclerosis complex 2; mTORC1, mechanistic target of rapamycin complex 1; 4EBP1, eukaryotic translation initiation factor 4E-binding protein; S6K1, ribosomal protein S6 kinase 1; HIF-1α, hypoxia-inducible factor 1 alpha subunit; SREBP1, sterol regulatory element binding protein 1; FASN, fatty acid synthase; HSP90A, heat shock protein 90 alpha.
Ijms 25 09896 g004
Figure 5. O-GlcNAcylation influences p53 functions. Abbreviations: GLUT: glucose transporters; HK2: hexokinase 2; PGAM1: phosphoglycerate mutase 1; CPT1C: carnitine palmitoyltransferase; SREBP1: sterol regulatory element-binding protein 1; R5P: ribose-5-phosphate; PTEN: phosphatidylinositol-3,4,5-trisphosphate 3-phosphatase; FAS: Fas receptor; DR5: death receptor 5; ISCU: iron–sulfur cluster assembly enzyme; SLC7A11: solute carrier family 7, member 11; GLS2: glutaminase 2; SAT1: spermidine/spermine N1-acetyltransferase 1; SCO2: synthesis of cytochrome c oxidase 2; AIF: mitochondrial protein apoptosis-inducing factor; PDK2: pyruvate dehydrogenase kinase 2; G6PD: glucose-6-phosphate dehydrogenase; MCD: malonyl-CoA decarboxylase; BAX: Bcl-2-associated X protein; PUMA: P53 up-regulated modulator of apoptosis.
Figure 5. O-GlcNAcylation influences p53 functions. Abbreviations: GLUT: glucose transporters; HK2: hexokinase 2; PGAM1: phosphoglycerate mutase 1; CPT1C: carnitine palmitoyltransferase; SREBP1: sterol regulatory element-binding protein 1; R5P: ribose-5-phosphate; PTEN: phosphatidylinositol-3,4,5-trisphosphate 3-phosphatase; FAS: Fas receptor; DR5: death receptor 5; ISCU: iron–sulfur cluster assembly enzyme; SLC7A11: solute carrier family 7, member 11; GLS2: glutaminase 2; SAT1: spermidine/spermine N1-acetyltransferase 1; SCO2: synthesis of cytochrome c oxidase 2; AIF: mitochondrial protein apoptosis-inducing factor; PDK2: pyruvate dehydrogenase kinase 2; G6PD: glucose-6-phosphate dehydrogenase; MCD: malonyl-CoA decarboxylase; BAX: Bcl-2-associated X protein; PUMA: P53 up-regulated modulator of apoptosis.
Ijms 25 09896 g005
Figure 6. O-GlcNAcylation, a crosstalk between hemostasis, inflammation, and cancer. Abbreviations: ADP: adenosine diphosphate; VEGF: vascular endothelial growth factor; PDGF: platelet-derived growth factor; TGFβ: transforming growth factor β; CXCL: C-X-C motif chemokine 13; MAPK: mitogen-activated protein kinases; PAI-1: plasminogen activator inhibitor 1; TF: tissue factor; PMPs: platelet-derived microparticles; CD40L: CD40 ligand; PLR: index platelet to lymphocyte ratio.
Figure 6. O-GlcNAcylation, a crosstalk between hemostasis, inflammation, and cancer. Abbreviations: ADP: adenosine diphosphate; VEGF: vascular endothelial growth factor; PDGF: platelet-derived growth factor; TGFβ: transforming growth factor β; CXCL: C-X-C motif chemokine 13; MAPK: mitogen-activated protein kinases; PAI-1: plasminogen activator inhibitor 1; TF: tissue factor; PMPs: platelet-derived microparticles; CD40L: CD40 ligand; PLR: index platelet to lymphocyte ratio.
Ijms 25 09896 g006
Table 1. Infectious agents, types of cancer, and virulence factors associated with chronic inflammation.
Table 1. Infectious agents, types of cancer, and virulence factors associated with chronic inflammation.
Infectious Agents Associated with CancerCancer Types with Increased Protein O-GlcNAcylation [106]Virulence FactorsAction through Modulation of Immune–Inflammatory PathwaysReferences
Helicobacter pyloriGastric, primary gastric non-Hodgkin’s lymphomaCytotoxin-associated gene A (CagA), vacuolating cytotoxin A (VacA), cytotoxin-associated gene pathogenicity island (cagPAI), neutrophil-activating protein (NAP)↑ Smad7
↑ROS, RNS
↑ IL-12, IL-17, IL-21
↑ IFN-γ, TNF-α
↑ MMP-2, MMP-9
↑ NF-κB activity
[107,108,109]
Hepatitis B virus (HBV)Hepatocellular carcinomaHBV surface antigen (HBsAg), HBV x protein (HBx)↑ NF-κB, MAPKs, JAK/STAT
↑ IFN-α, TNF-α, IL-15
↑ IL-1β, IL-18
↑ TAK1 activity
↑ IL-34 by activation STAT3
↑ TGF-β, IL-10
↑ miR-1269b by NF-κB dependent
[110,111,112]
Human papillomavirus (HPV)Skin, cervical, penile, vulvovaginal, analE proteins/genes↑ COX-1/COX-2, MMP, TNF-α
↑ ROS, RNS
↓ TLR 9, IFN-γ
↑ MDSCs
↓ cytotoxic T lymphocytes
↑ IL-10, TGF-β
EGFR/NF-κB activation
[113,114,115]
Propionibacterium acnesProstate, gastricAdhesive Flp pili encoded by a gene locus for tight adherence (tad)↑ M2 polarization of macrophages by TLR4/PI3K/Akt signaling
↑ IL-4, IL-10
↑ IL-6, IL-8
↑ JAK/STAT, IL-17
[116,117,118]
Schistosoma haematobiumBladder squamous cell carcinomaSoluble egg antigens, IPSE/alpha-1 protein↑ STAT4 (Th1), GATA3 (Th2), FOXP3 (T regulatory), CD8 (T cytotoxic)
↑ IL-10, TGF-β
[119,120,121,122]
Klebsiella pneumoniaeColorectalFimbriae virulence factors (mrkD, mrkF, mrkC)↑ JAK/STAT
NF-κB activation via the Akt/IκB kinase
↑ COX-2, ROS
[123,124,125]
Escherichia coliFimbriae virulence factors (FimE, FimB, FimG, erpC, FimA)
Yersinia pestisYersiniabactin (ybtE, fyuA, Ybtp, Irp1, Irp2)
Abbreviations: Smad7: suppressor of mothers against decapentaplegic 7; ROS: reactive oxygen species; RNS: reactive nitrogen species; IL: interleukin; MMP: matrix metalloproteinase; NF-κB: nuclear factor kappa-light-chain-enhancer of activated B cells; JAK: Janus kinase; STAT: signal transducer and activator of transcription; TAK1: factor-β-activated kinase 1; TGF-β: transforming growth factor-β; TLR: Toll-like receptor; MDSCs: immunosuppressive myeloid-derived suppressor cells; EGFR: epidermal growth factor receptor; ↑ increase; ↓ decreases.
Table 4. Inhibitors of O-GlcNAcylation and effect on platelets activity.
Table 4. Inhibitors of O-GlcNAcylation and effect on platelets activity.
Drugs/MoleculeMechanism of Action and PathwayEffect on Platelets ActivityReferences
OSMI-1Inhibition of OGT activates tumor-suppressor gene expression in tamoxifen-resistant breast cancer cells.
Enhances TRAIL-induced apoptosis through ER stress and NF-κB signaling in colon cancer cells.
Also, decreased expression of NKG2D and NKG2A receptors; cytokines including TNF-α and IFN-γ; cytotoxic mediators perforin, granzyme B, soluble Fas ligand, and granulysin, resulting in reduced cytotoxic function of NK cells.
Inhibition of OGT and O-GlcNAcylation induces megakaryocyte differentiation and platelet production through c-Myc stabilization and integrin perturbation. Absolute platelet counts revealed an increase in the number of CD41a+ platelet-like particles (PLPs) and CD42b+ PLPs, thereby validating the platelet-promoting role of OGT inhibition by OSMI-1.[69,209,210,211,212,213]
ST 045849Depletes intracellular alanine and decreases glucose consumption by cancer cells; reduces proliferation and viability of prostate cancer cells. [214,215]
L01Specific inhibitor of OGT and has low toxicity in cellular and zebrafish models. Inhibit cell proliferation by reducing the O-GlcNAcylation of proteins related to proliferation. Also, it could bind to N557, located near the UDP-binding pocket of OGT, and might contribute to the specificity of L01.Inhibition of OGT by small molecule inhibitors facilitates differentiation of hematopoietic stem and progenitor cells into megakaryocytes and stimulates platelet production, suggesting that reduced O-GlcNAcylation promotes megakaryopoiesis and thrombopoiesis.[216,217]
Benzyl-2-acetamido-2-deoxy-α-d-galactopyranoside (BADGP)BADGP acts as a GalNAc-α-1-O-serine/threonine mimic and, thus, as a competitive inhibitor of O-glycan chain extension by blocking the β1,3-galactosyltransferase involved in O-glycosylation elongation. Also, reduces O-GlcNAc level, but lacks specificity for OGT. [218,219,220]
Streptozotocin (STZ)STZ is primarily used to treat neuroendocrine tumors. STZ is incorporated into pancreatic β-cells through glucose transporter type 2, thereby promoting DNA damage, reactive oxygen species production, mitochondrial dysfunction, and subsequent apoptosis. STZ inhibits O-GlcNAcase via the production of a transition state analog.STZ is known to trigger a slow activation of coagulation, resulting in the consumption of circulating fibrinogen and loss of platelet aggregation.[221,222,223,224]
Alloxan Alloxan is an inhibitor of O-GlcNAc-selective N-acetyl-beta-d-glucosaminidase, with inhibition corresponding to an altered tryptic digest pattern of N-terminal active site peptides.The clotting time of whole blood and
recalcified plasma was slightly prolonged. Tissue plasmin inhibition showed no significant difference.
[225,226,227]
Thiamet GThiamet-G binds to OGA in competition with 4-methylumbelliferyl N-acetyl-β-d-glucosaminide dehydrate, an analogue of O-GlcNAc UDP, thereby suppressing the activity of OGA; inhibition of OGA by thiamet-G decreased the phosphorylation of microtubule-associated protein Tau and caused alterations of microtubule network in cells.
Also, thiamet-G-mediated inhibition of O-GlcNAcase sensitizes human leukemia cells to microtubule-stabilizing agent paclitaxel.
[228]
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

Vásquez Martínez, I.P.; Pérez-Campos, E.; Pérez-Campos Mayoral, L.; Cruz Luis, H.I.; Pina Canseco, M.d.S.; Zenteno, E.; Bazán Salinas, I.L.; Martínez Cruz, M.; Pérez-Campos Mayoral, E.; Hernández-Huerta, M.T. O-GlcNAcylation: Crosstalk between Hemostasis, Inflammation, and Cancer. Int. J. Mol. Sci. 2024, 25, 9896. https://doi.org/10.3390/ijms25189896

AMA Style

Vásquez Martínez IP, Pérez-Campos E, Pérez-Campos Mayoral L, Cruz Luis HI, Pina Canseco MdS, Zenteno E, Bazán Salinas IL, Martínez Cruz M, Pérez-Campos Mayoral E, Hernández-Huerta MT. O-GlcNAcylation: Crosstalk between Hemostasis, Inflammation, and Cancer. International Journal of Molecular Sciences. 2024; 25(18):9896. https://doi.org/10.3390/ijms25189896

Chicago/Turabian Style

Vásquez Martínez, Itzel Patricia, Eduardo Pérez-Campos, Laura Pérez-Campos Mayoral, Holanda Isabel Cruz Luis, María del Socorro Pina Canseco, Edgar Zenteno, Irma Leticia Bazán Salinas, Margarito Martínez Cruz, Eduardo Pérez-Campos Mayoral, and María Teresa Hernández-Huerta. 2024. "O-GlcNAcylation: Crosstalk between Hemostasis, Inflammation, and Cancer" International Journal of Molecular Sciences 25, no. 18: 9896. https://doi.org/10.3390/ijms25189896

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