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 Letters
Join Editorial Board Propose a Special Issue
Print ISSN: 1792-1074 Online ISSN: 1792-1082
Journal Cover
November-2025 Volume 30 Issue 5

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
November-2025 Volume 30 Issue 5

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.xlsx
    • Supplementary_Data3.pdf
Article Open Access

Machine learning‑based model identifies a novel cuproptosis‑related mitochondrial gene signature with a key role in the prognosis and treatment of lung adenocarcinoma

  • Authors:
    • Yi-Hao Liu
    • Wen-Hao Zhao
    • Ze-Xia Zhao
    • Zhi-Xuan Duan
    • Hua Huang
    • Cheng Ding
    • Ming-Hui Liu
    • Hong-Bing Zhang
    • Yong-Wen Li
    • Min Wang
    • Jun Chen
    • Hong-Lin Zhao
  • View Affiliations / Copyright

    Affiliations: Department of Lung Cancer Surgery, Tianjin Medical University General Hospital, Tianjin 300052, P.R. China
    Copyright: © Liu et al. This is an open access article distributed under the terms of Creative Commons Attribution License.
  • Article Number: 494
    |
    Published online on: August 21, 2025
       https://doi.org/10.3892/ol.2025.15240
  • 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

Lung adenocarcinoma (LUAD) remains one of the most prevalent and lethal cancers globally, making it critical to understand the mechanisms driving its progression and improve prognosis. Moreover, cuproptosis and mitochondrial dysfunction may be involved in lung cancer. Therefore, the present study aimed to identify mitochondrial genes associated with cuproptosis to develop a prognostic model for patients with LUAD, with the potential to predict survival outcomes and responses to treatment. Bulk RNA‑sequencing data was utilized from The Cancer Genome Atlas and the Gene Expression Omnibus (GEO), and Pearson correlation analysis was employed to identify mitochondrial genes associated with cuproptosis. A prognostic model was constructed using univariate Cox regression combined with least absolute shrinkage and selection operator analysis, and a nomogram was developed to predict survival with clinical relevance. The accuracy of the model was evaluated using two independent GEO datasets. Additionally, the clinical value of the risk score model was assessed using immune infiltration analysis, tumor mutational burden and drug sensitivity predictions. Furthermore, the effects of superoxide dismutase 2 (SOD2) gene knockdown on tumor metastasis and proliferation were experimentally evaluated. A set of 22 mitochondrial genes associated with cuproptosis were identified: Metabolism of cobalamin associated D, SOD2, human immunodeficiency virus‑1 Tat interactive protein 2, cytochrome C somatic, mitochondrial pyruvate carrier 1, adenylate kinase 2, mitochondrial ribosomal protein L44, transforming growth factor β regulator 4, mitochondrial transcription factor A, tetratricopeptide repeat domain 19, coiled‑coil‑helix‑­coiled‑coil‑helix domain containing 4, sideroflexin 1, ATP binding cassette subfamily D member 1, NADH:ubiquinone oxidoreductase complex assembly factor 7, NOP2/Sun RNA methyltransferase 4, NME/NM23 nucleoside diphosphate kinase 6, X‑Prolyl aminopeptidase 3, lipoyltransferase 1, mitochondrial methionyl aminopeptidase type 1D, carbonic anhydrase 5B, kynurenine 3‑monooxygenase and alcohol dehydrogenase iron containing 1. The model was validated as an independent predictor of overall survival, dividing patients into high‑ and low‑risk groups. Immune infiltration analysis revealed that tumors in the low‑risk group displayed more active immune responses and improved immune function. Drug sensitivity analysis suggested that high‑risk patients may be more responsive to specific drug treatments. Finally, knockdown of the SOD2 gene suppressed tumor cell metabolism, proliferation and metastasis. In conclusion, the present study successfully established a prognostic model based on cuproptosis‑related mitochondrial genes and developed a nomogram to predict LUAD prognosis with high accuracy, thereby providing improved tools for treatment decision‑making and enhancing patient outcomes.

Introduction

Non-small cell lung cancer (NSCLC) is primarily divided into two major subtypes: Lung adenocarcinoma (LUAD) and lung squamous cell carcinoma (LUSC). Among these, LUAD is the more common form, representing 40–50% of NSCLC cases. The incidence of LUAD has increased significantly, with prevalence rates reaching 39% in males and 57% in females in 2020, according to a global population-based study (1), particularly among non-smokers (2). By contrast, LUSC, which is associated with smoking, accounts for 25–30% of NSCLC cases (3), and its occurrence has decreased in line with lower smoking rates (4). For early-stage LUAD (stages I and II), surgery remains the primary treatment modality, whereas patients with EGFR mutations may benefit from targeted therapies (5). However, for more advanced stages (III and IV), LUAD typically requires a combination of chemotherapy, targeted therapy and immunotherapy (6,7). Furthermore, the 5-year survival rate for patients with advanced LUAD is generally <30%, with certain cases at <5% (8). This highlights the need for the development of genetic signatures that can guide treatment decisions and predict patient outcomes.

Mitochondrial genes associated with cuproptosis are key factors in the progression of LUAD, making them potential prognostic biomarkers (9). Cuproptosis is a form of programmed cell death (PCD) induced by the accumulation of copper ions (Cu2+), and there is an increased demand for Cu2+ during tumor development and metastasis (10). Excessive copper accumulation leads to oxidative stress, lipid peroxidation and damage to essential cellular components, ultimately resulting in cell death. However, tumor cells require metal ions for metabolic processes, and targeting metal ions presents notable potential for future therapeutic research (11). A total of three primary mechanisms underlie cuproptosis: i) Oxidative stress, where copper ions catalyze the Fenton reaction, generating reactive oxygen species (ROS) that damage cell membranes, proteins and DNA; ii) mitochondrial dysfunction, where excess copper ions impair mitochondrial function, reduce ATP production and release cytokines, triggering cell death; and iii) lipid peroxidation, where copper ions induce lipid peroxidation, disrupting the cell membrane and leading to cell death (12). Mitochondria and their associated genes serve pivotal roles in cellular energy metabolism and apoptosis, impacting tumor cell proliferation, migration and invasion (13). Moreover, mitochondria are crucial for LUAD progression, particularly under hypoxic conditions where they support rapid tumor growth through mitochondrial glycolysis, a phenomenon known as the Warburg effect (14). In LUAD cells, mitochondrial dysfunction enhances resistance to oxidative stress and disrupts mitochondrial membrane integrity, which in turn affects the balance of Bcl-2 family proteins and facilitates evasion of apoptosis, ultimately contributing to an uncontrolled cell cycle (15). Cuproptosis influences mitochondrial membrane permeability and electron transport chain activity, altering oxidative stress tolerance. Additionally, copper-induced cell death leads to mitochondrial dysfunction by binding copper ions to specific mitochondrial proteins, such as fatty acid synthase and proteins involved in the tricarboxylic acid cycle (16). Therefore, both cuproptosis and mitochondrial function serve central roles in the invasion, metastasis and immune evasion of LUAD cells, markedly influencing patient prognosis.

Despite the aforementioned findings, no studies to date have explored the prognostic impact of cuproptosis-related mitochondrial genes in LUAD, to the best of our knowledge. Therefore, the present study aimed to identify such genes to construct a prognostic model, and to further assess the prognostic value of these genes in patients with LUAD by analyzing immune cell infiltration, tumor gene mutations and drug sensitivity.

Materials and methods

Data acquisition

LUAD sample data were obtained from The Cancer Genome Atlas (TCGA) database (https://portal.gdc.cancer.gov), which provides gene expression profiles and clinical data for 489 patients with LUAD and 59 adjacent normal tissue samples (17). To retrieve TCGA-LUAD dataset, the official TCGA website was accessed and the data repository page was entered by selecting the ‘Repository’ tab and clicking on ‘Explore Our Cancer Datasets’. Using the ‘Cohort Builder’, the ‘Program’ field was expanded, and TCGA and LUAD were selected. Further filtering was then performed by re-entering the ‘Repository’ section and selecting ‘RNA-Seq’, ‘STAR Counts’ and ‘Open access’ under the ‘Experimental Strategy’ filter. All relevant files were subsequently added to the cart and downloaded using the ‘Download’ option. After acquisition, data cleaning and preprocessing were performed, including gene symbol conversion. For duplicate genes, the average expression value was calculated, and the data were transformed from Fragments Per Kilobase of exon Model per Million Mapped Fragments (FPKM) to transcripts per kilobase million followed by log2 transformation. The final dataset included gene expression profiles for >50,000 genes and comprehensive clinical information for 489 patients with LUAD, which included sex, age, survival time, survival status, clinical stage and tumor (T)-node (N)-metastasis (M) stage. The GSE26939, GSE31210 and GSE72094 datasets were also retrieved from the Gene Expression Omnibus (GEO; http://www.ncbi.nih.gov/geo), which provide gene expression profiles and clinical survival data for 115, 226 and 398 patients, respectively. TCGA-LUAD dataset was used as the primary training set, whilst the three GEO datasets were used for external validation. The GSE26939, GSE31210, and GSE72094 datasets were selected as they provide sufficient gene expression profiles and comprehensive clinical information, especially survival time and survival status: GSE26939 consists of gene expression data measured using Agilent 44K microarrays and DNA copy number measurements using Affymetrix 250K Sty and SNP6 microarrays from samples from 115 patients with LUAD (18); GSE31210 involves the detection of EGFR, KRAS and ALK mutation statuses in 226 patients with pathological stages I–II (19); and GSE72094 provides expression sequencing of mutation-associated genes from tumor samples of 398 patients with LUAD, as well as data for tumor proliferation and immune surveillance (20). Data processing and visualization were performed using R software (version 4.3.2; The R Foundation).

Acquisition of cuproptosis-related mitochondrial and differential gene analysis

Based on previous studies, 13 cuproptosis-related genes and 1,136 mitochondrial-related genes were identified (21,22). Pearson correlation analysis was performed using the ‘limma’ R package (version 3.54.1), which is available from bioconductor.org/packages/limma (23,24). To explore more potential cuproptosis-related mitochondrial genes, genes with a correlation (cor) of >0.3 were selected as cuproptosis-related mitochondrial genes (cor >0.3 was considered to be strongly correlated) (25,26). Differentially expressed genes (DEGs) were filtered based on the criteria of |logFoldChange|≥1 and P<0.05. A Venn diagram was created using the online tool jvenn (version 1.2.0, jvenn.toulouse.inrae.fr/app/index.html) to show the overlap between DEGs and mitochondrial genes in TCGA dataset.

Machine learning and risk score model construction

A univariate Cox analysis was performed, and the results were visualized using a forest plot. TCGA, GSE26939, GSE31210 and GSE72094 datasets were combined for machine learning analysis. A total of 10 machine learning algorithms and 117 algorithm combinations were used, and the best algorithm combination was determined based on 10-fold cross-validation. The best model was selected based on the highest average C-index value. The 10 machine learning algorithms included the following: Least Absolute Shrinkage and Selection Operator (LASSO) (26), Random Survival Forest (RSF) (27), StepCox (28), Elastic Net (Enet) (29), Ridge (30), Generalized Boosted Regression Modeling (GBM) (31), CoxBoost (32), Cox Partial Least Squares Regression (plsRcox) (33), Supervised Principal Components (SuperPC) (34) and survival-Support Vector Machine (survival-SVM) (35). Finally, a risk score model was constructed using the LASSO + survival-SVM algorithm based on the C-index value, and the optimal cutoff point was determined (22). The machine learning algorithms used are listed in Table I. A total of 22 prognostic genes were identified. The λ.min value was used as the penalty parameter for the LASSO machine learning model as λ.min ensures that more prognostic genes are retained without compromising predictive performance, allowing the exploration of the roles of more relevant prognostic genes in LUAD prognosis and treatment (36). The code for the 117 machine learning methods is provided in Data S1. The risk score was calculated by applying the standardized expression values of the identified genes and their respective coefficients, with the following formula: Risk score=∑ (Coef × Expr), where Coef is the coefficient and Expr is the FPKM for each gene. The code used to calculate the risk coefficients is provided in Data S2. Survival analysis was performed using the ‘survival’ R package (version 3.8–3, CRAN.R-project.org/package=survival) (37), the ‘survminer’ package (version 0.5.0, available at http://CRAN.R–project.org/package=survminer) (38), and the ‘ggplot2’ package (version 3.5.2, available at http://CRAN.R–project.org/package=ggplot2) (39). Kaplan-Meier analysis was performed based on risk grouping to evaluate patient prognosis. Patients were categorized into high- and low-risk groups based on the median risk score.

Table I.

Types of machine learning.

Table I.

Types of machine learning.

AlgorithmR package(Refs.)
LASSOglmnet(26)
RSF RandomForestSRC(27)
StepCoxsurvival(28)
Enetglmnet(29)
Ridgeglmnet(30)
GBMgbm(31)
CoxBoostCoxBoost(32)
plsRcoxplsRcox(33)
SuperPCsuperpc(34)
survival-SVMsurvivalsvm(35)

[i] LASSO, least absolute shrinkage and selection operator; RSF, Random Survival Forest; Enet, Elastic Net; GBM, Generalized boosted regression modeling; plsRox, Cox Partial Least Squares Regression; SuperPC, Supervised Principal Components; survival-SVM, survival support vector machine.

Nomogram construction based on risk score model

A forest plot was used to present the impact of each variable on the model, including P-values, hazard ratios (HRs) and 95% confidence intervals (CIs). A nomogram was constructed using both univariate and multivariate Cox regression analyses of clinical features and risk scores, using the ‘rms’ R package (version 6.7-0, CRAN.R-project.org/package=rms) (40). Calibration curves were used to assess the clinical applicability of the model. Receiver operating characteristic (ROC) curves were employed to evaluate the predictive accuracy of the model (22).

Tumor microenvironment (TME) and tumor mutational burden (TMB) estimation based on risk score

Single-sample gene set enrichment analysis (ssGSEA) normalized the gene expression profile within each sample and then calculated the ssGSEA score for each gene set. ssGSEA analysis of the gene expression profile from TCGA-LUAD dataset was performed to quantitatively analyze the levels of immune cells and immune-related pathways. The ssGSEA analysis code is provided in Data S3. In this way, ssGSEA transforms the gene expression profile of an individual sample into a gene set enrichment score matrix. The gene sets used for immune infiltration analysis mainly came from previously published studies and were downloaded from the supplementary file of He et al (25). Additionally, these gene sets were obtained by searching for the target pathway genes in the ‘Quick Search’ section on the homepage of the TISIDB database (http://cis.hku.hk/TISIDB/index.php). These gene sets are provided in Table SI. The expression of interleukin (IL) and tumor necrosis factor (TNF) gene families were compared between the high- and low-risk groups (26). Additionally, a correlation heatmap analysis of 22 model genes with immune cells and pathways was generated. Moreover, the Tumor Immune Dysfunction and Exclusion (TIDE) analysis was performed using the TIDE web platform (version 1.0, tide.dfci.harvard.edu) to predict the response of patients with LUAD to immune checkpoint therapy by calculating immune dysfunction, with a higher TIDE score indicating greater resistance to immunotherapy (41). The ESTIMATE algorithm (available at: bioinformatics.mdanderson.org/estimate/) and the ‘estimate’ R package (version 1.0.13) were used to evaluate stromal, immune, and estimate scores to infer TME in each LUAD patient (42). The CIBERSORT algorithm (http://cibersort.stanford.edu/) and the ‘xCell’ R package (version 1.1.0, xcell.ucsf.edu/) were used to analyze immune cell infiltration and estimate the expression of 67 immune cell subtypes in each patient (43,44). In addition, tumor mutational burden (TMB), an important biomarker for predicting tumor immune response, was calculated using the ‘maftools’ R package (version 2.24.0, available at http://bioconductor.org/packages/maftools) (36,45).

Drug sensitivity analysis (IC50)

The association between prognostic genes and 198 drugs was assessed. Drugs were selected for analysis based on their correlation with risk scores and prognostic genes, and a heatmap was created. Drug sensitivity was assessed using the ‘oncoPredict’ R package (version 1.2, available at http://github.com/HuangLabUMN/oncoPredict) (46), whilst the ‘ggpubr’ package (version 0.6.1, rpkgs.datanovia.com/ggpubr/) was used for correlation analysis (47). Subsequently, the ‘pheatmap’ (version 1.0.13, available at http://github.com/raivokolde/pheatmap) and ‘psych’ (version 2.5.6, available at personality-project.org/r/psych/) packages were used to generate heatmaps (48,49).

Cell culture

A549 and PC9 human lung cancer cell lines, were purchased from the Cell Bank of the Chinese Academy of Sciences (Shanghai, China). The catalog numbers are as follows: A549 (Cat. No. TCHu150) and PC9 (Cat. No. TCHu142). The cells were cultured in RPMI 1640 medium (Gibco; Thermo Fisher Scientific, Inc.) supplemented with 10% FBS (NEWZERUM; FBS-UE500, Uruguay) and 1% penicillin/streptomycin solution. They were maintained in a humidified incubator at 37°C with 5% CO2, and the culture medium was replaced every 2–3 days. When cells reached 90% confluence, they were digested with 0.25% trypsin-EDTA solution for subsequent experiments.

Small interfering (si)RNA transfection

siRNA was used to knockdown gene expression, with a negative control (siNC). Superoxide dismutase 2 (SOD2) siRNA and siNC were purchased from IBSBIO. The sequences used were as follows: si-SOD2, 5′-GGUUCCUUUGACAAGUUUAAG-3′ and si-NC, 5′-UUCUCCGAACGUGUCACGUTT-3′. According to the manufacturer's instructions, si-SOD2 was transfected into A549 and PC9 cells using Lipofectamine™ 3000 Transfection Reagent (Invitrogen™; Thermo Fisher Scientific, Inc.). Transfection was performed using 50 nM siRNA/well. Cells were incubated at 37°C in a humidified incubator with 5% CO2 for 6 h, after which the transfection medium was replaced with fresh complete medium.

Subsequent experiments, including RT-qPCR, Western blotting, and migration/proliferation assays, were carried out 24 to 48 h post-transfection. RT-qPCR analysis confirmed that SOD2 expression was significantly reduced after siRNA transfection in both A549 and PC9 cells.

RT-qPCR

After transfection, A549 and PC9 cell lines were digested with trypsin according to the manufacturer's instructions. Total RNA was extracted using TRIzol™ reagent (Invitrogen; Thermo Fisher Scientific, Inc.), and its concentration was measured using a NanoDrop instrument (Thermo Fisher Scientific, Inc.). Total RNA was reverse transcribed into complementary DNA using the PrimeScript™ RT Reagent Kit (Takara Biotechnology Co., Ltd.) following the manufacturer's protocol. qPCR was performed using the SYBR Green Master Mix (Vazyme Biotech Co., Ltd.) to assess the expression of the SOD2 gene. The thermocycling conditions were as follows: initial denaturation at 95°C for 30 sec, followed by 40 cycles of 95°C for 10 sec and 60°C for 30 sec. β-actin was used as the internal control, and relative mRNA expression was calculated using the 2−ΔΔCq method (50). The primer sequences used were as follows: β-actin (forward), 5′-ATAGCACAGCCTGGATAGCAACGTAC-3′ and (reverse), 5′-CACCTTCTACAATGAGCTGCGTGTG-3′; and SOD2 (forward), 5′-GCTCCGGTTTTGGGGTATCTG-3′ and (reverse), 5′-GCGTTGATGTGAGGTTCCAG-3′.

Cell proliferation assay

For the Cell Counting Kit-8 (CCK-8) assay (51), A549 and PC9 cell lines were transfected with siRNA for 48–72 h. Subsequently, 5,000 cells/well were seeded into 96-well plates. Cells were incubated at 37°C with 5% CO2, and CCK-8 solution (80 µl complete medium and 20 µl CCK-8 reagent; Beyotime Biotechnology, Shanghai, China) was added at 2, 24, 48 and 72 h after incubation. After 2 h of incubation, absorbance was measured at 450 nm. Each sample was analyzed in triplicate.

Transwell migration and invasion assays

For the Transwell migration and invasion assays, A549 and PC9 cells were cultured in 6-well plates and transfected with siRNA for 24 h. Migration was assessed using 24-well Transwell chambers (BD Biosciences). A total of 5×104 cells were seeded into the upper chamber containing 150 µl serum-free medium, whilst the lower chamber contained 500 µl medium supplemented with 20% FBS. For the invasion assay, Matrigel was thawed from −20°C to liquid state at 4°C and diluted at a ratio of 1:8 with serum-free medium. Subsequently, 50 µl diluted Matrigel was added to the upper chamber and incubated at 37°C for 3 h to allow gelation before seeding the cells, to evaluate cell invasion. Following 24 h of incubation, the cells that had migrated or invaded to the lower surface of the membrane were fixed with 4% paraformaldehyde for 30 min and stained with 0.1% crystal violet for 30 min, both at room temperature. Images were captured under a light microscope (Olympus Corporation, Tokyo, Japan) for analysis.

Colony formation assay

A total of 500 A549 or PC9 cells 500 cells were seeded into 6-well plates and incubated at 37°C in a humidified atmosphere with 5% CO2 for 10 days (52). Once visible colonies (≥50 cells) formed, the cells were washed with PBS, stained with 0.1% crystal violet for 20 min at room temperature and washed with PBS. Images of the cells were then captured by light microscope(Olympus, Olympus Corporation, Tokyo, Japan). Colony counts were analyzed using GraphPad software (version 9.5.1; Dotmatics).

Mitochondrial immunofluorescence assay

The A549 and PC9 cell lines were cultured on sterile coverslips in 6-well plates and treated with siNC and si-SOD2. When the cells reached 60% confluence, they were stained with MitoTracker Deep Red mitochondrial probe (Invitrogen; Thermo Fisher Scientific, Inc.) for 30 min at 37°C. After staining, the cells were fixed with 4% paraformaldehyde for 15 min at room temperature, followed by permeabilization with 0.5% Triton X-100. The coverslips were then mounted with a fluorescence quenching mounting medium containing DAPI (1.5 µg/ml). Fluorescence imaging was performed using a laser confocal microscope to compare the mitochondrial density after treatment with siNC and si-SOD2.

Statistical analysis

All experiments were performed ≥3 times. For comparisons between two groups, the Wilcoxon rank-sum test or unpaired t-test was used. All data are presented as the mean ± standard deviation (SD) with n=3. Pearson correlation analysis was performed to assess related genes, and univariate and multivariate Cox analyses were performed to identify prognostic-related genes. Data were analyzed using GraphPad Prism (version 9.5.1). All analyses were performed using R software version 4.3.5 (The R Foundation). P<0.05 was considered to indicate a statistically signification difference.

Results

Screening and identification of cuproptosis-related mitochondrial genes

A total of 501 cuproptosis-related mitochondrial genes were identified using Pearson correlation analysis (cor >0.3; Fig. 1A). In the screening of cuproptosis and mitochondria-related genes, Pearson's linear correlation analysis was used. A general threshold of 0.5 was set as a strong correlation. However, the 1-, 3- and 5-year ROC values of the prognostic genes selected from this analysis were 0.653, 0.639 and 0.602, respectively, all of which were <0.7 (Fig. S1). Moreover, the correlation between genes is not necessarily linear and could involve indirect regulation. Therefore, to include more potential genes, a threshold of 0.3 was set according to the study by Jiang et al (26).

Identification of cuproptosis-related
mitochondrial genes in lung adenocarcinoma. (A) Sankey diagram
showing the positive correlation (>0.3) between 13
cuproptosis-related genes and mitochondrial-related genes based on
Pearson correlation analysis. (B) Volcano plot displaying all
differentially expressed genes. (C) Venn diagram showing the
intersection of 415 genes obtained from differentially expressed
genes and cuproptosis-related mitochondrial genes. (D) Forest plot
demonstrating the 67 cuproptosis-related mitochondrial genes
identified through univariate Cox regression analysis. (E)
Chromosomal localization of these 67 genes. FC, fold change; DEG,
differentially expressed gene; Cup-mt, cuproptosis-related
mitochondrial genes; HR, hazard ratio; CI, confidence interval.

Figure 1.

Identification of cuproptosis-related mitochondrial genes in lung adenocarcinoma. (A) Sankey diagram showing the positive correlation (>0.3) between 13 cuproptosis-related genes and mitochondrial-related genes based on Pearson correlation analysis. (B) Volcano plot displaying all differentially expressed genes. (C) Venn diagram showing the intersection of 415 genes obtained from differentially expressed genes and cuproptosis-related mitochondrial genes. (D) Forest plot demonstrating the 67 cuproptosis-related mitochondrial genes identified through univariate Cox regression analysis. (E) Chromosomal localization of these 67 genes. FC, fold change; DEG, differentially expressed gene; Cup-mt, cuproptosis-related mitochondrial genes; HR, hazard ratio; CI, confidence interval.

Differential gene expression analysis using TCGA dataset revealed 415 cuproptosis-related mitochondrial genes (Fig. 1B and C). Moreover, univariate Cox regression analysis identified 67 prognostic-related genes (Fig. 1D), with their chromosomal distribution mapped (Fig. 1E).

Construction of a risk scoring model based on machine learning

A total of >100 predictive models were tested using 10 distinct algorithms across multiple datasets, including TCGA LUAD and GEO datasets. The LASSO + survival-SVM algorithm achieved the highest C-index (0.676) and was selected as the optimal model (Fig. 2A-C). This model identified 22 key prognostic genes, including metabolism of cobalamin associated D (MMADHC), SOD2, human immunodeficiency virus-1 Tat interactive protein 2 (HTATIP2), cytochrome C somatic (CYCS), mitochondrial pyruvate carrier 1 (MPC1), adenylate kinase 2 (AK2). The distribution of patient risk scores in the training and testing sets is shown in Fig. 2D, G, J and M. Moreover, scatter plots were generated for the survival times of patients in the high- and low-risk groups (Fig. 2E, H, K and N). Kaplan-Meier analysis revealed that the high-risk group had significantly worse overall survival (OS) than the low-risk group (Fig. 2F, I, L and O). To develop a risk-score model, the following formula was applied: Risk score=(0.04223596 × MMADHC) + (0.0910999 × SOD2) + (0.08776809 × HTATIP2) + (0.04116991 × CYCS) + (−0.05948219 × MPC1) + (−0.09171416 × AK2) + (0.05753366 × MRPL44) + [0.01824556 × transforming growth factor β regulator 4 (TBRG4)] + [0.09307276 × mitochondrial transcription factor A (TFAM)] + (−0.03425415 × tetratricopeptide repeat domain 19) + (0.13106342 × coiled-coil-helix-coiled-coil-helix domain containing 4) + (0.09911277 × sideroflexin 1) + (0.06724052 × ATP binding cassette subfamily D member 1) + (−0.08541973 × NADH: ubiquinone oxidoreductase complex assembly factor 7) + (−0.11770888 × NOP2/Sun RNA methyltransferase 4) + (−0.19223805 × NME/NM23 nucleoside diphosphate kinase 6) + (0.18218013 × X-Prolyl aminopeptidase 3) + (−0.09327797 × lipoyltransferase 1) + (0.11149579 × mitochondrial methionyl aminopeptidase type 1D) + (−0.11764198 × carbonic anhydrase 5B) + (−0.19248092 × kynurenine 3-monooxygenase) + (−0.13196013 × alcohol dehydrogenase iron containing 1).

Machine learning-based prognostic
risk score model construction. (A) C-index for each machine
learning prediction model calculated for the training and test
sets, with >100 models included. (B) LASSO regression analysis
established a model containing cuproptosis-related mitochondrial
genes associated with prognosis. (C) Coefficients of the LASSO
analysis. (D) Distribution of risk scores, (E) survival status and
time distribution in high- and low-risk groups. (F) Kaplan-Meier
curves showing the OS of patients in high- and low-risk groups in
TCGA training set. (G) Distribution of risk scores, (H) survival
status and time distribution in high- and low-risk groups, and (I)
Kaplan-Meier curves showing the OS of patients in high- and
low-risk groups in the GSE26939 test set. (J) Distribution of risk
scores, (K) survival status and time distribution in high- and
low-risk groups, and (L) Kaplan-Meier curves showing the overall OS
of patients in high- and low-risk groups in the GSE31210 test set.
(M) Distribution of risk scores, (N) survival status and time
distribution in high- and low-risk groups, and (O) Kaplan-Meier
curves showing the overall OS of patients in high- and low-risk
groups in the GSE72094 test set. LASSO, least absolute shrinkage
and selection operator; TCGA, The Cancer Genome Atlas; OS, overall
survival.

Figure 2.

Machine learning-based prognostic risk score model construction. (A) C-index for each machine learning prediction model calculated for the training and test sets, with >100 models included. (B) LASSO regression analysis established a model containing cuproptosis-related mitochondrial genes associated with prognosis. (C) Coefficients of the LASSO analysis. (D) Distribution of risk scores, (E) survival status and time distribution in high- and low-risk groups. (F) Kaplan-Meier curves showing the OS of patients in high- and low-risk groups in TCGA training set. (G) Distribution of risk scores, (H) survival status and time distribution in high- and low-risk groups, and (I) Kaplan-Meier curves showing the OS of patients in high- and low-risk groups in the GSE26939 test set. (J) Distribution of risk scores, (K) survival status and time distribution in high- and low-risk groups, and (L) Kaplan-Meier curves showing the overall OS of patients in high- and low-risk groups in the GSE31210 test set. (M) Distribution of risk scores, (N) survival status and time distribution in high- and low-risk groups, and (O) Kaplan-Meier curves showing the overall OS of patients in high- and low-risk groups in the GSE72094 test set. LASSO, least absolute shrinkage and selection operator; TCGA, The Cancer Genome Atlas; OS, overall survival.

Prognostic curves for high and low risk scores in subgroups

To further assess the clinical relevance of the risk scoring model, it was applied it to survival prediction. The results revealed that the cuproptosis-related mitochondrial gene risk scoring system had strong predictive power and stability across several subgroups. Deceased patients had significantly higher risk scores compared with survivors (Fig. 3A). After age stratification, it was demonstrated that patients aged ≥65 years in the high-risk group had a significantly reduced OS compared with those in the low-risk group (Fig. 3B), and significant differences were also observed in patients aged <65 years (Fig. 3C). Sex stratification analysis revealed that male patients had significantly higher risk scores than female patients (Fig. 3D), and survival analysis (Fig. 3E and F) demonstrated worse OS for both male and female high-risk groups compared with low-risk groups. Regarding the M stage, the results revealed a statistically significant trend in risk scores between M0 and M1 patients (Fig. 3G); however, survival analysis demonstrated that high-risk M0 patients had significantly lower OS than low-risk patients (Fig. 3H), with similar differences observed in M1 patients (Fig. 3I). For the N stage, the risk scores in N1-3 stages were significantly higher than in the N0 stage (Fig. 3J). Moreover, survival analysis (Fig. 3K and L) indicated significantly worse OS in the high-risk group for both N0 and N1-3 stages compared with in the low-risk group. In the T stage analysis, patients with T3-4 stages had significantly higher risk scores compared with those with T1-2 stages (Fig. 3M). Survival analysis (Fig. 3N and O) further indicated significantly shorter OS in the high-risk group compared with in the low-risk group for T1-2 stages and T3-4 stages. In summary, the cuproptosis-related mitochondrial gene-based risk scoring system exhibited stable differentiation and clinical predictive value across diverse subgroups, offering robust support for patient stratification and risk management. Furthermore, this risk scoring system has the potential to identify high-risk patients.

Risk scores and survival curves for
different subgroups. (A) Risk score boxplot by survival status.
Kaplan-Meier curves for high- and low-risk groups in patients aged
(B) ≥65 years and (C) <65 years. (D) Risk score boxplot by sex.
Kaplan-Meier curves for high- and low-risk groups in (E) female and
(F) male patients. (G) Risk score boxplot by clinical stages M0 and
M1. Kaplan-Meier curves for high- and low-risk groups of patients
with clinical stage (H) M0 and (I) M1. (J) Risk score boxplot by
clinical stages N0 and N1-3. Kaplan-Meier curves for high- and
low-risk groups in patients with clinical stage (K) N0 and (L)
N1-3. (M) Risk score boxplot by stage T1-2 and T3-4. Kaplan-Meier
curves for high- and low-risk groups of patients with clinical
stage (N) T1-2 and (O) T3-4. M, metastasis; N, node; T, tumor.

Figure 3.

Risk scores and survival curves for different subgroups. (A) Risk score boxplot by survival status. Kaplan-Meier curves for high- and low-risk groups in patients aged (B) ≥65 years and (C) <65 years. (D) Risk score boxplot by sex. Kaplan-Meier curves for high- and low-risk groups in (E) female and (F) male patients. (G) Risk score boxplot by clinical stages M0 and M1. Kaplan-Meier curves for high- and low-risk groups of patients with clinical stage (H) M0 and (I) M1. (J) Risk score boxplot by clinical stages N0 and N1-3. Kaplan-Meier curves for high- and low-risk groups in patients with clinical stage (K) N0 and (L) N1-3. (M) Risk score boxplot by stage T1-2 and T3-4. Kaplan-Meier curves for high- and low-risk groups of patients with clinical stage (N) T1-2 and (O) T3-4. M, metastasis; N, node; T, tumor.

Construction of a prognostic nomogram

A univariate Cox regression analysis was performed, which evaluated the influence of several clinical variables and risk scores on the survival prognosis of patients with LUAD (Fig. 4A). The analysis revealed that age and sex were not significantly associated with survival prognosis. However, clinical stage (HR=1.652;), T stage (HR=1.580) and N stage (HR=1.679) were all significantly associated with a worse prognosis, with higher stages indicating an increased risk of death. Moreover, the risk score (HR=1.009) was confirmed as an independent prognostic factor for survival, with higher scores associated with a higher risk of death. Following the identification of these four significant factors, a multivariate Cox analysis was performed (Fig. 4B). Clinical stage and the cuproptosis-related mitochondrial gene risk score were independent prognostic factors for survival in patients with LUAD. A nomogram was then developed which combined clinical stage and risk score to predict 1-, 3- and 5-year survival (Fig. 4C). The nomogram enabled precise survival predictions at several time points by summing the total score based on the contributions from each parameter. To evaluate the nomogram, TCGA-LUAD training set and three independent validation datasets, GSE26939, GSE31210 and GSE72094, were used. The calibration curves demonstrated that the predictions of the nomogram for 1-, 3- and 5-year survival were highly consistent with actual observations, showing accuracy similar to that of an ideal model (Fig. 4D-G). Additionally, ROC curves were utilized to assess the predictive performance of the model (Fig. 4H-K). The results indicated that in TCGA training set, the area under the curve (AUC) values for 1-, 3- and 5-year survival were 0.731, 0.679 and 0.654, respectively. In the GSE26939 dataset, the AUC values for 1-, 3- and 5-year survival were 0.819, 0.787 and 0.753, respectively. The GSE31210 dataset demonstrated a more improved performance, with AUC values of 0.836, 0.806 and 0.779 for 1-, 3- and 5-year survival, respectively. Moreover, for the GSE72094 dataset, the AUC values for 1-, 3- and 5-year survival were 0.714, 0.678 and 0.670, respectively. These results highlight the high predictive accuracy and stability of the model across multiple datasets, with a particularly notable performance in the GSE31210 and GSE26939 datasets.

Nomogram construction to elucidate
lung adenocarcinoma prognosis. (A) Univariate and (B) multivariate
Cox regression analyses for clinical features and risk scores. (C)
Nomogram based on clinical features and risk scores. Calibration
curves showing the accuracy of predicted and actual values in (D)
TCGA training set, (E) the GSE26939 test set, (F) the GSE31210 test
set and (G) the GSE72094 test set. Receiver operating
characteristic curves evaluating the performance of the nomogram in
(H) TCGA training set, (I) the GSE26939 test sets, (J) the GSE31210
test set and (K) the GSE72094 test set. TCGA, The Cancer Genome
Atlas; T, tumor; N, node; OS, overall survival; FPR, false-positive
rate; TPR, true-positive rate; AUC, area under the curve.

Figure 4.

Nomogram construction to elucidate lung adenocarcinoma prognosis. (A) Univariate and (B) multivariate Cox regression analyses for clinical features and risk scores. (C) Nomogram based on clinical features and risk scores. Calibration curves showing the accuracy of predicted and actual values in (D) TCGA training set, (E) the GSE26939 test set, (F) the GSE31210 test set and (G) the GSE72094 test set. Receiver operating characteristic curves evaluating the performance of the nomogram in (H) TCGA training set, (I) the GSE26939 test sets, (J) the GSE31210 test set and (K) the GSE72094 test set. TCGA, The Cancer Genome Atlas; T, tumor; N, node; OS, overall survival; FPR, false-positive rate; TPR, true-positive rate; AUC, area under the curve.

TME, immune checkpoints and immune-related gene family analysis

Significant differences were observed in the levels of activated CD4 T cells, CD56dim natural killer cells, eosinophils, γδ T cells, immature dendritic cells, mast cells, natural killer T cells, neutrophils and type 2 T helper cells between the high- and low-risk groups (Fig. 5A). The high-risk group displayed higher levels of immunosuppressive cells, whilst the low-risk group had a higher proportion of immune-activating cells. Further analysis highlighted the presence of more immunosuppressive pathways in the high-risk group, with the low-risk group showing stronger immune activation (Fig. 5B). The expression levels of immune checkpoint genes [such as programmed cell death protein 1, programmed death-ligand 1 (PD-L1) and cytotoxic T-lymphocyte-associated protein 4] were significantly higher in the high-risk group than in the low-risk group (Fig. 5C), suggesting a more prominent immunosuppressive environment. Furthermore, the correlations between the 22 model genes, immune cell pathways and immune cells were identified (Fig. 5D).

Immune cell infiltration in high- and
low-risk groups. (A) Expression of 23 immune cell subtypes between
high- and low-risk groups. (B) Distribution and expression of
immune-related pathways in high- and low-risk groups. (C)
Expression of checkpoint-related genes in high- and low-risk
groups. (D) Correlation heatmap of risk scores and model genes with
immune cells and pathways. *P<0.05; **P<0.01; ***P<0.001.
ns, not significant. TNFRSF, tumor necrosis factor receptor
superfamily; LAG3, Lymphocyte activation gene 3; PDCD1LG2,
Programmed cell death 1 ligand 2; CD200R1, CD200 receptor 1; BTLA,
B and T lymphocyte attenuator; IDO2, indoleamine 2,3-dioxygenase 2;
BTNL2, Butyrophilin like 2; ADORA2A, Adenosine a2a receptor.

Figure 5.

Immune cell infiltration in high- and low-risk groups. (A) Expression of 23 immune cell subtypes between high- and low-risk groups. (B) Distribution and expression of immune-related pathways in high- and low-risk groups. (C) Expression of checkpoint-related genes in high- and low-risk groups. (D) Correlation heatmap of risk scores and model genes with immune cells and pathways. *P<0.05; **P<0.01; ***P<0.001. ns, not significant. TNFRSF, tumor necrosis factor receptor superfamily; LAG3, Lymphocyte activation gene 3; PDCD1LG2, Programmed cell death 1 ligand 2; CD200R1, CD200 receptor 1; BTLA, B and T lymphocyte attenuator; IDO2, indoleamine 2,3-dioxygenase 2; BTNL2, Butyrophilin like 2; ADORA2A, Adenosine a2a receptor.

The stromal score was significantly higher in the low-risk group, indicating a greater stromal component in these tumors. The immune score was also markedly higher in the low-risk group, suggesting enhanced immune cell infiltration. Similarly, the estimate score, which integrates both stromal and immune scores, was significantly elevated in the low-risk group. These findings suggested that low-risk patients possess a tumor microenvironment that is richer in stroma and more immunologically active, potentially contributing to better prognosis (Fig. 6A). Moreover, significant variations in the expression of major histocompatibility complex (MHC) gene family members [such as human leukocyte antigen (HLA)-B, HLA-C and HLA-DPB1] were revealed, with certain MHC genes downregulated in the high-risk group, potentially suppressing their antigen presentation function and promoting tumor immune evasion (Fig. 6B). Changes in the expression of inflammatory factor gene families (such as TNFSF10, TNFRSF1A and CD40) were also demonstrated (Fig. 6C), with certain inflammatory factors upregulated in the high-risk group, which may indicate a chronic inflammatory state that accelerates tumor progression. It was further revealed that the expression levels of cytotoxic molecule-related gene families (such as granzyme A, perforin 1, FAS and caspase 8) were generally lower in the high-risk group than in the low-risk group, suggesting a suppression of immune cytotoxicity, thus weakening immune killing ability (Fig. 6D). Patients in the high-risk group exhibited significantly higher TIDE and exclusion scores compared with those in the low-risk group, suggesting a greater likelihood of immune escape and decreased responsiveness to immunotherapy. Additionally, the dysfunction score was also higher in the high-risk group, indicating impaired T cell function (Fig. 6E). Finally, patients in the high-risk group exhibited a lower proportion of predicted responders and fewer potential benefits from immunotherapy compared to those in the low-risk group (Fig. 6F). In summary, the high-risk group demonstrated reduced antigen presentation, increased chronic inflammation and diminished immune killing capability, which may collectively contribute to tumor immune evasion and progression.

Immune scores and expression of
different immune gene families. (A) ESTIMATE scores. Boxplot of the
expression of (B) MHC gene families, (C) inflammatory cytokine gene
families and (D) cytotoxic molecule-related gene families between
high- and low-risk groups. (E) Difference in TIDE scores between
high- and low-risk groups. (F) Proportion of ‘No benefits’ and
‘Responder’ between high- and low-risk groups. *P<0.05;
**P<0.01; ***P<0.001. ns, not significant; MHC, major
histocompatibility complex; TIDE, Tumor Immune Dysfunction and
Exclusion.

Figure 6.

Immune scores and expression of different immune gene families. (A) ESTIMATE scores. Boxplot of the expression of (B) MHC gene families, (C) inflammatory cytokine gene families and (D) cytotoxic molecule-related gene families between high- and low-risk groups. (E) Difference in TIDE scores between high- and low-risk groups. (F) Proportion of ‘No benefits’ and ‘Responder’ between high- and low-risk groups. *P<0.05; **P<0.01; ***P<0.001. ns, not significant; MHC, major histocompatibility complex; TIDE, Tumor Immune Dysfunction and Exclusion.

In further analyses, immune cells with significant expression differences (Fig. 5A) were selected for survival prognosis analysis. In the high and low expression groups of eosinophils, immature dendritic cells and type 2 T helper cells, the survival differences were demonstrated to be significant (Fig. S2A-C). The relative proportions of 22 immune cell types estimated by the CIBERSORT algorithm showed distinct patterns between the high- and low-risk groups, with the low-risk group exhibiting a higher abundance of immune-active cells such as CD8+ T cells and activated NK cells, suggesting a more favorable immune microenvironment (Fig. S2D). In addition, single-sample GSEA (ssGSEA) analysis revealed significant differences in the infiltration scores of 67 immune cell subtypes between the two risk groups, with the low-risk group demonstrating stronger enrichment in immune-related signatures, such as antigen-presenting cells and cytotoxic lymphocytes (Fig. S2E).

Genetic variation and functional analysis

TMB was compared between the high- and low-risk groups (Fig. 7A). The analysis demonstrated that the high-risk group exhibited significantly higher TMB levels than the low-risk group. Survival analysis revealed that patients with elevated TMB had significantly improved survival outcomes than those with lower TMB (Fig. 7B and C), suggesting an association between TMB and prognosis. Moreover, when TMB was combined with the risk score, patients in the low-TMB and high-risk group (L-TMB + H-RISK) displayed the worst survival outcomes, whilst those in the high-TMB and low-risk group (H-TMB + L-RISK) showed the best survival results.

Relationship between TMB, risk scores
and gene mutations in different risk groups. (A) Comparison of TMB
between high- and low-risk groups. Kaplan-Meier survival curve for
overall survival stratified by (B) high and low TMB grouping, and
(C) TMB (high or low) and risk score (high or low). Mutation
spectrum of common gene mutations in the (D) high-risk group and
(E) low-risk group. TMB, tumor mutational burden; H-TMB, high TMB;
L-TMB, low TMB; H-RISK, high risk; L-RISK, low risk.

Figure 7.

Relationship between TMB, risk scores and gene mutations in different risk groups. (A) Comparison of TMB between high- and low-risk groups. Kaplan-Meier survival curve for overall survival stratified by (B) high and low TMB grouping, and (C) TMB (high or low) and risk score (high or low). Mutation spectrum of common gene mutations in the (D) high-risk group and (E) low-risk group. TMB, tumor mutational burden; H-TMB, high TMB; L-TMB, low TMB; H-RISK, high risk; L-RISK, low risk.

Additionally, in the high-risk cohort, 95.71% of the 223 samples had gene mutations (Fig. 7D). The most commonly mutated genes were tumor protein P53 (TP53; 59%), titin (TTN; 55%) and mucin 16 (MUC16; 45%), followed by other frequently mutated genes such as CUB and sushi multiple domains 3 (CSMD3), ryanodine receptor 2 (RYR2) and LDL receptor related protein 1B (LRP1B). The predominant mutation types included missense mutations and frameshift deletions. By contrast, in the low-risk group, 85.33% of the 225 samples had gene mutations, with TP53 (40%), TTN (33%) and MUC16 (33%) being the most frequently mutated genes, followed by CSMD3, RYR2 and LRP1B (Fig. 7E). Whilst both groups shared common mutated genes, the mutation frequency and TMB were markedly higher in the high-risk group than in the low-risk group.

Role of risk score in immunotherapy and drug sensitivity analysis

To assess the association between drug sensitivity, cuproptosis-related mitochondrial genes and the risk score, correlation analysis was performed. The results revealed significant correlations between these genes, the risk score and the sensitivity to several drugs (Fig. 8A). Several drugs demonstrated a significant negative correlation with the risk score, suggesting that patients in the high-risk group are more responsive to these treatments. The IC50 values for A770041, CGP.082996, Obatoclax.Mesylate, SL.0101.1, Thapsigargin and WZ.1.84 were notably lower in the high-risk group than in the low-risk group, indicating stronger inhibitory effects (Fig. 8B, D, F, H, J and L). Furthermore, significant negative correlations were demonstrated between risk scores and IC50 values for these drugs (Fig. 8C, E, G, I, K and M), suggesting that as the risk score increases, drug sensitivity significantly improves. Specifically, for high-risk patients, drugs such as A770041, CGP.082996, and Obatoclax Mesylate may offer more effective therapeutic outcomes, indicating that the risk score may serve as a crucial prognostic marker and a predictor of drug response.

Correlation between prognostic model
and drug sensitivity prediction. (A) Correlation between risk
score, model genes and drugs. Blue indicates a negative
correlation, whilst orange denotes a positive correlation. (B)
Boxplot showing the difference in IC50 of A.770041
between high- and low-risk groups. (C) Scatter plot showing the
correlation between risk score and A.770041. (D) Boxplot showing
the difference in IC50 of CGP.082996 between high- and
low-risk groups. (E) Scatter plot showing the correlation between
risk score and CGP.082996. (F) Boxplot showing the difference in
IC50 of Obatoclax.Mesylate between high- and low-risk
groups. (G) Scatter plot showing the correlation between risk score
and Obatoclax.Mesylate. (H) Boxplot showing the difference in
IC50 of SL.0101.1 between high- and low-risk groups. (I)
Scatter plot showing the correlation between risk score and
SL.0101.1. (J) Boxplot showing the difference in IC50 of
Thapsigargin between high- and low-risk groups. (K) Scatter plot
showing the correlation between risk score and Thapsigargin. (L)
Boxplot showing the difference in IC50 of WZ.1.84
between high- and low-risk groups. (M) Scatter plot showing the
correlation between risk score and WZ.1.84. *P<0.05;
**P<0.01; ***P<0.001.

Figure 8.

Correlation between prognostic model and drug sensitivity prediction. (A) Correlation between risk score, model genes and drugs. Blue indicates a negative correlation, whilst orange denotes a positive correlation. (B) Boxplot showing the difference in IC50 of A.770041 between high- and low-risk groups. (C) Scatter plot showing the correlation between risk score and A.770041. (D) Boxplot showing the difference in IC50 of CGP.082996 between high- and low-risk groups. (E) Scatter plot showing the correlation between risk score and CGP.082996. (F) Boxplot showing the difference in IC50 of Obatoclax.Mesylate between high- and low-risk groups. (G) Scatter plot showing the correlation between risk score and Obatoclax.Mesylate. (H) Boxplot showing the difference in IC50 of SL.0101.1 between high- and low-risk groups. (I) Scatter plot showing the correlation between risk score and SL.0101.1. (J) Boxplot showing the difference in IC50 of Thapsigargin between high- and low-risk groups. (K) Scatter plot showing the correlation between risk score and Thapsigargin. (L) Boxplot showing the difference in IC50 of WZ.1.84 between high- and low-risk groups. (M) Scatter plot showing the correlation between risk score and WZ.1.84. *P<0.05; **P<0.01; ***P<0.001.

Knockdown of SOD2 inhibits malignant behavior in LUAD

Among the aforementioned prognostic genes, SOD2, CYCS, MRPL44, TBRG4, and MMADHC exhibited stronger drug sensitivity, as indicated by a greater number of significantly associated compounds (Fig. 8A). The top 10 genes with high IC50 sensitivity and a significant correlation with several drugs were screened, and SOD2 demonstrated a significant correlation with the highest number of drugs. Therefore, we hypothesized that SOD2 has favorable prognostic and therapeutic significance in LUAD, and the role of SOD2 was further evaluated in vitro. First, SOD2 was significantly knocked down in A549 and PC9 cells (Fig. 9A). Subsequently, CCK-8 assays demonstrated a significant reduction in cell proliferation after SOD2 knockdown in both cell lines compared with controls (Fig. 9B and C). Furthermore, it was revealed that, compared with controls, there was a significant decrease in cell migration and invasion (Fig. 9D and E), and a significant reduction in cell proliferation (Fig. 9F) post-SOD2 knockdown. Mitochondrial immunofluorescence assay demonstrated that SOD2 knockdown markedly reduced mitochondrial expression compared with controls (Fig. 9G and H).

Expression levels and biological
functions of SOD2 in lung adenocarcinoma cell lines. (A) mRNA
levels of SOD2 in the NC and knockdown groups in A549 and PC9 cell
lines. Cell Counting Kit-8 analysis showing the effects of SOD2
knockdown on the (B) A549 and (C) PC9 cell lines. Comparison of (D)
migration, (E) invasion and (F) proliferation in A549 and PC9 cells
between NC and SOD2 knockdown groups. Comparison of mitochondrial
fluorescence between the si-NC and si-SOD2 groups in (G) A549 and
(H) PC9 cells. ****P<0.0001. SOD2, superoxide dismutase 2; NC,
negative control; si, small interfering; ns, not significant OD,
optical density.

Figure 9.

Expression levels and biological functions of SOD2 in lung adenocarcinoma cell lines. (A) mRNA levels of SOD2 in the NC and knockdown groups in A549 and PC9 cell lines. Cell Counting Kit-8 analysis showing the effects of SOD2 knockdown on the (B) A549 and (C) PC9 cell lines. Comparison of (D) migration, (E) invasion and (F) proliferation in A549 and PC9 cells between NC and SOD2 knockdown groups. Comparison of mitochondrial fluorescence between the si-NC and si-SOD2 groups in (G) A549 and (H) PC9 cells. ****P<0.0001. SOD2, superoxide dismutase 2; NC, negative control; si, small interfering; ns, not significant OD, optical density.

Discussion

The primary aim of the present study was to develop a prognostic model using cuproptosis-related mitochondrial genes for risk stratification and prognosis prediction in patients with LUAD, which may offer valuable insights for guiding the treatment of LUAD.

Cuproptosis is a newly identified form of PCD that is triggered by excessive intracellular accumulation of copper ions (53). The buildup of copper induces mitochondrial dysfunction, oxidative stress and damage to critical intracellular biomolecules, leading to cell death. Unlike other forms of PCD, such as apoptosis and necroptosis, copper-induced cell death does not involve the activation of caspases, particularly caspase-3, which is a hallmark of classical apoptosis. Instead, cuproptosis is characterized by copper accumulation that disrupts mitochondrial lipid metabolism and destabilizes iron-sulfur (Fe/S) cluster-containing proteins, resulting in mitochondrial membrane instability, oxidative damage and subsequent cell death (10).

In LUAD, a highly invasive and metastatic subtype of NSCLC, copper toxicity has been reported to promote mitochondrial dysfunction (54), a hallmark of cancer cell metabolic reprogramming. Tumor cells often exhibit the ‘Warburg effect’, favoring glycolysis over oxidative phosphorylation for energy production even in the presence of oxygen (55). This metabolic shift not only supports rapid tumor cell proliferation but also renders LUAD cells more susceptible to mitochondrial damage caused by excess copper accumulation (56). Elevated serum copper levels have been observed in patients with lung cancer, with significantly increased levels in individuals with advanced-stage disease compared with those diagnosed at earlier stages (57). Copper ions can promote tumor angiogenesis by activating hypoxia-inducible factor-1 (HIF-1) and upregulating VEGF expression, thereby facilitating the development and metastasis of lung cancer cells. Additionally, copper ions may activate the AKT signaling pathway, leading to anti-apoptotic effects in lung cancer cells. They also enhance glycolysis and alter amino acid metabolism, further reshaping the TME (58).

Elesclomol, an antitumor agent that targets mitochondrial metabolism, exerts part of its anticancer activity by inducing cuproptosis (59). Disruption of intracellular copper homeostasis can lead to metabolic imbalance and cell death. Disulfiram (DSF) has also been reported to elevate intracellular copper levels, thereby inducing death in lung cancer cells (60). Moreover, the interplay between cuproptosis and mitochondrial dysfunction provides critical insights into the progression and therapeutic resistance of LUAD. Therefore, targeting cuproptosis-related mechanisms, particularly by restoring and regulating mitochondrial function during treatment, may offer novel strategies for LUAD therapy. Therefore, the present study aimed to identify a distinct cuproptosis-related mitochondrial gene signature with significant prognostic value for LUAD, providing potential molecular targets for precision oncology (61).

Through the analysis of TCGA dataset, 501 genes related to cuproptosis and mitochondrial function were identified. Using univariate Cox regression, LASSO regression and multiple machine learning algorithms, 22 key genes significantly associated with the prognosis of LUAD were screened. These genes (for example, SOD2, MMADHC, CYCS, MPC1 and TFAM) are involved in mitochondrial energy metabolism, oxidative stress response and the regulation of apoptosis (62). The model was constructed using TCGA-LUAD dataset and assessed in three independent GEO datasets (GSE26939, GSE31210 and GSE72094). The results demonstrated that the risk score based on these 22 cuproptosis-related mitochondrial genes effectively distinguished between high- and low-risk patients. High-risk patients exhibited worse survival outcomes across several clinical subgroups (age, sex, clinical stage and TNM stage). Furthermore, both univariate and multivariate Cox regression analyses revealed the risk score as an independent prognostic factor for LUAD. A nomogram integrating the risk score and clinical variables (clinical stage, T stage and N stage) was constructed to predict 1-, 3- and 5-year survival probabilities, thereby validating the clinical utility of the model.

Research by Liu et al (63) employed LASSO regression to screen for cuproptosis-related genes and included only two external validation cohorts, without reporting ROC AUC values or performing phenotype-related experiments such as CCK-8 or Transwell assays. By contrast, the present study adopted a more comprehensive approach by utilizing 117 machine learning algorithms to select prognostic genes, incorporating three external GEO validation datasets, and calculating the corresponding ROC AUC values. Furthermore, extensive external validations were performed, including RT-qPCR, CCK-8, Transwell and colony formation assays, and immunofluorescence experiments. Similarly, the study by Yang et al (64) also relied on LASSO modeling to identify cuproptosis-related stemness genes, yet validation was limited to a single GEO dataset (GSE141569) with AUC values of <0.7, thus lacking persuasive power. The present study, however, included three external GEO cohorts (GSE26939: 0.811, 0.714 and 0.708; GSE31210: 0.824, 0.657 and 0.732; and GSE72094, 0.714, 0.671 and 0.769), most of which achieved AUC values of >0.7, highlighting the robustness of the prognostic model. Furthermore, research by You et al (65) only utilized LASSO for gene selection and performed RT-qPCR as the sole validation method, whereas the present work provided more comprehensive external validation encompassing RNA expression, phenotypical assays in two cell lines and immunofluorescence analysis. The study by Liang et al (66), which compared Deep Neural Network and Cox models, was limited by a single external validation dataset (GSE68465). The time-dependent ROC analysis for 1-, 3-, and 5-year survival prediction yielded AUC values of 0.606, 0.621, and 0.603, respectively, all of which were <0.7, indicating suboptimal prognostic performance. By contrast, the multi-model screening using 117 machine learning algorithms in the present study enabled the identification of optimal models with stronger validation performance across multiple GEO datasets. Finally, the study by Zhang et al (67) used GSE68465 as the training set and TCGA, GSE72094 and GSE37745 as validation cohorts. As TCGA is the largest and most widely used LUAD cohort, the authors believe it should serve as the training dataset, with the GEO cohorts more suited for external validation. Additionally, whilst the model by Zhang et al (67) yielded unsatisfactory time-dependent AUC values in the GSE37745 dataset (0.658, 0.646 and 0.630 for predicting 1-, 3-, and 5-year overall survival, respectively), the prognostic model in the present study demonstrated notably greater generalizability and predictive accuracy across multiple GEO datasets. Collectively, the present study integrated a more advanced machine learning framework, comprehensive external validations and robust ROC performance, thus offering a marked advancement over existing research in this field.

Zhao et al (21) established a prognostic model based on hypoxia- and mitochondria-related genes, which were selected using weighted gene co-expression network analysis. The model incorporated 16 genes [pyruvate kinase M1/2 (PKM), S100 calcium binding protein A16, related RAS viral (r-ras) oncogene homolog, tubulin α4a, plakophilin 3, potassium channel tetramerization domain containing 12, lysophosphatidylglycerol acyltransferase 1, ITPR interacting domain containing 2, mitotic spindle organizing protein 2A, leukemia inhibitory factor receptor alpha, protein tyrosine phosphatase receptor type M, large tumor suppressor kinase 2, PDLIM1 interacting kinase 1 like, golgin RAB6-interacting, protocadherin 7, and cadherin-like and PC-esterase domain containing 1). The prognostic performance was evaluated using ROC analysis, with AUC values of 0.721, 0.711 and 0.671 for 1-, 3- and 5-year survival in TCGA cohort, respectively. External validation was performed using GSE31210 (AUC, 0.756, 0.641 and 0.669, respectively) and GSE72094 (AUC, 0.672, 0.670 and 0.673, respectively). Although the study also assessed the TME and drug sensitivity, the model construction was based solely on a single LASSO analysis, and external validation was limited to two datasets without in vitro experimental validation. Moreover, Jiang et al (26) developed a prognostic model derived from cuproptosis- and anoikis-related genes [eukaryotic translation initiation factor 2 α kinase 3 (EIF2AK3), IKAROS family zinc finger 3 (IKZF3), integrin subunit alpha V (ITGAV), O-linked N-acetylglucosamine transferase (OGT), polo-like kinase 1 (PLK1), TNF receptor associated factor 2 (TRAF2), and X-ray repair cross complementing 5 (XRCC5)], which were selected using Pearson correlation analysis (threshold of >0.3). The model was constructed using the LASSO algorithm and further evaluated using immunohistochemistry data from the Human Protein Atlas database, comparing gene expression between LUAD and adjacent normal tissues. The ROC values for survival prediction in TCGA cohort were 0.732, 0.743 and 0.712 for 1-, 3- and 5-year survival, respectively. External validation in GSE26939 achieved AUCs of 0.663, 0.614 and 0.599, respectively. In comparison with the aforementioned models, the prognostic model in the present study was constructed using a more robust and systematic pipeline. First, prognostic genes were screened using Pearson correlation analysis (threshold of >0.3), followed by evaluation across 117 machine learning algorithms to select the optimal model, thereby ensuring optimum predictive performance. In terms of biological insight, the present study comprehensively analyzed the TME using ssGSEA for immune checkpoint and immune pathway enrichment analysis, combined with immune scoring, TIDE prediction and CIBERSORT for immune cell infiltration quantification. Furthermore, TMB analysis was performed, not only comparing mutation counts between high- and low-risk groups but also integrating survival analysis to reveal the prognostic impact of TMB. A detailed comparison between previous research and the present study is provided in Table SII.

The present study explored a mitochondrial gene signature related to cuproptosis, with a particular focus on SOD2. This gene signature shares a common foundation with two recently published signatures: Zhao et al (21), which focused on hypoxia and mitochondrial-related genes, especially PKM2, and Jiang et al (26), which identified a cuproptosis-anoikis-related gene signature, including EIF2AK3, IKZF3, ITGAV, OGT, TRAF2, XRCC5 and PLK1. The present study, along with the gene signatures proposed by Zhao et al (21) and Jiang et al (26), aimed to identify and evaluate mitochondria-related gene signatures with the potential to serve as diagnostic and therapeutic targeting biomarkers in LUAD. The study by Zhao et al (21) mainly linked mitochondrial dysfunction with hypoxia. PKM2 is a rate-limiting enzyme in glycolysis, catalyzing the conversion of phosphoenolpyruvate to pyruvate. In tumor cells, it predominantly exists in a low-activity dimeric form. PKM2 directly binds to integrin β1, activating the FAK/SRC/ERK axis to promote tumor metastasis and support angiogenesis. Furthermore, using integrin β1 inhibitors has been reported to markedly reduce tumor migration and invasion, indicating that PKM2 could serve as a potential diagnostic marker (68,69). Furthermore, in the signature by Jiang et al (26), genes such as EIF2AK3, IKZF3, ITGAV and OGT serve key roles in cell survival, adhesion and oxidative stress responses. EIF2AK3, also known as EIF2α kinase, is an important regulator of cellular stress responses. It modulates the cell cycle and helps cancer cells sustain growth, supporting the malignant phenotype of tumors (70). ITGAV, a cell adhesion molecule, promotes the interaction between cells and the extracellular matrix, helping cells resist anoikis (death induced by detachment), thereby assisting cancer cells in surviving and circulating in the bloodstream (71). OGT, by modulating the glycosylation of cell adhesion proteins, enhances the interaction between cells and the matrix, potentially serving a role in resisting anoikis, allowing cancer cells to survive after detaching from their primary site and migrate to new locations (72).

The present study investigated the cuproptosis-related mitochondrial gene signature. SOD2, identified through drug sensitivity-related screening, is a mitochondrial antioxidant enzyme that emphasizes the core role of oxidative stress and mitochondrial damage in cuproptosis and cancer progression. SOD2 protects cells from oxidative damage by clearing ROS in the mitochondria, which is particularly important in cancer cells, as they typically produce large amounts of ROS during metabolic processes. Overexpression of SOD2 may help tumor cells resist oxidative damage, promoting their survival in harsh microenvironments (73). Therefore, the development of SOD2 inhibitors or ROS modulators may increase ROS levels, promoting cancer cell death, which has clinical therapeutic implications. As such, SOD2 could serve as a potential therapeutic marker in clinical treatments.

The innovations of the present study were as follows: First, 117 machine learning algorithms (comprising 10 types of freely combinable machine learning methods) were applied to identify prognostic genes, and the optimal model was selected based on the highest concordance index (C-index). Second, three GEO datasets were used for external validation, whereas most existing studies used only one or two, thereby providing a more comprehensive validation of model reliability and generalizability. Third, the model in the present study achieved relatively high ROC values in both TCGA validation and all three GEO test sets, with several values reaching ≤0.8. The calibration curve closely aligned with the ideal diagonal, indicating notable predictive performance and strong clinical applicability (62). Moreover, the TME and TMB were analyzed, and the results demonstrated that patients in the high-risk group exhibited a greater number of gene mutations. This provides important insights for guiding targeted therapy and assessing prognosis. Finally, drug sensitivity analysis was performed for 198 molecular drugs and the top 40 statistically significant drugs were identified, offering clinical guidance for chemotherapy and targeted therapy selection. Using the ‘oncoPredict’ R package, the IC50 values of these 198 drugs were evaluated, providing a valuable reference for future personalized treatment strategies. Moreover, as the central theme of the present study was the cuproptosis-related mitochondrial gene signature, in vitro experiments were performed using immunofluorescence techniques to visualize mitochondrial changes after SOD2 gene knockdown, thereby validating the functional relevance of mitochondria in this context.

The TME serves a key role in tumor progression, immune evasion and treatment response. A study by Isomoto et al (74) reported that the immune landscape surrounding EGFR-mutant tumors in NSCLC exhibited a lack of T cell infiltration, along with a reduced PD-L1+/CD8+ tumor-infiltrating lymphocytes ratio, underscoring the critical role of the TME in immune therapy (75). The present study identified significant differences in immune cell infiltration between high- and low-risk patients with LUAD. Specifically, the high-risk group had a higher proportion of immunosuppressive cells, including regulatory T cells and myeloid-derived suppressor cells, which suppress antitumor immune responses. By contrast, the low-risk group exhibited a higher proportion of immune-activated cells, such as CD4+ T cells and natural killer cells, suggesting a more favorable immune microenvironment.

The present study also assessed the association between the cuproptosis-related mitochondrial gene signature and drug sensitivity. The correlation between risk scores and the sensitivity of 198 different drugs was evaluated. The findings indicated that high-risk patients had increased sensitivity to specific chemotherapeutic and targeted therapeutic agents, suggesting that these patients may benefit from certain drug treatments. Additionally, the results revealed that drugs such as A770041 and Obatoclax Mesylate were significantly associated with risk scores, implying that these drugs might be more effective in high-risk patients with LUAD compared with low-risk patients. In the study by Chen et al (75), Obatoclax Mesylate, a Bcl-2 family antagonist, was used in the treatment of hematologic malignancies and solid tumors. The nanoparticle-targeted delivery system of this drug improved its circulation time and accelerated delivery to tumor sites, increasing tumor-targeting specificity.

Furthermore, Fig. 8A illustrates the strong correlation between MMADHC gene expression and the IC50 values of several drugs. However, the role of MMADHC in lung cancer or other malignancies remains unexplored. Bhat et al (76) identified a homozygous pathogenic mutation in MMADHC, which is associated with homocystinuria. By contrast, the SOD2 gene, strongly correlated with the IC50 values of several drugs, has been associated with tumor development and progression. SOD2 is a crucial antioxidant enzyme in humans and is part of the superoxide dismutase family (77). Located on chromosome 6, the SOD2 gene encodes the SOD2 protein found in mitochondria, responsible for neutralizing superoxide anions (O2−) and converting them into hydrogen peroxide (H2O2), thus mitigating oxidative stress and cellular damage. SOD2 can eliminate excessive mitochondrial ROS. Deficiency in the sirtuin 3/SOD2 signaling pathway leads to increased mitochondrial ROS and mitochondrial damage. The underlying mechanisms include loss of mitochondrial membrane potential, abnormal mitochondrial ultrastructure and oxidative damage to mitochondrial DNA (78). Hernandez-Saavedra and McCord (79) suggested that genetic variations in SOD2 may increase the risk of cancer by creating potential glucocorticoid response elements. Additionally, Zhao et al (80) reported that overexpression of microRNA-512-5p increased SOD2 levels, promoting tumorigenesis and progression in NSCLC. At present, the specific regulatory role of SOD2 in the cuproptosis pathway has not been fully elucidated. However, recent findings by Li et al (81) indirect supported a potential association. In the present study, PX478, a HIF-1α inhibitor, was reported to enhance oxidative stress in A549 cells treated with DSF, as indicated by increased levels of ROS and malondialdehyde, accompanied by a reduction in SOD levels. This elevated oxidative stress subsequently upregulated ATPase copper transporting β (ATP7B) expression, thereby promoting cuproptosis and ultimately inhibiting the progression and metastasis of NSCLC cells. Moreover, DSF treatment markedly decreased the viability of A549 cells, and was associated with simultaneous upregulation of ATP7B and PD-L1. Notably, when the dosage of DSF was insufficient to induce robust cuproptosis, the increased PD-L1 expression potentially contributed to enhanced immune suppression and immune evasion, which may increase the risk of tumor cell proliferation and metastasis. Notably, combination therapy with anti-PD-L1 agents further augmented the antitumor effects of DSF, highlighting the potential synergistic benefits of integrating immune checkpoint blockade with cuproptosis-based therapeutic strategies in NSCLC.

Copper ions can catalyze the formation of ROS, leading to cellular damage, including mitochondrial dysfunction. This is particularly evident in LUAD cells, as these cells are often already in a high-oxidative state due to their rapid metabolic activity. Cuproptosis, a form of cell death induced by copper, takes advantage of this vulnerability (82). By increasing copper ion levels in tumor cells, it can push them beyond their capacity to handle ROS, ultimately leading to cell death. This mechanism may selectively target cancer cells that are already primed to handle oxidative stress while sparing normal cells (83). Although copper homeostasis is critical for tumor progression, copper overload disrupts cellular processes and induces cell death. Inducing cuproptosis in LUAD is a potential strategy to exploit the inherent oxidative stress and metabolic dysregulation of cancer cells, thereby selectively killing them. By targeting copper transport mechanisms (such as ATP7B) or using copper ionophores (such as DSF), it is possible to selectively accumulate copper in cancer cells to induce cuproptosis, thereby reducing tumor proliferation, migration and metastasis (84). Copper ions also serve a role in angiogenesis by activating VEGF. However, in cancer cells, the regulation of copper uptake, distribution and efflux is often impaired, leading to excessive intracellular copper accumulation that drives the cells into a programmed death state (cuproptosis). There is growing interest in exploiting this imbalance in copper homeostasis to achieve therapeutic effects (85–87). Therefore, increasing copper levels beyond physiological thresholds could be a way to ‘push’ cancer cells into an irreversible cell death state. In summary, whilst copper can promote tumor growth by supporting cellular functions, targeting copper overload via cuproptosis is an emerging strategy to selectively induce cancer cell death in LUAD. This approach aims to exploit the vulnerability of cancer cells to excess copper, making cuproptosis a promising mechanism to consider in LUAD treatment.

The study by Tsvetkov et al (88) reported that excessive Cu ions led to ROS generation via the Fenton reaction. When intracellular copper ion levels increase tenfold, it triggers cuproptosis, causing lipid peroxidation and mitochondrial dysfunction. DSF can induce apoptosis in tumor cells through copper ions, as reported in vitro by Li et al (81), where DSF promoted overexpression of ATP7B and PD-L1, inhibiting the cell viability of A549 cells and enhancing oxidative stress, which led to tumor cell apoptosis. The cuproptosis inducer DSF combined with sulfasalazine (SAS) has been proposed as a new antitumor drug. DSF regulates intracellular copper ion levels, promoting cuproptosis in NSCLC tumor cells, and enhances mitochondrial reactive oxygen species-mediated oxidative stress-induced cytotoxicity. In a mouse model, SAS combined with DSF-Cu markedly reduced both tumor size and number. In a mouse model, SAS combined with DSF-Cu markedly reduced both tumor size and number (89). Moreover, in vitro research by Liu et al (90) demonstrated that DSF-Cu exhibited anti-angiogenic activity by inhibiting EGFR, fibroblast growth factor receptor 1 and IGF-1Rβ. It also notably suppressed stem cell transcription factors such as SRY-box 2, thereby inhibiting tumor stem cell proliferation and invasion. Additionally, it enhanced cuproptosis in tumor cells. DSF-Cu can synergize with the WEE1 inhibitor Adavosertib to induce G2/M phase arrest and promote cuproptosis in NSCLC tumor cells (91). Furthermore, in in vitro mouse models, the combination of DSF-Cu and Adavosertib markedly reduced tumor size and weight, especially in a p53-deficient xenograft model (91).

While the present study offers valuable insights into the prognostic significance of the cuproptosis-mitochondrial gene signature in LUAD, several limitations exist. First, despite the high predictive accuracy of the model across multiple datasets, its clinical applicability requires further validation in prospective clinical trials. Second, although based on extensive RNA sequencing data, future studies could include single-cell RNA sequencing and spatial transcriptomics to explore cellular heterogeneity and expression patterns in different tissue areas. Third, although the machine learning model predicts survival outcomes effectively, its predictive power could be improved by incorporating additional clinical variables (such as treatment history and molecular subtypes).

In conclusion, the present study established a comprehensive prognostic model based on the copper death-mitochondrial-related gene signature, which not only provides new insights into the molecular mechanisms of LUAD but also offers valuable guidance for clinical decision-making. Future research should further validate the findings of the present study, explore novel therapies targeting the copper death pathway, and improve treatment outcomes for patients with LUAD through personalized therapeutic strategies.

Supplementary Material

Supporting Data
Supporting Data
Supporting Data

Acknowledgements

Not applicable.

Funding

The present study was funded by the Tianjin Municipal Health Commission, the Tianjin Key Medical Discipline Sub-project (grant no. TJLCMS2021-06), the Tianjin Municipal Education Commission through the General Project of the Natural Science Foundation (grant no. 2020KJ162), the Wu Jieping Medical Foundation (grant no. 320.6750.2022-11-43), the National Natural Science Foundation of China (grant no. 82172569), the Natural Science Foundation of Tianjin (grant no. 23JCYBJC01010) and the Tianjin Key Medical Discipline (Specialty) Construction Project (grant no. TJYXZDXK-061B).

Availability of data and materials

The data generated in the present study may be requested from the corresponding author.

Authors' contributions

YHL, WHZ, ZXZ and ZXD conceived and designed the study, analyzed and interpreted data, drafted and critically revised the manuscript, approved the final version, and agree to be accountable for all aspects of the work. HH, CD, YWL, MHL, HBZ and MW analyzed and interpreted data; assisted with figure preparation, visualization, reference organization, and manuscript drafting or revision; reviewed and approved the final version. JC supervised the entire project, contributed to study design and data interpretation, critically reviewed the manuscript, approved the final version. HLZ performed the experiments, supervised laboratory work, obtained funding and resources, critically reviewed the manuscript. All authors have read and approved the final manuscript. YHL and WHZ confirm the authenticity of all the raw data.

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.

References

1 

Zhang Y, Vaccarella S, Morgan E, Li M, Etxeberria J, Chokunonga E, Manraj SS, Kamate B, Omonisi A and Bray F: Global variations in lung cancer incidence by histological subtype in 2020: A population-based study. Lancet Oncol. 24:1206–1218. 2023. View Article : Google Scholar : PubMed/NCBI

2 

Florez N, Kiel L, Riano I, Patel S, DeCarli K, Dhawan N, Franco I, Odai-Afotey A, Meza K, Swami N, et al: Lung cancer in women: The past, present, and future. Clin Lung Cancer. 25:1–8. 2024. View Article : Google Scholar : PubMed/NCBI

3 

Carter-Harris L: Lung cancer stigma as a barrier to medical help-seeking behavior: Practice implications. J Am Assoc Nurse Pract. 27:240–245. 2015. View Article : Google Scholar : PubMed/NCBI

4 

Liu HI, Chiang CJ, Su SY, Jhuang JR, Tsai DR, Yang YW, Lin LJ, Wang YC and Lee WC: Incidence trends and spatial distributions of lung adenocarcinoma and squamous cell carcinoma in Taiwan. Sci Rep. 13:16552023. View Article : Google Scholar : PubMed/NCBI

5 

Nakagawa K, Yoshida Y, Yotsukura M and Watanabe SI: Minimally invasive open surgery (MIOS) for clinical stage I lung cancer: Diversity in minimally invasive procedures. Jpn J Clin Oncol. 51:1649–1655. 2021. View Article : Google Scholar : PubMed/NCBI

6 

Duma N, Santana-Davila R and Molina JR: Non-small cell lung cancer: Epidemiology, screening, diagnosis, and treatment. Mayo Clin Proc. 94:1623–1640. 2019. View Article : Google Scholar : PubMed/NCBI

7 

Doval DC, Desai CJ and Sahoo TP: Molecularly targeted therapies in non-small cell lung cancer: The evolving role of tyrosine kinase inhibitors. Indian J Cancer. 56 (Suppl 1):S23–S30. 2019. View Article : Google Scholar : PubMed/NCBI

8 

Laerum D, Strand TE, Brustugun OT, Gallefoss F, Falk R, Durheim MT and Fjellbirkeland L: Evaluation of sex inequity in lung-cancer-specific survival. Acta Oncol. 63:343–350. 2024. View Article : Google Scholar : PubMed/NCBI

9 

Lyu G, Dai L, Deng X, Liu X, Guo Y, Zhang Y, Wang X, Huang Y, Wu S, Guo JC and Liu Y: Integrative analysis of cuproptosis-related mitochondrial depolarisation genes for prognostic prediction in non-small cell lung cancer. J Cell Mol Med. 29:e704382025. View Article : Google Scholar : PubMed/NCBI

10 

Tong X, Tang R, Xiao M, Xu J, Wang W, Zhang B, Liu J, Yu X and Shi S: Targeting cell death pathways for cancer therapy: Recent developments in necroptosis, pyroptosis, ferroptosis, and cuproptosis research. J Hematol Oncol. 15:1742022. View Article : Google Scholar : PubMed/NCBI

11 

Wang W, Lu K, Jiang X, Wei Q, Zhu L, Wang X, Jin H and Feng L: Ferroptosis inducers enhanced cuproptosis induced by copper ionophores in primary liver cancer. J Exp Clin Cancer Res. 42:1422023. View Article : Google Scholar : PubMed/NCBI

12 

Chen L, Min J and Wang F: Copper homeostasis and cuproptosis in health and disease. Signal Transduct Target Ther. 7:3782022. View Article : Google Scholar : PubMed/NCBI

13 

Faubert B, Li KY, Cai L, Hensley CT, Kim J, Zacharias LG, Yang C, Do QN, Doucette S, Burguete D, et al: Lactate metabolism in human lung tumors. Cell. 171:358–371. 2017. View Article : Google Scholar : PubMed/NCBI

14 

He Y, Ji Z, Gong Y, Fan L, Xu P, Chen X, Miao J, Zhang K, Zhang W, Ma P, et al: Numb/Parkin-directed mitochondrial fitness governs cancer cell fate via metabolic regulation of histone lactylation. Cell Rep. 42:1120332023. View Article : Google Scholar : PubMed/NCBI

15 

Elnaggar GN, El-Hifnawi NM, Ismail A, Yahia M and Elshimy RAA: Micro RNA-148a targets Bcl-2 in patients with non-small cell lung cancer. Asian Pac J Cancer Prev. 22:1949–1955. 2021. View Article : Google Scholar : PubMed/NCBI

16 

Ran XM, Xiao H, Tang YX, Jin X, Tang X, Zhang J, Li H, Li YK and Tang ZZ: The effect of cuproptosis-relevant genes on the immune infiltration and metabolism of gynecological oncology by multiply analysis and experiments validation. Sci Rep. 13:194742023. View Article : Google Scholar : PubMed/NCBI

17 

Huang H, Shi Z, Li Y, Zhu G, Chen C, Zhang Z, Shi R, Su L, Cao P, Pan Z, et al: Pyroptosis-related LncRNA signatures correlate with lung adenocarcinoma prognosis. Front Oncol. 12:8509432022. View Article : Google Scholar : PubMed/NCBI

18 

Wilkerson MD, Yin X, Walter V, Zhao N, Cabanski CR, Hayward MC, Miller CR, Socinski MA, Parsons AM, Thorne LB, et al: Differential pathogenesis of lung adenocarcinoma subtypes involving sequence mutations, copy number, chromosomal instability, and methylation. PLoS One. 7:e365302012. View Article : Google Scholar : PubMed/NCBI

19 

Okayama H, Kohno T, Ishii Y, Shimada Y, Shiraishi K, Iwakawa R, Furuta K, Tsuta K, Shibata T, Yamamoto S, et al: Identification of genes upregulated in ALK-positive and EGFR/KRAS/ALK-negative lung adenocarcinomas. Cancer Res. 72:100–111. 2012. View Article : Google Scholar : PubMed/NCBI

20 

Schabath MB, Welsh EA, Fulp WJ, Chen L, Teer JK, Thompson ZJ, Engel BE, Xie M, Berglund AE, Creelan BC, et al: Differential association of STK11 and TP53 with KRAS mutation-associated gene expression, proliferation and immune surveillance in lung adenocarcinoma. Oncogene. 35:3209–3216. 2016. View Article : Google Scholar : PubMed/NCBI

21 

Zhao W, Huang H, Zhao Z, Ding C, Jia C, Wang Y, Wang G, Li Y, Liu H and Chen J: Identification of hypoxia and mitochondrial-related gene signature and prediction of prognostic model in lung adenocarcinoma. J Cancer. 15:4513–4526. 2024. View Article : Google Scholar : PubMed/NCBI

22 

Zhang Z, Zhang P, Xie J, Cui Y, Shuo W and Yue D: Five-gene prognostic model based on autophagy-dependent cell death for predicting prognosis in lung adenocarcinoma. Sci Rep. 14:264492024. View Article : Google Scholar : PubMed/NCBI

23 

Huang J, Zhang J, Zhang F, Lu S, Guo S, Shi R, Zhai Y, Gao Y, Tao X, Jin Z, et al: Identification of a disulfidptosis-related genes signature for prognostic implication in lung adenocarcinoma. Comput Biol Med. 165:1074022023. View Article : Google Scholar : PubMed/NCBI

24 

Ritchie ME: limma: Linear models for microarray and RNA-Seq data. The R Foundation for Statistical Computing. 2024.

25 

He Y, Jiang Z, Chen C and Wang X: Classification of triple-negative breast cancers based on Immunogenomic profiling. J Exp Clin Cancer Res. 37:3272018. View Article : Google Scholar : PubMed/NCBI

26 

Jiang G, Song C, Wang X, Xu Y, Li H, He Z, Cai Y, Zheng M and Mao W: The multi-omics analysis identifies a novel cuproptosis-anoikis-related gene signature in prognosis and immune infiltration characterization of lung adenocarcinoma. Heliyon. 9:e140912023. View Article : Google Scholar : PubMed/NCBI

27 

Wang B, Yin Y, Wang A, Liu W, Chen J and Li T: SMR-guided molecular subtyping and machine learning model reveals novel prognostic biomarkers and therapeutic targets in non-small cell lung adenocarcinoma. Sci Rep. 15:16402025. View Article : Google Scholar : PubMed/NCBI

28 

Ma Y, Li J, Xiong C, Sun X and Shen T: Development of a prognostic model for NSCLC based on differential genes in tumour stem cells. Sci Rep. 14:209382024. View Article : Google Scholar : PubMed/NCBI

29 

Wang Z, Zhang J, Zhang H, Dai Z, Liang X, Li S, Peng R, Zhang X, Liu F, Liu Z, et al: CMTM family genes affect prognosis and modulate immunocytes infiltration in grade II/III glioma patients by influencing the tumor immune landscape and activating associated immunosuppressing pathways. Front Cell Dev Biol. 10:7408222022. View Article : Google Scholar : PubMed/NCBI

30 

Chen Z and Zhang Y: Development of an immune-related gene signature applying ridge method for improving immunotherapy responses and clinical outcomes in lung adenocarcinoma. PeerJ. 13:e191212025. View Article : Google Scholar : PubMed/NCBI

31 

Ghisai SA, Barin N, van Hijfte L, Verhagen K, de Wit M, van den Bent MJ, Hoogstrate Y and French PJ: Transcriptomic analysis of EGFR co-expression and activation in glioblastoma reveals associations with its ligands. Neurooncol Adv. 7:vdae2292025.PubMed/NCBI

32 

Han AX, Long BY, Li CY, Huang DD, Xiong EQ, Li FJ, Wu GL, Liu Q, Yang GB and Hu HY: Machine learning framework develops neutrophil extracellular traps model for clinical outcome and immunotherapy response in lung adenocarcinoma. Apoptosis. 29:1090–1108. 2024. View Article : Google Scholar : PubMed/NCBI

33 

Chen T, Yang Y, Huang Z, Pan F, Xiao Z, Gong K, Huang W, Xu L, Liu X and Fang C: Prognostic risk modeling of endometrial cancer using programmed cell death-related genes: A comprehensive machine learning approach. Discov Oncol. 16:2802025. View Article : Google Scholar : PubMed/NCBI

34 

Feng Q, Lu H and Wu L: Identification of M2-like macrophage-related signature for predicting the prognosis, ecosystem and immunotherapy response in hepatocellular carcinoma. PLoS One. 18:e02916452023. View Article : Google Scholar : PubMed/NCBI

35 

Xu L, Wu J, Tian J, Zhang B, Zhao Y, Zhao Z, Luo Y and Li Y: Machine Learning unveils sphingolipid Metabolism's role in tumour microenvironment and immunotherapy in lung cancer. J Cell Mol Med. 29:e704352025. View Article : Google Scholar : PubMed/NCBI

36 

Meng WJ, Guo JM, Huang L, Zhang YY, Zhu YT, Tang LS, Wang JL, Li HS and Liu JY: Anoikis-related long non-coding RNA signatures to predict prognosis and immune infiltration of gastric cancer. Bioengineering (Basel). 11:8932024. View Article : Google Scholar : PubMed/NCBI

37 

Therneau TM: Survival: Survival analysis. The R Foundation for Statistical Computing. 2024.

38 

Kassambara A: Survminer: Drawing survival curves using ‘ggplot2’. The R Foundation for Statistical Computing. 2024.

39 

Wickham H: ggplot2: Create elegant data visualisations using the grammar of graphics. The R Foundation for Statistical Computing. 2025.

40 

Harrell FE Jr: rms: Regression modeling strategies. The R Foundation for Statistical Computing. 2024.

41 

Jiang P, Gu S, Pan D, Fu J, Sahu A, Hu X, Li Z, Traugh N, Bu X, Li B, et al: Signatures of T cell dysfunction and exclusion predict cancer immunotherapy response. Nat Med. 24:1550–1558. 2018. View Article : Google Scholar : PubMed/NCBI

42 

Yoshihara K, Shahmoradgoli M, Martinez E, Vegesna R, Kim H, Torres-Garcia W, Treviño V, Shen H, Laird PW, Levine DA, et al: Inferring tumour purity and stromal and immune cell admixture from expression data. Nat Commun. 4:26122013. View Article : Google Scholar : PubMed/NCBI

43 

Newman AM, Liu CL, Green MR, Gentles AJ, Feng W, Xu Y, Hoang CD, Diehn M and Alizadeh AA: Robust enumeration of cell subsets from tissue expression profiles. Nat Methods. 12:453–457. 2015. View Article : Google Scholar : PubMed/NCBI

44 

Aran D, Hu Z and Butte AJ: xCell: Digitally portraying the tissue cellular heterogeneity landscape. Genome Biol. 18:2202017. View Article : Google Scholar : PubMed/NCBI

45 

Mayakonda A: maftools: Summarize, analyze and visualize MAF files. The R Foundation for Statistical Computing. 2025.

46 

Maeser D: oncoPredict: Drug response modeling and biomarker discovery. The R Foundation for Statistical Computing. 2024.

47 

Kassambara A: ggpubr: ‘ggplot2’-Based Publication Ready Plots. The R Foundation for Statistical Computing. 2025.

48 

Kolde R: pheatmap: Pretty Heatmaps. The R Foundation for Statistical Computing. 2025.

49 

Revelle W: Psych: Procedures for psychological, psychometric, and personality research. The R Foundation for Statistical Computing. 2025.

50 

Livak KJ and Schmittgen TD: Analysis of relative gene expression data using real-time quantitative PCR and the 2(−Delta Delta C(T)) method. Methods. 25:402–408. 2001. View Article : Google Scholar : PubMed/NCBI

51 

Pan J, Liu F, Xiao X, Xu R, Dai L, Zhu M, Xu H, Xu Y, Zhao A, Zhou W, et al: METTL3 promotes colorectal carcinoma progression by regulating the m6A-CRB3-Hippo axis. J Exp Clin Cancer Res. 41:192022. View Article : Google Scholar : PubMed/NCBI

52 

Li J, Xie H, Ying Y, Chen H, Yan H, He L, Xu M, Xu X, Liang Z, Liu B, et al: YTHDF2 mediates the mRNA degradation of the tumor suppressors to induce AKT phosphorylation in N6-methyladenosine-dependent way in prostate cancer. Mol Cancer. 19:1522020. View Article : Google Scholar : PubMed/NCBI

53 

Wen H, Qu C, Wang Z, Gao H, Liu W, Wang H, Sun H, Gu J, Yang Z and Wang X: Cuproptosis enhances docetaxel chemosensitivity by inhibiting autophagy via the DLAT/mTOR pathway in prostate cancer. FASEB J. 37:e231452023. View Article : Google Scholar : PubMed/NCBI

54 

Zhu S, Wu H, Cui H, Guo H, Ouyang Y, Ren Z, Deng Y, Geng Y, Ouyang P, Wu A, et al: Induction of mitophagy via ROS-dependent pathway protects copper-induced hypothalamic nerve cell injury. Food Chem Toxicol. 181:1140972023. View Article : Google Scholar : PubMed/NCBI

55 

Vander Heiden MG, Cantley LC and Thompson CB: Understanding the Warburg effect: The metabolic requirements of cell proliferation. Science. 324:1029–1033. 2009. View Article : Google Scholar : PubMed/NCBI

56 

Tsvetkov P, Detappe A, Cai K, Keys HR, Brune Z, Ying W, Thiru P, Reidy M, Kugener G, Rossen J, et al: Mitochondrial metabolism promotes adaptation to proteotoxic stress. Nat Chem Biol. 15:681–689. 2019. View Article : Google Scholar : PubMed/NCBI

57 

Su Y, Zhang X, Li S, Xie W and Guo J: Emerging roles of the copper-CTR1 axis in tumorigenesis. Mol Cancer Res. 20:1339–1353. 2022. View Article : Google Scholar : PubMed/NCBI

58 

Wu Z, Zhang W and Kang YJ: Copper affects the binding of HIF-1α to the critical motifs of its target genes. Metallomics. 11:429–438. 2019. View Article : Google Scholar : PubMed/NCBI

59 

Zheng P, Zhou C, Lu L, Liu B and Ding Y: Elesclomol: A copper ionophore targeting mitochondrial metabolism for cancer therapy. J Exp Clin Cancer Res. 41:2712022. View Article : Google Scholar : PubMed/NCBI

60 

Li Q, Wang T, Zhou Y and Shi J: Cuproptosis in lung cancer: Mechanisms and therapeutic potential. Mol Cell Biochem. 479:1487–1499. 2024. View Article : Google Scholar : PubMed/NCBI

61 

Li H, Fu Y, Xu Y, Miao H, Wang H, Zhang T, Mei X, He Y, Zhang A and Ge X: Cuproptosis associated cytoskeletal destruction contributes to podocyte injury in chronic kidney disease. Am J Physiol Cell Physiol. 327:C254–C269. 2024. View Article : Google Scholar : PubMed/NCBI

62 

Zhao W, Ding C, Zhao M, Li Y, Huang H, Li X, Cheng Q, Shi Z, Gao W, Liu H and Chen J: Identification and validation of a hypoxia and glycolysis prognostic signatures in lung adenocarcinoma. J Cancer. 15:1568–1582. 2024. View Article : Google Scholar : PubMed/NCBI

63 

Liu Y, Lin W, Yang Y, Shao J, Zhao H, Wang G and Shen A: Role of cuproptosis-related gene in lung adenocarcinoma. Front Oncol. 12:10809852022. View Article : Google Scholar : PubMed/NCBI

64 

Yang J, Liu K, Yang L, Ji J, Qin J, Deng H and Wang Z: Identification and validation of a novel cuproptosis-related stemness signature to predict prognosis and immune landscape in lung adenocarcinoma by integrating single-cell and bulk RNA-sequencing. Front Immunol. 14:11747622023. View Article : Google Scholar : PubMed/NCBI

65 

You J, Yu Q, Chen R, Li J, Zhao T and Lu Z: A prognostic model for lung adenocarcinoma based on cuproptosis and disulfidptosis related genes revealing the key prognostic role of FURIN. Sci Rep. 15:60572025. View Article : Google Scholar : PubMed/NCBI

66 

Liang P, Chen J, Yao L, Hao Z and Chang Q: A deep learning approach for prognostic evaluation of lung adenocarcinoma based on cuproptosis-related genes. Biomedicines. 11:14792023. View Article : Google Scholar : PubMed/NCBI

67 

Zhang W, Qu H, Ma X, Li L, Wei Y, Wang Y, Zeng R, Nie Y, Zhang C, Yin K, et al: Identification of cuproptosis and immune-related gene prognostic signature in lung adenocarcinoma. Front Immunol. 14:11797422023. View Article : Google Scholar : PubMed/NCBI

68 

Wang C, Zhang S, Liu J, Tian Y, Ma B, Xu S, Fu Y and Luo Y: Secreted pyruvate kinase M2 promotes lung cancer metastasis through activating the integrin Beta1/FAK signaling pathway. Cell Rep. 30:1780–1797. 2020. View Article : Google Scholar : PubMed/NCBI

69 

Yin L, Shi J, Zhang J, Lin X, Jiang W, Zhu Y, Song Y, Lu Y and Ma Y: PKM2 is a potential prognostic biomarker and related to immune infiltration in lung cancer. Sci Rep. 13:222432023. View Article : Google Scholar : PubMed/NCBI

70 

Liu Y, Liang X, Zhang H, Dong J, Zhang Y, Wang J, Li C, Xin X and Li Y: ER stress-related genes EIF2AK3, HSPA5, and DDIT3 polymorphisms are associated with risk of lung cancer. Front Genet. 13:9387872022. View Article : Google Scholar : PubMed/NCBI

71 

Cheuk IW, Siu MT, Ho JC, Chen J, Shin VY and Kwong A: ITGAV targeting as a therapeutic approach for treatment of metastatic breast cancer. Am J Cancer Res. 10:211–223. 2020.PubMed/NCBI

72 

Wu D, Jin J, Qiu Z, Liu D and Luo H: Functional analysis of O-GlcNAcylation in cancer metastasis. Front Oncol. 10:5852882020. View Article : Google Scholar : PubMed/NCBI

73 

Crawford A, Fassett RG, Geraghty DP, Kunde DA, Ball MJ, Robertson IK and Coombes JS: Relationships between single nucleotide polymorphisms of antioxidant enzymes and disease. Gene. 501:89–103. 2012. View Article : Google Scholar : PubMed/NCBI

74 

Isomoto K, Haratani K, Hayashi H, Shimizu S, Tomida S, Niwa T, Yokoyama T, Fukuda Y, Chiba Y, Kato R, et al: Impact of EGFR-TKI treatment on the tumor immune microenvironment in EGFR mutation-positive non-small cell lung cancer. Clin Cancer Res. 26:2037–2046. 2020. View Article : Google Scholar : PubMed/NCBI

75 

Chen S, Ren Y and Duan P: Biomimetic nanoparticle loading obatoclax mesylate for the treatment of non-small-cell lung cancer (NSCLC) through suppressing bcl-2 signaling. Biomed Pharmacother. 129:1103712020. View Article : Google Scholar : PubMed/NCBI

76 

Bhat V, Narayanan DL and Shukla A: Report of rapid diagnosis and precise management of MMADHC-related intracellular cobalamin defect. BMJ Case Rep. 14:e2397552021. View Article : Google Scholar : PubMed/NCBI

77 

Wu D and Casey PJ: GPCR-Gα13 involvement in mitochondrial function, oxidative stress, and prostate cancer. Int J Mol Sci. 25:71622024. View Article : Google Scholar : PubMed/NCBI

78 

Mukherjee S, Forde R, Belton A and Duttaroy A: SOD2, the principal scavenger of mitochondrial superoxide, is dispensable for embryogenesis and imaginal tissue development but essential for adult survival. Fly (Austin). 5:39–46. 2011. View Article : Google Scholar : PubMed/NCBI

79 

Hernandez-Saavedra D and McCord JM: Association of a new intronic polymorphism of the SOD2 gene (G1677T) with cancer. Cell Biochem Funct. 27:223–227. 2009. View Article : Google Scholar : PubMed/NCBI

80 

Zhao C, Zhang Z, Wang Z and Liu X: Circular RNA circRANGAP1/miR-512-5p/SOD2 axis regulates cell proliferation and migration in non-small cell lung cancer (NSCLC). Mol Biotechnol. 66:3608–3617. 2024. View Article : Google Scholar : PubMed/NCBI

81 

Li P, Sun Q, Bai S, Wang H and Zhao L: Combination of the cuproptosis inducer disulfiram and anti-PD-L1 abolishes NSCLC resistance by ATP7B to regulate the HIF-1 signaling pathway. Int J Mol Med. 53:192024. View Article : Google Scholar : PubMed/NCBI

82 

Wang X, Liu Z and Lin C: Metal ions-induced programmed cell death: How does oxidative stress regulate cell death? Life Sci. 374:1236882025. View Article : Google Scholar : PubMed/NCBI

83 

Guo Z, Chen D, Yao L, Sun Y, Li D, Le J, Dian Y, Zeng F, Chen X and Deng G: The molecular mechanism and therapeutic landscape of copper and cuproptosis in cancer. Signal Transduct Target Ther. 10:1492025. View Article : Google Scholar : PubMed/NCBI

84 

Ning X, Chen X, Li R, Li Y, Lin Z and Yin Y: Identification of a novel cuproptosis inducer that induces ER stress and oxidative stress to trigger immunogenic cell death in tumors. Free Radic Biol Med. 229:276–288. 2025. View Article : Google Scholar : PubMed/NCBI

85 

Sun W, Lu H, Zhang P, Zeng L, Ye B, Xu Y, Chen J, Xue P, Yu J, Chen K, et al: Localized propranolol delivery from a copper-loaded hydrogel for enhancing infected burn wound healing via adrenergic β-receptor blockade. Mater Today Bio. 30:1014172025. View Article : Google Scholar : PubMed/NCBI

86 

Ge EJ, Bush AI, Casini A, Cobine PA, Cross JR, DeNicola GM, Dou QP, Franz KJ, Gohil VM, Gupta S, et al: Connecting copper and cancer: From transition metal signalling to metalloplasia. Nat Rev Cancer. 22:102–113. 2022. View Article : Google Scholar : PubMed/NCBI

87 

Oliveri V: Selective targeting of cancer cells by copper ionophores: An overview. Front Mol Biosci. 9:8418142022. View Article : Google Scholar : PubMed/NCBI

88 

Tsvetkov P, Coy S, Petrova B, Dreishpoon M, Verma A, Abdusamad M, Rossen J, Joesch-Cohen L, Humeidi R, Spangler RD, et al: Copper induces cell death by targeting lipoylated TCA cycle proteins. Science. 375:1254–1261. 2022. View Article : Google Scholar : PubMed/NCBI

89 

Bagherpoor AJ, Shameem M, Luo X, Seelig D and Kassie F: Inhibition of lung adenocarcinoma by combinations of sulfasalazine (SAS) and disulfiram-copper (DSF-Cu) in cell line models and mice. Carcinogenesis. 44:291–303. 2023. View Article : Google Scholar : PubMed/NCBI

90 

Liu X, Wang L, Cui W, Yuan X, Lin L, Cao Q, Wang N, Li Y, Guo W, Zhang X, et al: Targeting ALDH1A1 by disulfiram/copper complex inhibits non-small cell lung cancer recurrence driven by ALDH-positive cancer stem cells. Oncotarget. 7:58516–58530. 2016. View Article : Google Scholar : PubMed/NCBI

91 

Liu D, Cao J, Ding X, Xu W, Yao X, Dai M, Tai Q, Shi M, Fei K, Xu Y and Su B: Disulfiram/copper complex improves the effectiveness of the WEE1 inhibitor Adavosertib in p53 deficient non-small cell lung cancer via ferroptosis. Biochim Biophys Acta Mol Basis Dis. 1870:1674552024. View Article : Google Scholar : PubMed/NCBI

Related Articles

  • Abstract
  • View
  • Download
  • Twitter
Copy and paste a formatted citation
Spandidos Publications style
Liu Y, Zhao W, Zhao Z, Duan Z, Huang H, Ding C, Liu M, Zhang H, Li Y, Wang M, Wang M, et al: Machine learning‑based model identifies a novel cuproptosis‑related mitochondrial gene signature with a key role in the prognosis and treatment of lung adenocarcinoma. Oncol Lett 30: 494, 2025.
APA
Liu, Y., Zhao, W., Zhao, Z., Duan, Z., Huang, H., Ding, C. ... Zhao, H. (2025). Machine learning‑based model identifies a novel cuproptosis‑related mitochondrial gene signature with a key role in the prognosis and treatment of lung adenocarcinoma. Oncology Letters, 30, 494. https://doi.org/10.3892/ol.2025.15240
MLA
Liu, Y., Zhao, W., Zhao, Z., Duan, Z., Huang, H., Ding, C., Liu, M., Zhang, H., Li, Y., Wang, M., Chen, J., Zhao, H."Machine learning‑based model identifies a novel cuproptosis‑related mitochondrial gene signature with a key role in the prognosis and treatment of lung adenocarcinoma". Oncology Letters 30.5 (2025): 494.
Chicago
Liu, Y., Zhao, W., Zhao, Z., Duan, Z., Huang, H., Ding, C., Liu, M., Zhang, H., Li, Y., Wang, M., Chen, J., Zhao, H."Machine learning‑based model identifies a novel cuproptosis‑related mitochondrial gene signature with a key role in the prognosis and treatment of lung adenocarcinoma". Oncology Letters 30, no. 5 (2025): 494. https://doi.org/10.3892/ol.2025.15240
Copy and paste a formatted citation
x
Spandidos Publications style
Liu Y, Zhao W, Zhao Z, Duan Z, Huang H, Ding C, Liu M, Zhang H, Li Y, Wang M, Wang M, et al: Machine learning‑based model identifies a novel cuproptosis‑related mitochondrial gene signature with a key role in the prognosis and treatment of lung adenocarcinoma. Oncol Lett 30: 494, 2025.
APA
Liu, Y., Zhao, W., Zhao, Z., Duan, Z., Huang, H., Ding, C. ... Zhao, H. (2025). Machine learning‑based model identifies a novel cuproptosis‑related mitochondrial gene signature with a key role in the prognosis and treatment of lung adenocarcinoma. Oncology Letters, 30, 494. https://doi.org/10.3892/ol.2025.15240
MLA
Liu, Y., Zhao, W., Zhao, Z., Duan, Z., Huang, H., Ding, C., Liu, M., Zhang, H., Li, Y., Wang, M., Chen, J., Zhao, H."Machine learning‑based model identifies a novel cuproptosis‑related mitochondrial gene signature with a key role in the prognosis and treatment of lung adenocarcinoma". Oncology Letters 30.5 (2025): 494.
Chicago
Liu, Y., Zhao, W., Zhao, Z., Duan, Z., Huang, H., Ding, C., Liu, M., Zhang, H., Li, Y., Wang, M., Chen, J., Zhao, H."Machine learning‑based model identifies a novel cuproptosis‑related mitochondrial gene signature with a key role in the prognosis and treatment of lung adenocarcinoma". Oncology Letters 30, no. 5 (2025): 494. https://doi.org/10.3892/ol.2025.15240
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