v
Search
Advanced

Publications > Journals > Journal of Clinical and Translational Hepatology> Article Full Text

  • OPEN ACCESS

Oleanolic Acid Restores Drug Sensitivity in Sorafenib-resistant Hepatocellular Carcinoma: Evidence from In Vitro and In Vivo Studies

  • Tongtong Li1,2,
  • Xuan Shen1,2,
  • Tao Zhang1,
  • Jiaheng Ren1,
  • Wang Wang1,
  • Didi Wang1 and
  • Pengxia Zhang1,* 
 Author information 

Abstract

Background and Aims

Hepatocellular carcinoma (HCC) remains challenging to treat in advanced stages, primarily due to the development of resistance to sorafenib. There is an urgent need for novel therapeutic strategies to overcome this resistance. This study aimed to investigate the potential of oleanolic acid (OA), a natural hepatoprotective compound, in mitigating sorafenib resistance and elucidate its underlying molecular mechanisms.

Methods

Sorafenib-resistant Huh7 and HepG2 cell lines were established to mimic the resistant phenotype. The effects of OA on these cells were evaluated by assessing cell invasion, migration, and sensitivity to sorafenib. Gene expression analysis was conducted to identify molecular changes induced by OA treatment, with a focus on fabp3 expression.

Results

Oleanolic acid significantly inhibited the invasive and migratory capabilities of sorafenib-resistant Huh7 and HepG2 cells (p < 0.01). Furthermore, OA treatment downregulated fabp3 expression and restored the cells’ sensitivity to sorafenib.

Conclusions

Oleanolic acid shows promise as an adjunct therapy for overcoming sorafenib resistance in HCC. By reducing cell aggressiveness and restoring drug sensitivity, OA may enhance the therapeutic efficacy of current treatments for advanced HCC.

Graphical Abstract

Keywords

Hepatocellular carcinoma, Sorafenib, Drug resistance, Folic acid, Fatty acid-binding protein 3, Lipid metabolism

Introduction

Hepatocellular carcinoma (HCC) is the most common primary liver cancer, accounting for over 85% of cases and ranking as the fourth leading cause of cancer-related deaths worldwide.1,2 Despite advancements in HCC treatment, prognosis remains poor, particularly in advanced-stage disease.3,4 For HCC caused by hepatitis B virus (HBV) and other etiological factors, surgical resection and liver transplantation are primary treatment options.5 However, many patients are diagnosed at an advanced stage, limiting curative interventions and reducing the five-year survival rate to less than 5%.6–8

Sorafenib, the first FDA-approved systemic therapy for advanced HCC, functions as a multi-targeted tyrosine kinase inhibitor that inhibits tumor growth and angiogenesis.9,10 However, its clinical efficacy is limited by the rapid development of drug resistance.11–13 HCC cells acquire resistance to sorafenib through altered cellular signaling, enhanced drug metabolism, and inhibition of apoptosis.14–16 Overcoming this resistance is critical for improving treatment outcomes.17,18 Therefore, identifying sorafenib sensitizers that counteract resistance mechanisms is an urgent clinical need.

Lipid metabolism plays a key role in HCC progression, as cancer cells reprogram lipid pathways to sustain rapid proliferation.19–21 Fatty acid-binding proteins facilitate lipid transport and metabolism, with fatty acid-binding protein 3 (fabp3) linked to HCC growth, invasion, and metastasis.22–24fabp3 influences cancer cell behavior by directing fatty acid deposition into lipid droplets, which may contribute to sorafenib resistance by mitigating lipid peroxidation-induced cell death.25–28. Additionally, reactive oxygen species (ROS), as metabolic by-products, play a key role in tumor progression and therapy responses.29–31

Oleanolic acid (OA), a natural pentacyclic triterpenoid, exhibits antioxidant, anti-inflammatory, and anticancer properties.32–34 OA has demonstrated anti-tumor effects in various cancers, including HCC.35–37 However, its role in sorafenib resistance and lipid metabolism remains poorly understood. This study investigates OA’s potential to modulate fabp3 expression, thereby influencing lipid metabolism and sorafenib sensitivity in resistant HCC cells. Our findings suggest that OA significantly inhibits migration and invasion in sorafenib-resistant HCC cells, highlighting its potential as a therapeutic adjunct to overcome drug resistance.

This study fills a crucial gap by elucidating OA’s impact on sorafenib resistance in HCC (HBV) through fabp3 modulation. By addressing this resistance mechanism, we aimed to enhance the therapeutic efficacy of sorafenib, offering novel treatment strategies for patients with advanced HCC (HBV) who respond poorly to current therapies.

Methods

Cell culture

The human HCC cell lines HepG2 (derived from hepatoblastoma) and Huh7 (derived from well-differentiated hepatocellular carcinoma, HBV-negative) were obtained from the Chinese Academy of Sciences Cell Bank (Shanghai). The cells were cultured in Dulbecco’s Modified Eagle Medium (G4650, Servicebio, China) supplemented with 10% fetal bovine serum (v/v) and maintained at 37°C with 5% CO2.

Generation of drug-resistant cells

Huh7 and HepG2 cell lines were treated with sorafenib (53468ES70, Yeasen Biotech, Shanghai, China) at a concentration of 10 µM for 48 h. Following the initial treatment, the sorafenib concentration was gradually increased. The cell lines’ resistance to sorafenib was assessed by determining the resistance drug index, with a value of ≥2.5 considered significant.

Transfection and stable cell line generation

The human fabp3 plasmid was cloned into the pcDNA3.1-HA-C vector (QP1397, Qiyunbio, China). The empty plasmid and pcDNA3.1-fabp3 plasmid were transfected into HCC cells using the PEI 40K transfection reagent (G1802, Servicebio, China), following the Lipofectamine 2000 transfection protocol (11668-019, Invitrogen, CA, USA). Transfected cells were used for subsequent experiments after 24 h or 48 h.

To establish stable cell lines, HCC cells were transfected with lentivirus-shfabp3 and lentivirus-shNC (China). Cells were selected with 1 µg/mL puromycin for 14 days. After 12 h of infection, the medium was replaced with fresh culture medium and incubated for 96 h. Cells were then selected with 10 µg/mL puromycin (MABE341, Sigma-Aldrich, USA) for at least one week to establish stable transfected cell lines.

Cell viability assay

Cell viability of Huh7 and HepG2 cells was assessed using the CCK-8 assay (C0037, Beyotime, China). Following treatment, cells (2 × 103) were seeded into 96-well plates and cultured at 37°C with 5% CO2. On days 1, 2, and 3, 10 µL of CCK-8 reagent and 90 µL of serum-free medium were added to each well. After 1 h of incubation, absorbance was measured at 450 nm.

Quantitative real-time polymerase chain reaction (qRT-PCR)

Total RNA was extracted using a universal RNA extraction kit (R219-50, GeneBetter, China) and reverse transcribed using the PrimeScript™ RT reagent kit (Takara, China). The relative quantification of the fabp3 gene was standardized to the GAPDH gene (Sangon Biotech, China). Primer sequences are provided in Supplementary Table 1.

After recording the Ct values for the target gene in each well, the relative expression of the product was calculated using the 2−ΔΔCt method, with GAPDH as the internal control. The formula used was: ΔΔCt = (average Ct value of target gene in experimental group - average Ct value of housekeeping gene in experimental group) - (average Ct value of target gene in control group - average Ct value of housekeeping gene in control group).

Flow cytometry analysis

Treated cells were resuspended in PBS. The Annexin V-FITC Apoptosis Detection Kit (C1062S, Beyotime, China) was utilized to assess apoptotic cell populations. Gating and analysis were conducted using FlowJo v10 software (FlowJo, USA).

Colony formation assay

Cells were seeded at a density of 1 × 104 cells per well in six-well plates and cultured in Dulbecco’s Modified Eagle Medium supplemented with 10% fetal bovine serum and specific concentrations of sorafenib. After one week, visible colonies were formed. Cells were then washed with PBS, fixed with 4% formaldehyde for 15 m, and stained with crystal violet for 25 m.

Transwell assay

Invasion and migration experiments were performed using Transwell chambers with 8.0 µm pore size membranes. The upper chamber was filled with 200 µL of serum-free medium containing suspended cells, while the lower chamber contained 600 µL of medium with 10% fetal bovine serum. For invasion assays, chambers were pre-coated with Matrigel. After 24 h for migration assays and 48 h for invasion assays, non-migratory cells on the top membrane surface were gently washed with PBS. Cells that penetrated the bottom membrane were stained with 1% crystal violet solution and observed under an optical microscope.

Wound healing experiment

Cells were seeded in six-well culture plates and incubated overnight until they reached approximately 95% confluence. The next day, a controlled scratch was made on the cell monolayer using a sterile 1 mL pipette tip, ensuring the desired scratch intensity. After removing cell debris, fresh medium was added for continued cultivation. Wound images were captured at 0, 12, and 24 h.

Antibodies and Western blot (WB)

The cells were lysed, protein extraction was performed using RIPA lysis buffer (KGB5203, Keygen Biotech, Jiangsu, China), and protein concentration was measured using the enhanced BCA protein assay kit (KGB5203, Keygen Biotech, Jiangsu, China). Protein transfer was conducted using SDS-PAGE electrophoresis (G2003, Servicebio, China) and nitrocellulose membranes. Specific primary antibodies included anti-fabp3 (10676-1-AP, 1:1,000, Proteintech, China), anti-AKT (60203-2-Ig, 1:1,000, Proteintech, China), anti-phospho-AKT (28731-1-AP, 1:1,000, Proteintech, China), anti-PI3K (60225-1-Ig, 1:1,000, Proteintech, China), anti-SNAIL (13099-1-AP, 1:1,000, Proteintech, China), anti-E-cadherin (20874-1-AP, 1:1,000, Proteintech, China), anti-N-cadherin (22018-1-AP, 1:1,000, Proteintech, China), anti-PARP1 antibody (Cat# ab191217, Abcam, USA), anti-cleaved PARP1 antibody (Cat# ab32064, Abcam, USA), anti-Caspase-3 antibody (Cat# ab32351, Abcam, USA), anti-cleaved Caspase-3 antibody (Cat# E83-77, Abcam, USA), and β-actin (81115-1-RR, 1:5,000, Proteintech, China), incubated overnight at 4°C. Following primary antibody incubation, rabbit or mouse HRP-conjugated secondary antibodies (KFA025, 1:10,000, Proteintech, China) were applied and incubated at room temperature for two hours. Finally, WB images were detected using ChemistarTM High-sig ECL WB Substrate. Uncropped WB gels are provided in Supplementary File 1.

Lipid analysis

This study employed the common BODIPY 493/503 fluorescence dye, commonly used for neutral lipid staining. Following cell washing with PBS, cells were co-incubated with 4 mg/mL BODIPY 493/503 solution (HY-W090090, MedChemExpress, China) at room temperature for 10 m. Nuclear staining was performed using DAPI (C1002, Beyotime, China) to visualize the cells, and observations were conducted using a fluorescence microscope. Intracellular triglyceride (TG) content was quantitatively measured using a TG assay kit (A110-1-1, Nanjing Jiacheng Bioengineering Institute, Nanjing, China).

Mitochondrial membrane potential analysis using JC-1 staining

JC-1 staining (G1515, Servicebio, China) was performed by incubating cells for 20 m in 200 nM JC-1 medium. The ratio of red fluorescence (indicating high mitochondrial membrane potential) to green fluorescence (representing total mitochondria) was calculated to assess mitochondrial membrane potential.

Detection of ROS

Mitochondrial ROS levels were assessed using DCFH-DA (G1515, Servicebio, China). After specific treatments, cells were washed three times with PBS and incubated with 5 µM DCFH-DA (1:1,000) in Dulbecco’s Modified Eagle Medium for 30 m. Fluorescence signals were recorded using the IBE2000 inverted fluorescence microscope, with an excitation wavelength of 510 nm and an emission wavelength of 580 nm.

Xenograft model

All animal care and experimental procedures were approved by the Animal Ethics Committee of Jiamusi University (2022-500-190). Four- to six-week-old BALB/c nude mice (HFK Bioscience Co., Ltd, Beijing, China) were randomly divided into four groups (N = 10 per group). The first group received a subcutaneous injection of 2 × 106 HepG2-SR cells. The cell suspension was injected subcutaneously into the right axillary region of the mice. The second group received a subcutaneous injection of 2 × 106 HepG2-SR cells and was administered sorafenib orally at 30 mg/kg daily, divided into two doses.

The third group received a subcutaneous injection of 2 × 106 HepG2-SR cells, oral administration of sorafenib at 30 mg/kg daily, divided into two doses, along with nobiletin at 4 mg/kg daily, divided into two doses.38 Nobiletin, a flavonoid compound typically found in the peel of citrus fruits, was dissolved in the aqueous solvent DMSO. The fourth group received a subcutaneous injection of 2 × 106 sh-fabp3-HepG2-SR cells, followed by oral administration of sorafenib at 30 mg/kg daily, divided into two doses.

After two months, the mice were euthanized by cervical dislocation, and tumor samples were subjected to immunohistochemical staining. All animal experiments were conducted following the ARRIVE guidelines.

Immunohistochemistry

Tissue sections were first defatted at 65°C for 1.5 h and dewaxed with xylene. Antigen retrieval was achieved by heating sections in boiling 0.01 M sodium citrate buffer for two minutes. To prevent nonspecific staining, tissue microarrays were blocked with 10% normal goat serum for one hour. Then, the primary antibody, rabbit anti-fabp3 (PA5-13461; 1:2,000; Thermo, USA), was incubated with the sections overnight at 4°C. After three washes with 0.1 M PBS (5 m each), a goat anti-rabbit IgG secondary antibody (ab6721, 1:1,000, Abcam, Cambridge, UK) was applied at 37°C for 20 m, followed by horseradish peroxidase-labeled streptavidin working solution (0343-10000U, Easybio, Beijing, China) for another 20 m at 37°C. Staining was performed using the DAB detection kit (G1212, Servicebio, China), and the sections were counterstained with hematoxylin and sealed with neutral resin.

After staining, tissue sections were mounted and photographed using a Nikon digital microscope camera. The expression levels of the staining results were scored on a proportional scale: the area of immunopositive staining (0%, 0; 1–25%, 1; 26–50%, 2; 51–75%, 3; 76–100%, 4) was multiplied by staining intensity (0 = negative, 1 = weakly positive, 2 = moderately positive, 3 = strongly positive). Two pathologists independently assigned scores. The staining intensity results of different groups were statistically analyzed to calculate the mean and standard deviation.

RNA-seq

Total RNA was first treated with the Ribo-Zero™ Magnetic Kit (MRZH116, Epicentre Technologies, USA) to remove ribosomal RNA. Sequencing libraries were then prepared using the NEB Next Ultra RNA Library Prep Kit (E7770, NEB, USA) according to the manufacturer’s protocol. RNA fragments (∼300 base pairs) were generated using the NEB Next First Strand Synthesis Reaction Buffer (5x). First-strand cDNA was synthesized using reverse transcriptase and random primers, followed by second-strand cDNA synthesis using the Second Strand Synthesis Reaction Buffer (10x) with dUTP Mix.

Subsequently, cDNA fragments were end-repaired, including adding polyA tails and sequencing adapters. After ligation of the Illumina sequencing adapters, the second strand of cDNA was digested using USER Enzyme (M5505S, NEB, USA) to construct strand-specific libraries. The library DNA was then polymerase chain reaction-amplified, purified, and enriched. The libraries were quantified and validated using the Agilent 2100 and the KAPA Library Quantification Kit (TAQDKB, KAPA Biosystems). Finally, paired-end sequencing was performed on the Illumina NextSeq CN500 sequencer.

Raw data was pre-processed for quality using the Trimmomatic tool, which included: (1) removal of adapters; (2) removal of low-quality reads; (3) trimming of low-quality bases from the 3′ and 5′ ends; (4) statistics on raw sequencing volume, effective sequencing volume, Q30, and GC content, with a summary of read counts throughout the quality control process. Filtered high-quality reads were then aligned to the genome in the NCBI database using hisat2 software.

Differential expression analysis of mRNA was performed using the R package “edgeR” based on mRNA read counts. Differential genes were screened with criteria set at |log2FC|>1 and p-value < 0.05.

Gene Ontology/Kyoto Encyclopedia of Genes and Genomes (KEGG) Analysis

Differentially expressed genes were identified using the DESeq tool. After clustering and filtering, Gene Ontology and KEGG pathway analyses were performed. Pathways with a false discovery rate < 0.05 were retained for further analysis.

Data analysis

Statistical analysis and image acquisition were conducted using GraphPad Prism 8. Inter-group differences were assessed using a paired two-tailed Student’s t-test or one-way analysis of variance (ANOVA). All results were based on at least three independent experiments, with qRT-PCR performed in triplicate. Data are presented as mean ± standard deviation. Normality and homogeneity of variance were tested first. For normally distributed data with equal variance, an unpaired t-test was used for two-group comparisons. For multiple group comparisons, one-way ANOVA or repeated measures ANOVA was applied, followed by Tukey’s post hoc test. Repeated measures ANOVA was employed for behavioral experiments, with Bonferroni corrections applied post hoc when appropriate. In omics data, p-values were adjusted using the false discovery rate method. Statistical significance was set at p < 0.05, with the following notations: *p < 0.05, **p < 0.01, ***p < 0.001, ****p < 0.0001, and “ns” (not significant).

Results

OA inhibits the migration and invasion capabilities of sorafenib-resistant HCC cells and enhances the effectiveness of sorafenib

Huh7 and HepG2 cells were chronically exposed to sorafenib, successfully establishing sorafenib-resistant cell lines Huh7-R and HepG2-R. Cell viability assays were performed to determine the IC50 values of sorafenib and OA in both resistant and parental cell lines. The results showed that the IC50 of sorafenib was significantly higher in Huh7-R and HepG2-R compared to their parental counterparts. Similarly, the IC50 of OA was also elevated in the resistant cell lines (Fig. 1A). Further analysis assessed the sensitivity of Huh7-R and HepG2-R to different concentrations of OA (0 µM, 20 µM, 40 µM, 60 µM) and the combinatorial effects of OA with sorafenib (10 µM, 20 µM, 30 µM). The results indicated that at 40 µM OA, sorafenib-resistant cells exhibited enhanced sensitivity to sorafenib, with a significant decrease in cell viability. This suggests that OA exerts an inhibitory effect on resistant cells and acts synergistically with sorafenib (Fig. 1B). This observation was further validated by colony formation assays (Fig. 1C) and flow cytometry analysis (Fig. 1D). Additionally, Transwell and wound healing assays demonstrated that OA significantly inhibited the migration and invasion of sorafenib-resistant cells (Fig. 1E, F).

The effect of OA on the sensitivity of sorafenib-resistant HCC.
Fig. 1  The effect of OA on the sensitivity of sorafenib-resistant HCC.

(A) The half-maximal inhibitory concentration (IC50) under different concentrations of sorafenib over 48 h. (B, C, D) The impact of OA on cell viability, plate cloning, and apoptosis of sorafenib-resistant cells and their parental cells. (E) The migration and invasion abilities of sorafenib-resistant cells and their parental cells after OA treatment. (F) Wound healing assay to measure cell motility of sorafenib-resistant cells and their parental cells post-OA treatment. Cell experiments were repeated three times, N = 3. *p < 0.05; **p < 0.01; ***p < 0.001. OA, oleanolic acid; IC50, half-maximal inhibitory concentration.

OA treatment decreases fabp3 expression, potentially contributing to the development of drug resistance

To explore the mechanisms underlying OA’s effect on resistance development, RNA sequencing was performed on HepG2-R cells, comparing OA-treated (60 µM, two days) and untreated groups. A total of 344 differentially expressed genes were identified, with 211 genes downregulated in the OA-treated group (Fig. 2A). KEGG pathway analysis revealed significant enrichment in ferroptosis, mineral absorption, and PPAR signaling pathways (Fig. 2B, see Supplementary File 2). Among these, six genes (pck2, fabp3, angptl4, plin4, sorbs1, and acadl) were functionally linked within the PPAR signaling pathway and associated with lipid metabolism.

Altered gene expression in sorafenib-resistant HCC regulated by OA.
Fig. 2  Altered gene expression in sorafenib-resistant HCC regulated by OA.

(A) RNA sequencing heatmap. (B) Enrichment of KEGG signaling pathways. (C) Impact of OA treatment on triglyceride levels in sorafenib-resistant cells and their parental cells. (D) Immunofluorescence staining in sorafenib-resistant cells and parental cells post-OA treatment. (E) Expression pattern of fabp3 in normal tissue and HCC tissue. (F) Expression of fabp3 in sorafenib-resistant cells and parental cells. (G) Protein expression of fabp3 in sorafenib-resistant cells and parental cells following OA treatment. All cell experiments were repeated three times, N = 3. *p < 0.05, **p < 0.01, ***p < 0.001. TG, triglyceride; KEGG, Kyoto Encyclopedia of Genes and Genomes.

TG levels were higher in resistant cells than in parental cells and decreased with increasing OA concentration (Fig. 2C). A similar trend was observed using BODIPY 493/503 staining (Fig. 2D). Bioinformatics analysis indicated that fabp3 was highly expressed in HCC tissues, particularly in HBV-related cases, and correlated with poor prognosis (Supplementary Fig. 1A). Based on these observations, fabp3 was selected for further investigation (Fig. 2E). qRT-PCR and Western blot analysis confirmed that fabp3 expression was significantly upregulated in sorafenib-resistant cells compared to parental cells. However, OA treatment reversed this upregulation (Fig. 2F, G). These findings suggest that OA may contribute to reversing sorafenib resistance in hepatocellular carcinoma cells by modulating the PPAR signaling pathway and downregulating fabp3 expression.

fabp3 mitigates sorafenib-induced apoptosis in HCC cells in vitro

Previous studies have identified fabp3 as a specific marker of adipocytes, with its fluorescence primarily localized in the cytoplasm, where lipid droplets (LDs) accumulate.39 Functionally, fabp3 facilitates fatty acid (FA) uptake and storage, promoting triglyceride synthesis and LD accumulation,40,41 which aligns with the findings of this study (Fig. 3A, B). To investigate the potential correlation between fabp3 and sorafenib resistance in HCC, Huh7 and HepG2 cells were transfected with a fabp3 plasmid, and its expression was confirmed by Western blot analysis (Supplementary Fig. 1B). The cells were then treated with Triacsin C (an acyl-CoA synthetase inhibitor that suppresses lipid synthesis), OA (which induces LD formation), or left untreated. Cell viability assays revealed that fabp3 overexpression attenuated sorafenib (10 µM)-induced cell death, an effect that was reversed by Triacsin C. However, OA treatment did not significantly affect the inhibitory action of Triacsin C (Fig. 3C). In hepatic lipid metabolism, triglycerides synthesized from FAs primarily serve as energy reserves in LDs. In this study, OA treatment did not affect cell viability but significantly increased LDs and lipid accumulation. This suggests that fabp3 may enhance sorafenib resistance in HCC by regulating intracellular FA transport rather than directly affecting triglyceride synthesis and LD accumulation. These findings were further supported by colony formation assays and flow cytometry apoptosis assays (Fig. 3D, E).

Impact of <italic>fabp3</italic> overexpression on lipid metabolism and sorafenib sensitivity in HCC cells.
Fig. 3  Impact of fabp3 overexpression on lipid metabolism and sorafenib sensitivity in HCC cells.

(A) Effect of fabp3 on triglyceride levels in Huh-7 and HepG2 cells. (B) Immunofluorescence staining in Huh-7 and HepG2 cells. (C, D, E) Influence of fabp3 on cell viability, plate cloning, and apoptosis in Huh-7 and HepG2 cells. All cell experiments were repeated three times, N = 3. *p < 0.05, **p < 0.01, ***p < 0.001. LDs, lipid droplets; FA, fatty acids.

Inhibiting fabp3 in sorafenib-resistant HCC can restore sensitivity to sorafenib

fabp3 upregulation in sorafenib-resistant cells was assessed using shRNA plasmid interference, with WB analysis confirming the knockdown efficiency (Fig. 1C). Both fabp3 knockdown and Triacsin C treatment significantly reduced TG synthesis and LD accumulation (Fig. 4A, B). Cell viability and colony formation assays demonstrated that shfabp3 knockdown significantly decreased cell survival and colony formation compared to resistant cells, with minimal difference from the Triacsin C-treated group (Fig. 4C, D). Additionally, shfabp3 knockdown induced apoptosis in sorafenib-treated resistant cells, similar to Triacsin C treatment (Fig. 4E). Western blot analysis showed that shfabp3 knockdown increased the protein levels of total and cleaved Caspase-3 and PARP in sorafenib-treated resistant cells, consistent with the trend observed in the Triacsin C-treated group (Fig. 4F).

Impact of <italic>fabp3</italic> interference on sorafenib-resistant HCC cells.
Fig. 4  Impact of fabp3 interference on sorafenib-resistant HCC cells.

(A) Effect of fabp3 interference on triglyceride levels in Huh-7 and HepG2 cells. (B) Immunofluorescence staining in Huh-7 and HepG2 cells. (C, D, E) Influence of fabp3 interference on cell viability, colony formation, and apoptosis in Huh-7 and HepG2 cells. (F) Western blot analysis of PARP, cleaved PARP, Caspase-3, and cleaved Caspase-3 expression in each group. All cell experiments were repeated three times, N = 3. *p < 0.05, **p < 0.01, ***p < 0.001. TG, triglyceride; WB, Western blot.

fabp3 alleviates sorafenib-induced oxidative stress in HCC cells

Following sorafenib treatment, JC-1 monomers (green fluorescence) increased, while JC-1 aggregates (red fluorescence) decreased in Huh7 and HepG2 cells, indicating enhanced mitochondrial depolarization. In resistant cells, JC-1 monomers remained the dominant fluorescent form regardless of sorafenib treatment, suggesting that mitochondrial depolarization is a characteristic feature of sorafenib-resistant cells (Fig. 5A). Further analysis of ROS levels showed that ROS levels in resistant cells were unaffected by sorafenib treatment, whereas OA combined with sorafenib significantly increased ROS-mediated cell death (Fig. 5A). Additionally, OA treatment increased ROS levels in Huh7-R and HepG2-R cells, while fabp3 overexpression suppressed ROS production. However, OA alone did not influence OA-induced ROS generation (Fig. 5B). Consistently, fabp3 overexpression inhibited sorafenib-induced ROS generation, whereas fabp3 knockdown or Triacsin C treatment enhanced ROS production (Fig. 5C).

Effects of sorafenib and OA on mitochondrial function in drug-resistant HCC cells.
Fig. 5  Effects of sorafenib and OA on mitochondrial function in drug-resistant HCC cells.

(A) JC-1 and intracellular ROS levels in sorafenib-resistant cells and their parental cells; JC-1 aggregates (red) indicate normal mitochondrial membrane potential, while JC-1 monomers (green) indicate depolarized mitochondrial membrane potential. (B) Levels of ROS in Huh-7-R and HepG2-R cells. (C) ROS levels in sorafenib-resistant cells and their parental cells. All cell experiments were repeated three times, N = 3. *p < 0.05, **p < 0.01, ***p < 0.001. JC-1, a mitochondrial membrane potential indicator; ROS, reactive oxygen species.

fabp3 promotes the migration and invasion of human HCC by modulating the PI3K/AKT/Snail signaling pathway

Sorafenib resistance is closely associated with increased invasiveness in HCC cells. Given that OA treatment effectively suppressed fabp3 expression and the migratory capacity of resistant cells, fabp3 is likely to promote HCC cell migration and invasion.

Transwell migration and invasion assays revealed that fabp3 overexpression significantly enhanced migration and invasion in HepG2 and Huh7 cells (Fig. 6A), whereas fabp3 knockdown reduced the invasive capacity in HepG2-R and Huh7-R cells (Fig. 6B). Wound healing assays further confirmed that fabp3 overexpression accelerated wound closure, while fabp3 knockdown delayed the process in resistant cells (Fig. 6C,D). However, Triacsin C treatment did not significantly affect fabp3-mediated migration and invasion (Fig. 6A-D), suggesting the involvement of additional regulatory mechanisms.

Regulatory role of <italic>fabp3</italic> on migration and invasion abilities of HCC cells.
Fig. 6  Regulatory role of fabp3 on migration and invasion abilities of HCC cells.

(A, B) Migration and invasion capabilities of Huh-7 and HepG2 cells after transfection with fabp3 plasmid. (C, D) Measurement of cell migration ability based on wound healing assay in sorafenib-resistant cells and their parental cells. (E) Impact of fabp3 overexpression in Huh-7 and HepG2 cells or interference in Huh-7-R and HepG2-R cells on the PI3K/AKT/Snail signaling pathway and EMT markers. All cell experiments were repeated three times, N = 3. **p < 0.01, ***p < 0.001. EMT, epithelial–mesenchymal transition; PI3K, phosphoinositide 3-kinase; AKT, protein kinase B. “+” indicates treatment or gene overexpression/interference was applied; “–” indicates no treatment or vector control.

TCGA data analysis indicated a correlation between fabp3 expression and the PI3K/AKT pathway (Supplementary Fig. 1D). Additionally, treatment with the AKT inhibitor GSK2141795 effectively suppressed migration and invasion induced by fabp3 overexpression or sorafenib resistance (Fig. 6A-D), suggesting that fabp3 promotes HCC cell migration and invasion through the PI3K/AKT pathway.

Western blot analysis further demonstrated that fabp3 overexpression upregulated EMT-related proteins (N-cadherin and Snail) while downregulating E-cadherin. These effects were reversed by GSK2141795 treatment (Fig. 6E). In resistant cells, fabp3 knockdown increased E-cadherin expression and reduced N-cadherin and Snail levels, whereas Triacsin C treatment had no significant effect on EMT markers (Fig. 6E).

In vivo, OA enhances the sensitivity of HCC cells to sorafenib by targeting fabp3

In the HepG2-R xenograft model, the effects of OA and fabp3 on sorafenib resistance in HCC were evaluated. Tumors were excised for analysis 27 days post-implantation (Fig. 7A). Tumor growth monitoring demonstrated that OA or fabp3 knockdown significantly inhibited tumor volume and weight increase, whereas sorafenib treatment alone failed to effectively suppress tumor growth (Fig. 7B,C). Immunohistochemical analysis revealed reduced fabp3 expression in tumors treated with OA or fabp3 knockdown. Additionally, N-cadherin and Snail expression were downregulated, while E-cadherin levels were increased (Fig. 7D), consistent with in vitro findings.

Effects of OA and <italic>fabp3</italic> intervention on xenograft HCC model.
Fig. 7  Effects of OA and fabp3 intervention on xenograft HCC model.

(A) Representative images of tumors excised 27 days post-implantation. (B) Monitoring tumor volume every three days. (C) Tumor weight. (D) Representative immunohistochemical staining images of fabp3, Snail, E-cadherin, and N-cadherin in the four groups. All cell experiments were repeated three times, N = 3. nsp < 0.05, ***p < 0.001. IHC, immunohistochemistry. “+” indicates the presence of drug treatment or gene knockdown; “–” indicates absence of treatment or wild-type control.

Discussion

This study reveals that fabp3 binds and transports fatty acids susceptible to ROS damage to LDs, protecting them from peroxidation and thereby reducing lipotoxicity in sorafenib (SOR)-resistant HCC cells induced by HBV and other pathogenic factors. Sorafenib inhibits tumor growth by increasing intracellular ROS levels27,42,43; however, fabp3, through its fatty acid-binding function, may mitigate ROS toxicity induced by SOR, making it a potential mechanism of SOR resistance.44,45 OA effectively inhibits fabp3 expression, overcoming SOR resistance and inducing apoptosis in HCC cells through ROS-mediated mechanisms.46 These findings provide new insights into HCC resistance mechanisms and emphasize the critical role of lipid metabolism in HCC progression and resistance formation.

fabp3, a fatty acid-binding protein, has been identified as a specific marker of adipocytes and is primarily localized in the cytoplasm, where it is closely associated with lipid droplet accumulation.25,47fabp3 facilitates fatty acid uptake and storage, promoting triglyceride synthesis and lipid droplet formation,40,48,49 which aligns with our study’s findings. Moreover, fabp3 promotes HCC cell migration and invasion through the PI3K/AKT/Snail signaling pathway, enhancing the invasive ability of SOR-resistant cells. Under normal physiological conditions, the PPAR signaling pathway regulates glucose and lipid metabolism to maintain energy homeostasis, but in cancer cells, its abnormal activation promotes tumor growth.50 Previous research indicates that simvastatin restores SOR sensitivity in HCC cells by inhibiting HIF-1α/PPAR-γ/PKM2-mediated glycolysis.51 This study confirms that fabp3 may regulate SOR resistance, migration, and invasion in HCC cells through the PPAR and PI3K/AKT/Snail signaling pathways, further emphasizing the clinical significance of fabp3 as a potential therapeutic target.

Our findings demonstrate that OA inhibits fabp3 expression, restores SOR sensitivity in resistant HCC cells, suppresses cancer cell migration and invasion, and reduces the risk of metastasis.46 Although this study did not directly assess the effects of OA and SOR combination therapy, previous research suggests that OA enhances SOR sensitivity in HCC cells by inhibiting HIF-1α/PPAR-γ/PKM2-mediated glycolysis.51 Thus, OA may serve as an important adjunctive therapy for SOR-resistant patients, warranting further investigation into its combined therapeutic potential with SOR. Additionally, in vitro studies indicate that OA enhances tumor cell radiosensitivity by inhibiting glutathione synthesis.52–54 The efficacy of radiotherapy in HCC is limited by tumor cell antioxidant capacity,55,56 but OA reduces GSH levels, weakening tumor cell antioxidant defenses and enhancing radiotherapy effectiveness. These findings suggest that OA may also be a valuable therapeutic option for radiotherapy-resistant HCC patients, a hypothesis that warrants further investigation.

fabp3 may be a predictive biomarker for SOR resistance, yet its expression levels in real patient samples require further validation. Previous studies have shown that fabp3 influences tumor cell metabolism and is closely associated with treatment response.57–59 Confirmation of fabp3 as a reliable biomarker for SOR resistance requires future studies to analyze fabp3 expression patterns in SOR-resistant patients using TCGA databases and other clinical samples, assessing its correlation with patient prognosis. If fabp3 levels are significantly upregulated in resistant patients, it could serve as a potential biomarker for predicting SOR treatment response in HCC and may help tailor personalized therapeutic strategies.

Although OA has demonstrated anticancer potential in various cancer cell lines,60,61 its application in humans remains challenging due to issues related to absorption, bioavailability, metabolism, dosing, and safety. Currently, there is no standardized recommended dose for OA, as most studies remain at the preclinical or early clinical trial stage.62 OA is primarily metabolized in the liver via the CYP450 system, with its metabolites excreted through urine.63,64 However, its metabolic pathways and clearance rates require further clinical validation. Additionally, OA exhibits lipophilicity, which enhances its bioavailability,62 but its oral absorption may be influenced by dietary intake, necessitating further research to optimize its administration.65 An animal study demonstrated that a single subcutaneous injection of 300 mg/kg OA did not cause significant toxicity,66 but data on human absorption and pharmacokinetics remain limited.

OA is generally considered safe, with no widespread reports of severe adverse effects,67 but long-term use may pose a risk of hepatotoxicity. Given that OA is predominantly metabolized in the liver, future research should focus on its potential impact on liver function (ALT/AST levels) and evaluate its long-term safety.68,69 Further studies should explore the safety of OA at different doses and administration routes and assess its potential synergy with SOR and other standard HCC treatments, ultimately advancing its clinical translation.

Despite providing new insights into the role of OA in reversing HCC resistance, this study has several limitations. First, the study is primarily based on in vitro cell models. Although the xenograft model provides some in vivo evidence, it does not fully replicate the real clinical environment. Future studies should incorporate in vivo animal models to validate the effects of OA in different microenvironments and assess its long-term impact on HCC treatment. Additionally, this study did not directly test the efficacy of OA+SOR combination therapy, although our results indicate that OA restores SOR sensitivity by suppressing fabp3 expression. The exact mechanisms underlying their combined therapeutic effect require further investigation through in vivo combination therapy studies or preclinical models to verify whether OA can enhance SOR efficacy in HCC.

Moreover, this study did not include a diverse range of HCC patient-derived cell lines. The current conclusions are primarily based on the Huh7 and HepG2 cell lines and their resistant sublines, which may not fully represent the biological characteristics of HCC patients. Future studies should incorporate a broader range of HCC cell lines from different pathological subtypes and resistance backgrounds to further validate the generalizability of OA’s effects on fabp3 inhibition. Furthermore, the role of fabp3 as a biomarker for SOR resistance still requires further validation, and future studies should leverage real patient datasets (such as TCGA) to assess fabp3 expression in HCC and its correlation with SOR resistance.

Future research should explore the clinical application of OA, optimizing its dosage, absorption, bioavailability, and long-term safety profile to facilitate its translation into practical treatment. To ensure its efficacy and safety, further investigations should evaluate OA’s impact on the HCC tumor microenvironment and metabolic pathways. Additionally, further validation of fabp3 as a predictive biomarker for SOR resistance is necessary to determine its role in guiding personalized treatment strategies. Ultimately, clinical trials will be crucial in confirming OA’s therapeutic potential in HCC, advancing its transition from experimental research to clinical application.

Conclusions

Our study demonstrates that OA effectively inhibits fabp3-mediated drug resistance of HCC cells to sorafenib. The upregulation of fabp3 in the sorafenib-resistant subset is strongly associated with the ability of HCC cells to evade the lethal effects of sorafenib (Graphical Abstract). Therefore, evaluating fabp3 levels can be useful in predicting sorafenib tolerance, and the use of OA holds promise for developing personalized treatment strategies for HCC patients. Our study has limitations, primarily focusing on HCC cell lines rather than real clinical tumors, as sorafenib is mainly used for late-stage HCC patients with limited surgical options.

Supporting information

Supplementary Table 1

qPCR Primer Sequences.

(DOCX)

Supplementary File 1

Uncropped WB gels.

(DOCX)

Supplementary File 2

KEGG pathway analysis.

(XLSX)

Supplementary Fig. 1

Expression patterns of pck2, fabp3, angptl4, plin4, sorbs1, acadl genes in HCC tissues.

(A) Relevance of the six genes with HCC demonstrated through survival and differential expression analyses, respectively presented with Kaplan-Meier survival curves (left panel) and Wilcoxon test comparing expression levels in tumor and normal tissues (right panel). (B) Validation of fabp3 plasmid transfection effect on HepG2 and Huh7 cell lines using WB. Comparison between empty vector group (−) and transfected group (+) demonstrates protein expression levels. (C) Verification of interference effect after fabp3 knockdown using shRNA through WB analysis in HepG2-R and Huh7-R sorafenib-resistant cell lines. Comparison between empty vector group (−) and interference group (+) shows protein expression levels. (D) Spearman correlation analysis between FABP3 gene expression and pathway scores in liver cancer samples. In the figure, the X-axis represents the expression level of the gene, while the Y-axis represents the pathway score.The density curve on the right side indicates the distribution trend of the pathway score; the density curve on the top shows the distribution of gene expression.The values at the top (blue curve within the coordinate axes) indicate the p-value, correlation coefficient, and the correlation method used (Spearman) for the association between the single gene and the pathway score. All cell experiments were repeated three times. N = 3. Data were analyzed using independent sample t-tests between two groups, with statistical significance levels indicated as *p < 0.05, **p < 0.01, ***p < 0.001, ****p < 0.0001.

(JPG)

Declarations

Ethical statement

All animal experiments were approved by the Animal Ethics Committee of Jiamusi University (2022-500-190). All animals received human care.

Data sharing statement

The raw data supporting the conclusions of this article will be made available by the authors upon reasonable request.

Funding

This study was supported by the Horizontal Project of Jiangsu Medical College (2021010401), the collaborative innovation project for “Double First-Class” Discipline Heilongjiang Province (LJGXCG2023-089), the Joint Guidance Project of Natural Science Foundation of Heilongjiang Province (LH2022H090, PL2024H002), the Innovation Team of Heilongjiang Provincial Department of Education (2024-KYYWF-0617),and the he East-Extreme Discipline Construction Team of Jiamusi University (DJXSTD202404).

Conflict of interest

The authors have no conflict of interests related to this publication.

Authors’ contributions

Study design, experiment performance, data analysis (TL, XS), establishment of sorafenib-resistant cell lines, molecular analyses (TZ, JR), data interpretation, statistical analysis (WW, DW), supervision, critical insights, and revision (PZ). All authors reviewed and approved the final version of the manuscript.

References

  1. Ganesan P, Kulik LM. Hepatocellular Carcinoma: New Developments. Clin Liver Dis 2023;27(1):85-102 View Article PubMed/NCBI
  2. Bruni A, Castellana C, Dajti E, Barbara G, Marasco G, Maida M, et al. Epidemiological, diagnostic, therapeutic and prognostic impact of hepatitis B and D virus infection on hepatocellular carcinoma: A review of the literature. Virology 2024;600:110273 View Article PubMed/NCBI
  3. Abdelhamed W, El-Kassas M. Hepatitis B virus as a risk factor for hepatocellular carcinoma: There is still much work to do. Liver Res 2024;8(2):83-90 View Article PubMed/NCBI
  4. Hu X, Wen Y. The hepatic steatosis, steatohepatitis and metabolic dysfunction: distinct roles in hepatocellular carcinoma occurrence in chronic hepatitis B patients. Hepatol Int 2024;18(4):1338-1339 View Article PubMed/NCBI
  5. Yang X, Wang H, Yu C. The Mechanism of APOBEC3B in Hepatitis B Virus Infection and HBV Related Hepatocellular Carcinoma Progression, Therapeutic and Prognostic Potential. Infect Drug Resist 2024;17:4477-4486 View Article PubMed/NCBI
  6. Zhu D, Tulahong A, Abuduhelili A, Liu C, Aierken A, Lin Y, et al. Machine Learning Prognostic Model for Post-Radical Resection Hepatocellular Carcinoma in Hepatitis B Patients. J Hepatocell Carcinoma 2025;12:353-365 View Article PubMed/NCBI
  7. Hwang H, Kim J, Kim TH, Han Y, Choi D, Cho S, et al. Exosomal miR-6126 as a novel therapeutic target for overcoming resistance of anti-cancer effect in hepatocellular carcinoma. BMC Cancer 2024;24(1):1557 View Article PubMed/NCBI
  8. Wei H, Dong C, Li X. Treatment Options for Hepatocellular Carcinoma Using Immunotherapy: Present and Future. J Clin Transl Hepatol 2024;12(4):389-405 View Article PubMed/NCBI
  9. Lai HC, Chung WM, Chang CM, Liao PY, Su YT, Yeh CC, et al. Androgen Receptor Enhances the Efficacy of Sorafenib Against Hepatocellular Carcinoma Through Enriched EpCAM Stemness. Anticancer Res 2020;40(3):1285-1295 View Article PubMed/NCBI
  10. Chava S, Ekmen N, Ferraris P, Aydin Y, Moroz K, Wu T, et al. Mechanisms of Sorafenib Resistance in HCC Culture Relate to the Impaired Membrane Expression of Organic Cation Transporter 1 (OCT1). J Hepatocell Carcinoma 2024;11:839-855 View Article PubMed/NCBI
  11. Zheng Y, Ye S, Huang S, Cheng Y, Zhang Y, Leng Y, et al. Lefamulin Overcomes Acquired Drug Resistance via Regulating Mitochondrial Homeostasis by Targeting ILF3 in Hepatocellular Carcinoma. Adv Sci (Weinh) 2024;11(30):e2401789 View Article PubMed/NCBI
  12. Mir IH, Jyothi KC, Thirunavukkarasu C. The prominence of potential biomarkers in the diagnosis and management of hepatocellular carcinoma: Current scenario and future anticipation. J Cell Biochem 2022;123(10):1607-1623 View Article PubMed/NCBI
  13. Liao S, Wang Q, Chen S, Huang Q, Zhou L, Liu H, et al. Mito-LND and (E)-Akt inhibitor-IV: novel compounds inducing endoplasmic reticulum stress and ROS accumulation against hepatocellular carcinoma. J Transl Med 2024;22(1):792 View Article PubMed/NCBI
  14. Zheng R, Weng S, Xu J, Li Z, Wang Y, Aizimuaji Z, et al. Autophagy and biotransformation affect sorafenib resistance in hepatocellular carcinoma. Comput Struct Biotechnol J 2023;21:3564-3574 View Article PubMed/NCBI
  15. Shueng PW, Chan HW, Lin WC, Kuo DY, Chuang HY. Orlistat Resensitizes Sorafenib-Resistance in Hepatocellular Carcinoma Cells through Modulating Metabolism. Int J Mol Sci 2022;23(12):6501 View Article PubMed/NCBI
  16. Sun T, Liu H, Ming L. Multiple Roles of Autophagy in the Sorafenib Resistance of Hepatocellular Carcinoma. Cell Physiol Biochem 2017;44(2):716-727 View Article PubMed/NCBI
  17. Marin JJG, Macias RIR, Monte MJ, Romero MR, Asensio M, Sanchez-Martin A, et al. Molecular Bases of Drug Resistance in Hepatocellular Carcinoma. Cancers (Basel) 2020;12(6):1663 View Article PubMed/NCBI
  18. Kong FH, Ye QF, Miao XY, Liu X, Huang SQ, Xiong L, et al. Current status of sorafenib nanoparticle delivery systems in the treatment of hepatocellular carcinoma. Theranostics 2021;11(11):5464-5490 View Article PubMed/NCBI
  19. Du D, Liu C, Qin M, Zhang X, Xi T, Yuan S, et al. Metabolic dysregulation and emerging therapeutical targets for hepatocellular carcinoma. Acta Pharm Sin B 2022;12(2):558-580 View Article PubMed/NCBI
  20. Xu X, Poulsen KL, Wu L, Liu S, Miyata T, Song Q, et al. Targeted therapeutics and novel signaling pathways in non-alcohol-associated fatty liver/steatohepatitis (NAFL/NASH). Signal Transduct Target Ther 2022;7(1):287 View Article PubMed/NCBI
  21. Ockner RK, Kaikaus RM, Bass NM. Fatty-acid metabolism and the pathogenesis of hepatocellular carcinoma: review and hypothesis. Hepatology 1993;18(3):669-676 PubMed/NCBI
  22. Amiri M, Yousefnia S, Seyed Forootan F, Peymani M, Ghaedi K, Nasr Esfahani MH. Diverse roles of fatty acid binding proteins (FABPs) in development and pathogenesis of cancers. Gene 2018;676:171-183 View Article PubMed/NCBI
  23. Agirre-Lizaso A, Huici-Izagirre M, Urretabizkaia-Garmendia J, Rodrigues PM, Banales JM, Perugorria MJ. Targeting the Heterogeneous Tumour-Associated Macrophages in Hepatocellular Carcinoma. Cancers (Basel) 2023;15(20):4977 View Article PubMed/NCBI
  24. Chen Y, Zhu Y, Dong Y, Li H, Gao C, Zhu G, et al. A pyroptosis-related gene signature for prognosis prediction in hepatocellular carcinoma. Front Oncol 2023;13:1085188 View Article PubMed/NCBI
  25. Liu RZ, Godbout R. An Amplified Fatty Acid-Binding Protein Gene Cluster in Prostate Cancer: Emerging Roles in Lipid Metabolism and Metastasis. Cancers (Basel) 2020;12(12):3823 View Article PubMed/NCBI
  26. Petan T, Jarc E, Jusović M. Lipid Droplets in Cancer: Guardians of Fat in a Stressful World. Molecules 2018;23(8):1941 View Article PubMed/NCBI
  27. Wei L, Lee D, Law CT, Zhang MS, Shen J, Chin DW, et al. Genome-wide CRISPR/Cas9 library screening identified PHGDH as a critical driver for Sorafenib resistance in HCC. Nat Commun 2019;10(1):4681 View Article PubMed/NCBI
  28. Li Y, Yang W, Zheng Y, Dai W, Ji J, Wu L, et al. Targeting fatty acid synthase modulates sensitivity of hepatocellular carcinoma to sorafenib via ferroptosis. J Exp Clin Cancer Res 2023;42(1):6 View Article PubMed/NCBI
  29. Liese J, Hinrichs TM, Lange M, Fulda S. Cotreatment with sorafenib and oleanolic acid induces reactive oxygen species-dependent and mitochondrial-mediated apoptotic cell death in hepatocellular carcinoma cells. Anticancer Drugs 2019;30(3):209-217 View Article PubMed/NCBI
  30. Lange M, Abhari BA, Hinrichs TM, Fulda S, Liese J. Identification of a novel oxidative stress induced cell death by Sorafenib and oleanolic acid in human hepatocellular carcinoma cells. Biochem Pharmacol 2016;118:9-17 View Article PubMed/NCBI
  31. Mishra JS, Zhao H, Hattis S, Kumar S. Elevated Glucose and Insulin Levels Decrease DHA Transfer across Human Trophoblasts via SIRT1-Dependent Mechanism. Nutrients 2020;12(5):1271 View Article PubMed/NCBI
  32. Chen C, Ai Q, Shi A, Wang N, Wang L, Wei Y. Oleanolic acid and ursolic acid: therapeutic potential in neurodegenerative diseases, neuropsychiatric diseases and other brain disorders. Nutr Neurosci 2023;26(5):414-428 View Article PubMed/NCBI
  33. Wang Y, Zhai J, Yang D, Han N, Liu Z, Liu Z, et al. Antioxidant, Anti-Inflammatory, and Antidiabetic Activities of Bioactive Compounds from the Fruits of Livistona chinensis Based on Network Pharmacology Prediction. Oxid Med Cell Longev 2021;2021:7807046 View Article PubMed/NCBI
  34. Bai SR, Zhao BX, Zhao Q, Ge YC, Li M, Zhao CG, et al. Oleanolic acid improves 5-fluorouracil-induced intestinal damage and inflammation by alleviating intestinal senescence. Sci Rep 2024;14(1):21852 View Article PubMed/NCBI
  35. Wiemann J, Heller L, Csuk R. Targeting cancer cells with oleanolic and ursolic acid derived hydroxamates. Bioorg Med Chem Lett 2016;26(3):907-909 View Article PubMed/NCBI
  36. Gu Z, Deng M, Luo Q, Lin S, Yu J, Li C, et al. Harnessing HDAC-targeted oleanolic acid derivatives for combined anti-cancer and hepatoprotective effects. Int J Biol Macromol 2025;297:139761 View Article PubMed/NCBI
  37. Wang X, Bai H, Zhang X, Liu J, Cao P, Liao N, et al. Inhibitory effect of oleanolic acid on hepatocellular carcinoma via ERK-p53-mediated cell cycle arrest and mitochondrial-dependent apoptosis. Carcinogenesis 2013;34(6):1323-1330 View Article PubMed/NCBI
  38. Abdelgalil AA, Alkahtani HM, Al-Jenoobi FI. Sorafenib. Profiles Drug Subst Excip Relat Methodol 2019;44:239-266 View Article PubMed/NCBI
  39. Xu Y, Mirmalek-Sani SH, Yang X, Zhang J, Oreffo RO. The use of small interfering RNAs to inhibit adipocyte differentiation in human preadipocytes and fetal-femur-derived mesenchymal cells. Exp Cell Res 2006;312(10):1856-1864 View Article PubMed/NCBI
  40. Pascual G, Domínguez D, Benitah SA. The contributions of cancer cell metabolism to metastasis. Dis Model Mech 2018;11(8):dmm032920 View Article PubMed/NCBI
  41. Li X, Li Y, Lu W, Chen M, Ye W, Zhang D. The Tumor Vessel Targeting Strategy: A Double-Edged Sword in Tumor Metastasis. Cells 2019;8(12):1602 View Article PubMed/NCBI
  42. Arlauckas SP, Popov AV, Delikatny EJ. Choline kinase alpha-Putting the ChoK-hold on tumor metabolism. Prog Lipid Res 2016;63:28-40 View Article PubMed/NCBI
  43. Yu L, Wang M, Yang Y, Xu F, Zhang X, Xie F, et al. Predicting therapeutic drugs for hepatocellular carcinoma based on tissue-specific pathways. PLoS Comput Biol 2021;17(2):e1008696 View Article PubMed/NCBI
  44. Singh MK, Han S, Kim S, Kang I. Targeting Lipid Metabolism in Cancer Stem Cells for Anticancer Treatment. Int J Mol Sci 2024;25(20):11185 View Article PubMed/NCBI
  45. Zhang H, Chen N, Ding C, Zhang H, Liu D, Liu S. Ferroptosis and EMT resistance in cancer: a comprehensive review of the interplay. Front Oncol 2024;14:1344290 View Article PubMed/NCBI
  46. Shan C, Liang Y, Wang K, Li P. Noncoding RNAs in cancer ferroptosis: From biology to clinical opportunity. Biomed Pharmacother 2023;165:115053 View Article PubMed/NCBI
  47. Lee MJ, Jash S, Jones JEC, Puri V, Fried SK. Rosiglitazone remodels the lipid droplet and britens human visceral and subcutaneous adipocytes ex vivo. J Lipid Res 2019;60(4):856-868 View Article PubMed/NCBI
  48. Zhang JF, Choi SH, Li Q, Wang Y, Sun B, Tang L, et al. Overexpression of DGAT2 Stimulates Lipid Droplet Formation and Triacylglycerol Accumulation in Bovine Satellite Cells. Animals (Basel) 2022;12(14):1847 View Article PubMed/NCBI
  49. Guo P, Yao X, Jin X, Xv Y, Zhang J, Li Q, et al. Interference with DGAT Gene Inhibited TAG Accumulation and Lipid Droplet Synthesis in Bovine Preadipocytes. Animals (Basel) 2023;13(13):2223 View Article PubMed/NCBI
  50. Panasyuk G, Espeillac C, Chauvin C, Pradelli LA, Horie Y, Suzuki A, et al. PPARγ contributes to PKM2 and HK2 expression in fatty liver. Nat Commun 2012;3:672 View Article PubMed/NCBI
  51. Feng J, Dai W, Mao Y, Wu L, Li J, Chen K, et al. Simvastatin re-sensitizes hepatocellular carcinoma cells to sorafenib by inhibiting HIF-1α/PPAR-γ/PKM2-mediated glycolysis. J Exp Clin Cancer Res 2020;39(1):24 View Article PubMed/NCBI
  52. Sappington DR, Siegel ER, Hiatt G, Desai A, Penney RB, Jamshidi-Parsian A, et al. Glutamine drives glutathione synthesis and contributes to radiation sensitivity of A549 and H460 lung cancer cell lines. Biochim Biophys Acta 2016;1860(4):836-843 View Article PubMed/NCBI
  53. Li X, Zhang Y, Liu A, Li L, Yang X, Wang Y, et al. Nanozyme as tumor energy homeostasis disruptor mediated ferroptosis for high-efficiency radiotherapy. J Colloid Interface Sci 2025;688:44-58 View Article PubMed/NCBI
  54. Singh B, Patwardhan RS, Jayakumar S, Sharma D, Sandur SK. Oxidative stress associated metabolic adaptations regulate radioresistance in human lung cancer cells. J Photochem Photobiol B 2020;213:112080 View Article PubMed/NCBI
  55. Liu R, Bian Y, Liu L, Liu L, Liu X, Ma S. Molecular pathways associated with oxidative stress and their potential applications in radiotherapy (Review). Int J Mol Med 2022;49(5):65 View Article PubMed/NCBI
  56. Hanot M, Boivin A, Malésys C, Beuve M, Colliaux A, Foray N, et al. Glutathione depletion and carbon ion radiation potentiate clustered DNA lesions, cell death and prevent chromosomal changes in cancer cells progeny. PLoS One 2012;7(11):e44367 View Article PubMed/NCBI
  57. Deng Z, Wang J, Xu B, Jin Z, Wu G, Zeng J, et al. Mining TCGA Database for Tumor Microenvironment-Related Genes of Prognostic Value in Hepatocellular Carcinoma. Biomed Res Int 2019;2019:2408348 View Article PubMed/NCBI
  58. Seo J, Jeong DW, Park JW, Lee KW, Fukuda J, Chun YS. Fatty-acid-induced FABP5/HIF-1 reprograms lipid metabolism and enhances the proliferation of liver cancer cells. Commun Biol 2020;3(1):638 View Article PubMed/NCBI
  59. Seo J, Yun JE, Kim SJ, Chun YS. Lipid metabolic reprogramming by hypoxia-inducible factor-1 in the hypoxic tumour microenvironment. Pflugers Arch 2022;474(6):591-601 View Article PubMed/NCBI
  60. Tang ZY, Li Y, Tang YT, Ma XD, Tang ZY. Anticancer activity of oleanolic acid and its derivatives: Recent advances in evidence, target profiling and mechanisms of action. Biomed Pharmacother 2022;145:112397 View Article PubMed/NCBI
  61. D’Mello RS, Mendon V, Pai P, Das I, Sundara BK. Exploring the therapeutic potential of oleanolic acid and its derivatives in cancer treatment: a comprehensive review. 3 Biotech 2025;15(3):56 View Article PubMed/NCBI
  62. Staicu V, Tomescu JA, Calinescu I. Bioavailability of ursolic/oleanolic acid, with therapeutic potential in chronic diseases and cancer. Adv Clin Exp Med 2024;33(11):1173-1178 View Article PubMed/NCBI
  63. Verma N, Raghuvanshi DS, Singh RV. Recent advances in the chemistry and biology of oleanolic acid and its derivatives. Eur J Med Chem 2024;276:116619 View Article PubMed/NCBI
  64. Castellano JM, Ramos-Romero S, Perona JS. Oleanolic Acid: Extraction, Characterization and Biological Activity. Nutrients 2022;14(3):623 View Article PubMed/NCBI
  65. Pan D, Qu Y, Shi C, Xu C, Zhang J, Du H, et al. Oleanolic acid and its analogues: promising therapeutics for kidney disease. Chin Med 2024;19(1):74 View Article PubMed/NCBI
  66. Yuan M, Wan Y, Wang Y, Li S, Tang J, Liang X, et al. Ursodeoxycholic acid grafted chitosan oligosaccharide self-assembled micelles with enhanced oral absorption and antidiabetic effect of oleanolic acid. Food Chem 2025;470:142708 View Article PubMed/NCBI
  67. Chen P, Zeng H, Wang Y, Fan X, Xu C, Deng R, et al. Low dose of oleanolic acid protects against lithocholic acid-induced cholestasis in mice: potential involvement of nuclear factor-E2-related factor 2-mediated upregulation of multidrug resistance-associated proteins. Drug Metab Dispos 2014;42(5):844-852 View Article PubMed/NCBI
  68. Liu J, Lu YF, Wu Q, Xu SF, Shi FG, Klaassen CD. Oleanolic acid reprograms the liver to protect against hepatotoxicants, but is hepatotoxic at high doses. Liver Int 2019;39(3):427-439 View Article PubMed/NCBI
  69. Liao S, Fu X, Huang J, Wang Y, Lu Y, Zhou S. Suppression of SIRT1/FXR signaling pathway contributes to oleanolic acid-induced liver injury. Toxicol Appl Pharmacol 2023;467:116509 View Article PubMed/NCBI

About this Article

Cite this article
Li T, Shen X, Zhang T, Ren J, Wang W, Wang D, et al. Oleanolic Acid Restores Drug Sensitivity in Sorafenib-resistant Hepatocellular Carcinoma: Evidence from In Vitro and In Vivo Studies. J Clin Transl Hepatol. Published online: Apr 18, 2025. doi: 10.14218/JCTH.2024.00369.
Copy        Export to RIS        Export to EndNote
Article History
Received Revised Accepted Published
October 11, 2024 March 19, 2025 March 25, 2025 April 18, 2025
DOI http://dx.doi.org/10.14218/JCTH.2024.00369
  • Journal of Clinical and Translational Hepatology
  • pISSN 2225-0719
  • eISSN 2310-8819
Back to Top

Oleanolic Acid Restores Drug Sensitivity in Sorafenib-resistant Hepatocellular Carcinoma: Evidence from In Vitro and In Vivo Studies

Tongtong Li, Xuan Shen, Tao Zhang, Jiaheng Ren, Wang Wang, Didi Wang, Pengxia Zhang
  • Reset Zoom
  • Download TIFF