Open Access

Hypoxia studies in non‑small cell lung cancer: Pathogenesis and clinical implications (Review)

  • Authors:
    • Sirui Zhou
    • Jiazheng Sun
    • Weijian Zhu
    • Zhiying Yang
    • Ping Wang
    • Yulan Zeng
  • View Affiliations

  • Published online on: December 30, 2024     https://doi.org/10.3892/or.2024.8862
  • Article Number: 29
  • Copyright: © Zhou et al. This is an open access article distributed under the terms of Creative Commons Attribution License.

Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

Non‑small cell lung cancer (NSCLC) is one of the most prevalent and lethal types of cancers worldwide and its high incidence and mortality rates pose a significant public health challenge. Despite significant advances in targeted therapy and immunotherapy, the overall prognosis of patients with NSCLC remains poor. Hypoxia is a critical driving factor in tumor progression, influencing the biological behavior of tumor cells through complex molecular mechanisms. The present review systematically examined the role of the hypoxic microenvironment in NSCLC, demonstrating its crucial role in promoting tumor cell growth, invasion and metastasis. Additionally, it has been previously reported that the hypoxic microenvironment enhances tumor cell resistance by activating hypoxia‑inducible factor and regulating exosome secretion. The hypoxic microenvironment also enables tumor cells to adapt to low oxygen and nutrient‑deficient conditions by enhancing metabolic reprogramming, such as through upregulating glycolysis. Further studies have shown that the hypoxic microenvironment facilitates immune escape by modulating tumor‑associated immune cells and suppressing the antitumor response of the immune system. Moreover, the hypoxic microenvironment increases tumor resistance to radiotherapy, chemotherapy and other types of targeted therapy through various pathways, significantly reducing the therapeutic efficacy of these treatments. Therefore, it could be suggested that early detection of cellular hypoxia and targeted therapy based on hypoxia may offer new therapeutic approaches for patients with NSCLC. The present review not only deepened the current understanding of the mechanisms of action and role of the hypoxic microenvironment in NSCLC but also provided a solid theoretical basis for the future development of precision treatments for patients with NSCLC.

Introduction

The International Agency for Research on Cancer (IARC) reported that in 2020, ~2.2 million new cases of lung cancer (LC) were diagnosed globally, leading to ~1.8 million related deaths. These figures represent 11.4% of all newly diagnosed cancer cases and 18.0% of all cancer-related mortalities (1). LC remains a significant global health concern, with ~28% of cases being diagnosed at an early stage and >50% of patients exhibiting distant metastasis at the time of the initial diagnosis (1,2). Furthermore, LC is characterized by high recurrence and metastasis rates, low cure rates and a poor overall 5-year survival rate of 22% (2,3). Histologically, LC is primarily classified into two primary types: Small cell lung cancer (SCLC) and non-SCLC (NSCLC), with NSCLC comprising ~80–85% of all LC cases. Therefore, the importance of early detection and prompt intervention for patient is important, as these factors are crucial for enhancing patient prognosis and improving survival rates.

Under normal physiological conditions, the body obtains energy through the respiratory process to maintain the normal functioning of its structure and functions, with oxygen serving an indispensable role in metabolic processes (4). The oxygen content in ambient air is ~21%, equivalent to an atmospheric pressure of 150 mmHg, whereas the oxygen level in most healthy mammalian tissues is maintained between 2–9%, averaging ~40 mmHg. By contrast, the oxygen concentration within tumor tissue typically ranges from 1–2% (5,6). Hypoxia is generally defined as an oxygen concentration in tissues of ≤2%, while severe hypoxia refers to an oxygen content of ≤0.02% (6). Hypoxia is a characteristic of the solid tumor microenvironment (TME) (5). Solid tumors often exhibit acute, chronic or cyclic hypoxia, all of which stem from an imbalance between oxygen demand and supply (7). Acute hypoxia, or perfusion-limited hypoxia, is caused by irregular dilation of the tumor vasculature or excessive aggregation of tumor cells leading to vascular occlusion. Chronic hypoxia arises from the continuous proliferation of tumor cells, which consume a large amount of oxygen near the vasculature, resulting in inadequate oxygen supply to distal cells (8). Intermittent hypoxia is caused by transient occlusion of the immature and unevenly distributed tumor vasculature, lasting from several minutes to several days (5). The hypoxic TME is a critical factor driving tumor progression, affecting tumor metabolism and its surrounding microenvironment through various mechanisms, thereby influencing tumor biological characteristics and treatment efficacy.

Given the pivotal role of hypoxia in the progression of NSCLC, exploring its potential molecular mechanisms and possible therapeutic intervention strategies is important for improving patient prognosis and survival rates. While there is increasing awareness of the impact of hypoxia on NSCLC, the precise localization and interaction of hypoxic conditions with tumor biology are currently unknown. The present review integrated previously published research findings and aimed to elucidate how hypoxia affects the behavior of cancer cells through a series of mechanisms, offering a solid theoretical basis and practical operational guidance for early detection and precise targeted therapy of NSCLC.

Mechanisms of hypoxia in NSCLC

The TME is composed of multiple components, including tumor cells, blood vessels, immune cells, cancer-associated fibroblasts, signaling molecules and the extracellular matrix (ECM). All of these elements serve crucial roles in tumor progression (9,10). Tumor cells can actively alter the composition of the TME by secreting extracellular signals to adapt to or respond to changes in the host environment (11). In the hypoxic microenvironment, the host promotes the progression of NSCLC through various pathways.

Hypoxia-inducible factor (HIF)

Cells adapt to hypoxic environments by modulating the expression of certain genes. HIF, a key transcription factor regulating gene expression, serves a central role in the hypoxic response (5). HIF is a heterodimer composed of an unstable, but critical, α subunit and a stable β subunit. The alpha subunit has three subtypes: HIF-1α, HIF-2α and HIF-3α. Although all subtypes are involved in regulating inflammatory responses, typically only HIF-1α is expressed in vivo (5). The structure of HIF-1α and HIF-1β are key to their interactions and activation during the hypoxic response (Fig. 1A). Under hypoxic conditions, HIF-α degradation is inhibited, which allows HIF-α to translocate to the nucleus and bind with HIF-β to form an active HIF transcription factor complex. This complex can interact with specific sequences in >70 target gene promoters, known as hypoxia response elements (HREs), to regulate the transcription of protein-coding and non-coding RNA genes (12) (Fig. 1B). These regulated genes serve critical roles in multiple key biological processes, including glucose uptake, tumor metabolism, angiogenesis, cell proliferation and apoptosis, helping cells effectively adapt to hypoxic stress (13).

HIF-1α can promote the expression of various cell proliferation factors, including insulin-like growth factor-2 and TGF-α. Upregulation of these factors stimulates the growth and proliferation of NSCLC cells, thereby promoting tumor development (14). HIF-1α can also activate the transcription of angiogenesis-related genes, including VEGF-A and angiopoietin-2. The increased expression of these factors enhances vasculogenesis, supplying essential oxygen and nutrients to support both local growth and metastasis of NSCLC cells. Under the influence of HIF-1α, the expression of proteases such as MMPs is significantly upregulated. These enzymes can degrade the ECM, facilitating tumor cell migration and invasion of surrounding tissues, thereby promoting NSCLC spread and metastasis (15). Additionally, hypoxia enhances the expansion and motility of NSCLC cells by inhibiting the expression of connexin (CX) proteins, thereby impairing intercellular communication. Under hypoxic conditions, the expression of certain connexin proteins, particularly CX26 and CX43, decreases, making these proteins more susceptible to degradation or internalization. The disruption of intercellular communication not only disrupts signal synchronization and metabolic exchange but also further promotes the proliferation and migration of tumor cells. Moreover, this communication barrier also activates the P53/MDM2 proto-oncogene signaling pathway, accelerating the malignant progression of NSCLC (16). Under hypoxic conditions, the activation of HIF-1α can induce EMT in NSCLC cells. During this process, epithelial cells lose their polarity and intercellular adhesion properties, transforming into cells with mesenchymal features. This transformation enhances their migration and invasion abilities, further promoting the malignant progression of tumors (17,18).

Release of exosomes

Exosomes are nanoscale extracellular vesicles enclosed by a lipid bilayer membrane and actively secreted by cells. Their biogenesis is a tightly regulated and complex process. The formation of exosomes is initiated by the invagination of the plasma membrane, leading to the generation of early endosomes. These early endosomes undergo intracellular maturation to form late endosomes, which subsequently give rise to multivesicular bodies (MVBs) through intraluminal vesicle budding (19). The fusion of MVBs with the plasma membrane facilitates the release of these vesicles into the extracellular milieu, a process defined as exosome secretion (Fig. 2) (19). Exosomes serve an essential role in intercellular communication by transporting a wide range of bioactive molecules, including mRNAs, circular RNAs (circRNAs), long non-coding RNAs (lncRNAs), microRNAs (miRNAs; miRs), proteins, lipids and other molecular entities (20). It has been previously reported that under hypoxic conditions, HIF-1 facilitates the release of tumor-derived exosomes by upregulating pyruvate kinase M2 (PKM2) mRNA expression levels. PKM2, upon upregulation and phosphorylation in tumors, functions as a protein kinase, phosphorylating synaptosome-associated protein 23. This phosphorylation event promotes the exocytic release of exosomes (21,22). Exosomes, in turn, drive the growth and progression of NSCLC by transmitting key signals that influence processes such as glycolysis, angiogenesis, tumor cell migration, invasion and immune infiltration (23). Tumor cells in hypoxic environments exhibit abnormal proliferation and altered intercellular communication, often accompanied by increased exosome secretion. During this process, the hypoxic environment specifically regulates the expression of certain molecular components in exosomes, adapting to the needs of the TME by upregulating or downregulating key molecule regulators, such as miRNAs (23,24).

For example, miR-21 upregulates the expression and activity of HIF-1α, further activating glycolysis-related genes and increasing radiation resistance in NSCLC (25). HIF-1α also promotes endothelial cell (EC) angiogenesis by downregulating miR-186-5p and upregulating protein kinase C-α (26). Under chronic hypoxic conditions, miR-191 accelerates the proliferation and migration of NSCLC cells by reducing the levels of nuclear factor I A and its downstream oncogene CCAAT enhancer binding protein α (27). Under low oxygen conditions, lncRNA pvt1 is overexpressed in lung cancer cells and upregulates HIF-1α by binding to miR-199a-5p, thereby promoting cell proliferation and malignant transformation (28). However, intermittent hypoxia increases the expression level of miR-31-5p, activating WD repeat-containing protein 5, thereby enhancing EMT in tumor cells (29). Hypoxia can also stabilize HIF-1α by promoting its binding with mutant p53 and forming a complex that downregulates miR-129-1-3p. This, in turn, upregulates protein phosphatase, Mg2+/Mn2+ dependent 1D, which dephosphorylates and inhibits p53, promoting the growth, survival and metastasis of NSCLC cells (30).

Macrophages are key immune cells that serve a pivotal role in early immune surveillance and response to tumors. They recognize and eliminate tumor cells and also activate other immune cells through antigen presentation, triggering specific immune responses. However, tumor cells interact with macrophages by secreting molecules such as exosomes that regulate macrophage polarization, thereby affecting their function. Tumor-associated macrophages (TAMs) can exhibit two distinct phenotypes: M1 and M2 (31). M1 phenotype macrophages promote inflammatory responses and inhibit tumor growth, while M2 phenotype macrophages have anti-inflammatory effects, suppress immune responses and promote tumor development (32). Under hypoxic conditions, LC cells further promote M2 polarization of TAMs by upregulating miR-21 and miR-1290. This polarization enhances angiogenesis and promotes tumor invasion and metastasis, thereby accelerating LC progression (33,34). Therefore, hypoxia is not only a result of tumorigenesis but also an important driving factor for tumor infiltration and growth (23,35). Depending on the duration and severity of hypoxia, its impact on tissues may be beneficial or harmful (36).

Metabolic reprogramming

Under hypoxic conditions, NSCLC cells undergo metabolic reprogramming to adapt to the hypoxic environment and maintain their growth and survival. This process is related to the Warburg effect, where tumor cells convert glucose to lactate via glycolysis, even when oxygen is abundant and mitochondrial function is normal, rather than through the more efficient oxidative phosphorylation pathways (37). This metabolic transformation not only improves the efficiency of energy production but also provides necessary biosynthetic precursors for tumor cells to support their rapid proliferation needs (38). In NSCLC cells, the expression of key glycolytic enzymes, such as hexokinase, phosphoglycerate kinase 1, lactate dehydrogenase and glucose transporter 1 (GLUT-1) are increased, thereby promoting increased glucose uptake and lactate production (39). Lactate is not only a product of glycolysis but also serves multiple roles in the TME. It can create an acidic microenvironment that is conducive to the survival and proliferation of NSCLC cells and can also regulate certain signaling pathways to promote tumor progression (40). Specifically, lactate can activate the ERK/STAT3 signaling pathway, induce TAMs to polarize towards the M2 phenotype and stimulate angiogenesis, which together promote tumor growth and metastasis (41,42). Additionally, lactate also has immunosuppressive effects, which can regulate the function of immune cells, thereby limiting the effectiveness of antitumor immune responses (43,44). Although glycolysis is the primary energy source for NSCLC cells, fatty acid oxidation and mitochondrial oxidative phosphorylation also serve important roles in cellular metabolism, signal transduction and tumor progression (45). Mitochondria not only produce ATP through oxidative phosphorylation but also participate in fatty acid oxidation, redox regulation and the production of reactive oxygen species (ROS), which affect cell proliferation, survival and migration. Thus, NSCLC cells achieve adaptive growth in complex TMEs and promote tumor growth and metastasis by integrating multiple metabolic pathways such as glycolysis, lactate accumulation and mitochondrial metabolism (45).

Antitumor treatment

Hypoxia, a common characteristic of tumor growth, poses a significant challenge to cancer therapy. Hypoxic environments weaken the efficacy of antitumor therapy through various mechanisms. These mechanisms mainly include decreasing DNA damage repair and promoting the development of invasive phenotypes in tumor cells, thereby enhancing the survival rate of tumor cells (46). Therefore, the hypoxic microenvironment serves a critical role in determining the effectiveness of radiotherapy, chemotherapy, immunotherapy and targeted therapy.

Radiotherapy

Radiotherapy is a crucial approach for treating tumors and its mechanism of action involves utilizing high-energy electrons photons, such as X-rays or γ-rays, to interact with atoms in the body and release high-energy electrons. These electrons can trigger cellular damage, particularly DNA damage, thereby inhibiting the proliferation and survival of tumor cells (47). The mechanism of action of radiotherapy primarily encompasses two pathways: direct and indirect. In the direct mechanism, radiation photons directly interact with DNA molecules, causing single-strand or double-strand breaks, which can prevent cell division and may induce cell necrosis or apoptosis (48). The indirect mechanism involves the interaction between radiation and water molecules, generating free radicals such as hydrogen atoms, hydroxyl radicals and hydrogen peroxide. These free radicals further react with oxygen to generate ROS, which can damage macromolecules such as DNA, proteins and lipids, leading to cellular dysfunction, alterations in signal transduction and long-term cellular damage (49). The efficacy of radiotherapy is closely linked to the oxygen level of tissues. In hypoxic tumor cells, due to the insufficient oxygen available to support redox reactions, DNA repair mechanisms may not effectively repair radiation-induced damage (50). Therefore, to achieve similar damaging effects as in oxygen-rich cells, hypoxic tumor cells require an increase in radiation dose by 2–3 times, a phenomenon known as the oxygen enhancement ratio (OER) (51).

Chemotherapy

To support their rapid proliferation, tumors form an invasive and irregular vascular system. However, these neovessels often have abnormal structures and impaired functions, leading to low drug delivery efficiency and reduced permeability and distribution of chemotherapeutic agents. This significantly reduces drug exposure to tumor cells (52). Under hypoxic conditions, the HIF-1α and p53 genes play key roles in regulating the resistance of NSCLC to cisplatin. Specifically, HIF-1α promotes metabolic adaptation by enhancing glycolysis and limiting the accumulation of ROS, thereby increasing resistance to cisplatin. Simultaneously, p53 inhibits the progression of the cell cycle from the G1 phase to the S phase by activating the downstream target gene p21, further reducing the sensitivity of tumor cells to cisplatin (53). Additionally, cell cycle arrest under hypoxic conditions triggers metabolic reprogramming, including upregulation of glycolysis pathways and antioxidant responses, which further enhance cisplatin resistance (41).

Immunotherapy

Hypoxia can also impair the antitumor immune response. Under hypoxic conditions, HIF-1α is activated, leading to the upregulation of CD47 expression. CD47 is a transmembrane protein that interacts with the macrophage signal regulatory protein α, sending a signal to macrophages to prevent degradation. This signaling inhibits the recognition and phagocytosis of tumor cells by macrophages, promoting immune evasion and thus facilitating tumor growth and metastasis (54). Furthermore, dendritic cells (DCs) serve a crucial role in activating CD4+ and CD8+ T cells via antigen presentation, which is hindered by chronic hypoxia (55). Key factors such as IL-10 and VEGF induced by hypoxia can suppress the differentiation and maturation of DCs, thereby impairing T cell responses and reducing antitumor immunity (56,57). In the hypoxic microenvironment, activation of the mTOR-HIF-1α pathway also induces the upregulation of extracellular nucleotide enzymes CD39 and CD73 in myeloid-derived suppressor cells, converting extracellular ATP/ADP to adenosine. Conversely, adenosine suppresses immunotherapy, induces cell apoptosis and further impairs immune response by binding to adenosine A2A receptors on T cells and natural killer cells (58,59).

Targeted therapies

The hypoxic microenvironment can also affect the efficacy of targeted therapy. For example, gefitinib is a tyrosine kinase inhibitor commonly used to treat NSCLC. However, under hypoxic conditions, gefitinib upregulates IL-6 expression and activates inflammatory pathways such as the TNF, NF-κB and JAK-STAT pathways, leading to the enrichment of LC stem cells and initiation of the EMT, ultimately enhancing resistance to gefitinib (60). In cases of NSCLC with EMAP like 4-ALK receptor tyrosine kinase (ALK) rearrangement, although ALK inhibitors usually show good efficacy, some patients develop resistance to agents such as crizotinib in hypoxic environments. The emergence of this resistance is closely related to the upregulation of HIF-1α and its downstream regulatory factor Slug under hypoxic conditions, which promote the EMT process and significantly enhance the migration and invasion ability of cancer cells, thereby conferring tumors with resistance to ALK inhibitors (61).

Hypoxia detection in NSCLC

Hypoxia serves a critical role in the progression of NSCLC. Early detection of the hypoxic state of tumor cells can facilitate timely intervention measures and effectively correct these conditions (62). Currently, hypoxia detection in laboratory cells and human tumors can be categorized based on the types of substances measured: physical substances, such as oxygen partial pressure and oxygen saturation, biological substances, such as enzymes and proteins, and chemical substances, such as signaling pathway products. Detection technology can be categorized as direct or indirect based on the measurement approach and can also be categorized as transient or dynamic based on the nature of the detection process (Table I) (63).

Table I.

Comprehensive summary of methods for detecting hypoxia in non-small cell lung cancer.

Table I.

Comprehensive summary of methods for detecting hypoxia in non-small cell lung cancer.

A, Direct detection

Method of detectionAnalyte/markerTechniques or toolsPrinciplesAdvantagesDisadvantages(Refs.)
PhysicalPO2Electrochemical probeMeasurement of current generated by electrode redoxDirect measurement of tissue PO2 and is classed as the gold standard in clinical testingProbes may damage tumor tissue and instantaneous detection cannot track changes over time(64,65)
Optical probeDetecting the intensity or lifetime of probesHigh sensitivity and resolution, non-invasive and enables real-time monitoringProbes measure oxygen levels in blood rather than tissues and it necessitates particular environmental conditions(66,67)

B, Indirect detection

Method of detection Analyte/markerTechniques or tools Principles Advantages Disadvantages(Refs.)

Vascular system testingVascular function and morphology analysisCT perfusion imaging, MRI and near-infrared spectroscopyEvaluating tumor vascular abnormalities and blood flowNon-invasive, analyses vascular function and allows early detectionUneven blood flow and individual variations make blood flow parameters inadequate for assessing tissue oxygen supply(6871)
Hypoxia imagingOxygen metabolism tracers imagingPositron emission tomographyRadiotracer probe-based oxygen metabolism evaluationHigh sensitivity and specificity and can analyze the whole tumor tissueCan cause radiation hazards and has a longer half-life of the tracer(8,73,74)
MRIBlood oxygen level-dependent-MRI uses the magnetic property comparison of oxy-vs. deoxyhemoglobin whereas tissue oxygen level-dependent-MRI measures tissue oxygenation via longitudinal relaxation time changesNo radiation hazard and enhanced clarity of soft tissue visualizationTime-consuming and cannot be monitored in real-time(7578)
Electron paramagnetic resonance imagingBased on the spin distribution of unpaired electrons in free oxygen moleculesHigh sensitivity, experiences minimal interference from external factors and is effective in detecting low concentrations of free radicalsLow spatial resolution, rapid signal decay with tissue depth and high cost of imaging equipment(79,80)
Measurement of gene expression levelsHypoxia inducible factor-1α, solute carrier family 2 member 1, Egl-9 family hypoxia inducible factor 2, carbonic anhydrase 9 and VEGFQuantitative PCR, RNA sequencing and Northern blottingHypoxia-induced gene expression for tumor hypoxia assessmentCan measure molecular changes in hypoxia and is useful for tissue sectioningTissue samples are required, is unable to assess dynamic changes in vivo and has a high cost(8184)
Measurement of protein expression levelsSolute carrier family 2 member 1, carbonic anhydrase 9, osteopontin and nitroreductaseWestern blotting, immunohistochemistry, immunofluorescence and ELISAHypoxia-related protein expression for tumor hypoxia assessmentHigh sensitivity and is appropriate for clinical and laboratory analysisTissue samples are required, is unable to assess dynamic changes in vivo and has a high cost(78,8588)

[i] PO2, oxygen partial pressure.

Direct detection of oxygen concentration

Physical sensors are among the most direct methods for detecting cellular hypoxia. Electrochemical probes and optical sensors can directly measure the oxygen partial pressure in tumor tissue, providing accurate tissue oxygenation values. The working principle of an electrochemical probe is to measure the oxygen partial pressure by detecting the current generated by the oxygen reduction reaction at the working electrode and the oxidation reaction at the anode (64). This technique remains the gold standard in the clinical field and with the assistance of CT-guided oxygen partial pressure measurement, it is now possible to detect hypoxia in deeper tumor regions (65). However, needle-based detection methods may cause damage to tumor tissue, making it difficult to distinguish between necrotic and hypoxic areas and these methods can only provide transient measurements, limiting the possibility of continuous monitoring (65). Alternatively, optical probes detect oxygen partial pressure by assessing the effect of oxygen partial pressure on the luminescent intensity of fluorescent substances, a process known as quenching (63). Metal-organic complexes, such as Pt(II), Pd(II), Ru(II) and Ir(III), are commonly used as phosphorescent probes (66). These probes interact with oxygen molecules, shortening their phosphorescence lifetime and leading to quenching phenomena. By measuring changes in phosphorescence intensity or lifetime, in combination with the Stern-Volmer equation, the oxygen partial pressure in tumor tissue can be quantified (67). These phosphorescent probes offer high sensitivity, high resolution and the ability to achieve real-time, non-invasive monitoring, thereby demonstrating significant advantages in effectively identifying tissue hypoxia.

Indirect detection of hypoxic microenvironment
Vascular system testing

The hypoxic state of tumors is closely linked to abnormalities in their vascular system. The structural and functional defects of the tumor vasculature lead to insufficient oxygen supply, making it particularly important to evaluate these vascular parameters that are critical for detecting hypoxia. The generation of hypoxia is generally believed to be caused by three mechanisms: i) Diffusion, due to distance from perfusing blood vessels; ii) intermittent, fluctuations in vessel opening and closing; and iii) perfusion-related, low blood flow efficiency (68). Various imaging techniques, such as cryophotometry, near-infrared spectroscopy and MRI, have been used to evaluate tumor vascular features such as distance, density, distribution and oxygenation status (69). Additionally, CT perfusion imaging, a non-invasive functional imaging method, can evaluate tumor morphology and also assess microcirculation parameters such as blood volume, blood flow, mean transit time and permeability surface, which provide valuable insights for understanding hypoxia (70,71). However, due to uneven blood flow distribution, variations in oxygen consumption between individuals and increased hemoglobin levels caused by chronic hypoxic adaptation, relying solely on blood flow parameters cannot fully capture tissue-level oxygen supply and demand dynamics (72). Therefore, combining multiple parameters to evaluate the TME is essential for a comprehensive assessment of hypoxia.

Hypoxia imaging detection

Optical imaging provides a non-invasive and dynamic approach to assessing tissue oxygenation. Positron emission tomography (PET) is one of the most widely adopted methods of optical imaging. PET assesses oxygen metabolism in tissues and identifies preclinical or clinical hypoxia through the use of radioactive tracer probes (73). Unlike histological testing, PET can monitor the entire tumor comprehensively and boasts higher sensitivity and specificity compared with MRI, with fewer artifacts (8). Presently, two primary tracers are employed for hypoxia detection: 18F-labeled nitroimidazole and Cu-labeled diacetylbis (N4-methylaminothiourea) analogs (74). The most widely used tracer is 18F-Fluoromisonidazole (18F-FMISO), a highly lipophilic 2-nitroimidazole analog. However, its slow plasma clearance rate necessitates prolonged imaging times. To overcome this, new hydrophilic tracers such as fluazomycin cytarabine and fluorinated ethyl imidazole derivatives have been developed. These tracers have faster plasma clearance rates, which enable quicker imaging and improve the efficiency and accuracy of hypoxia detection (75).

MRI includes techniques like blood oxygen level-dependent (BOLD) and tissue oxygen level-dependent (TOLD) imaging. Compared with PET, MRI has the advantages of high soft tissue contrast and no radiation exposure, making it particularly suitable for individuals with conditions that contraindicate radiation. These include pregnant women, children, patients with thyroid disorders and individuals with suppressed immune systems (75). In BOLD-MRI, contrast arises from the differing magnetic properties of oxyhemoglobin (diamagnetic) and deoxyhemoglobin (paramagnetic), reflecting local concentrations of oxygenated and deoxygenated hemoglobin (76). TOLD-MRI detects tissue hypoxia by measuring free oxygen molecules within tissues. As mitochondria consume oxygen, the remaining oxygen dissolves in the plasma, leading to an increase in longitudinal relaxation time (T1), thereby evaluating tissue oxygenation (77). While MRI is highly effective in the 3D mapping of tumor hypoxia, it is time-consuming and does not allow for real-time monitoring (78).

Another method for detecting oxygen levels is electron paramagnetic resonance imaging (EPRI). Analogous to MRI which depicts proton distribution, EPRI measures the spin distribution of unpaired electrons in free oxygen molecules (76). Oxygen molecules are paramagnetic and possess two unpaired electrons in their ground state. When oxygen molecules interact with probes, which are typically paramagnetic substances, the collision between the unpaired electrons of oxygen and the probe alters the energy state of the probe (79). Specifically, this interaction alters the probe's spin state, which consequently shifts its resonance frequency. EPRI detects these frequency shifts, enabling precise the recording of oxygen levels (80). EPRI exhibits high sensitivity and is less influenced by external factors, making it a promising tool in clinical research.

Gene and protein markers

In the hypoxic microenvironment of NSCLC, HIF-1α serves as a key transcription factor that regulates the expression of various hypoxia-responsive genes. Due to the short half-life of HIF-1α and its difficulty for use in direct labeling, studies typically indirectly evaluate hypoxic status by measuring gene expression levels regulated by HIF-1α. For instance, solute carrier family 2 member 1 (SLC2A1) is a target gene of HIF-1α, which enhances glycolytic activity by upregulating the expression of SLC2A1, contributing to energy generation and biosynthesis in tumor cells. The upregulation of SLC2A1 expression is strongly associated with the prognosis of patients with lung squamous cell carcinoma and can also serve as a key hypoxia biomarker (81). Similarly, Egl-9 family hypoxia inducible factor 2 (EGLN2) is a HIF-1α hydroxylase that regulates hypoxia by promoting the hydroxylation and degradation of HIF-1α. During hypoxia, the decrease in EGLN2 activity leads to the stabilization of HIF-1α, thereby activating the hypoxic pathway. The expression of EGLN2 may offer diagnostic and prognostic value in lung adenocarcinoma (82). Additionally, genes such as carbonic anhydrase 9 (CA9) and VEGF, which are widely expressed in NSCLC, are closely correlated with the degree of cellular hypoxia, tumor invasiveness and the effectiveness of radiotherapy and immunotherapy (83,84).

To detect hypoxia in tumors, studies commonly employ immunohistochemistry and immunofluorescence techniques to assess protein expression associated with hypoxia, primarily involving biomarkers like GLUT-1 (SLC2A1), CA9 (CA-IX), osteopontin (OPN) and nitroreductase (NTR). SLC2A1 supports glycolysis by enhancing glucose uptake, providing tumor cells with the energy and raw materials necessary for metabolism (85). CA-IX catalyzes the formation of carbonic acid from carbon dioxide and water, promoting extracellular acidification and thereby enhancing tumor invasiveness (86). OPN, as a hypoxia biomarker, is closely related to increased invasion and metastasis potential in NSCLC (87,88). NTR is an enzyme expressed only under hypoxic conditions and can reflect the degree of tumor hypoxia based on its activity level (78). In hypoxic environments, nitroimidazole is reduced by intracellular NTRs to generate reactive intermediates that disrupt DNA repair enzymes and cell cycle regulators, thereby inhibiting tumor cell survival (76). Currently, derivatives of 2-nitroimidazole, such as pimonidazole and EF-5, have become key tools for identifying and quantifying hypoxic regions within tumor tissues. By detecting specific antibodies like hypoxyprobe and ELK3-51, it is possible to efficiently map the distribution of hypoxia in tumors (76).

Targeted hypoxia in NSCLC

In NSCLC, the hypoxic microenvironment serves a crucial role in regulating tumor growth and invasion. Targeted therapy strategies aimed at this microenvironment have attracted widespread attention, as hypoxia not only affects the proliferation, invasion and metastasis of tumor cells but also activates a series of adaptive responses, such as angiogenesis, metabolic reprogramming and anti-apoptotic mechanisms. These responses enhance tumor invasiveness and lead to the development of therapeutic resistance. Therefore, targeted hypoxia therapy not only directly inhibits tumor progression but also has the potential to improve the efficacy of conventional treatment methods. Currently, several targeted therapies aimed at tumor hypoxia are being actively researched and developed (Table II).

Table II.

Summary of targeted hypoxia therapies for non-small cell lung cancer.

Table II.

Summary of targeted hypoxia therapies for non-small cell lung cancer.

Therapeutic approachesMethods and drug treatmentsMechanism of actionAdvantagesDisadvantages(Refs.)
Enhancing tissue oxygenationHyperbaric oxygen therapy, normobaric hyperoxia and aerobic exerciseIncreases oxygen levels in tumor tissues, inhibits cell proliferation and enhances therapeutic effectsIncreases oxygen supply and enhances the effect of radiotherapy and chemotherapyHigh-concentration oxygen induces oxygen toxicity, tissue damage and oxidative stresss and the efficacy of aerobic exercise remains unclear(7,8994)
Reduction of oxygen consumptionMetformin, Atovaquone, Papaverine, heme-segregating protein 2Inhibition of oxidative phosphorylation reduces the oxygen demand of tumor cellsImproves the tumor hypoxic environment and enhances treatment sensitivityCan cause metabolic toxicity and research on this therapy is currently underdeveloped(98102)
Restoration of blood flowNitroglycerin, and Cilengitide in combination with verapamilImproved tumor blood flow and enhanced drug and oxygen deliveryReduces glycolysis and improves distribution of chemotherapy and targeted drugsIndividual differences exist and excessive blood flow may promote tumor spread(104107)
Targeting pathways of HIF-1αAcriflavine, Polyamides, Temsirolimus, Trametinib and BevacizumabInhibits HIF-1α activity, targets HIF-1α signaling and inhibits angiogenesisInhibits tumor growth and improves the tumor hypoxic microenvironmentMonotherapy may induce resistance and the clinical efficacy and safety require further validation(108117)
Hypoxia-activated prodrugsTPZ, TH-302, PR-104, and EO9Targeting hypoxic tumor regionsPrecisely kills hypoxic tumor cells while minimizing normal tissue damageLimited reach to avascular areas and poor clinical performance of some drugs(121126)
Improvement of radiotherapyRadiosensitizers, carbon ion radiotherapy and dose paintingEnhances radiotherapy sensitivity, higher linear energy transfer and precise dose distributionMore precise and higher-dose radiotherapy efficacy and minimizes impact on normal cellsFurther clinical validation needed(127130)

[i] HIF, hypoxia inducible factor.

Enhancing tissue oxygenation

The goal of oxygen therapy is to increase the levels of dissolved oxygen in plasma and enhance tissue oxygenation, thereby slowing down tumor progression and improving treatment efficacy. Currently, various key strategies including hyperbaric oxygen therapy (HBO), normobaric hyperoxia (NBO) and aerobic exercise are being explored to augment tissue oxygen levels. HBO involves administering 100% oxygen at a pressure of 1.5–2.5 atmospheres absolute. Previous studies have shown that HBO can significantly reduce the expression level of HIF-1α and its downstream target genes, exert anti-angiogenic effects, reduce vascular density, inhibit tumor cell proliferation and thereby improve chemotherapy efficacy (7,89). NBO involves the delivery of high concentrations of oxygen via masks or shields, primarily by generating ROS to inhibit tumor growth (90). Although NBO is not as effective as HBO in enhancing the oxygenation of arteries and tumor tissues, it is considered more practically valuable due to its increased safety (91). However, the clinical use of HBO therapy faces certain challenges, mainly due to the possibility of causing some adverse reactions. These reactions include barotrauma to the middle ear, paranasal sinuses and lungs, ROS-induced damage to alveolar epithelium and pulmonary vascular endothelium and excessive production of nitric oxide (NO) due to aberrant activation of NO synthase. These mechanisms trigger pro-inflammatory pathways, such as the NF-κB pathway, amplifying local and systemic inflammation and worsening tissue damage (92).

Additionally, aerobic exercise has been shown to regulate the tumor vasculature and promote the relative normalization of dysfunctional tumor blood vessels (93). Jo et al (94) reported that sustained aerobic exercise improved lung function in NSCLC mice, increased oxygen saturation within tumor cells and elevated HIF-1α and ROS levels, thereby enhancing the efficacy of radiotherapy. In conclusion, oxygen-based therapies have potential in improving tumor oxygenation and enhancing treatment efficacy. However, a detailed assessment of the potential risks and benefits of this approach is particularly important for optimizing its clinical application.

Reduction of oxygen consumption rate

Tumor cells partially produce ATP through the mitochondrial oxidative phosphorylation (OXPHOS) pathway, in which hemoglobin serves a key role in the mitochondrial electron transport chain (ETC) complexes II–IV, thereby promoting ATP synthesis (95). In recent years, certain drugs approved by the US Food and Drug Administration (FDA), such as metformin, atovaquone and papaverine, have been shown to improve the hypoxic TME by inhibiting the OXPHOS pathway (96,97). Metformin binds to triphenylphosphine to generate MitoMet molecules, allowing the drug to accumulate in the mitochondria of tumor cells. MitoMet can inhibit mitochondrial complex I, thereby disrupting the function of ETC, increasing ROS production, inducing oxidative stress, leading to lipid peroxidation and ultimately causing tumor cell death (98). Atovaquone is an antimalarial drug that reduces hypoxia in NSCLC and improves cellular oxygenation by inhibiting complex III of OXPHOS (99). Papaverine is a smooth muscle relaxant and its derivative, papaverine pyrazole (PPV), temporarily reduces tumor hypoxia by reversibly inhibiting complex I, thereby enhancing the sensitivity of tumor cells to radiotherapy (100). Additionally, the inhibition mechanism of OXPHOS is closely associated with the action of heme-segregating proteins (HSPs), particularly HSP2. These proteins inhibit mitochondrial OXPHOS by limiting heme uptake in tumor cells, reducing oxygen consumption and ATP synthesis (101,102). Overall, targeting OXPHOS and related metabolic pathways offers a promising strategy for the treatment of NSCLC. Although these therapies have shown some potential, their toxicity and long-term efficacy still require further in-depth research to ensure their safety and effectiveness in clinical applications.

Restoration of blood flow

Acute hypoxia in tumors is often caused by vascular occlusion. Restoring blood flow can not only reduce the glycolytic metabolism of tumor cells but also improve the distribution of chemotherapy or targeted therapy, thereby significantly enhancing the therapeutic effect (103). Nitroglycerin, as a vasodilator, is primarily converted to NO through the mitochondrial aldehyde dehydrogenase and PI3K pathways, thereby inducing vasodilation, improving the oxygenation status of tumor tissue and enhancing the efficacy of anticancer drugs (104,105). Additionally, the Arg-Gly-Asp analogue cilengitide, can target αVβ3 and αVβ5 integrins on tumor ECs, regulate VEGFR2 expression and promote VEGF-dependent angiogenesis. Verapamil, a calcium channel blocker with vasodilatory effects, can enhance chemotherapy efficacy. When used in combination with cilengitide and gemcitabine, low-dose cilengitide significantly enhances the cytotoxic activity of gemcitabine against NSCLC cells (106). However, an excessive increase in blood flow may promote tumor growth, so the efficacy of this approach is highly dependent on tumor type and individual differences among patients. It is crucial to dynamically adjust treatment strategies within the time frame of vascular normalization to achieve optimal treatment outcomes (107).

Targeting upstream and downstream pathways of HIF-1α

HIF-1α serves a critical role in inflammation and hypoxia and strategies targeting HIF-1α have shown potential in reducing inhibiting the proliferation and invasion of NSCLC cells. Studies have demonstrated that Acriflavine binds to the Per-ARNT-Sim domain B subdomain of HIF-1α, thereby preventing the dimerization of the α and β subunits. In addition, polyamides and Echinomycin can inhibit the binding of the HIF-1 dimer to the HRE sequence of target genes. Compounds such as Chetomin, Bortezomib and Amphotericin B inhibit the transcriptional activity of HIF-1α through different mechanisms, while EZN-2968 suppresses the synthesis of the HIF-1α protein by blocking the translation process of HIF-1α mRNA (108). Collectively, these compounds effectively inhibit tumor growth and angiogenesis by modulating the activity of HIF-1α (Fig. 3B) (109).

Upstream pathways

Under hypoxic conditions, signaling pathways like PI3K-AKT-mTOR, RAS-RAF-MEK-ERK and JAK-STAT3 are activated, thereby promoting the synthesis of the HIF-1α protein (110). Specifically, the PI3K-AKT-mTOR pathway can enhance mRNA stability and protein synthesis efficiency, significantly increasing the transcription level of HIF-1α. Conversely, the RAS-RAF-MEK-ERK pathway phosphorylates the C-terminal activation domain of HIF-1α, facilitating its interaction with CBP/p300 and further enhancing HIF-1α-mediated gene transcription (111). Additionally, the STAT3 protein can be directly translocated to the nucleus to promote the transcription of HIF-1α and it can also work synergistically with the PI3K-AKT-mTOR pathway to trigger the transcription of HIF-1α through AKT-mediated mechanisms (110,112). Clinical trials have demonstrated that temsirolimus and rapamycin inhibit mTORC1 activity by binding to the FKBP prolyl isomerase 1A protein, while Voxtalisib exhibits broad antitumor activity by simultaneously inhibiting PI3K and mTOR targets (113). Additionally, the use of MEK inhibitors alone did not significantly improve the survival rate of patients with NSCLC; however, when combined with molecular targets such as VEGF, EGFR, ALK and BRAF, it shows significant clinical efficacy. The combination therapy of trametinib and dabrafenib has been approved by the FDA for the treatment of these tumors (Fig. 3A) (114).

Downstream pathways

VEGF is the only angiogenic factor consistently expressed throughout the entire tumor cycle and serves as a key downstream target gene of HIF-1α. When VEGF-A binds to VEGFR-2, it initiates angiogenesis signals, fostering the survival and proliferation of ECs and enhancing vascular permeability, all aiding in improving tumor resistance (115). Consequently, anti-angiogenesis has become a crucial strategy in the treatment of NSCLC. The primary focus of this approach is to disrupt the interactions between VEGF and VEGFR or interfere with the transmission of angiogenesis signals to normalize the tumor vasculature. Currently, anti-angiogenic therapies encompass a variety of drugs, including anti-VEGF antibodies (such as bevacizumab), anti-VEGFR antibodies (such as ramucirumab), VEGF trap receptors (such as aflibercept) and small-molecule VEGFR tyrosine kinase inhibitors (such as nintedanib, axitinib, sorafenib, sunitinib, vandetanib and cediranib) (116). However, the sole use of anti-angiogenic drugs may prompt tumor cells to develop hypoxia-tolerant phenotypes, thereby promoting vascular remodeling and tumor invasion, without significantly improving patient prognosis (117). Therefore, in clinical practice, it is more common to combine anti-angiogenic therapy with other treatments such as radiotherapy, chemotherapy or anti-EGFR therapy to enhance antitumor efficacy and prolong patient survival (Fig. 3C) (116).

Hypoxia-activated prodrugs (HAPs)

HAPs, or bioreductive alkylating agents, typically consist of five chemical groups: Nitro, quinone, aromatic, aliphatic and transition metals (118). Due to the rarity of chronic or severe hypoxia in normal tissues, HAPs can selectively target tumor cells by disrupting DNA replication in the hypoxic tumor environment (119). In vivo, HAPs are activated by a single electron reductase system, generating free radical intermediates. Under normoxic conditions, these intermediates are rapidly oxidized, lose their activity or transform into non-toxic products. However, in hypoxic environments, the lifespan of these intermediates is prolonged, leading to DNA alkylation, inducing cell apoptosis or necrosis and selectively inducing tumor cell death (75,120).

Although HAPs were originally designed for hypoxic tumor regions, they cannot reach areas far from the vascular network, which limits their clinical efficacy (121). For instance, although the classic HAP tirapazamine demonstrated effective hypoxic cell-killing effects in early experiments, the overall survival rate and response rate of patients with advanced or metastatic NSCLC did not significantly improve in phase III trials (122). This limitation was mainly attributed to the insufficient efficacy of tirapazamine against normoxic tumor cells (123). Currently, researchers are developing several new HAPs, such as TH-302 (evofosfamide), PR-104 and Alpraziquinone (EO9), to overcome these limitations and improve clinical efficacy (124). Combining HAPs with conventional chemotherapy or radiotherapy can improve treatment efficacy. For example, the combination of ezetophosamide and erlotinib has been shown to reduce the resistance of patients with NSCLC to erlotinib (125), while the combination of HAPs and radiotherapy may increase the sensitivity of hypoxic tumor cells, thereby improving overall antitumor efficacy (126).

Improvement of radiotherapy in hypoxic NSCLC

Radiotherapy is a key component of comprehensive treatment for various stages of NSCLC. However, to overcome its limitations, particularly in hypoxic tumors, there is a need to develop effective radiosensitization strategies that enhance accuracy and improve outcomes. Radiosensitizers primarily augment the effectiveness of radiotherapy by inhibiting radiation-induced DNA repair, disrupting cell cycle progression, affecting organelle function or modulating gene expression (49). For example, compounds containing nitrogen oxides and NO act as oxygen mimetics, thereby increasing tumor sensitivity to radiation in low-oxygen environments. These reagents intensify the interaction between free radicals and DNA, exacerbating DNA damage (127,128). Nanomaterials serve a significant role in enhancing radiosensitization through multiple mechanisms, including amplifying ROS production, modulating the cell cycle and stimulating apoptosis (129). Furthermore, the natural compound deguelin has shown promising radiosensitizing effects in NSCLC. It achieves this by inhibiting FBXO22 expression, downregulating the FOXM1/Rad51 pathway and consequently impairing the DNA damage repair capability of tumor cells (130).

Although radiosensitizers have enhanced the efficacy of conventional radiotherapy, their effectiveness is still constrained by the inability of conventional photon radiation to effectively target radiation-resistant tumor areas. To address this challenge, carbon ion radiotherapy (12C-ion) exhibits heightened tumor cell-killing effects due to its higher linear energy transfer (LET). The elevated LET of carbon ions leads to more localized deposition of radiation energy, resulting in complex and difficult-to-repair double-stranded DNA breaks, thereby increasing cell lethality (131). The advantages of carbon ion radiotherapy include: i) Increased relative biological effectiveness with radiation-killing effects 1.2–3.5 times greater than conventional photon or proton therapy at equivalent physical doses; ii) lower OER ensures effective therapeutic outcomes even in hypoxic TMEs; and iii) its weak dependence on the cell cycle renders it effective against S-phase tumor cells, which are typically resistant to photon radiation (132).

Despite the enhanced tumor-killing efficacy of carbon ion radiotherapy, its therapeutic potential needs to be optimized through precise dose distribution. Dose painting is an advanced technique that utilizes functional imaging techniques, such as PET and MRI, to personalize dose allocation based on specific tumor characteristics (133). The two primary strategies for dose painting include dose painting by contours (DPBC) and dose painting by numbers (DPBN). DPBC classifies tumors into high-risk, requiring higher doses, and low-risk areas, using conventional doses, based on imaging thresholds such as PET SUV values. Although clinically feasible, DPBC is constrained by variations in imaging thresholds, which affect treatment consistency. DPBN assigns continuous dose values to each voxel based on functional imaging data, more accurately reflecting tumor heterogeneity. However, this approach demands high image alignment precision and complex treatment planning (134).

In summary, the development of radiosensitization strategies and precise dose allocation techniques presents new opportunities for improving radiotherapy for NSCLC. The combination of efficient radiosensitizers and advanced radiotherapy approaches, such as carbon ion therapy and dose painting, may potentially optimize treatment strategies and enhance patient outcomes.

Discussion and conclusions

The present review summarized the role of the hypoxic microenvironment in the progression, metabolic reprogramming, immune escape and therapeutic resistance of NSCLC. Hypoxia is a key driving factor for tumor progression through multiple pathways. Specifically, HIF, as a central regulatory factor, serves a crucial role in the proliferation, angiogenesis and migration of tumor cells by regulating EMT and ECM degradation. Additionally, the hypoxic environment promotes intercellular communication through exosomes, thereby regulating processes such as glycolysis, angiogenesis, tumor invasion and immune evasion, which collectively promote the progression of NSCLC. At the metabolic level, hypoxia induces metabolic reprogramming, making NSCLC cells more inclined to rely on glycolytic pathways to generate lactate, thus meeting their energy demands and enhancing their adaptability to hypoxic conditions. The accumulation of lactate not only promotes tumor cell proliferation and immune escape but also supports tumor growth and metastasis by modulating the TME. Moreover, under hypoxic conditions, the DNA damage repair capability of tumor cells is reduced, and invasive phenotypes are more likely to form, thereby increasing resistance to radiotherapy, chemotherapy, immunotherapy and targeted therapy.

To address the challenges posed by hypoxic microenvironments, further studies should focus on hypoxia detection and intervention strategies. Existing methods for hypoxia detection include electrochemical probes and optical sensors for directly measuring oxygen partial pressure, as well as imaging techniques such as PET, MRI and EPRI for evaluating hypoxic regions within tumors. Additionally, gene markers, such as SLC2A1, EGLN2, VEGF and CA9, and protein markers, such as GLUT-1, CA9 and OPN, are currently used to detect tissue hypoxia. However, there are currently still a number of challenges in overcoming the heterogeneity of oxygen distribution within tumors and achieving real-time and accurate hypoxia monitoring. Future research should focus on developing more precise, non-invasive and real-time hypoxia detection technologies, which may have profound implications for clinical treatment. Regarding hypoxia treatment, recent studies have proposed various strategies to improve TME and enhance treatment efficacy. These strategies include increasing tissue oxygenation, reducing tissue oxygen consumption, restoring blood perfusion, downregulating HIF-related signaling pathways, using HAPs and improving radiotherapy efficacy. Although these strategies have shown some promise in preclinical studies, they still face challenges such as inadequate targeting and potential adverse reactions in clinical applications.

The current hypoxia detection and therapeutic strategies have brought new perspectives to tumor treatment, but the development of more accurate, real-time and non-invasive monitoring techniques remains the key research area. For example, combining high-resolution imaging technology with the detection of hypoxia-related biomarkers can provide a dynamic and effective method for monitoring tumor hypoxia in clinical settings. Additionally, integrating artificial intelligence and machine learning into image processing and data analysis is expected to further advance real-time monitoring of the tumor hypoxic microenvironment. In terms of treatment strategies, a single intervention may not be sufficient to improve tumor hypoxia effectively. Future research should explore multi-target approaches, combined with various therapeutic approaches, to enhance therapeutic response, alleviate tumor hypoxia and thus improve overall treatment efficacy. However, the clinical application of these strategies still faces many challenges, requiring larger-scale interdisciplinary collaboration and more clinical trials of hypoxia-targeted therapy to achieve true therapeutic breakthroughs.

In a hypoxic microenvironment, NSCLC cells undergo a series of adaptive changes, including enhanced growth and metabolism, angiogenesis to maintain nutrient supply, increased cell energy through activation of glycolysis-related genes and enhanced invasion and migration capabilities. These adaptive mechanisms enable tumor cells to escape the primary site and spread to distant organs, thereby making treatment more complex and contributing to the emergence of treatment resistance. Thus, identifying hypoxic signals in tissues is crucial for early diagnosis and disease monitoring of NSCLC. The best detection method should have the characteristics of non-invasiveness, reproducibility, high sensitivity and specificity and be able to accurately assess hypoxia status to guide clinical decision-making. Future therapeutic strategies should be combined with conventional therapies and optimized for hypoxia. By improving tissue oxygenation, modulating hypoxic genes and signaling pathways and applying selective drug delivery methods, it is expected to enhance the clinical treatment efficacy for patients with NSCLC.

Acknowledgments

Not applicable.

Funding

Funding: No funding was received.

Availability of data and materials

Not applicable.

Authors' contributions

All authors were involved in the research design and manuscript writing. SZ and JZ wrote the manuscript and prepared the figures and tables. WZ and ZY were involved in collecting and analyzing the literature for this review. PW and YZ critically revised the article for important ideological content. Data authentication is not applicable. All authors read and approved the final version of the manuscript.

Ethics approval and consent to participate

Not applicable.

Patient consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

References

1 

Sung H, Ferlay J, Siegel RL, Laversanne M, Soerjomataram I, Jemal A and Bray F: Global cancer statistics 2020: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J Clin. 71:209–249. 2021. View Article : Google Scholar : PubMed/NCBI

2 

Steeg PS: Targeting metastasis. Nat Rev Cancer. 16:201–218. 2016. View Article : Google Scholar : PubMed/NCBI

3 

Siegel RL, Miller KD, Fuchs HE and Jemal A: Cancer statistics, 2022. CA Cancer J Clin. 72:7–33. 2022. View Article : Google Scholar : PubMed/NCBI

4 

Xu FX, Zhang YL, Liu JJ, Zhang DD and Chen HB: Hypoxic markers in non-small cell lung cancer (NSCLC)-A review. Eur Rev Med Pharmacol Sci. 20:849–852. 2016.PubMed/NCBI

5 

Saxena K and Jolly MK: Acute vs. chronic vs. cyclic hypoxia: Their differential dynamics, molecular mechanisms, and effects on tumor progression. Biomolecules. 9:3392019. View Article : Google Scholar : PubMed/NCBI

6 

Bertout JA, Patel SA and Simon MC: The impact of O2 availability on human cancer. Nat Rev Cancer. 8:967–975. 2008. View Article : Google Scholar : PubMed/NCBI

7 

Meng X, Kong FM and Yu J: Implementation of hypoxia measurement into lung cancer therapy. Lung Cancer. 75:146–150. 2012. View Article : Google Scholar : PubMed/NCBI

8 

Challapalli A, Carroll L and Aboagye EO: Molecular mechanisms of hypoxia in cancer. Clin Transl Imaging. 5:225–253. 2017. View Article : Google Scholar : PubMed/NCBI

9 

Hinshaw DC and Shevde LA: The tumor microenvironment innately modulates cancer progression. Cancer Res. 79:4557–4566. 2019. View Article : Google Scholar : PubMed/NCBI

10 

Desai P, Takahashi N, Kumar R, Nichols S, Malin J, Hunt A, Schultz C, Cao Y, Tillo D, Nousome D, et al: Microenvironment shapes small-cell lung cancer neuroendocrine states and presents therapeutic opportunities. Cell Rep Med. 5:1016102024. View Article : Google Scholar : PubMed/NCBI

11 

Tam FF, Ning KL, Lee M, Dumlao JM and Choy JC: Cytokine induction of HIF-1α during normoxia in A549 human lung carcinoma cells is regulated by STAT1 and JNK signalling pathways. Mol Immunol. 160:12–19. 2023. View Article : Google Scholar : PubMed/NCBI

12 

Tirpe AA, Gulei D, Ciortea SM, Crivii C and Berindan-Neagoe I: Hypoxia: Overview on hypoxia-mediated mechanisms with a focus on the role of HIF genes. Int J Mol Sci. 20:61402019. View Article : Google Scholar : PubMed/NCBI

13 

Rankin EB and Giaccia AJ: The role of hypoxia-inducible factors in tumorigenesis. Cell Death Differ. 15:678–685. 2008. View Article : Google Scholar : PubMed/NCBI

14 

Masoud GN and Li W: HIF-1α pathway: Role, regulation and intervention for cancer therapy. Acta Pharm Sin B. 5:378–389. 2015. View Article : Google Scholar : PubMed/NCBI

15 

Della Rocca Y, Fonticoli L, Rajan TS, Trubiani O, Caputi S, Diomede F, Pizzicannella J and Marconi GD: Hypoxia: molecular pathophysiological mechanisms in human diseases. J Physiol Biochem. 78:739–752. 2022. View Article : Google Scholar : PubMed/NCBI

16 

Zeng SG, Lin X, Liu JC and Zhou J: Hypoxia-induced internalization of connexin 26 and connexin 43 in pulmonary epithelial cells is involved in the occurrence of non-small cell lung cancer via the P53/MDM2 signaling pathway. Int J Oncol. 55:845–859. 2019.PubMed/NCBI

17 

Hapke RY and Haake SM: Hypoxia-induced epithelial to mesenchymal transition in cancer. Cancer Lett. 487:10–20. 2020. View Article : Google Scholar : PubMed/NCBI

18 

Musleh Ud Din S, Streit SG, Huynh BT, Hana C, Abraham AN and Hussein A: Therapeutic targeting of hypoxia-inducible factors in cancer. Int J Mol Sci. 25:20602024. View Article : Google Scholar : PubMed/NCBI

19 

Liu S, Zhan Y, Luo J, Feng J, Lu J, Zheng H, Wen Q and Fan S: Roles of exosomes in the carcinogenesis and clinical therapy of non-small cell lung cancer. Biomed Pharmacother. 111:338–346. 2019. View Article : Google Scholar : PubMed/NCBI

20 

Wang C, Xu S and Yang X: Hypoxia-driven changes in tumor microenvironment: Insights into exosome-mediated cell interactions. Int J Nanomedicine. 19:8211–8236. 2024. View Article : Google Scholar : PubMed/NCBI

21 

Luo W, Hu H, Chang R, Zhong J, Knabel M, O'Meally R, Cole RN, Pandey A and Semenza GL: Pyruvate kinase M2 is a PHD3-stimulated coactivator for hypoxia-inducible factor 1. Cell. 145:732–744. 2011. View Article : Google Scholar : PubMed/NCBI

22 

Wei Y, Wang D, Jin F, Bian Z, Li L, Liang H, Li M, Shi L, Pan C, Zhu D, et al: Pyruvate kinase type M2 promotes tumour cell exosome release via phosphorylating synaptosome-associated protein 23. Nat Commun. 8:140412017. View Article : Google Scholar : PubMed/NCBI

23 

Ji X, Zhu R, Gao C, Xie H, Gong X and Luo J: Hypoxia-derived exosomes promote lung adenocarcinoma by regulating HS3ST1-GPC4-mediated glycolysis. Cancers (Basel). 16:6952024. View Article : Google Scholar : PubMed/NCBI

24 

Jiang H, Zhao H, Zhang M, He Y, Li X, Xu Y and Liu X: Hypoxia induced changes of exosome cargo and subsequent biological effects. Front Immunol. 13:8241882022. View Article : Google Scholar : PubMed/NCBI

25 

Jiang S, Wang R, Yan H, Jin L, Dou X and Chen D: MicroRNA-21 modulates radiation resistance through upregulation of hypoxia-inducible factor-1α-promoted glycolysis in non-small cell lung cancer cells. Mol Med Rep. 13:4101–4107. 2016. View Article : Google Scholar : PubMed/NCBI

26 

Becker V, Yuan X, Boewe AS, Ampofo E, Ebert E, Hohneck J, Bohle RM, Meese E, Zhao Y, Menger MD, et al: Hypoxia-induced downregulation of microRNA-186-5p in endothelial cells promotes non-small cell lung cancer angiogenesis by upregulating protein kinase C alpha. Mol Ther Nucleic Acids. 31:421–436. 2023. View Article : Google Scholar : PubMed/NCBI

27 

Zhao J, Qiao CR, Ding Z, Sheng YL, Li XN, Yang Y, Zhu DY, Zhang CY, Liu DL, Wu K and Zhao S: A novel pathway in NSCLC cells: miR-191, targeting NFIA, is induced by chronic hypoxia, and promotes cell proliferation and migration. Mol Med Rep. 15:1319–1325. 2017. View Article : Google Scholar : PubMed/NCBI

28 

Wang C, Han C, Zhang Y and Liu F: LncRNA PVT1 regulate expression of HIF1α via functioning as ceRNA for miR-199a-5p in non-small cell lung cancer under hypoxia. Mol Med Rep. 17:1105–1110. 2018.PubMed/NCBI

29 

Ren J: Intermittent hypoxia BMSCs-derived exosomal miR-31-5p promotes lung adenocarcinoma development via WDR5-induced epithelial mesenchymal transition. Sleep Breath. 27:1399–1409. 2023. View Article : Google Scholar : PubMed/NCBI

30 

Yin HL, Xu HW and Lin QY: miR129-1 regulates protein phosphatase 1D protein expression under hypoxic conditions in non-small cell lung cancer cells harboring a TP53 mutation. Oncol Lett. 20:2239–2247. 2020. View Article : Google Scholar : PubMed/NCBI

31 

Leone RD and Powell JD: Metabolism of immune cells in cancer. Nat Rev Cancer. 20:516–531. 2020. View Article : Google Scholar : PubMed/NCBI

32 

Yu Z, Zou J and Xu F: Tumor-associated macrophages affect the treatment of lung cancer. Heliyon. 10:e293322024. View Article : Google Scholar : PubMed/NCBI

33 

Jin J and Yu G: Hypoxic lung cancer cell-derived exosomal miR-21 mediates macrophage M2 polarization and promotes cancer cell proliferation through targeting IRF1. World J Surg Oncol. 20:2412022. View Article : Google Scholar : PubMed/NCBI

34 

Gu J, Yang S, Wang X, Wu Y, Wei J and Xu J: Hypoxic lung adenocarcinoma-derived exosomal miR-1290 induces M2 macrophage polarization by targeting SOCS3. Cancer Med. 12:12639–12652. 2023. View Article : Google Scholar : PubMed/NCBI

35 

Jackson AL, Zhou B and Kim WY: HIF, hypoxia and the role of angiogenesis in non-small cell lung cancer. Expert Opin Ther Targets. 14:1047–1057. 2010. View Article : Google Scholar : PubMed/NCBI

36 

Eales KL, Hollinshead KER and Tennant DA: Hypoxia and metabolic adaptation of cancer cells. Oncogenesis. 5:e1902016. View Article : Google Scholar : PubMed/NCBI

37 

Dai E, Wang W and Li Y, Ye D and Li Y: Lactate and lactylation: Behind the development of tumors. Cancer Lett. 591:2168962024. View Article : Google Scholar : PubMed/NCBI

38 

Courtnay R, Ngo DC, Malik N, Ververis K, Tortorella SM and Karagiannis TC: Cancer metabolism and the Warburg effect: the role of HIF-1 and PI3K. Mol Biol Rep. 42:841–851. 2015. View Article : Google Scholar : PubMed/NCBI

39 

Luo F, Liu X, Yan N, Li S, Cao G, Cheng Q, Xia Q and Wang H: Hypoxia-inducible transcription factor-1α promotes hypoxia-induced A549 apoptosis via a mechanism that involves the glycolysis pathway. BMC Cancer. 6:262006. View Article : Google Scholar : PubMed/NCBI

40 

Shang S, Wang MZ, Xing Z, He N and Li S: Lactate regulators contribute to tumor microenvironment and predict prognosis in lung adenocarcinoma. Front Immunol. 13:10249252022. View Article : Google Scholar : PubMed/NCBI

41 

Nisar H, Sanchidrián González PM, Brauny M, Labonté FM, Schmitz C, Roggan MD, Konda B and Hellweg CE: Hypoxia changes energy metabolism and growth rate in non-small cell lung cancer cells. Cancers (Basel). 15:24722023. View Article : Google Scholar : PubMed/NCBI

42 

Guo Z, Hu L, Wang Q, Wang Y, Liu XP, Chen C, Li S and Hu W: Molecular characterization and prognosis of lactate-related genes in lung adenocarcinoma. Curr Oncol. 30:2845–2861. 2023. View Article : Google Scholar : PubMed/NCBI

43 

Zhang X, Liang C, Wu C, Wan S and Xu L, Wang S, Wang J, Huang X and Xu L: A rising star involved in tumour immunity: Lactylation. J Cell Mol Med. 28:e701462024. View Article : Google Scholar : PubMed/NCBI

44 

Zhang C, Zhou L, Zhang M, Du Y, Li C, Ren H and Zheng L: H3K18 lactylation potentiates immune escape of non-small cell lung cancer. Cancer Res. 84:3589–3601. 2024. View Article : Google Scholar : PubMed/NCBI

45 

Porporato PE, Filigheddu N, Pedro JMB, Kroemer G and Galluzzi L: Mitochondrial metabolism and cancer. Cell Res. 28:265–280. 2018. View Article : Google Scholar : PubMed/NCBI

46 

Brustugun OT: Hypoxia as a cause of treatment failure in non-small cell carcinoma of the lung. Semin Radiat Oncol. 25:87–92. 2015. View Article : Google Scholar : PubMed/NCBI

47 

Huang RX and Zhou PK: DNA damage response signaling pathways and targets for radiotherapy sensitization in cancer. Signal Transduct Target Ther. 5:602020. View Article : Google Scholar : PubMed/NCBI

48 

Chen H, Han Z, Luo Q, Wang Y, Li Q, Zhou L and Zuo H: Radiotherapy modulates tumor cell fate decisions: A review. Radiat Oncol. 17:1962022. View Article : Google Scholar : PubMed/NCBI

49 

Gong L, Zhang Y, Liu C, Zhang M and Han S: Application of radiosensitizers in cancer radiotherapy. Int J Nanomedicine. 16:1083–1102. 2021. View Article : Google Scholar : PubMed/NCBI

50 

Zhou T, Zhang LY, He JZ, Miao ZM, Li YY, Zhang YM, Liu ZW, Zhang SZ, Chen Y, Zhou GC and Liu YQ: Review: Mechanisms and perspective treatment of radioresistance in non-small cell lung cancer. Front Immunol. 14:11338992023. View Article : Google Scholar : PubMed/NCBI

51 

Gray LH, Conger AD, Ebert M, Hornsey S and Scott OC: The concentration of oxygen dissolved in tissues at the time of irradiation as a factor in radiotherapy. Br J Radiol. 26:638–648. 1953. View Article : Google Scholar : PubMed/NCBI

52 

Herrera-Campos AB, Zamudio-Martinez E, Delgado-Bellido D, Fernández-Cortés M, Montuenga LM, Oliver FJ and Garcia-Diaz A: Implications of hyperoxia over the tumor microenvironment: An overview highlighting the importance of the immune system. Cancers (Basel). 14:27402022. View Article : Google Scholar : PubMed/NCBI

53 

Guo Q, Lan F, Yan X, Xiao Z, Wu Y and Zhang Q: Hypoxia exposure induced cisplatin resistance partially via activating p53 and hypoxia inducible factor-1α in non-small cell lung cancer A549 cells. Oncol Lett. 16:801–808. 2018.PubMed/NCBI

54 

Roy S, Kumaravel S, Sharma A, Duran CL, Bayless KJ and Chakraborty S: Hypoxic tumor microenvironment: Implications for cancer therapy. Exp Biol Med (Maywood). 245:1073–1086. 2020. View Article : Google Scholar : PubMed/NCBI

55 

Hu CY, Hung CF, Chen PC, Hsu JY, Wang CT, Lai MD, Tsai YS, Shiau AL, Shieh GS and Wu CL: Oct4 and hypoxia dual-regulated oncolytic adenovirus armed with shRNA-targeting dendritic cell immunoreceptor exerts potent antitumor activity against bladder cancer. Biomedicines. 11:25982023. View Article : Google Scholar : PubMed/NCBI

56 

Mancino A, Schioppa T, Larghi P, Pasqualini F, Nebuloni M, Chen IH, Sozzani S, Austyn JM, Mantovani A and Sica A: Divergent effects of hypoxia on dendritic cell functions. Blood. 112:3723–3734. 2008. View Article : Google Scholar : PubMed/NCBI

57 

Peng X, He Y, Huang J, Tao Y and Liu S: Metabolism of dendritic cells in tumor microenvironment: For immunotherapy. Front Immunol. 12:6134922021. View Article : Google Scholar : PubMed/NCBI

58 

Eltzschig HK, Thompson LF, Karhausen J, Cotta RJ, Ibla JC, Robson SC and Colgan SP: Endogenous adenosine produced during hypoxia attenuates neutrophil accumulation: Coordination by extracellular nucleotide metabolism. Blood. 104:3986–3992. 2004. View Article : Google Scholar : PubMed/NCBI

59 

Li J, Wang L, Chen X, Li L, Li Y, Ping Y, Huang L, Yue D, Zhang Z, Wang F, et al: CD39/CD73 upregulation on myeloid-derived suppressor cells via TGF-β-mTOR-HIF-1 signaling in patients with non-small cell lung cancer. Oncoimmunology. 6:e13200112017. View Article : Google Scholar : PubMed/NCBI

60 

An SM, Lei HM, Ding XP, Sun F, Zhang C, Tang YB, Chen HZ, Shen Y and Zhu L: Interleukin-6 identified as an important factor in hypoxia- and aldehyde dehydrogenase-based gefitinib adaptive resistance in non-small cell lung cancer cells. Oncol Lett. 14:3445–3454. 2017. View Article : Google Scholar : PubMed/NCBI

61 

Kogita A, Togashi Y, Hayashi H, Sogabe S, Terashima M, De Velasco MA, Sakai K, Fujita Y, Tomida S, Takeyama Y, et al: Hypoxia induces resistance to ALK inhibitors in the H3122 non-small cell lung cancer cell line with an ALK rearrangement via epithelial-mesenchymal transition. Int J Oncol. 45:1430–1436. 2014. View Article : Google Scholar : PubMed/NCBI

62 

Karan S, Cho MY, Lee H, Lee H, Park HS, Sundararajan M, Sessler JL and Hong KS: Near-infrared fluorescent probe activated by nitroreductase for in vitro and in vivo hypoxic tumor detection. J Med Chem. 64:2971–2981. 2021. View Article : Google Scholar : PubMed/NCBI

63 

Cheng MHY, Mo Y and Zheng G: Nano versus molecular: Optical imaging approaches to detect and monitor tumor hypoxia. Adv Healthc Mater. 10:e20015492021. View Article : Google Scholar : PubMed/NCBI

64 

Clark LC Jr, Wolf R, Granger D and Taylor Z: Continuous recording of blood oxygen tensions by polarography. J Appl Physiol. 6:189–193. 1953. View Article : Google Scholar : PubMed/NCBI

65 

Sun X, Niu G, Chan N, Shen B and Chen X: Tumor hypoxia imaging. Mol Imaging Biol. 13:399–410. 2011. View Article : Google Scholar : PubMed/NCBI

66 

Shirmanova MV, Lukina MM, Sirotkina MA, Shimolina LE, Dudenkova VV, Ignatova NI, Tobita S, Shcheslavskiy VI and Zagaynova EV: Effects of photodynamic therapy on tumor metabolism and oxygenation revealed by fluorescence and phosphorescence lifetime imaging. Int J Mol Sci. 25:17032024. View Article : Google Scholar : PubMed/NCBI

67 

Vanderkooi JM, Maniara G, Green TJ and Wilson DF: An optical method for measurement of dioxygen concentration based upon quenching of phosphorescence. J Biol Chem. 262:5476–5482. 1987. View Article : Google Scholar : PubMed/NCBI

68 

Koch CJ and Evans SM: Optimizing hypoxia detection and treatment strategies. Semin Nucl Med. 45:163–176. 2015. View Article : Google Scholar : PubMed/NCBI

69 

Horsman MR, Sørensen BS, Busk M and Siemann DW: Therapeutic modification of hypoxia. Clin Oncol (R Coll Radiol). 33:e492–e509. 2021. View Article : Google Scholar : PubMed/NCBI

70 

Huang C, Liang J, Lei X, Xu X, Xiao Z and Luo L: Diagnostic performance of perfusion computed tomography for differentiating lung cancer from benign lesions: A meta-analysis. Med Sci Monit. 25:3485–3494. 2019. View Article : Google Scholar : PubMed/NCBI

71 

Hu X, Gou J, Wang L, Lin W, Li W and Yang F: Diagnostic accuracy of low-dose dual-input computed tomography perfusion in the differential diagnosis of pulmonary benign and malignant ground-glass nodules. Sci Rep. 14:170982024. View Article : Google Scholar : PubMed/NCBI

72 

Liao C, Liu X, Zhang C and Zhang Q: Tumor hypoxia: From basic knowledge to therapeutic implications. Semin Cancer Biol. 88:172–186. 2023. View Article : Google Scholar : PubMed/NCBI

73 

Nasri D, Manwar R, Kaushik A, Er EE and Avanaki K: Photoacoustic imaging for investigating tumor hypoxia: A strategic assessment. Theranostics. 13:3346–3367. 2023. View Article : Google Scholar : PubMed/NCBI

74 

Krohn KA, Link JM and Mason RP: Molecular imaging of hypoxia. J Nucl Med. 49 (Suppl 2):129S–148S. 2008. View Article : Google Scholar : PubMed/NCBI

75 

Brender JR, Saida Y, Devasahayam N, Krishna MC and Kishimoto S: Hypoxia imaging as a guide for hypoxia-modulated and hypoxia-activated therapy. Antioxid Redox Signal. 36:144–159. 2022. View Article : Google Scholar : PubMed/NCBI

76 

Godet I, Doctorman S, Wu F and Gilkes DM: Detection of hypoxia in cancer models: Significance, challenges, and advances. Cells. 11:6862022. View Article : Google Scholar : PubMed/NCBI

77 

Matsumoto K, Mitchell JB and Krishna MC: Multimodal functional imaging for cancer/tumor microenvironments based on MRI, EPRI, and PET. Molecules. 26:16142021. View Article : Google Scholar : PubMed/NCBI

78 

Qin W, Xu C, Zhao Y, Yu C, Shen S, Li L and Huang W: Recent progress in small molecule fluorescent probes for nitroreductase. Chin Chem Lett. 29:1451–1455. 2018. View Article : Google Scholar

79 

Vikram DS, Zweier JL and Kuppusamy P: Methods for noninvasive imaging of tissue hypoxia. Antioxid Redox Signal. 9:1745–1756. 2007. View Article : Google Scholar : PubMed/NCBI

80 

Epel B, Bowman MK, Mailer C and Halpern HJ: Absolute oxygen R1e imaging in vivo with pulse electron paramagnetic resonance. Magn Reson Med. 72:362–368. 2014. View Article : Google Scholar : PubMed/NCBI

81 

Hao B, Dong H, Xiong R, Song C, Xu C, Li N and Geng Q: Identification of SLC2A1 as a predictive biomarker for survival and response to immunotherapy in lung squamous cell carcinoma. Comput Biol Med. 171:1081832024. View Article : Google Scholar : PubMed/NCBI

82 

Zhang R, Lai L, He J, Chen C, You D, Duan W, Dong X, Zhu Y, Lin L, Shen S, et al: EGLN2 DNA methylation and expression interact with HIF1A to affect survival of early-stage NSCLC. Epigenetics. 14:118–129. 2019. View Article : Google Scholar : PubMed/NCBI

83 

Ostheimer C, Bache M, Güttler A, Kotzsch M and Vordermark D: A pilot study on potential plasma hypoxia markers in the radiotherapy of non-small cell lung cancer. Osteopontin, carbonic anhydrase IX and vascular endothelial growth factor. Strahlenther Onkol. 190:276–282. 2014. View Article : Google Scholar : PubMed/NCBI

84 

Giatromanolaki A, Harris AL, Banham AH, Contrafouris CA and Koukourakis MI: Carbonic anhydrase 9 (CA9) expression in non-small-cell lung cancer: Correlation with regulatory FOXP3+T-cell tumour stroma infiltration. Br J Cancer. 122:1205–1210. 2020. View Article : Google Scholar : PubMed/NCBI

85 

Geng H, Chen L, Lv S, Li M, Huang X, Li M and Liu C and Liu C: Photochemically controlled release of the glucose transporter 1 inhibitor for glucose deprivation responses and cancer suppression research. J Proteome Res. 23:653–662. 2024. View Article : Google Scholar : PubMed/NCBI

86 

Kim SJ, Rabbani ZN, Vollmer RT, Schreiber EG, Oosterwijk E, Dewhirst MW, Vujaskovic Z and Kelley MJ: Carbonic anhydrase IX in early-stage non-small cell lung cancer. Clin Cancer Res. 10:7925–7933. 2004. View Article : Google Scholar : PubMed/NCBI

87 

Coppola D, Szabo M, Boulware D, Muraca P, Alsarraj M, Chambers AF and Yeatman TJ: Correlation of osteopontin protein expression and pathological stage across a wide variety of tumor histologies. Clin Cancer Res. 10((1 Pt 1)): 184–190. 2004. View Article : Google Scholar : PubMed/NCBI

88 

Ostheimer C, Bache M, Güttler A, Reese T and Vordermark D: Prognostic information of serial plasma osteopontin measurement in radiotherapy of non-small-cell lung cancer. BMC Cancer. 14:8582014. View Article : Google Scholar : PubMed/NCBI

89 

Stępień K, Ostrowski RP and Matyja E: Hyperbaric oxygen as an adjunctive therapy in treatment of malignancies, including brain tumours. Med Oncol. 33:1012016. View Article : Google Scholar : PubMed/NCBI

90 

Kim SW, Kim IK, Ha JH, Yeo CD, Kang HH, Kim JW and Lee SH: Normobaric hyperoxia inhibits the progression of lung cancer by inducing apoptosis. Exp Biol Med (Maywood). 243:739–748. 2018. View Article : Google Scholar : PubMed/NCBI

91 

Thews O and Vaupel P: Spatial oxygenation profiles in tumors during normo- and hyperbaric hyperoxia. Strahlenther Onkol. 191:875–882. 2015. View Article : Google Scholar : PubMed/NCBI

92 

Heyboer M III, Sharma D, Santiago W and McCulloch N: Hyperbaric oxygen therapy: Side effects defined and quantified. Adv Wound Care (New Rochelle). 6:210–224. 2017. View Article : Google Scholar : PubMed/NCBI

93 

Jones LW, Viglianti BL, Tashjian JA, Kothadia SM, Keir ST, Freedland SJ, Potter MQ, Moon EJ, Schroeder T, Herndon JE II and Dewhirst MW: Effect of aerobic exercise on tumor physiology in an animal model of human breast cancer. J Appl Physiol (1985). 108:343–348. 2010. View Article : Google Scholar : PubMed/NCBI

94 

Jo S, Jeon J, Park G, Do HK, Kang J, Ahn KJ, Ma SY, Choi YM, Kim D, Youn B and Ki Y: Aerobic exercise improves radiation therapy efficacy in non-small cell lung cancer: Preclinical study using a xenograft mouse model. Int J Mol Sci. 25:27572024. View Article : Google Scholar : PubMed/NCBI

95 

Ghosh P, Vidal C, Dey S and Zhang L: Mitochondria targeting as an effective strategy for cancer therapy. Int J Mol Sci. 21:33632020. View Article : Google Scholar : PubMed/NCBI

96 

Ashton TM, McKenna WG, Kunz-Schughart LA and Higgins GS: Oxidative phosphorylation as an emerging target in cancer therapy. Clin Cancer Res. 24:2482–2490. 2018. View Article : Google Scholar : PubMed/NCBI

97 

Kalyanaraman B, Cheng G, Hardy M and You M: OXPHOS-targeting drugs in oncology: New perspectives. Expert Opin Ther Targets. 27:939–952. 2023. View Article : Google Scholar : PubMed/NCBI

98 

Shameem M, Bagherpoor AJ, Nakhi A, Dosa P, Georg G and Kassie F: Mitochondria-targeted metformin (mitomet) inhibits lung cancer in cellular models and in mice by enhancing the generation of reactive oxygen species. Mol Carcinog. 62:1619–1629. 2023. View Article : Google Scholar : PubMed/NCBI

99 

Skwarski M, McGowan DR, Belcher E, Di Chiara F, Stavroulias D, McCole M, Derham JL, Chu KY, Teoh E, Chauhan J, et al: Mitochondrial inhibitor atovaquone increases tumor oxygenation and inhibits hypoxic gene expression in patients with non-small cell lung cancer. Clin Cancer Res. 27:2459–2469. 2021. View Article : Google Scholar : PubMed/NCBI

100 

Benej M, Hong X, Vibhute S, Scott S, Wu J, Graves E, Le QT, Koong AC, Giaccia AJ, Yu B, et al: Papaverine and its derivatives radiosensitize solid tumors by inhibiting mitochondrial metabolism. Proc Natl Acad Sci USA. 115:10756–10761. 2018. View Article : Google Scholar : PubMed/NCBI

101 

Sohoni S, Ghosh P, Wang T, Kalainayakan SP, Vidal C, Dey S, Konduri PC and Zhang L: Elevated heme synthesis and uptake underpin intensified oxidative metabolism and tumorigenic functions in non-small cell lung cancer cells. Cancer Res. 79:2511–2525. 2019. View Article : Google Scholar : PubMed/NCBI

102 

Wang T, Ashrafi A, Konduri PC, Ghosh P, Dey S, Modareszadeh P, Salamat N, Alemi PS, Berisha E and Zhang L: Heme sequestration as an effective strategy for the suppression of tumor growth and progression. Mol Cancer Ther. 20:2506–2518. 2021. View Article : Google Scholar : PubMed/NCBI

103 

Dewhirst MW and Secomb TW: Transport of drugs from blood vessels to tumour tissue. Nat Rev Cancer. 17:738–750. 2017. View Article : Google Scholar : PubMed/NCBI

104 

Sukhatme V, Bouche G, Meheus L, Sukhatme VP and Pantziarka P: Repurposing drugs in oncology (ReDO)-nitroglycerin as an anti-cancer agent. Ecancermedicalscience. 9:5682015. View Article : Google Scholar : PubMed/NCBI

105 

Reymen BJT, van Gisbergen MW, Even AJG, Zegers CML, Das M, Vegt E, Wildberger JE, Mottaghy FM, Yaromina A, Dubois LJ, et al: Nitroglycerin as a radiosensitizer in non-small cell lung cancer: Results of a prospective imaging-based phase II trial. Clin Transl Radiat Oncol. 21:49–55. 2019.PubMed/NCBI

106 

Wong PP, Bodrug N and Hodivala-Dilke KM: Exploring novel methods for modulating tumor blood vessels in cancer treatment. Curr Biol. 26:R1161–R1166. 2016. View Article : Google Scholar : PubMed/NCBI

107 

Guelfi S, Hodivala-Dilke K and Bergers G: Targeting the tumour vasculature: From vessel destruction to promotion. Nat Rev Cancer. 24:655–675. 2024. View Article : Google Scholar : PubMed/NCBI

108 

Wigerup C, Påhlman S and Bexell D: Therapeutic targeting of hypoxia and hypoxia-inducible factors in cancer. Pharmacol Ther. 164:152–169. 2016. View Article : Google Scholar : PubMed/NCBI

109 

Xia Y, Choi HK and Lee K: Recent advances in hypoxia-inducible factor (HIF)-1 inhibitors. Eur J Med Chem. 49:24–40. 2012. View Article : Google Scholar : PubMed/NCBI

110 

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

111 

Albadari N, Deng S and Li W: The transcriptional factors HIF-1 and HIF-2 and their novel inhibitors in cancer therapy. Expert Opin Drug Discov. 14:667–682. 2019. View Article : Google Scholar : PubMed/NCBI

112 

Ma J, Cao K, Ling X, Zhang P and Zhu J: LncRNA HAR1A suppresses the development of non-small cell lung cancer by inactivating the STAT3 pathway. Cancers (Basel). 14:28452022. View Article : Google Scholar : PubMed/NCBI

113 

Deng H, Chen Y, Li P, Hang Q, Zhang P, Jin Y and Chen M: PI3K/AKT/mTOR pathway, hypoxia, and glucose metabolism: Potential targets to overcome radioresistance in small cell lung cancer. Cancer Pathog Ther. 1:56–66. 2022. View Article : Google Scholar : PubMed/NCBI

114 

Lara MS, Blakely CM and Riess JW: Targeting MEK in non-small cell lung cancer. Curr Probl Cancer. 49:1010652024. View Article : Google Scholar : PubMed/NCBI

115 

Zhu H and Zhang S: Hypoxia inducible factor-1α/vascular endothelial growth factor signaling activation correlates with response to radiotherapy and its inhibition reduces hypoxia-induced angiogenesis in lung cancer. J Cell Biochem. 119:7707–7718. 2018. View Article : Google Scholar : PubMed/NCBI

116 

Tian W, Cao C, Shu L and Wu F: Anti-angiogenic therapy in the treatment of non-small cell lung cancer. Onco Targets Ther. 13:12113–12129. 2020. View Article : Google Scholar : PubMed/NCBI

117 

Mahdi A, Darvishi B, Majidzadeh-A K, Salehi M and Farahmand L: Challenges facing antiangiogenesis therapy: The significant role of hypoxia-inducible factor and MET in development of resistance to anti-vascular endothelial growth factor-targeted therapies. J Cell Physiol. 234:5655–5663. 2019. View Article : Google Scholar : PubMed/NCBI

118 

Yuan X, Xie Z and Zou T: Recent advances in hypoxia-activated compounds for cancer diagnosis and treatment. Bioorg Chem. 144:1071612024. View Article : Google Scholar : PubMed/NCBI

119 

Bryant JL, Meredith SL, Williams KJ and White A: Targeting hypoxia in the treatment of small cell lung cancer. Lung Cancer. 86:126–132. 2014. View Article : Google Scholar : PubMed/NCBI

120 

Wilson WR and Hay MP: Targeting hypoxia in cancer therapy. Nat Rev Cancer. 11:393–410. 2011. View Article : Google Scholar : PubMed/NCBI

121 

Chen SX, Zhang J, Xue F, Liu W, Kuang Y, Gu B, Song S and Chen H: In situ forming oxygen/ROS-responsive niche-like hydrogel enabling gelation-triggered chemotherapy and inhibition of metastasis. Bioact Mater. 21:86–96. 2022.PubMed/NCBI

122 

Shepherd F, Koschel G, Von Pawel J, Gatzmeier U, Van Zandwiyk N, Woll P, Van Klavren R, Krasko P, Desimone P, Nicolson M, et al: Comparison of Tirazone (Tirapazamine) and cisplatin vs. etoposide and cisplatin in advanced non-small cell lung cancer (NSCLC): Final results of the International phase III CATAPULT II trial. Lung Cancer. 29:282000. View Article : Google Scholar

123 

Marcu L and Olver I: Tirapazamine: From bench to clinical trials. Curr Clin Pharmacol. 1:71–79. 2006. View Article : Google Scholar : PubMed/NCBI

124 

Graham K and Unger E: Overcoming tumor hypoxia as a barrier to radiotherapy, chemotherapy and immunotherapy in cancer treatment. Int J Nanomedicine. 13:6049–6058. 2018. View Article : Google Scholar : PubMed/NCBI

125 

Lindsay D, Garvey CM, Mumenthaler SM and Foo J: Leveraging hypoxia-activated prodrugs to prevent drug resistance in solid tumors. PLoS Comput Biol. 12:e10050772016. View Article : Google Scholar : PubMed/NCBI

126 

Meaney C, Powathil GG, Yaromina A, Dubois LJ, Lambin P and Kohandel M: Role of hypoxia-activated prodrugs in combination with radiation therapy: An in silico approach. Math Biosci Eng. 16:6257–6273. 2019. View Article : Google Scholar : PubMed/NCBI

127 

Oronsky BT, Knox SJ and Scicinski JJ: Is nitric oxide (NO) the last word in radiosensitization? A review. Transl Oncol. 5:66–71. 2012. View Article : Google Scholar : PubMed/NCBI

128 

Oronsky B, Scicinski J, Ning S, Peehl D, Oronsky A, Cabrales P, Bednarski M and Knox S: RRx-001, A novel dinitroazetidine radiosensitizer. Invest New Drugs. 34:371–377. 2016. View Article : Google Scholar : PubMed/NCBI

129 

Zhao L, Li M, Shen C and Luo Y, Hou X, Qi Y, Huang Z, Li W, Gao L, Wu M and Luo Y: Nano-assisted radiotherapy strategies: New opportunities for treatment of non-small cell lung cancer. Research (Wash D C). 7:04292024.PubMed/NCBI

130 

Chen Y, Zhou Y, Feng X, Wu Z, Yang Y, Rao X, Zhou R, Meng R, Dong X, Xu S, et al: Targeting FBXO22 enhances radiosensitivity in non-small cell lung cancer by inhibiting the FOXM1/Rad51 axis. Cell Death Dis. 15:1042024. View Article : Google Scholar : PubMed/NCBI

131 

Dey P, Das R, Chatterjee S, Paul R and Ghosh U: Combined effects of carbon ion radiation and PARP inhibitor on non-small cell lung carcinoma cells: Insights into DNA repair pathways and cell death mechanisms. DNA Repair (Amst). 144:1037782024. View Article : Google Scholar : PubMed/NCBI

132 

Demizu Y, Fujii O, Iwata H and Fuwa N: Carbon ion therapy for early-stage non-small-cell lung cancer. Biomed Res Int. 2014:7279622014. View Article : Google Scholar : PubMed/NCBI

133 

Bentzen SM and Gregoire V: Molecular-imaging-based dose painting: A novel paradigm for radiation therapy prescription. Semin Radiat Oncol. 21:101–110. 2011. View Article : Google Scholar : PubMed/NCBI

134 

Meijer G, Steenhuijsen J, Bal M, De Jaeger K, Schuring D and Theuws J: Dose painting by contours versus dose painting by numbers for stage II/III lung cancer: Practical implications of using a broad or sharp brush. Radiother Oncol. 100:396–401. 2011. View Article : Google Scholar : PubMed/NCBI

Related Articles

Journal Cover

February-2025
Volume 53 Issue 2

Print ISSN: 1021-335X
Online ISSN:1791-2431

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Zhou S, Sun J, Zhu W, Yang Z, Wang P and Zeng Y: Hypoxia studies in non‑small cell lung cancer: Pathogenesis and clinical implications (Review). Oncol Rep 53: 29, 2025.
APA
Zhou, S., Sun, J., Zhu, W., Yang, Z., Wang, P., & Zeng, Y. (2025). Hypoxia studies in non‑small cell lung cancer: Pathogenesis and clinical implications (Review). Oncology Reports, 53, 29. https://doi.org/10.3892/or.2024.8862
MLA
Zhou, S., Sun, J., Zhu, W., Yang, Z., Wang, P., Zeng, Y."Hypoxia studies in non‑small cell lung cancer: Pathogenesis and clinical implications (Review)". Oncology Reports 53.2 (2025): 29.
Chicago
Zhou, S., Sun, J., Zhu, W., Yang, Z., Wang, P., Zeng, Y."Hypoxia studies in non‑small cell lung cancer: Pathogenesis and clinical implications (Review)". Oncology Reports 53, no. 2 (2025): 29. https://doi.org/10.3892/or.2024.8862