Spandidos Publications Logo
  • About
    • About Spandidos
    • Aims and Scopes
    • Abstracting and Indexing
    • Editorial Policies
    • Reprints and Permissions
    • Job Opportunities
    • Terms and Conditions
    • Contact
  • Journals
    • All Journals
    • Oncology Letters
      • Oncology Letters
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • International Journal of Oncology
      • International Journal of Oncology
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • Molecular and Clinical Oncology
      • Molecular and Clinical Oncology
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • Experimental and Therapeutic Medicine
      • Experimental and Therapeutic Medicine
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • International Journal of Molecular Medicine
      • International Journal of Molecular Medicine
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • Biomedical Reports
      • Biomedical Reports
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • Oncology Reports
      • Oncology Reports
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • Molecular Medicine Reports
      • Molecular Medicine Reports
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • World Academy of Sciences Journal
      • World Academy of Sciences Journal
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • International Journal of Functional Nutrition
      • International Journal of Functional Nutrition
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • International Journal of Epigenetics
      • International Journal of Epigenetics
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • Medicine International
      • Medicine International
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
  • Articles
  • Information
    • Information for Authors
    • Information for Reviewers
    • Information for Librarians
    • Information for Advertisers
    • Conferences
  • Language Editing
Spandidos Publications Logo
  • About
    • About Spandidos
    • Aims and Scopes
    • Abstracting and Indexing
    • Editorial Policies
    • Reprints and Permissions
    • Job Opportunities
    • Terms and Conditions
    • Contact
  • Journals
    • All Journals
    • Biomedical Reports
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • Experimental and Therapeutic Medicine
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • International Journal of Epigenetics
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • International Journal of Functional Nutrition
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • International Journal of Molecular Medicine
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • International Journal of Oncology
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • Medicine International
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • Molecular and Clinical Oncology
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • Molecular Medicine Reports
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • Oncology Letters
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • Oncology Reports
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • World Academy of Sciences Journal
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
  • Articles
  • Information
    • For Authors
    • For Reviewers
    • For Librarians
    • For Advertisers
    • Conferences
  • Language Editing
Login Register Submit
  • This site uses cookies
  • You can change your cookie settings at any time by following the instructions in our Cookie Policy. To find out more, you may read our Privacy Policy.

    I agree
Search articles by DOI, keyword, author or affiliation
Search
Advanced Search
presentation
Oncology Reports
Join Editorial Board Propose a Special Issue
Print ISSN: 1021-335X Online ISSN: 1791-2431
Journal Cover
June-2025 Volume 53 Issue 6

Full Size Image

Sign up for eToc alerts
Recommend to Library

Journals

International Journal of Molecular Medicine

International Journal of Molecular Medicine

International Journal of Molecular Medicine is an international journal devoted to molecular mechanisms of human disease.

International Journal of Oncology

International Journal of Oncology

International Journal of Oncology is an international journal devoted to oncology research and cancer treatment.

Molecular Medicine Reports

Molecular Medicine Reports

Covers molecular medicine topics such as pharmacology, pathology, genetics, neuroscience, infectious diseases, molecular cardiology, and molecular surgery.

Oncology Reports

Oncology Reports

Oncology Reports is an international journal devoted to fundamental and applied research in Oncology.

Experimental and Therapeutic Medicine

Experimental and Therapeutic Medicine

Experimental and Therapeutic Medicine is an international journal devoted to laboratory and clinical medicine.

Oncology Letters

Oncology Letters

Oncology Letters is an international journal devoted to Experimental and Clinical Oncology.

Biomedical Reports

Biomedical Reports

Explores a wide range of biological and medical fields, including pharmacology, genetics, microbiology, neuroscience, and molecular cardiology.

Molecular and Clinical Oncology

Molecular and Clinical Oncology

International journal addressing all aspects of oncology research, from tumorigenesis and oncogenes to chemotherapy and metastasis.

World Academy of Sciences Journal

World Academy of Sciences Journal

Multidisciplinary open-access journal spanning biochemistry, genetics, neuroscience, environmental health, and synthetic biology.

International Journal of Functional Nutrition

International Journal of Functional Nutrition

Open-access journal combining biochemistry, pharmacology, immunology, and genetics to advance health through functional nutrition.

International Journal of Epigenetics

International Journal of Epigenetics

Publishes open-access research on using epigenetics to advance understanding and treatment of human disease.

Medicine International

Medicine International

An International Open Access Journal Devoted to General Medicine.

Journal Cover
June-2025 Volume 53 Issue 6

Full Size Image

Sign up for eToc alerts
Recommend to Library

  • Article
  • Citations
    • Cite This Article
    • Download Citation
    • Create Citation Alert
    • Remove Citation Alert
    • Cited By
  • Similar Articles
    • Related Articles (in Spandidos Publications)
    • Similar Articles (Google Scholar)
    • Similar Articles (PubMed)
  • Download PDF
  • Download XML
  • View XML

  • Supplementary Files
    • Supplementary_Data1.pdf
    • Supplementary_Data2.pdf
Article Open Access

Glucose transporter 2‑transported glucosamine inhibits glycolysis in cancer cell lines through competition with glucose for hexokinase II

  • Authors:
    • Se Yong Park
    • Ki-Hoon Song
    • Ju-Hee Kang
    • Seung Hyun Oh
  • View Affiliations / Copyright

    Affiliations: College of Veterinary Medicine, Seoul National University, Seoul 08826, Republic of Korea, ViroCure Inc., Guro, Seoul 08381, Republic of Korea
    Copyright: © Park et al. This is an open access article distributed under the terms of Creative Commons Attribution License.
  • Article Number: 73
    |
    Published online on: April 30, 2025
       https://doi.org/10.3892/or.2025.8906
  • Expand metrics +
Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Metrics: Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )
Cited By (CrossRef): 0 citations Loading Articles...

This article is mentioned in:



Abstract

Antiproliferative effects of glucosamine, a glucose derivative with a similar structure to glucose, have been discovered, but the molecular mechanisms are not yet fully understood. Since glucosamine and glucose not only have similar structures but also are catalyzed by the same enzyme, hexokinase (HK), the present study delved into determining whether the antiproliferative effect of glucosamine involved the inhibition of glycolysis by competition with glucose. Whole‑genome screening analysis showed that a number of the gene pathways controlled by glucosamine were directly and indirectly involved in glycolysis. In vitro experiments revealed that as more glucose was added, the antiproliferative effect of glucosamine decreased. Also, it was found that glucosamine was transported into cells mainly through glucose transporter (GLUT) 2 which was responsible for the antiproliferative effects of glucosamine. In addition, the present study found that cancer cell lines with low expression level of HKII show high sensitivity to glucosamine and a HK inhibitor, 3‑bromopyruvate, enhanced the antiproliferative effect of glucosamine. Under hypoxic conditions, activated hypoxia‑inducible factor 1α (HIF‑1α) inducing glucose uptake and glycolysis hampered glucosamine‑induced cell death and HIF1A knockdown or HK inhibitors restored the antiproliferative effects of glucosamine. These findings demonstrated that glucosamine is an efficient glycolysis inhibitor and that GLUT2 and HKII play important roles as biomarkers for determining sensitivity to glucosamine. Moreover, the results suggested that the antiproliferative effect of glucosamine may be more efficient when administered in combination with other glycolytic agents or inhibitors targeting HIF‑1α.

Introduction

Glucosamine is a natural amino sugar that is used as one of the building blocks of the body, such as in cartilage and tendon. Therefore, it has been used to treat patients with painful knee osteoarthritis and in complementary and alternative medicine as an osteoarthritis cure (1,2). In addition, a number of studies have reported that glucosamine inhibits cell proliferation and induces cell cycle arrest and apoptosis in various types of cancer (3,4). Other studies have described target proteins of glucosamine, such as transglutaminase 2, p70S6K, hypoxia-inducible factor 1α (HIF-1α), cyclooxygenase-2 (COX-2) and the insulin-like growth factor 1 receptor (IGF-1R)/Akt pathway (5–9). These studies propose that glucosamine has potential value as an antiproliferative drug.

Facilitative glucose transporters (GLUTs) are integral membrane proteins that have 12 transmembrane domains and a binding site for hexose substrates such as glucose, fructose and glucosamine (10). A total of 13 GLUT isoforms with highly conserved amino acid sequences and 12 hydrophobic α-helical domains have been identified and characterized in mammalian cells. Each GLUT isoform has a different expression level in various tissues and organs and plays a specific role in the energy-independent uptake of hexoses. For instance, GLUT1, which is highly expressed in all tissues, is in charge of basal glucose transport; GLUT2, which is abundant in liver, pancreatic islet and retinal cells, has a higher affinity for glucosamine than glucose (11).

Cancer cells undergo metabolic reprogramming to support their rapid growth and proliferation. One hallmark of this reprogramming is the Warburg effect, which describes the preference of cancer cells for aerobic glycolysis, even in the presence of sufficient oxygen (12). This phenomenon allows cancer cells to generate ATP quickly while producing metabolic intermediates essential for biosynthesis. A number of previous studies have also demonstrated that GLUTs are overexpressed in various types of cancer, facilitating increased glucose uptake to sustain the heightened glycolytic flux. This overexpression of GLUT proteins has been closely associated with aggressive tumor behavior, including metastasis and poor prognosis (13,14).

Once hexoses are transported into the cytoplasm, they are catalyzed by hexokinases (HKs), which carry out the first and rate-limiting step of the hexose metabolic pathway. In addition to glycolysis, HK also performs the irreversible step in the hexosamine pathway that converts hexose to hexose-6-phosphate, using an ATP molecule to start the process (15). In humans and other mammals, four HK isoforms have identified; HKI, II, III and IV (glucokinase). HKI, II and III have a high affinity for glucose and a molecular mass of ~100 kDa, whereas HKIV has a molecular mass of ~50 kDa and can only phosphorylate glucose, giving it a higher Km than the other HKs (16). During cancer progression, among those HKs, HKII, which has low affinity for glucose and is only slightly expressed in normal tissues, is markedly upregulated (17), suggesting that HKII plays an important role in malignancy.

HIF-1, a heterodimer consisting of α and β subunits, is stabilized under hypoxic conditions and plays a pivotal role in the glycolytic phenotype in human cancers (18). It has been reported that HIF-1α acts as a transcription factor and increases expressions of GLUTs and HKs and thus, tumor cells which are in hypoxic conditions activate HIF-1α to facilitate glucose uptake and glycolysis to satisfy their energy demands (19). Furthermore, since the metabolic reprogramming induced by HIF-1α helps cancer cells to grow, cancer cells turn on HIF-1α even with enough oxygens, a phenomenon called pseudohypoxia (20). These features make cancer cells more resistant to glycolysis inhibitors, such as 2-deoxy-D-glucose (2-DG), by increasing the number of glycolytic enzymes (21).

Malignant cancer cells have an elevated and accelerated glucose metabolism due to increased requirements for glucose, which is essential as an energy source (17). In addition, dysfunction of the Krebs cycle caused by mitochondrial mutations leads to increased dependence on glycolysis by cancer cells for ATP production (22). Therefore, it is not surprising that the increased aerobic glycolysis in cancer is closely related to overexpression of GLUTs and HKII in various types of cancer. A number of researchers have used glucose analogs such as fluorodeoxyglucose and 2-DG and HKII inhibitors such as 3-bromopyruvate (3-BrPA) to study interference of glucose uptake and glycolysis (15,23). As stated for the aforementioned molecules, these results indicate that glucosamine also may be a glycolysis inhibitor.

Based on these preliminary studies, the present study sought to determine whether the antiproliferative activity of glucosamine occurs because it is a glycolysis inhibitor in various cancer cells and whether GLUTs, HK isoforms and HIF-1α affect the sensitivity to glucosamine of cancer cells. It showed a new mechanism of glucosamine to specifically inhibit the glycolytic pathway to suppress cancer progression.

Materials and methods

Cell lines and agents

Human liver cancer cell lines (HepG2 and Hep3B), non-small cell lung cancer cell lines (A549, H1299 and H460) and breast cancer cell lines (MCF-7, MDA-MB-231 and T47D) were purchased from the American Type Culture Collection and cultured in RPMI 1640 (PAA Laboratories GmbH) supplemented with 10% fetal bovine serum (FBS; PAA Laboratories GmbH) and 100 U/ml penicillin-streptomycin (Gibco; Thermo Fisher Scientific, Inc.) at 37°C in a humidified atmosphere with 5% CO2. The authenticity of all eight cell lines used in the present study was validated by short tandem repeat (STR) DNA profiling, which was performed using PowerPlex 18D system by commercial service (Cosmo Genetech).

D-(+)-glucose, D-(+)-glucosamine hydrochloride, cytochalasin B and phloretin were purchased from MilliporeSigma, while 2-DG and 3-BrPA were purchased from Tokyo Chemical Industry. In the competition experiment between glucose and glucosamine, glucose-free RPMI 1640 medium (MilliporeSigma) was used and cells were pre-incubated for 24 h with the indicated doses of glucose prior to being treated with glucosamine.

Semiquantitative RT-PCR

Total RNA was isolated from cells using TRIzol® reagent (Thermo Fisher Scientific, Inc.) and chloroform extraction. First-strand cDNA was synthesized from 5 µg of total RNA using Superscript II reverse transcriptase (Invitrogen; Thermo Fisher Scientific, Inc.) according to the manufacturer's instructions. Semiquantitative PCR was performed in a thermal cycler (MyCycler; Bio-Rad Laboratories, Inc.) using gene-specific primer sets and sequences of primers are listed in the supplementary Table SI. The thermocycling conditions were as follows: Initial denaturation at 95°C for 5 min, followed by 25–45 cycles of denaturation at 95°C, annealing at 42–60°C for 30 sec (depending on the target gene), and elongation at 72°C for 1 min, with a final extension at 72°C for 5 min. The reactions were stopped during the exponential phase to ensure accurate comparisons. ACTB was used as a reference gene. PCR products were electrophoresed on 1.2–1.5% agarose gel and visualized using ethidium bromide.

siRNA transfection

A549 and HepG2 cells (2.5×105 cells/well) were seeded in 6-well plates and transfected with double-stranded small interfering RNA (siRNA). Target-specific siRNAs were designed for knockdown of GLUT1, GLUT2 and HIF1A were from Bioneer Corporation and MBiotech, respectively. A scrambled siRNA (siSCR) was synthesized and validated by Bioneer Corporation. The sense and anti-sense sequences of siRNAs are listed in the Table SII. Following 25 pmole of siRNA transfection for 24 h at 37°C using Lipofectamine® reagent (Invitrogen; Thermo Fisher Scientific, Inc.) according to the manufacturer's instructions, subsequent experiments were conducted. Knockdown of GLUT1, GLUT2 and HIF1A expression were validated by semiquantitative RT-PCR. The thermocycling conditions were as follows: Initial denaturation at 95°C for 5 min, followed by 25–45 cycles of denaturation at 95°C, annealing at 42–60°C for 30 sec (depending on the target gene), and elongation at 72°C for 1 min, with a final extension at 72°C for 5 min. The reactions were stopped during the exponential phase to ensure accurate comparisons. ACTB was used as a reference gene. PCR products were electrophoresed on 1.2–1.5% agarose gel and visualized using ethidium bromide.

Western blotting

Whole cell lysates were prepared using RIPA-B lysis buffer containing 150 mM NaCl, 1 mM EDTA (pH 8.0), 20 mM Tris-Cl (pH 7.4), 1% NP-40, 0.5% sodium deoxycholate, 1% Triton X-100, 1 mM Na3VO4, 5 mM NaF, 10% glycerol and protease inhibitor cocktail (Roche Diagnostics GmbH). Protein concentration was determined using BCA protein assay kit (Thermo Fisher Scientific, Inc.). Equivalent amounts of proteins (10–20 µg per lane, depending on the target protein) were separated by electrophoresis on 6, 8, 10 or 12% SDS-PAGE and then transferred to a polyvinylidene fluoride membrane (Pall Corporation). Membranes were blocked with 5% skimmed milk at room temperature for 1 h. The membranes were then incubated overnight at 4°C with primary antibodies against phosphorylated (p-)Akt (1:1,000, cat. no. 9271, Cell Signaling Technology, Inc.), Akt (1:1,000, cat. no. 4691, Cell Signaling Technology, Inc.), a/β-tubulin (1:1,000, cat. no. 2148, Cell Signaling Technology, Inc.), IGF-1R (1:1,000; cat. no. sc-713, Santa Cruz Biotechnology, Inc.) and COX-2 (1:1,000; cat. no. sc-1745, Santa Cruz Biotechnology, Inc.). After washing with Tris-buffered saline containing 0.1% Tween 20, the membranes were incubated with horseradish peroxidase-conjugated secondary antibodies (1:10,000; cat. no. AbC-5001 for anti-mouse antibody and cat. no. AbC-5003 for anti-rabbit antibody, Abclon) for 1 h at room temperature. Immunoreactive bands were detected using enhanced chemiluminescence (AbSignal; Abclon).

MTT assay

Cells (3.0×103 cells/well) were seeded into 96-well plates, incubated overnight at 37°C and treated with glucosamine or other agents. Control groups were given dimethyl sulfoxide (DMSO; 0.1% final concentration) or phosphate-buffered saline (PBS) vehicle. Cell viability was measured by MTT assay. An insoluble purple formazan precipitate was solubilized in DMSO and the absorbance was determined by microplate spectrometer at 562 nm.

Cell cycle analysis

A549 and HepG2 cells (1.5×105) were plated onto 60×15 mm dishes for 1 day and then cells were treated with glucose (1, 10, 20, or 50 mM) and glucosamine (1, 2, or 5 mM) at 37°C. Glucose was administered for 48 h and glucosamine was co-administered with glucose during the last 24 h of that period. After treatment, cells were washed once with PBS, trypsinized and harvested. All the harvested cells were fixed in 70% ethanol for 1 h at 4°C. After centrifugation for 5 min. at 1,000 × g and 4°C, cell pellets were washed twice with PBS and then stained with PI containing RNase A (40 µg/ml) for 30 min at 4°C in the dark. Total DNA contents were detected by BD FACSCalibur flow cytometer (BD Biosciences), and analyzed using FlowJo 10.10.0 (FlowJo LLC).

Apoptosis analysis

A549 and HepG2 cells (1.5×105) were plated onto 60×15 mm dishes for 1 day and after treated with glucosamine for 48 h, cells were harvested and washed twice with PBS on ice. Then, cells were stained using the FITC Annexin V Apoptosis Detection Kit I (BD Pharmingen; BD Biosciences). Cells were resuspended in 1× binding buffer containing 5 µl of fluorescein isothiocyanate (FITC)-Annexin V and 5 µl of PI. After staining, cells were detected by BD FACSCalibur flow cytometer (BD Biosciences) at 488 and 633 nm, and analyzed using FlowJo 10.10.0 (FlowJo LLC). The apoptotic rate was calculated as the percentage of early apoptotic cells and the percentage of late apoptotic cells. All procedures were carried out according to the manufacturer's instructions.

Measurement of glucosamine consumption rate

Cells (2.5×105 cells/well) were seeded in 6-well plates and incubated under RPMI 1640 medium containing the indicated concentrations of glucosamine. Glucosamine in harvested culture medium was reacted using D-Glucosamine Assay Kits (K-GAMINE; Megazyme) and optical density was measured with a spectrophotometer at 340 nm. All procedures were carried out according to the manufacturer's instructions.

Measurement of lactate production

The cellular glycolytic efficiency could be estimated by measuring the cellular lactate generation rate. Cells (2.5×105 cells/well) in the exponential growth phase were plated in 6-well plates for 1 day and cells were washed and treated with indicated concentrations of glucosamine for an additional 24 h. Aliquots of the culture medium were withdrawn and used for analysis of secreted lactic acid using the YSI 2300D STAT Plus Glucose and Lactate Analyzer (YSI Inc.).

Whole genome gene expression analysis

A549 cells were treated with DMSO or 5 mM of glucosamine for 24 h at 37°C and RNA was isolated. From each sample, 500 ng of total RNA was biotinylated and amplified using the Illumina TotalPrep RNA Amplification Kit (Ambion; Thermo Fisher Scientific, Inc.) according to the manufacturer's instructions. The cRNA yield was measured using RiboGreen RNA quantitation kit (Invitrogen; Thermo Fisher Scientific, Inc. and 750 ng of the cRNA sample was hybridized on a human HT-12 expression bead chip (Illumina, Inc.) profiling 48,804 transcripts per sample. The chips were stained with streptavidin and scanned using an Illumina BeadArray Reader (Illumina, Inc.). BeadStudio v3 (Illumina, Inc.) was used to quantile-normalize the data.

Gene ontology (GO) analysis was performed on genes that were increased or decreased by more than 2-fold in the glucosamine-treated sample compared with the control. The analysis was conducted using the gseGO function in the clusterProfiler package (ver. 4.14.3) (24) in R software (ver. 4.4.1, R Core Team, 2024) (25), based on the Gene Set Enrichment Analysis approach. GO terms which have adjust P-value <0.01 were considered markedly regulated GO. Up- or downregulated GO terms were featured by R software.

Kyoto Encyclopedia of Genes and Genomes (KEGG) pathway enrichment analysis was performed on differentially expressed genes using the enrichKEGG function from the clusterProfiler package in R software. The result of glycolysis pathway (hsa00010) was visualized using the Pathview package (Release 3.20) (26).

Statistical analysis

All statistical analyses were conducted using GraphPad Prism 8.0 (Dotmatics) and Microsoft Excel (Microsoft Corporation). For comparisons involving more than three groups, one-way ANOVA was performed, followed by Tukey's post hoc test. For comparisons between two groups, the Student's t-test was employed. P<0.05 was considered to indicate a statistically significant difference.

Results

Glucosamine regulates some gene sets related to glycolysis

In 1953, J.H. Quastel and A. Cantero demonstrated, for the first time, that D-glucosamine inhibits the tumor growth in a xenograft mouse model (27). After their findings, a number of studies have sought to discover the mechanism and related molecules in this antiproliferative effect. A number of studies have demonstrated that glucosamine decreases the ATP level in cancer cells and inhibits oncogenes, such as IGF-1R, HIF-1α and COX-2 (8,9,21). However, the mechanism has not been described in detail. Therefore, the present study used a microarray to screen genes that are regulated directly or affected indirectly by glucosamine. Genes induced by glucosamine more than two folds were categorized mainly into pro-apoptosis and cell death (Fig. 1A), while genes downregulated by glucosamine more than two folds mostly belonged to cell cycle and DNA damage responses (Fig. 1B). Next, the present study examined whether the genes involved in carbohydrate metabolism, biomolecules biosynthesis and glycosylation, which pathways are highly associated with glycolysis (28–30), are regulated by glucosamine. Notably, as shown in Table SIII, several genes upregulated by glucosamine more than twofold were involved in these pathways. Also, pathway enrichment of differential genes was used and then the KEGG pathway analysis results viewed to see whether the differential genes are enriched in glycolysis. As a result, differential genes are represented in the glycolysis pathway and their expression is primarily downregulated by glucosamine (Fig. 1C). With all these findings, glucosamine's similar structure to glucose and its reduction of ATP levels in cancer cells (4), as demonstrated by previous studies, it was assumed that glycolysis is the main target pathway of glucosamine.

Figure 1.

Whole genome gene expression analysis revealed that glucosamine regulates apoptosis and cell cycle. A549 cells were treated with 5 mM of glucosamine for 24 h and Whole genome gene expression analysis was performed. (A) Upregulated and (B) downregulated GO terms in the glucosamine treated sample compared with the control. GO terms possessing adjust P-value <0.01 were listed. (C) KEGG pathway enrichment analysis was performed and the result of the glycolysis pathway (hsa00010) was visualized with Pathview. GO, Gene Ontology; KEGG, Kyoto Encyclopedia of Genes and Genomes.

Competitive relationship between glucose and glucosamine constrains the inhibitory effects of glucosamine in cancer cells

Other than its NH2 group, glucosamine has a structure very similar to that of glucose. In fact, several studies have demonstrated that glucosamine competes with glucose in carbohydrate metabolism (31–33). Based on this, whether its antiproliferative effect is linked to this competitive relationship was investigated. A recent study has shown that glucosamine exhibits antiproliferative effects on a liver cancer cell line (34), while our previous research demonstrated similar effects on lung cancer cell lines (9). Therefore, the present study initially conducted experiments on HepG2 liver cancer cell line and A549 lung cancer cell line. Under low level of glucose, glucosamine inhibited proliferation of A549 and HepG2 cancer cells, which anti-proliferative effects of glucosamine were reversed by increasing glucose concentration (Fig. 2A). Glucose-treated cells also showed a dose-dependent recovery of the G0/G1 arrest induced by glucosamine (Fig. 2B). Since previous studies showed that glucosamine inhibits the IGF-1R/Akt pathway (9,35), the present study next investigated the effect of glucose on these target molecules. As expected, IGF-1R and p-Akt expression levels that had been reduced by glucosamine were restored as concentrations of glucose increased (Fig. 2C). To determine whether the competitive effect of glucose and glucosamine occurs during glycolysis in cancer cells, the level of lactate, which is produced from pyruvate during glycolysis was investigated. Glucosamine markedly suppressed lactate production, but this suppression was nearly reversed with a high dose of glucose (Fig. 2D). These findings suggested that the antiproliferative effect of glucosamine was closely related to competition between glucose and glucosamine, which occurs during glycolysis.

Figure 2.

GlcN-induced anticancer effect occurs through competition with Glc. (A) Cell viability of A549 and HepG2 cells after treatment with Glc and GlcN for 48 h. Cell viability was measured by MTT assay and results are presented as mean ± standard deviation. (B) Cell population by cell cycle of A549 and HepG2 cells after treated with Glc and GlcN. Cells were treated with Glc for 24 h, followed by additional 24 h treatment with GlcN. Cell cycle analysis was performed after propidium-iodide staining. (C) Western blot analysis of IGF-1R, p-Akt and Akt expression levels in A549 and HepG2 cells. Cells were treated with indicated concentrations of Glc and GlcN for 24 h. (D) Lactate production in A549 and HepG2 cell lines after treatment with Glc and GlcN for 24 h. Results are presented as mean ± standard deviation. *P<0.05, **P<0.01, ***P<0.001, ns, no significance; Glc, glucose; GlcN, glucosamine; p-, phosphorylated; IGF-1R, insulin-like growth factor 1 receptor.

Antiproliferative effect of glucosamine is related to basal expression level of GLUT2 and basal activity of glucosamine consumption

Glucosamine enters the cell mostly via GLUT2, along with other isoforms of GLUT (11). Although several studies have examined the kinetics of glucosamine transport, there is no direct evidence on how glucosamine, particularly its antiproliferative effects, influences these kinetics. Thus, it was hypothesized that cancer cells with higher expression levels of GLUT2 were more sensitive to the antiproliferative effect of glucosamine. First, the expression level of GLUT1 and GLUT2 in cancer cell lines originating from various types of cancer including liver cancer, lung cancer and breast cancer was screened. For this purpose, two liver cancer cell lines (HepG2 and Hep3B), three lung cancer cell lines (A549, H1299 and H460) and three breast cancer cell lines (MCF-7, MDA-MB-231 and T47D) were used. Among them, liver cancer cell lines showed higher levels of GLUT2 compared with the others, while most of the cell lines showed similar levels of GLUT1 (Fig. 3A). Then, to assess the correlation between glucosamine uptake and GLUT2 expression levels, a glucosamine consumption assay was performed and the results showed that liver cancer cell lines, which showed highly expressed GLUT2, consumed glucosamine more actively than others (Fig. 3B). Next, whether the expression level of GLUT2 and the rate of glucosamine uptake are related to the antiproliferative effect of glucosamine was investigated. As a result, liver cancer cell lines, which showed higher expressions of GLUT2 and actively consumed glucosamine, were more sensitive to glucosamine than others (Fig. 3C). Along with these results, HepG2 and Hep3B cells showed greater decreases in IGF-1R and p-Akt expression levels in the presence of glucosamine than A549 cells, which were the most sensitive among the remaining cell lines (Fig. 3D). Taken together, these results implied that the antiproliferative effect of glucosamine has a strong relationship with the expression level of GLUT2 and the uptake rate of glucosamine.

Figure 3.

GLUT2 basal level is associated with GlcN sensitivity in liver, lung and breast cancer cell lines. (A) Semiquantitative PCR of GLUT1 and GLUT2 in various cancer cell lines. Total mRNA from the following cancer cell lines was isolated and used for analysis: HepG2, Hep3B, A549, H1299, H460, MCF-7, MDA-MB-231 and T47D. ACTB was included as a loading control. (B) GlcN consumption rates in the eight cell lines. Each cell line was incubated in Glc-free RPMI 1640 containing both 1 mM Glc and 5 mM GlcN for 6 h. Results are presented as mean ± standard deviation. (C) Cell viability of the eight cell lines after treated with GlcN for 48 h. Cell viability was measured by MTT assay and results are presented as mean ± standard deviation. (D) Western blot analysis of IGF-1R, p-Akt and Akt expression levels in HepG2, Hep3B and A549 cells after 1mM of glucosamine treatment for 12 h. GLUT, glucose transporter; GlcN, glucosamine; IGF-1R, insulin-like growth factor 1 receptor.

Antiproliferative effect of glucosamine is dependent on the absorption process of glucosamine into cancer cell lines through GLUT2, not GLUT1

Then it was explored whether cancer cells uptake glucosamine through GLUT2, not GLUT1, and whether the uptake of glucosamine is an essential step for its antiproliferative effects. A549 and HepG2 cells were treated with cytochalasin B, a GLUT1-specific inhibitor, or phloretin, a GLUT2-specific inhibitor, in combination with glucosamine, respectively. Cytochalasin B and phloretin concentrations were taken from previously reported studies (36,37). As a result, phloretin recovered glucosamine-induced downregulation of IGF-1R and p-Akt expressions in both cancer cell lines, while cytochalasin B showed no recovery effects in A549 cells and partial recovery effects in HepG2 cells (Fig. 4A). In addition, glucosamine shifted the molecular mass of COX-2 as previously reported (38), which were markedly reversed by phloretin, but not by cytochalasin B in A549 cells (Fig. S1). Following these changes in molecular features, phloretin-mediated GLUT2 inhibition sufficiently rescued A549 and HepG2 cells from glucosamine-induced cell death, while the effects of cytochalasin B were not significant (Fig. 4B). Along with the results using inhibitors, GLUT2-knockdowned HepG2 cells by transfection of siGLUT2 showed recovery of IGF-1R and p-Akt expression levels decreased by glucosamine, while there were no changes in those expressions in GLUT1-knockdowned HepG2 cells (Fig. 4C). The present study also investigated whether treatment with phloretin or siGLUT2, which have an inhibitory effect on glucosamine transportation, could decrease consumption of glucosamine in both cancer cell lines and it was found that glucosamine consumption rate was reduced when GLUT2 was inhibited by phloretin and siGLUT2 (Fig. 4D). These findings demonstrate that the absorption of glucosamine via GLUT2 is the basic and necessary process for the antiproliferative effect of glucosamine.

Figure 4.

GlcN entrance into cancer cells through GLUT2 leads to its anticancer activities. (A) Western blot analysis of IGF-1R, p-Akt and Akt expression levels in A549 and HepG2 cells treated with cytochalasin B or phloretin for 24 h. Cytochalasin B was used as a GLUT1-specific inhibitor, while phloretin was used as a GLUT2-specific inhibitor. (B) Cell viability of A549 and HepG2 cells after treatment with glucosamine, cytochalasin B and phloretin for 48 h. Cell viability was measured by MTT assay and results are presented as mean ± standard deviation. (C) Left: Semiquantitative PCR of GLUT1 and GLUT2 in HepG2 cells after transfection with siSCR, siGLUT1, or siGLUT2. ACTB served as a loading control. Right: Western blot analysis of IGF-1R, p-Akt and Akt expression levels in GLUT1 or GLUT2 knockdown HepG2 cells after 1 mM of GlcN treatment for 24 h. (D) GlcN consumption rates in A549 and HepG2 cells. GLUT2 was inhibited by phloretin or siGLUT2 and then GlcN treatment for 6 h. Results are presented as mean ± standard deviation. *P<0.05, **P<0.01, ***P<0.001. GlcN, glucosamine; GLUT, glucose transporter; IGF-1R, insulin-like growth factor 1 receptor; si, small interfering.

Inhibition of glycolysis by glucosamine is greater in cancer cells with lower HKII levels

Although the other six cell lines expressed lower levels of GLUT2 compared with HepG2 and Hep3B cells, there were differences in the antiproliferative effect of glucosamine among those six cell lines (Fig. 3B and C), suggesting that cellular sensitivity to glucosamine is influenced by not only uptake through GLUT2 but also by post-uptake processes. Glucosamine can be phosphorylated by yeast hexokinase (39,40). In addition, yeast hexokinase is similar to rat hexokinase, human hexokinase N-terminal and C-terminal in respect of its sequence and structure amino acids, especially in the active site (41). Thus, it could be hypothesized that glucosamine serves as a substrate for hexokinase in mammalian cells including those of human origin (42–44). Thus, the present study examined the basic expression levels of HK isoforms (HK1-4) in the studied eight cell lines. Most cell lines showed similar levels of HK1 expression and rarely expressed HK3. Notably, the six cell lines showed lower GLUT2 expression and reduced glucosamine consumption rates compared with HepG2 and Hep3B cells, A549 cells exhibited the lowest HK2 expression level compared with the other five cell lines (Fig. 5A). Next, since HK is the first enzyme to be activated during glycolysis and this activation is a rate-limiting step that regulates the speed of glycolysis (45), the differences in glycolytic rates and the extent of glycolysis inhibition by glucosamine was assessed. Notably, the rate of lactate production was most markedly reduced by glucosamine in A549 cells which were the most sensitive to glucosamine after liver cancer cell lines. It was observed that 3-BrPA or 2-DG, which are used as HKII inhibitors (46,47), sufficiently inhibited the proliferation of those six cell lines (Fig. 5C and D). In case of A549 cells, the anti-proliferative effect of 3-BrPA was relatively weak compared with other lung cancer cell lines, H1299 and H460, and this was probably due to the lower level of sodium-coupled monocarboxylate transporter 1 (SMCT1), which is responsible for the transportation of 3-BrPA into cells (48), than other cancer cell lines tested (Fig. 5E). Taken together, these results showed that glycolysis inhibition by glucosamine is negatively related to the level of HKs, especially HKII.

Figure 5.

Basal levels of HKs are associated with GlcN sensitivity in human lung and breast cancer cell lines. (A) Semiquantitative PCR of HK isoforms in a variety of cancer cell lines. In the case of HK3, HL-60 which have high basal levels of HKIII, were used as a positive control. ACTB served as a loading control. (B) Reduction rate of lactate production by GlcN in the indicated cancer cell lines. Cells were treated with PBS or 2 mM or GlcN for 24 h and lactate levels in cultured medium were quantified. Results are presented as mean ± standard deviation. Cell viability of lung and breast cancer cells treated with (C) 3-BrPA or (D) 2-DG for 72 h. Cell viability was measured by MTT assay and results are presented as mean ± standard deviation. (E) Semiquantitative PCR of SMCT1 in the eight cell lines. ACTB served as a loading control. HK, hexokinase; GlcN, glucosamine.

Glucosamine and 3-BrPA enhance the antiproliferative activity of one another

3-BrPA, 2-DG and glucosamine, which are all glycolysis inhibitors, take different pathways to enter cells and work along different mechanisms. For example, 2-DG enters cells via GLUT1 and GLUT4 like glucose and works as a competitor of glucose, which eventually lowers the efficiency of glycolysis (23), whereas glucosamine is transported into cells efficiently via GLUT2 rather than GLUT1 and competes with glucose for HK. By contrast, 3-BrPA is taken up into cells by SMCT1 and directly suppresses the activity of HKII (47,48). Therefore, it was examined whether the antiproliferative effect of glucosamine is elevated when co-treated with 2-DG or 3-BrPA. Co-treatment of 3-BrPA with glucosamine increased the anti-proliferative effects of glucosamine in both A549 and HepG2 cells (Fig. 6A), whereas 2-DG showed a slight combinational effect and no combinational effect in A549 cells and HepG2 cells, respectively (Fig. 6B). Along with these results, glucosamine-induced decreased in IGF-1R and p-Akt expression level were tended to be enhanced by co-treatment of 3-BrPA (Fig. 6C), but not by co-treatment of 2-DG (Fig. 6D). These results suggested that the mechanism of glucosamine inhibiting HK is similar to that of 2-DG and different from that of 3-BrPA and that 3-BrPA may be a suitable HK inhibitor for co-administration with glucosamine.

Figure 6.

Combined effects of glycolysis inhibitors and GlcN. Cell viability of A549 and HepG2 cells after treated with GlcN combined with (A) 3-BrPA or (B) 2-DG for 48 h. Cell viability was measured by MTT assay and results are presented as mean ± standard deviation. (C, D) Western blot analysis of IGF-1R, p-Akt and Akt expression levels in HepG2, Hep3B and A549 cells after treatment with GlcN combined with 3-BrPA (C) or 2-DG (D) for 24 h. **P<0.01, ***P<0.001, ns, no significance; GlcN, glucosamine; IGF-1R, insulin-like growth factor 1 receptor; p-, phosphorylated.

HIF-1α/HKII signaling is responsible for hypoxia-induced glucosamine resistance

Glycolysis is mainly regulated by the quantity of glucose taken up into cells and the activity of HK, a rate-limiting step of the glycolysis pathway. Therefore, the expression level of GLUTs and HKs is an important factor in the glycolysis process. HIF-1α, one of the upstream transcriptional factors of GLUTs and HKs, plays an important role in controlling glycolysis (19). Based on this knowledge, it was first investigated whether the expression of GLUTs and HKs changes under hypoxic conditions in A549 and HepG2 cells and whether this is regulated by HIF-1α. As expected, the expression of GLUT1, HK1, HK2 and HK4 markedly increased under hypoxic conditions and it was confirmed that knocking down HIF-1 α markedly reduced their expression again (Fig. 7A). The expression of GLUT2 was constant in the examined conditions and the expression of HK1 was induced by hypoxic stress in HepG2 cells, but not in A549 cells. Next, the effects of hypoxia on the antiproliferative effects of glucosamine was examined. Notably, glucosamine reduced HIF-1α expression under hypoxic conditions in cancer cells (Fig. S2) and it was found that glucosamine-induced apoptosis was diminished under hypoxic conditions and this hypoxia-induced glucosamine resistance were restored by HIF-1α knockdown (Fig. 7B). Furthermore, cell cycle analysis found that the subG1 hypodiploid population, which was increased by glucosamine, was reduced by hypoxic stress and this hypoxic stress-induced change was not observed in HIF-1α depleted cells (Fig. 7C). Along with these results obtained from flow cytometry analysis, MTT assays showed that the anti-proliferation effects of glucosamine were reduced by hypoxic stress and this hypoxia-induced resistance was markedly ablated by HIF-1α downregulation in both A549 and HepG2 cells (Fig. 7D). Lastly, it was found that 3-BrPA, HKII inhibitor, also effectively reversed hypoxia-induced glucosamine resistance in both A549 and HepG2 cells (Fig. 7E). Collectively, these results suggested that glucosamine exerts its antiproliferative activity by inhibiting HKII within the cell and that the induction of HIF-1α and transactivation of its target genes, particularly HKII, are responsible for hypoxia-induced glucosamine resistance (Fig. 8).

Figure 7.

Anticancer effects of GlcN are regulated through changing of HK levels during hypoxia. (A) Semiquantitative PCR of HIF1A, GLUT1, GLUT2 and HK isoforms of A549 and HepG2 cells. Cells were transfected with siSCR or siHIF1A and then cultured under normoxic (20% O2) or hypoxic (1% O2) conditions for 24 h. HepG2 sample (C*) in Fig. 3A was used as a positive control for GLUT2 expression in the left panel. ACTB served as a loading control. Flow cytometric analysis of (B) apoptosis and (C) cell cycle in A549 cells. Cells were transfected with siSCR or siHIF1A for 24 h and further incubated under normoxic or hypoxic conditions for 24h followed by glucosamine treatment for additional 48 h. Subsequently, cells were stained with Annexin V-FITC and PI. Apoptosis (%) indicates sum of Annexin V single positive proportion and Annexin V-PI double positive proportion. Cell cycle analysis was performed with linearized value of PI intensity and population of subG1 cells is presented. (D) Cell viability of A549 and HepG2 cells. Cells were transfected with siSCR or siHIF1A for 24 h and further incubated under normoxic or hypoxic conditions for 24 h followed by GlcN treatment for additional 48 h. Cell viability was measured by MTT assay and results are presented as mean ± standard deviation. (E) Cell viability of A549 and HepG2 cells. Cells were incubated under normoxic or hypoxic conditions for 24 h followed by GlcN and 3-BrPA (10 or 20 µM for A549 or HepG2 cells, respectively) treatment for additional 48 h. Cell viability was measured by MTT assay and results are presented as mean ± standard deviation. *P<0.05, **P<0.01, ***P<0.001, ns, no significance; GlcN, glucosamine; HK, hexokinase; si, small interfering.

Figure 8.

Schematic representation of the glycolysis inhibition effect by GlcN in cancer cells. GlcN is mainly transported into cancer cells through GLUT2. Within cytoplasm, GlcN can inhibit the glycolysis of cancer cells by acting as competitor to glucose. As HIF-1α enhances glycolysis through increasing GLUT1 and HK expression, inhibition of HIF-1α by glucosamine can also contribute to the reduction in glycolysis. Eventually, ablated glycolysis by glucosamine leads cancer cell death. GlcN, glucosamine; GLUT, glucose transporter; hypoxia-inducible factor 1α; HK, hexokinase; Glc, glucose.

Discussion

A number of cancer cells exhibit elevated glycolysis even in the presence of oxygen, relying more heavily on this process to produce ATP compared with normal cells. This phenomenon, known as the Warburg effect, has become a key focus in cancer treatment strategies, particularly through the development of glycolytic inhibitors (15,23). Meanwhile, glycolysis is more than just an energy-generating pathway; it plays a crucial role in regulating various cellular processes necessary for cancer cell growth and physiology (49,50). Hence, glycolysis inhibition has been studied as an antiproliferative strategy. Although the inhibitory effects of glucosamine on glycolysis and tumor growth were reported in the late 1900s (27,32), the precise mechanisms of action of glucosamine driving antiproliferative effects have not been fully understood and thus, the present study investigated the antiproliferative mechanism of glucosamine.

At first, to gain a broader understanding of the mechanism behind glucosamine's antiproliferative effects, microarray technology, a powerful tool for high-throughput screening was employed. By using a whole-genome chip, the present study aimed to identify changes in the gene expression profile of A549 cells following the introduction of glucosamine. Genes upregulated by glucosamine by more than two-fold were primarily associated with pro-apoptosis and cell death, while those downregulated by more than two-fold were mostly linked to cell cycle regulation and DNA damage response. Additionally, glucosamine was found to influence genes involved in carbohydrate metabolism, biomolecule biosynthesis and glycosylation; processes related to glycolysis. In particular, the result showing that genes related to glycosylation were regulated by glucosamine is supported by previous reports demonstrating that glucosamine induces abnormal glycosylation of the IGF-1R prototype and COX-2 and eventually decreases their molecular mass (9,35). KEGG pathway analysis revealed that glucosamine downregulated the overall glycolysis pathway, while some enzymes regulating the entry into glycolysis were upregulated, possibly as a compensatory response to glycolysis inhibition. This compensatory mechanism is consistent with findings from previous studies using other glycolysis inhibitors (51,52). Given these findings, along with glucosamine's structural similarity to glucose and its previously reported effect of reducing ATP levels in cancer cells, it is suggested that glycolysis is the main target pathway of glucosamine.

Glucosamine, structurally similar to glucose except for its NH2 group, has been shown to compete with glucose in carbohydrate metabolism (23). According to this, glucose competitively inhibited glucosamine-induced antiproliferative effects and restored glycolysis as indicated by the recovery of lactate production that had been suppressed by glucosamine. Since glucosamine is known to inhibit the IGF-1R/Akt pathway (9), it was found that increasing glucose restored IGF-1R and p-Akt expression levels reduced by glucosamine. The results supported that glucosamine inhibited the rate of proliferation of cancer cell lines mainly by competing with glucose during glycolysis.

Based on the structural similarity of glucosamine and glucose, it was predicted that plasma membranes, especially where glucose transports exist, would be the battlefield of glucosamine and glucose. Among GLUT isoforms, expression levels in liver cancer cell lines of GLUT2, which is responsible for transportation of glucosamine as well (11), were much higher compared with other cell lines, while GLUT1 expression levels were similar between types of cancer. These GLUT2-rich liver cancer cell lines (HepG2 and Hep3B) consumed more glucosamine and were more sensitive to glucosamine than other GLUT2-poor cell lines. In addition, glucosamine consumption and the antiproliferative effects of glucosamine were abrogated by inhibiting GLUT2 with its specific inhibitor (Phloretin) or siGLUT2 transfection. Thus, so far, it can be said that GLUT2 is the main gate for glucosamine to get into cells and high dose of glucose can keep glucosamine from entering into cells at GLUT2. Previous studies report that cancer cells have higher GLUT2 expression levels compared with normal cells (37,53). These facts suggested that the amount of GLUT2 expression in each cancer cell can be a crucial biomarker in determining the sensitivity against glucosamine and also proves that glucosamine is more effective against cancer cells than normal cells.

Glucosamine and glucose not only have a similar structure but are also catalyzed by the same enzyme, HKs and are converted to GlcN-6-phosphate and Glc-6-phosphate, respectively, after transport into the cell (33). The lower the expression of HKII, the greater the inhibitory effect of glucosamine on glycolysis. Compared with the known glycolysis inhibitors, 3-BrPA and 2-DG, the effective dose range of glucosamine was similar to 2-DG, not to 3-BrPA which is entered into cells by SMCT rather than GLUTs, demonstrating that glucosamine inhibits glycolysis by competing with glucose as does 2-DG, rather than by directly inhibiting HKs. The synergistic effect observed when two drugs with different mechanisms of action are combined suggests distinct pathways (54). While the combination of glucosamine and 3-BrPA showed enhanced antiproliferative effects, the same was not observed with 2-DG. Considering these results and the structure of glucosamine, it can be inferred that the mechanism of action of glucosamine is probably different from that of 3-BrPA and similar to that of 2-DG.

Various former studies reveal that HIF-1α regulates the expression level of HKs and GLUT1 efficiently by working as a transcription factor (17,19,55). Furthermore, the increased stability of HIF-1α under hypoxic conditions has been reported to influence sensitivity of types of cancer to a number of anticancer agents (56,57). In particular, studies have demonstrated that the antiproliferative effect of 2-DG is decreased when there are high levels of HIF-1α and that this phenomenon is related to increased levels of HKs, which are induced by HIF-1α (21). The present study showed that A549 cells, which have low basal levels of HKII and lactate, inhibited glycolysis and decreased cell proliferation when treated with glucosamine. Under hypoxic conditions, GLUT2 expression levels were relatively constant in HepG2 cells, suggesting that the uptake amount of glucosamine would not be changed by HIF-1α activation. However, activated HIF-1α induced HKII expressions in lung and liver cancer cell lines and this activated HIF-1α/HKII pathway made cancer cell lines more resistant to glucosamine. In addition, genetically depleting HIF1A and pharmacologically inhibiting HKII markedly restored glucosamine-induced antiproliferative effects.

In humans, cohort studies found that regularly taking glucosamine as a supplement reduced the incidence and mortality of colorectal cancer, lung cancer and kidney cancer and the overall cancer mortality could be 6% lower compared with non-users (58,59). We previously reported in vivo anti-tumor effects of glucosamine using a xenograft mouse model (9). Furthermore, the present results supported the relationship between glucosamine intake and lower cancer mortality with the molecular mechanism of glucosamine-induced antiproliferative effects. Collectively, glucosamine can act like a roadblock on the path of glycolysis, competing with glucose for HK, slowing down the process like a rival runner in a race for the same baton. Nevertheless, the present study was limited by its exclusive use of cell line models, which may not fully reflect in vivo conditions. Further studies should focus on validating these findings in more physiologically relevant models and exploring the broader therapeutic potential of glucosamine under diverse biological conditions. Such efforts could clarify its efficacy across different tumor types and microenvironments, ultimately strengthening its clinical applicability.

In conclusion, the present study indicated that glucosamine reduced growth, decreased cell viability and impaired glucose metabolism in the studied cancer cell lines. Moreover, GLUT2 and HKII were identified as biomarkers determining sensitivity to glucosamine. While the precise molecular mechanisms underlying the antiproliferative effects of glucosamine remain to be fully elucidated, the present study provided significant insights by demonstrating its mode of action as a glucose metabolism inhibitor in cancer cells. Furthermore, the findings suggested potential therapeutic markers for identifying types of cancer or characteristics that are most responsive to glucosamine, thereby offering a rationale for its use in combination therapy. These findings contributed to expanding the potential of glucosamine as an effective anticancer agent.

Supplementary Material

Supporting Data
Supporting Data

Acknowledgements

Not applicable.

Funding

The present study was supported by the New Faculty Startup Fund (grant no. 550-20240054) from Seoul National University (Seoul, Korea).

Availability of data and materials

Microarray datasets generated in the present study has been submitted to ArrayExpress (accession number: E-MTAB-14871), and publicly accessible under this URL: https://www.ebi.ac.uk/biostudies/arrayexpress/studies/E-MTAB-14871. Other data generated in the present study may be requested from the corresponding author.

Authors' contributions

SP, KS and JK participated in the study design, conducted the experiments, collected and analyzed the data, and drafted the initial manuscript. JK and SO supervised the study, contributed to the study design and data interpretation, and critically revised the manuscript for important intellectual content. All authors confirm the authenticity of all the raw data. All authors read and approved the final manuscript.

Ethics approval and consent to participate

Not applicable.

Patient consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests. ViroCure Inc. had no commercial interest in this study.

References

1 

Sanders M and Grundmann O: The use of glucosamine, devil's claw (Harpagophytum procumbens), and acupuncture as complementary and alternative treatments for osteoarthritis. Altern Med Rev. 16:228–238. 2011.PubMed/NCBI

2 

Zhu X, Sang L, Wu D, Rong J and Jiang L: Effectiveness and safety of glucosamine and chondroitin for the treatment of osteoarthritis: A meta-analysis of randomized controlled trials. J Orthop Surg Res. 13:1702018. View Article : Google Scholar : PubMed/NCBI

3 

Wang LS, Chen SJ, Zhang JF, Liu MN, Zheng JH and Yao XD: Anti-proliferative potential of Glucosamine in renal cancer cells via inducing cell cycle arrest at G0/G1 phase. BMC Urol. 17:382017. View Article : Google Scholar : PubMed/NCBI

4 

Zahedipour F, Dalirfardouei R, Karimi G and Jamialahmadi K: Molecular mechanisms of anticancer effects of Glucosamine. Biomed Pharmacother. 95:1051–1058. 2017. View Article : Google Scholar : PubMed/NCBI

5 

Kim DS, Park KS, Jeong KC, Lee BI, Lee CH and Kim SY: Glucosamine is an effective chemo-sensitizer via transglutaminase 2 inhibition. Cancer Lett. 273:243–249. 2009. View Article : Google Scholar : PubMed/NCBI

6 

Oh HJ, Lee JS, Song DK, Shin DH, Jang BC, Suh SI, Park JW, Suh MH and Baek WK: D-glucosamine inhibits proliferation of human cancer cells through inhibition of p70S6K. Biochem Biophys Res Commun. 360:840–845. 2007. View Article : Google Scholar : PubMed/NCBI

7 

Jo JR, Park YK and Jang BC: Short-term treatment with glucosamine hydrochloride specifically downregulates hypoxia-inducible factor-1α at the protein level in YD-8 human tongue cancer cells. Int J Oncol. 44:1699–1706. 2014. View Article : Google Scholar : PubMed/NCBI

8 

Chou WY, Chuang KH, Sun D, Lee YH, Kao PH, Lin YY, Wang HW and Wu YL: Inhibition of PKC-Induced COX-2 and IL-8 expression in human breast cancer cells by glucosamine. J Cell Physiol. 230:2240–2251. 2015. View Article : Google Scholar : PubMed/NCBI

9 

Song KH, Kang JH, Woo JK, Nam JS, Min HY, Lee HY, Kim SY and Oh SH: The novel IGF-IR/Akt-dependent anticancer activities of glucosamine. BMC Cancer. 14:312014. View Article : Google Scholar : PubMed/NCBI

10 

Cura AJ and Carruthers A: Role of monosaccharide transport proteins in carbohydrate assimilation, distribution, metabolism, and homeostasis. Compr Physiol. 2:863–914. 2012. View Article : Google Scholar : PubMed/NCBI

11 

Uldry M, Ibberson M, Hosokawa M and Thorens B: GLUT2 is a high affinity glucosamine transporter. FEBS Lett. 524:199–203. 2002. View Article : Google Scholar : PubMed/NCBI

12 

Pouyssegur J, Marchiq I, Parks SK, Durivault J, Zdralevic M and Vucetic M: ‘Warburg effect’ controls tumor growth, bacterial, viral infections and immunity-Genetic deconstruction and therapeutic perspectives. Semin Cancer Biol. 86((Pt 2)): 334–346. 2022. View Article : Google Scholar : PubMed/NCBI

13 

Barron CC, Bilan PJ, Tsakiridis T and Tsiani E: Facilitative glucose transporters: Implications for cancer detection, prognosis and treatment. Metabolism. 65:124–139. 2016. View Article : Google Scholar : PubMed/NCBI

14 

Ancey PB, Contat C and Meylan E: Glucose transporters in cancer - from tumor cells to the tumor microenvironment. FEBS J. 285:2926–2943. 2018. View Article : Google Scholar : PubMed/NCBI

15 

Pelicano H, Martin DS, Xu RH and Huang P: Glycolysis inhibition for anticancer treatment. Oncogene. 25:4633–4646. 2006. View Article : Google Scholar : PubMed/NCBI

16 

Wilson JE: Isozymes of mammalian hexokinase: Structure, subcellular localization and metabolic function. J Exp Biol. 206((Pt 12)): 2049–2057. 2003. View Article : Google Scholar : PubMed/NCBI

17 

Guo D, Meng Y, Jiang X and Lu Z: Hexokinases in cancer and other pathologies. Cell Insight. 2:1000772023. View Article : Google Scholar : PubMed/NCBI

18 

Paredes F, Williams HC and San Martin A: Metabolic adaptation in hypoxia and cancer. Cancer Lett. 502:133–142. 2021. View Article : Google Scholar : PubMed/NCBI

19 

Denko NC: Hypoxia, HIF1 and glucose metabolism in the solid tumour. Nat Rev Cancer. 8:705–713. 2008. View Article : Google Scholar : PubMed/NCBI

20 

Iommarini L, Porcelli AM, Gasparre G and Kurelac I: Non-canonical mechanisms regulating hypoxia-inducible factor 1 alpha in cancer. Front Oncol. 7:2862017. View Article : Google Scholar : PubMed/NCBI

21 

Maher JC, Wangpaichitr M, Savaraj N, Kurtoglu M and Lampidis TJ: Hypoxia-inducible factor-1 confers resistance to the glycolytic inhibitor 2-deoxy-D-glucose. Mol Cancer Ther. 6:732–741. 2007. View Article : Google Scholar : PubMed/NCBI

22 

Hsu CC, Tseng LM and Lee HC: Role of mitochondrial dysfunction in cancer progression. Exp Biol Med (Maywood). 241:1281–1295. 2016. View Article : Google Scholar : PubMed/NCBI

23 

Zhang Y, Li Q, Huang Z, Li B, Nice EC, Huang C, Wei L and Zou B: Targeting glucose metabolism enzymes in cancer treatment: current and emerging strategies. Cancers (Basel). 14:45682022. View Article : Google Scholar : PubMed/NCBI

24 

Yu G, Wang LG, Han Y and He QY: clusterProfiler: An R package for comparing biological themes among gene clusters. OMICS. 16:284–287. 2012. View Article : Google Scholar : PubMed/NCBI

25 

R Core Team, . R: A language and environment for statistical computing. R Foundation for Statistical Computing; Vienna: 2024, http://www.R-project.org/

26 

Luo W and Brouwer C: Pathview: An R/Bioconductor package for pathway-based data integration and visualization. Bioinformatics. 29:1830–1831. 2013. View Article : Google Scholar : PubMed/NCBI

27 

Quastel JH and Cantero A: Inhibition of tumour growth by D-glucosamine. Nature. 171:252–254. 1953. View Article : Google Scholar : PubMed/NCBI

28 

Tamayo B, Kercher K, Vosburg C, Massimino C, Jernigan MR, Hasan DL, Harper D, Mathew A, Adkins S, Shippy T, et al: Annotation of glycolysis, gluconeogenesis, and trehaloneogenesis pathways provide insight into carbohydrate metabolism in the Asian citrus psyllid. GigaByte. 2022:gigabyte412022. View Article : Google Scholar : PubMed/NCBI

29 

Suginohara T, Wakabayashi K, Ato S and Ogasawara R: Effect of 2-deoxyglucose-mediated inhibition of glycolysis on the regulation of mTOR signaling and protein synthesis before and after high-intensity muscle contraction. Metabolism. 114:1544192021. View Article : Google Scholar : PubMed/NCBI

30 

Berthe A, Zaffino M, Muller C, Foulquier F, Houdou M, Schulz C, Bost F, De Fay E, Mazerbourg S and Flament S: Protein N-glycosylation alteration and glycolysis inhibition both contribute to the antiproliferative action of 2-deoxyglucose in breast cancer cells. Breast Cancer Res Treat. 171:581–591. 2018. View Article : Google Scholar : PubMed/NCBI

31 

Bertoni JM and Weintraub ST: Competitive inhibition of human brain hexokinase by metrizamide and related compounds. J Neurochem. 42:513–518. 1984. View Article : Google Scholar : PubMed/NCBI

32 

Tesoriere G, Vento R and Calvaruso G: Inhibitory effect of D-glucosamine on glycolysis in bovine retina. Biochim Biophys Acta. 385:58–67. 1975. View Article : Google Scholar : PubMed/NCBI

33 

Marshall S, Yamasaki K and Okuyama R: Glucosamine induces rapid desensitization of glucose transport in isolated adipocytes by increasing GlcN-6-P levels. Biochem Biophys Res Commun. 329:1155–1161. 2005. View Article : Google Scholar : PubMed/NCBI

34 

Samizu M and Iida K: Glucosamine inhibits the proliferation of hepatocellular carcinoma cells by eliciting apoptosis, autophagy, and the anti-warburg effect. Scientifica (Cairo). 2025:56858842025. View Article : Google Scholar : PubMed/NCBI

35 

Jang BC, Sung SH, Park JG, Park JW, Bae JH, Shin DH, Park GY, Han SB and Suh SI: Glucosamine hydrochloride specifically inhibits COX-2 by preventing COX-2 N-glycosylation and by increasing COX-2 protein turnover in a proteasome-dependent manner. J Biol Chem. 282:27622–27632. 2007. View Article : Google Scholar : PubMed/NCBI

36 

Kapoor K, Finer-Moore JS, Pedersen BP, Caboni L, Waight A, Hillig RC, Bringmann P, Heisler I, Müller T, Siebeneicher H and Stroud RM: Mechanism of inhibition of human glucose transporter GLUT1 is conserved between cytochalasin B and phenylalanine amides. Proc Natl Acad Sci USA. 113:4711–4716. 2016. View Article : Google Scholar : PubMed/NCBI

37 

Wu CH, Ho YS, Tsai CY, Wang YJ, Tseng H, Wei PL, Lee CH, Liu RS and Lin SY: In vitro and in vivo study of phloretin-induced apoptosis in human liver cancer cells involving inhibition of type II glucose transporter. Int J Cancer. 124:2210–2219. 2009. View Article : Google Scholar : PubMed/NCBI

38 

Kapoor M, Mineau F, Fahmi H, Pelletier JP and Martel-Pelletier J: Glucosamine sulfate reduces prostaglandin E(2) production in osteoarthritic chondrocytes through inhibition of microsomal PGE synthase-1. J Rheumatol. 39:635–644. 2012. View Article : Google Scholar : PubMed/NCBI

39 

Sols A, De La Fuente G, Villarpalasi C and Asensio C: Substrate specificity and some other properties of baker's yeast hexokinase. Biochim Biophys Acta. 30:92–101. 1958. View Article : Google Scholar : PubMed/NCBI

40 

Brown DH: The phosphorylation of D (+) glucosamine by crystalline yeast hexokinase. Biochim Biophys Acta. 7:487–493. 1951. View Article : Google Scholar : PubMed/NCBI

41 

Kuser PR, Krauchenco S, Antunes OA and Polikarpov I: The high resolution crystal structure of yeast hexokinase PII with the correct primary sequence provides new insights into its mechanism of action. J Biol Chem. 275:20814–20821. 2000. View Article : Google Scholar : PubMed/NCBI

42 

Chen YH and Cheng WH: Hexosamine biosynthesis and related pathways, protein N-glycosylation and O-GlcNAcylation: Their interconnection and role in plants. Front Plant Sci. 15:13490642024. View Article : Google Scholar : PubMed/NCBI

43 

Paneque A, Fortus H, Zheng J, Werlen G and Jacinto E: The hexosamine biosynthesis pathway: Regulation and function. Genes (Basel). 14:9932023. View Article : Google Scholar

44 

Salazar J, Bello L, Chavez M, Anez R, Rojas J and Bermudez V: Glucosamine for osteoarthritis: Biological effects, clinical efficacy, and safety on glucose metabolism. Arthritis. 2014:4324632014. View Article : Google Scholar : PubMed/NCBI

45 

Robey RB and Hay N: Mitochondrial hexokinases, novel mediators of the antiapoptotic effects of growth factors and Akt. Oncogene. 25:4683–4696. 2006. View Article : Google Scholar : PubMed/NCBI

46 

Xian H and Wang Y, Bao X, Zhang H, Wei F, Song Y and Wang Y, Wei Y and Wang Y: Hexokinase inhibitor 2-deoxyglucose coordinates citrullination of vimentin and apoptosis of fibroblast-like synoviocytes by inhibiting HK2/mTORC1-induced autophagy. Int Immunopharmacol. 114:1095562023. View Article : Google Scholar : PubMed/NCBI

47 

Rai Y, Yadav P, Kumari N, Kalra N and Bhatt AN: Hexokinase II inhibition by 3-bromopyruvate sensitizes myeloid leukemic cells K-562 to anti-leukemic drug, daunorubicin. Biosci Rep. 39:BSR201908802019. View Article : Google Scholar : PubMed/NCBI

48 

Thangaraju M, Karunakaran SK, Itagaki S, Gopal E, Elangovan S, Prasad PD and Ganapathy V: Transport by SLC5A8 with subsequent inhibition of histone deacetylase 1 (HDAC1) and HDAC3 underlies the antitumor activity of 3-bromopyruvate. Cancer. 115:4655–4666. 2009. View Article : Google Scholar : PubMed/NCBI

49 

Paul S, Ghosh S and Kumar S: Tumor glycolysis, an essential sweet tooth of tumor cells. Semin Cancer Biol. 86((Pt 3)): 1216–1230. 2022. View Article : Google Scholar : PubMed/NCBI

50 

Bose S, Zhang C and Le A: Glucose metabolism in cancer: The warburg effect and beyond. Adv Exp Med Biol. 1311:3–15. 2021. View Article : Google Scholar : PubMed/NCBI

51 

Zhao J, Ma Y, Zhang Y, Fu B, Wu X, Li Q, Cai G, Chen X and Bai XY: Low-dose 2-deoxyglucose and metformin synergically inhibit proliferation of human polycystic kidney cells by modulating glucose metabolism. Cell Death Discov. 5:762019. View Article : Google Scholar : PubMed/NCBI

52 

Wang Z, Zhang L, Zhang D, Sun R, Wang Q and Liu X: Glycolysis inhibitor 2-deoxy-D-glucose suppresses carcinogen-induced rat hepatocarcinogenesis by restricting cancer cell metabolism. Mol Med Rep. 11:1917–1924. 2015. View Article : Google Scholar : PubMed/NCBI

53 

Medina RA and Owen GI: Glucose transporters: Expression, regulation and cancer. Biol Res. 35:9–26. 2002. View Article : Google Scholar : PubMed/NCBI

54 

Lehar J, Krueger AS, Avery W, Heilbut AM, Johansen LM, Price ER, Rickles RJ, Short GF III, Staunton JE, Jin X, et al: Synergistic drug combinations tend to improve therapeutically relevant selectivity. Nat Biotechnol. 27:659–666. 2009. View Article : Google Scholar : PubMed/NCBI

55 

Sadlecki P, Bodnar M, Grabiec M, Marszalek A, Walentowicz P, Sokup A, Zegarska J and Walentowicz-Sadlecka M: The role of Hypoxia-inducible factor-1 alpha, glucose transporter-1, (GLUT-1) and carbon anhydrase IX in endometrial cancer patients. Biomed Res Int. 2014:6168502014. View Article : Google Scholar : PubMed/NCBI

56 

Wang P, Wan W, Xiong S, Wang J, Zou D, Lan C, Yu S, Liao B, Feng H and Wu N: HIF1α regulates glioma chemosensitivity through the transformation between differentiation and dedifferentiation in various oxygen levels. Sci Rep. 7:79652017. View Article : Google Scholar : PubMed/NCBI

57 

Sharma A, Sinha S and Shrivastava N: Therapeutic targeting hypoxia-inducible factor (HIF-1) in cancer: Cutting gordian knot of cancer cell metabolism. Front Genet. 13:8490402022. View Article : Google Scholar : PubMed/NCBI

58 

Zhou J, Wu Z, Lin Z, Wang W, Wan R and Liu T: Association between glucosamine use and cancer mortality: A large prospective cohort study. Front Nutr. 9:9478182022. View Article : Google Scholar : PubMed/NCBI

59 

Li ZH, Gao X, Chung VC, Zhong WF, Fu Q, Lv YB, Wang ZH, Shen D, Zhang XR, Zhang PD, et al: Associations of regular glucosamine use with all-cause and cause-specific mortality: A large prospective cohort study. Ann Rheum Dis. 79:829–836. 2020. View Article : Google Scholar : PubMed/NCBI

Related Articles

  • Abstract
  • View
  • Download
  • Twitter
Copy and paste a formatted citation
Spandidos Publications style
Park SY, Song K, Kang J and Oh SH: Glucose transporter 2‑transported glucosamine inhibits glycolysis in cancer cell lines through competition with glucose for hexokinase II. Oncol Rep 53: 73, 2025.
APA
Park, S.Y., Song, K., Kang, J., & Oh, S.H. (2025). Glucose transporter 2‑transported glucosamine inhibits glycolysis in cancer cell lines through competition with glucose for hexokinase II. Oncology Reports, 53, 73. https://doi.org/10.3892/or.2025.8906
MLA
Park, S. Y., Song, K., Kang, J., Oh, S. H."Glucose transporter 2‑transported glucosamine inhibits glycolysis in cancer cell lines through competition with glucose for hexokinase II". Oncology Reports 53.6 (2025): 73.
Chicago
Park, S. Y., Song, K., Kang, J., Oh, S. H."Glucose transporter 2‑transported glucosamine inhibits glycolysis in cancer cell lines through competition with glucose for hexokinase II". Oncology Reports 53, no. 6 (2025): 73. https://doi.org/10.3892/or.2025.8906
Copy and paste a formatted citation
x
Spandidos Publications style
Park SY, Song K, Kang J and Oh SH: Glucose transporter 2‑transported glucosamine inhibits glycolysis in cancer cell lines through competition with glucose for hexokinase II. Oncol Rep 53: 73, 2025.
APA
Park, S.Y., Song, K., Kang, J., & Oh, S.H. (2025). Glucose transporter 2‑transported glucosamine inhibits glycolysis in cancer cell lines through competition with glucose for hexokinase II. Oncology Reports, 53, 73. https://doi.org/10.3892/or.2025.8906
MLA
Park, S. Y., Song, K., Kang, J., Oh, S. H."Glucose transporter 2‑transported glucosamine inhibits glycolysis in cancer cell lines through competition with glucose for hexokinase II". Oncology Reports 53.6 (2025): 73.
Chicago
Park, S. Y., Song, K., Kang, J., Oh, S. H."Glucose transporter 2‑transported glucosamine inhibits glycolysis in cancer cell lines through competition with glucose for hexokinase II". Oncology Reports 53, no. 6 (2025): 73. https://doi.org/10.3892/or.2025.8906
Follow us
  • Twitter
  • LinkedIn
  • Facebook
About
  • Spandidos Publications
  • Careers
  • Cookie Policy
  • Privacy Policy
How can we help?
  • Help
  • Live Chat
  • Contact
  • Email to our Support Team