Abstract
Background
   High-altitude pulmonary edema (HAPE), a severe manifestation of
   hypoxia-induced pulmonary hypertension, continues to present a major
   health concern in high-altitude environments due to the absence of
   efficient preventive measures. This investigation explores the
   protective influence of ginsenoside Rg3 (G-Rg3), an active substance
   derived from the botanical drug Panax ginseng C.A.Mey., on the
   prevention of HAPE progression.
Methods
   A mouse model mimicking exposure to 6000-m altitude (n = 63 C57BL/6
   mice) was employed to evaluate the impact of G-Rg3 (15/30 mg/kg) using
   histopathological, biochemical, and multi-dimensional molecular
   assessments. Western blotting, network pharmacology and computational
   simulations were utilized to identify molecular targets of G-Rg3. The
   role of the PI3K/AKT signaling pathway was further validated through
   experiments using the PI3K/AKT inhibitor LY294002.
Results
   Pre-treatment with G-Rg3 effectively alleviated HAPE, maintained the
   stability of lung ultrastructure, and inhibited inflammatory mediators
   and oxidative stress indicators. Mechanistically, G-Rg3 prevented
   ferroptosis by stimulating the PI3K/AKT signaling pathway, as evidenced
   by the upregulation of protective proteins (GPX4, Nrf2, HO-1, SLC7A11,
   FTH1, FLC) and the downregulation of iron metabolism regulatory factors
   (TFRC, COX2). Network pharmacology and molecular docking analysis
   confirmed that PI3K/AKT is the core target of G-Rg3, and the protective
   effect disappeared when this pathway was inhibited. G-Rg3 uniquely
   regulated oxidative stress and inflammation by inhibiting ferroptosis,
   demonstrating adaptability to high-altitude environments.
Conclusion
   This research examined the pharmacological impacts and molecular
   pathways of ginseng active monomers on HAPE, suggesting the potential
   of G-Rg3 as a promising treatment option for this condition.
   Keywords: high-altitude pulmonary edema (HAPE), ginsenoside Rg3
   (G-Rg3), PI3K/Akt pathway, ferroptosis, hypoxic pulmonary hypertension
1 Introduction
   High-altitude pulmonary edema (HAPE) represents a critical form of
   hypoxic pulmonary hypertension, typically emerging in settings with
   diminished atmospheric pressure and oxygen levels ([42]Bärtsch and
   Swenson, 2013; [43]Persson and Bondke Persson, 2017). Extended exposure
   to high-altitude hypoxia is the primary driver of HAPE, which, without
   intervention, can evolve into chronic hypoxic pulmonary hypertension.
   Clinically, HAPE is marked by non-cardiogenic pulmonary edema and
   worsening hypoxemia, with an untreated fatality rate estimated at
   nearly 50% ([44]Bouzat et al., 2013; [45]Li et al., 2024; [46]Menon,
   1965). Individual susceptibility to hypoxic stress and the development
   of HAPE may be influenced by genetic variations, particularly in
   mitochondrial DNA and components of the renin-angiotensin-aldosterone
   system (RAAS) ([47]Bhagi et al., 2015; [48]Luo et al., 2012; [49]Sharma
   et al., 2019; [50]Srivastava et al., 2012). Furthermore, hypoxic
   conditions can induce oxidative stress via mitochondrial dysfunction
   and disruptions in metabolic pathways, thereby amplifying abnormalities
   in the pulmonary vasculature ([51]Ali et al., 2012; [52]Sharma et al.,
   2021). Importantly, in high-altitude environments, cold temperatures
   may interact synergistically with hypoxia to exacerbate cardiopulmonary
   dysfunction, thus elevating the risk of HAPE ([53]Eichstaedt et al.,
   2023). These underlying pathophysiological mechanisms ultimately result
   in abnormal pulmonary vascular reactions and sustained pulmonary
   hypertension, leading to the formation of edema.
   In recent years, there has been a growing emphasis on the use of
   natural medicines for treating HAPE, particularly those derived from
   botanical drugs with lung-protective active metabolites. Recent
   clinical studies have highlighted the significant potential of
   traditional botanical drugs, such as ginseng, in preventing and
   managing HAPE through multi-target mechanisms ([54]Huang et al., 2024;
   [55]Shen et al., 2023). Panax ginseng C.A.Mey. (known as Renshen in
   Chinese), a classic botanical drug historically used for respiratory
   enhancement, exemplifies this potential through its anti-inflammatory,
   antioxidant, and cellular protective properties ([56]Ratan et al.,
   2021). One notable metabolite, ginsenoside Rg3 (G-Rg3), demonstrates
   potential efficacy in mitigating hypoxic lung injury ([57]Bae et al.,
   2014; [58]Chen et al., 2010; [59]Wang et al., 2016). Its mechanism of
   action involves promoting cell survival by activating the PI3K/AKT
   signaling pathway ([60]Yang et al., 2018). Our previous study
   demonstrated that intraperitoneal administration of G-Rg3 (15/30 mg/kg)
   significantly mitigated acute mountain sickness (AMS) in C57BL/6 mice
   through ferroptosis regulation ([61]Liu et al., 2025a). These
   advancements not only underscore the scientific significance of
   traditional medicinal resources but also offer a theoretical foundation
   for developing plant-based therapies against high-altitude hypoxia.
   Oxidative stress plays a central role in the pathogenesis of HAPE.
   Under normoxic conditions, endogenous antioxidant systems, including
   catalase, glutathione (GSH), and superoxide dismutase (SOD), maintain
   redox balance by neutralizing reactive oxygen species (ROS)
   ([62]Diaz-Vivancos et al., 2015; [63]Sies, 2017). Hypoxia disrupts this
   balance through two main mechanisms: impaired electron transport chain
   function due to oxygen deficiency increases ROS production ([64]Gaur et
   al., 2021), while simultaneous depletion of antioxidant reserves
   exacerbates oxidative damage to lipids, proteins, and DNA ([65]Wang et
   al., 2022b). This dual disruption creates a self-perpetuating cycle of
   oxidative injury, eventually overwhelming cellular repair mechanisms
   and triggering redox imbalance ([66]Yu et al., 2021).
   The resulting oxidative stress induces ferroptosis, an iron-dependent
   form of programmed cell death marked by three key features: (1)
   membrane rupture via lipid peroxidation cascades, (2) intracellular
   iron overload, and (3) failure of the glutathione peroxidase 4 (GPX4)
   system ([67]Dixon et al., 2012; [68]Yang and Stockwell, 2016).
   Mechanistically, hypoxia-induced ROS overproduction initiates lipid
   peroxidation through Fenton reactions, while iron accumulation
   amplifies oxidative damage by catalyzing hydroxyl radical formation
   ([69]Djulbegovic and Uversky, 2019). These processes compromise
   membrane integrity and inactivate iron-regulatory proteins,
   establishing a pathological feedback loop. Recent evidence links
   ferroptosis to the activation of Phosphoinositide 3-kinase/Protein
   Kinase B (PI3K/AKT) and Mitogen-Activated Protein Kinase (MAPK)
   signaling pathways, which are involved in both cellular survival
   decisions and hypoxic adaptation ([70]Gao et al., 2025; [71]Guo et al.,
   2022). Notably, preclinical models demonstrate ferroptosis involvement
   in hypoxia-induced organ damage, including neurological impairment
   following acute altitude exposure ([72]Han et al., 2025), suggesting
   its potential as a therapeutic target in HAPE pathophysiology.
   In view of the aforementioned research context, our study puts forward
   the hypothesis that G-Rg3 may prevent HAPE by suppressing ferroptosis.
   To verify this hypothesis, we developed a HAPE model induced by
   hypobaric hypoxia in C57BL/6 mice. The therapeutic efficacy of G-Rg3 on
   HAPE was assessed using multiple approaches, including quantitative
   analysis of pulmonary edema, histopathological scoring, profiling of
   inflammatory cytokines, and evaluation of oxidative stress biomarkers.
   To investigate the mechanistic regulation of ferroptosis by G-Rg3, we
   employed an integrated strategy combining computational techniques
   (network pharmacology, molecular docking, and molecular dynamics
   simulations) with experimental validation methods: immunofluorescence
   detection of ferroptosis markers, transmission electron microscopy for
   mitochondrial ultrastructure analysis, and Western blot examination of
   PI3K/AKT signaling pathway components.
   This holistic approach not only facilitates the identification of
   therapeutic targets involved in ferroptosis-driven progression of HAPE
   but also establishes a theoretical foundation for G-Rg3 as a promising
   candidate for treating this condition.
2 Materials and methods
2.1 Materials
   Ginsenoside Rg3 (G-Rg3, ≥99.15% purity, Cat.14197-60-5) was procured
   from Must Biotechnology (Chengdu, China). The PI3K/AKT pathway
   inhibitor LY294002 (Cat.S1105) and all analytical-grade chemicals were
   obtained from Selleck Chemicals (United States) and commercial
   suppliers, respectively. Oxidative stress parameters were measured
   using Nanjing Jiancheng kits: Malondialdehyde (MDA, A003-1), GSH
   (A006-2-1), SOD (A001-3), and tissue iron quantification (A039-2-1).
   Proinflammatory cytokines Interleukin-1 beta (IL-1β, ZC-37974W),
   Interleukin 6 (IL-6, ZC-37988W), and Tumor necrosis factor-alpha
   (TNF-α, ZC-39024W) were analyzed with Zhucai Biotechnology ELISA kits
   (Shanghai).
   Antibody Specifications Immunoblotting employed the following primary
   antibodies: Abcam (UK): Vascular Endothelial Growth Factor (VEGF,
   ab32152), Hypoxia-inducible factor 1 alpha (HIF-1α, ab179483), Nuclear
   Factor Erythroid 2-Related Factor 2 (Nrf2, ab92946), Heme Oxygenase-1
   (HO-1, ab68477), GPX4 (ab125066), Ferritin light chain (FLC, ab75973),
   Transferrin receptor gene (TFRC, ab269513), PI3K/AKT pathway components
   (PI3K ab191606, AKT ab185633, p-PI3K ab182651, p-AKT ab192623), β-actin
   (ab227387). Cell Signaling Technology (United States): Ferritin Heavy
   Chain 1 (FTH1, #4393S), Cyclooxygenase-2 (COX2, #12282S). ABclonal
   (Wuhan): Solute Carrier Family 7 Member 11 (SLC7A11, A2413) Secondary
   detection used HRP-conjugated goat anti-rabbit IgG (BF03008, Biodragon
   Biotech).
   All supplementary chemicals meeting analytical-grade specifications.
2.2 Animals
   Sixty-three male C57BL/6 mice (6–8 weeks, 18–22 g, specific
   pathogen-free) were sourced from Chengdu Dasuo Biological Technology
   (Certification SCXK 2022–0345). Following 7-day acclimation under
   controlled conditions (22°C ± 1°C, 55% ± 5% humidity; 12-hr light/dark
   cycle), animals received standardized feeding with autoclaved water and
   chow ([73]Zhang et al., 2025a). The study received ethical approval
   (2022-18) from Chengdu University of Traditional Chinese Medicine’s
   Animal Welfare Experimental Center, with all procedures complying with
   institutional ethics guidelines and national welfare legislation.
2.3 Experimental design
   Animals were randomly divided into five experimental subgroups (n = 7
   per group): (1) Sham control, (2) G-Rg3 monotherapy (30 mg/kg), (3)
   HAPE model, (4) HAPE + G-Rg3-L (15 mg/kg), and (5) HAPE + G-Rg3-H
   (30 mg/kg), with G-Rg3 dosages validated by prior pharmacological
   studies ([74]Cheng and Li, 2016; [75]Heinrich et al., 2020; [76]Liu et
   al., 2025a). While Sham and HAPE groups received intraperitoneal (i.p.)
   phosphate-buffered saline (PBS), other cohorts were administered G-Rg3
   via i. p. injection for 72 consecutive days. To investigate
   PI3K/AKT-ferroptosis regulatory mechanisms, an additional cohort of 28
   mice underwent randomization into four groups: (1) HAPE baseline, (2)
   HAPE + G-Rg3-H (30 mg/kg), (3) HAPE + LY294002 (5 mg/kg PI3K/AKT
   inhibitor), and (4) HAPE + G-Rg3-H/LY294002 combinatorial therapy.
   LY294002 formulations utilized a solvent system containing 50%
   distilled water, 40% PEG300, 5% Tween 80, and 5% DMSO. All
   interventions employed standardized i. p. administration protocols
   (5 mL/kg injection volume, daily dosing over 3 days).
2.4 HAPE modelling
   To replicate the HAPE condition in mice, we employed a hypobaric
   hypoxic chamber model ProOx-830 from Tawang Intelligent Technology
   (Shanghai, China), following the protocols outlined in our previous
   research and relevant publications ([77]Ma et al., 2020; [78]Tan et
   al., 2020; [79]Wang et al., 2022b). After a 3-day administration
   period, the mice were transferred to a hypobaric hypoxia chamber set at
   an altitude of 6000 m, with an oxygen partial pressure of 9.6 kPa,
   humidity of 60%, and temperature of 20°C. The animals were rapidly
   elevated to this altitude within 5 min at a rate of 20 m per second and
   maintained there for 48 h.
   Following the 48-h modeling period, the altitude was gradually adjusted
   to normal levels, and then the mice were taken out of the chamber and
   euthanized via an i. p. injection of sodium pentobarbital. Following
   euthanasia, blood was collected from the abdominal aorta for serum
   extraction, bronchoalveolar lavage fluid (BALF) was obtained, and lung
   tissues were sectioned for additional analyses. The experimental
   protocol for the HAPE animal model is depicted in [80]Figure 1.
FIGURE 1.
   [81]Diagram illustrating a research process involving mice. Mice
   undergo G-Rg3 pretreatment for three days, subjected to hypobaric
   hypoxia, followed by sample analysis. Methods include H&E/TEM, Western
   blot for oxidative stress and cytokines, and IF for tissue iron.
   Samples are collected from lungs and analyzed post-euthanasia.
   [82]Open in a new tab
   The flowchart of HAPE animal experiments.
2.5 Histology of lung tissue
   Post-euthanasia pulmonary specimens underwent standardized
   histoprocessing: The right middle lung lobe was immediately harvested,
   cleared of extraneous tissue, and immersion-fixed in ice-cold 4%
   paraformaldehyde (12 h, 4°C). Sequential ethanol dehydration preceded
   paraffin embedding, with resultant blocks sectioned at 4 μm thickness.
   Following hematoxylin-eosin (H&E) staining, slides were imaged using an
   OLYMPUS BX41 microscope. Two professional pathologists blindly scored
   the degree of lung tissue damage using the McGuigan pathology scoring
   method ([83]Faller et al., 2012; [84]McGuigan et al., 2003).
2.6 Lung wet/dry (W/D) weight ratio
   The right upper lung lobe was surgically harvested post-euthanasia for
   hydrostatic evaluation. Fresh tissue mass (W) was immediately recorded
   before dehydration at 60°C until mass stabilization (D), enabling
   calculation of the W/D ratio - a validated indicator of alveolar fluid
   accumulation severity in hypobaric pulmonary edema models.
2.7 Cytokine content in BALF
   Following right lung ligation, the left lung underwent dual
   bronchoalveolar lavage procedures utilizing 0.2 mL ice-cold PBS
   administered via tracheal cannulation. Pooled lavage fluid was
   centrifuged (12,000×g, 10 min, 4°C) to separate cellular components,
   with the resultant supernatant aliquoted for cryopreservation at −80°C
   pending cytokine analysis. Quantitative assessment of IL-1β, IL-6, and
   TNF-α concentrations in BALF was conducted through enzyme-linked
   immunosorbent assay (ELISA) following manufacturer-specified protocols
   (Thermo Fisher Scientific).
2.8 Oxidative stress and iron content in lung tissues
   The right posterior lobe of the lung was surgically removed and
   perfused with sterile saline prior to mechanical homogenization in PBS.
   Cellular debris was eliminated by centrifugation at 12,000×g for 15 min
   at 4°C, allowing us to collect clear supernatants for further
   biochemical evaluation. Key markers of oxidative stress, including MDA
   levels, SOD activity, and GSH concentrations, as well as parameters
   related to iron metabolism, were quantified using commercially
   available assay kits. Absorbance values were measured using a Thermo
   Scientific Varioskan LUX microplate reader. To ensure accurate
   quantitative comparisons across samples, total protein content was
   normalized using the bicinchoninic acid (BCA) protein assay method.
2.9 Western blotting
   The right inferior pulmonary lobe was immediately excised from
   euthanized mice and flash-frozen in liquid nitrogen prior to
   cryostorage at −80°C for subsequent molecular characterization. Tissue
   homogenates prepared in RIPA buffer containing protease/phosphatase
   inhibitors (1 mM PMSF) underwent centrifugation (12,000×g, 10 min,
   4°C), with supernatants subjected to BCA protein quantification (Pierce
   Biotechnology) per manufacturer specifications. Aliquots (30 μg)
   underwent electrophoretic separation on 8%–15% gradient SDS-PAGE gels
   (Bio-Rad Laboratories) followed by semi-dry transfer to PVDF membranes.
   Post-blocking with 5% BSA/TBST (1 h, 25°C), membranes were probed with
   primary antibodies at 4°C (16 h) targeting: VEGF (1:2000), HIF-1α
   (1:1000), Nrf2 (1:1000), GPX4 (1:5000), HO-1 (1:10,000), FLC (1:1000),
   TFRC (1:5000), PI3K (1:1000), AKT (1:2000), p-PI3K (1:800), p-AKT
   (1:1000), FTH1 (1:1000), COX2 (1:1000), SLC7A11 (1:2000) and β-actin
   (1:10,000)as normalization control. Post-TBST washes, HRP-conjugated
   secondary antibodies (1:5000, 2 h, 25°C) enabled chemiluminescent
   detection via ECL substrate, with band visualization (GelView 6000Plus)
   and densitometric quantification (Image-Pro Plus 6.0) relative to
   β-actin expression.
2.10 Transmission electron microscopy (TEM)
   The excised apex of the right pulmonary lobe underwent dual fixation
   protocol involving primary stabilization in 3% glutaraldehyde (24 h,
   4°C) followed by secondary osmication with 1% osmium tetroxide (2 h,
   25°C). Processed through acetone-gradient dehydration and epoxy resin
   (Epon 812) polymerization, ultrastructural specimens were microtomed at
   70 nm thickness. Transmission electron microscopy was conducted using a
   JEOL JEM-1400-FLASH system operating at 120 kV, with contrast
   enhancement achieved through sequential uranyl acetate and lead citrate
   staining regimens.
2.11 Immunofluorescence staining
   Paraffin-embedded lung tissue sections (4 μm) underwent sequential
   processing through xylene deparaffinization and graded alcohol
   rehydration. For antigen retrieval, heat-mediated treatment was
   performed in 0.01 M citrate buffer (pH 6.0) for 10 min. Subsequent
   pre-treatment included 15-min endogenous peroxidase blockade with 0.3%
   H[2]O[2] and 20-min non-specific binding inhibition using 10% normal
   goat serum. Primary antibody incubation proceeded overnight at 4°C with
   anti-GPX4 (1:100) and anti-p-PI3K (1:100), followed by 30-min room
   temperature exposure to species-matched secondary antibodies. Nuclear
   counterstaining was achieved through 10-min DAPI application (1 μg/mL)
   at 25°C, with residual dye removal via triple PBS washing (5 min each).
   Fluorescent signal acquisition utilized a Nikon C2 confocal system
   (Tokyo, Japan), with subsequent quantitative analysis performed using
   ImageJ (NIH, United States) for fluorescence intensity measurements.
2.12 Network pharmacology
2.12.1 Target genes acquisition
   The identification of G-Rg3 target genes incorporated a multi-platform
   strategy utilizing both computational prediction tools and
   pharmacological databases. Initial screening was performed through the
   Encyclopedia of Traditional Chinese Medicine (ETCM,
   [85]http://www.tcmip.cn/ETCM), TargetNet
   ([86]http://targetnet.scbdd.com), Swiss Target Prediction
   ([87]http://www.swisstargetprediction.ch), and PharmMapper
   ([88]http://www.lilab-ecust.cn/pharmmapper). To compensate for
   potential database limitations, literature-curated targets from prior
   ligand-receptor interaction studies were integrated ([89]Wang et al.,
   2022a). Concurrently, HAPE-associated genes were systematically
   collated from four disease-specific repositories: GeneCards
   ([90]https://www.genecards.org), OMIM ([91]https://omim.org), CTD
   ([92]http://ctdbase.org), and DisGeNET ([93]http://www.disgenet.org).
   All identified genes underwent nomenclature standardization via a
   bioinformatics pipeline combining UniProt ([94]https://www.uniprot.org)
   for sequence annotation and STRING ([95]https://string-db.org) for
   orthology mapping, ensuring compliance with HUGO Gene Nomenclature
   Committee guidelines ([96]Yin et al., 2024).
2.12.2 Development of the protein-protein interaction (PPI) network
   Following deduplication, gene set intersections between pharmacological
   targets and pathological associations were graphically represented
   through Venn diagram construction. Hub gene identification employed PPI
   network modeling via the STRING platform (organism: Homo sapiens;
   confidence threshold ≥0.7). The resultant interactome data underwent
   advanced topological analysis using Cytoscape v3.6.2, facilitating
   three-dimensional visualization of the metabolite-target-disease triad
   and systematic interrogation of network architecture through integrated
   bioinformatics tools.
2.12.3 Gene pathway analysis
   Following the identification of key genes in the network through
   topological analysis, a multi-level functional assessment was carried
   out using an integrated bioinformatics approach (with the
   clusterProfiler package). This framework executed Gene Ontology (GO)
   enrichment analysis for cellular process annotation and Kyoto
   Encyclopedia of Genes and Genomes (KEGG) pathway mapping to
   contextualize biological systems simultaneously. As a result, it
   enabled a comprehensive functional evaluation connecting molecular
   functions, biological processes, and pathways.
2.13 Molecular docking
2.13.1 Molecular acquisition
   The PI3Kγ structural coordinates for molecular docking were acquired
   from the RCSB Protein Data Bank (accession code 3ML9) ([97]Burley et
   al., 2017), while the bioactive conformation of Ginsenoside Rg3 was
   obtained from PubChem (CID: 12855989) and energetically optimized
   through MMFF94 molecular mechanics refinement ([98]Sunghwan et al.,
   2016).
   Molecular docking utilized AutoDock Vina version 1.2, with receptor
   proteins preprocessed in PyMOL 2.5 to remove water, salt ions, and
   other small molecules beforehand ([99]Eberhardt et al., 2021). The
   central coordinates for the docking box were determined by the
   center_of_mass.py plugin, using the active site as a reference, with
   each side of the box being 22.5 Å long. Moreover, all preprocessed
   small molecules and receptor proteins were altered into the PDBQT
   format necessary for docking with AutoDock Vina 1.2, utilizing ADFR
   Suite 1.0 ([100]Pradeep et al., 2015). The global docking search
   exhaustiveness was configured to 32, whereas the remaining parameters
   were kept at their default values. The highest-ranked docking
   conformation was selected as the bound structure for follow-up
   molecular dynamics simulations in this investigation.
2.13.2 Molecule dynamics
   Informed by the docking results, complexes derived from small molecules
   and proteins served as the initial configurations for extensive
   molecular dynamics simulations, which were performed using the AMBER 22
   software ([101]Salomon-Ferrer et al., 2013). Prior to initiating the
   simulations, the partial charges of the small molecules were computed
   using the Antechamber module and the Hartree-Fock (HF) SCF/6-31G*
   method via the Gaussian 09 software ([102]Frisch et al., 2009;
   [103]Wang et al., 2005). For the small molecules, the GAFF2 force field
   was utilized, while the ff14SB force field was applied for the proteins
   ([104]Maier et al., 2015; [105]Wang J. et al., 2024; [106]Wang X. et
   al., 2004). For each system configuration, the LEaP module was employed
   to add hydrogen atoms. A truncated octahedral TIP3P water box, with a
   10 Å buffer around the system, was then positioned. Additionally,
   Na^+/Cl^− ions were introduced to neutralize any charge imbalances
   ([107]Mark and Nilsson, 2001). Ultimately, the topology and parameter
   files needed for the simulations were produced.
   Using the AMBER 22 software package, molecular dynamics simulations
   were executed ([108]Salomon-Ferrer et al., 2013). An energy
   minimization process was applied to the system before initiating the
   simulation that consisted of 2500 iterations of steepest descent,
   succeeded by a further 2500 iterations employing the conjugate gradient
   technique. Upon finishing the energy minimization, the system underwent
   a controlled temperature increase from 0 K to 298.15 K over 200
   picoseconds at a uniform rate, maintaining a constant volume throughout
   this process. To ensure a homogeneous dispersion of solvent molecules
   within the solvent box and maintain a temperature of 298.15 K, an NVT
   ensemble simulation was performed for a duration of 500 picoseconds.
   Subsequently, to stabilize the system, an equilibrium simulation under
   NPT conditions was carried out for an additional 500 picoseconds
   throughout the entire system. The composite system underwent an
   extended simulation lasting 100 nanoseconds within an NPT ensemble,
   with periodic boundary conditions applied consistently. During this
   period, non-bonded interactions were truncated at a cutoff distance of
   10 Å. Long-range electrostatic interactions were evaluated using the
   Particle Mesh Ewald (PME) method ([109]Sagui and Darden, 1999). The
   SHAKE algorithm was utilized to fix the lengths of hydrogen bonds
   ([110]Kräutler et al., 2015), while Langevin dynamics with with a
   collision frequency of γ = 2 ps^-1 was used for temperature control
   ([111]Larini et al., 2007). The pressure was maintained at 1 atm,
   integration time steps were established at 2 femtoseconds, and
   trajectory data were captured every 10 picoseconds for further
   analysis.
2.13.3 MM/GBSA binding free energy calculation
   The MM/GBSA method was applied to evaluate the binding free energy
   between the protein and ligand across all systems ([112]Chen et al.,
   2020; [113]Genheden and Ryde, 2015; [114]Hou et al., 2011;
   [115]Rastelli et al., 2010). Acknowledging that extended MD simulations
   can significantly affect the accuracy of MM/GBSA outcomes is crucial
   ([116]Hou et al., 2011). Hence, this investigation used an MD
   trajectory covering 90–100 nanoseconds for the calculations, based on
   the formula described below:
   [MATH:
   ΔGbi
   nd=ΔGcomplex 
   – ΔGreceptor+
   mo>ΔGligand
   mfenced> :MATH]
   [MATH: =ΔEinternal+
   mo>ΔEVDW+ΔEelec+ΔGGB+ΔGSA
   :MATH]
   In the formula, internal energy is denoted by ΔE[internal], van der
   Waals interaction by ΔE[VDW], and electrostatic interaction by
   ΔE[elec]. The internal energies, which include bond energy (E[bond]),
   angle energy (E[angle]), and torsion energy (E[torsion]), are
   collectively known as solvation free energy. Here, ΔG[GB] represents
   polar solvation free energy, while ΔG[SA] denotes non-polar solvation
   free energy. In this study, the ΔG[GB] model developed by Nguyen et al.
   (igb = 2) was utilized for calculations ([117]Nguyen et al., 2013). To
   calculate the non-polar solvation free energy (ΔG[SA]), the surface
   tension (γ) was multiplied by the solvent-accessible surface area
   (ΔSASA), given by the formula ΔG[SA] = 0.0072 × ΔSASA ([118]Weiser et
   al., 1999). Entropy change calculations were not included in this study
   because they require significant computational resources and have low
   accuracy ([119]Chen et al., 2020; [120]Hou et al., 2011).
2.14 Statistical analyses
   This research employed GraphPad Prism 8 (San Diego, California, United
   States) for data analysis and figure generation. For normally
   distributed datasets, one-way analysis of variance was conducted. The
   findings are expressed as mean ± SEM. A threshold of P < 0.05 was used
   to establish statistical significance.
3 Results
3.1 Effects of G-Rg3 pre-treatment on the prevention of HAPE
   In instances of acute high-altitude pulmonary edema, patients typically
   exhibit clinical symptoms such as breathing difficulties, cyanosis of
   the mucous membranes, delayed reactions, and reduced activity levels
   ([121]Gatterer et al., 2024). Consequently, we monitored the overall
   condition of mice to evaluate the pathological impacts of low-pressure
   hypoxia on the body, as well as the effectiveness of drug
   interventions. In their baseline state, all mice displayed normal
   physiological signs, characterized by glossy fur, steady respiration,
   and a heightened stress response. Following exposure to low-pressure
   hypoxia, distinct variations emerged among the groups: both the NC
   group and the G-Rg3 groups maintained normal activity levels, with
   regular breathing and feeding patterns; in contrast, the HAPE group
   demonstrated classic pathological features, including mucosal cyanosis,
   rapid breathing, and lethargic responses. The two dosage levels of
   G-Rg3 exhibited dose-related improvements, particularly the 30 mg/kg
   group, which showed markedly enhanced recovery in terms of activity,
   mucosal coloration, and respiratory rhythm.
   Histopathological analysis demonstrated that lung tissues in both the
   sham operation group and the G-Rg3 pre-treatment group remained
   structurally intact. Conversely, lung tissues in the HAPE group
   exhibited significant damage, including thickened alveolar walls,
   widened alveolar septa, alveolar cavity congestion, and prominent
   inflammatory cell infiltration ([122]Figure 2A). These observations
   confirm the successful establishment of the rat model for high-altitude
   pulmonary edema. Following G-Rg3 pre-treatment, the extent of
   inflammatory cell infiltration, hemorrhage, and alveolar wall
   thickening was dose-dependently attenuated. Semi-quantitative
   assessment of lung tissue injury revealed that G-Rg3 pre-treatment
   markedly decreased the lung injury score associated with HAPE
   ([123]Figure 2B) ([124]McGuigan et al., 2003). Additionally, the
   wet/dry weight ratio of lung tissue was substantially elevated in the
   HAPE group compared to the sham operation group. However, this ratio
   was significantly lowered following G-Rg3 pre-treatment ([125]Figure
   2C).
FIGURE 2.
   [126]Panel A shows lung tissue histological sections under different
   conditions: Sham, G-Rg3, HAPE, G-Rg3-L, and G-Rg3-H, at magnifications
   of 200x and 400x. Panels B and C present bar graphs illustrating the
   lung injury score and W/D ratio for these conditions. Panel D includes
   Western blot images for VEGF and HIF-1α with bar graphs indicating
   their relative expression levels.
   [127]Open in a new tab
   Effects of G-Rg3 pre-treatment on the prevention of HAPE. (A)
   H&E-stained lung sections (upper: ×200, 50 μm scale; lower: ×400, 20 μm
   scale). (B) Semi-quantified lung injury scores. (C) Hydrostatic
   imbalance via W/D weight ratio. (D) Hypoxic response markers VEGF and
   HIF-1α by immunoblotting. Data represent mean ± SEM (n = 6 biological
   replicates). One-way ANOVA with Tukey’s post hoc analysis: ^##/### P <
   0.01/0.001 vs. Sham; ^*/**/*** P < 0.05/0.01/0.001 vs. HAPE controls.
   As a pivotal transcription factor, HIF-1α modulates the expression of
   genes involved in hypoxic responses and is crucial for the
   physiological adaptation of organisms to hypoxic settings
   ([128]Diaz-Vivancos et al., 2015). Elevated levels of HIF-1α have been
   conspicuously observed in individuals prone to HAPE under normoxic
   conditions. Moreover, empirical evidence indicates that HIF-1α
   contributes to the progression of HAPE through the regulation of its
   target gene, VEGF. Western blotting showed increased expression of VEGF
   and HIF-1α in the HAPE group, which was significantly attenuated
   following G-Rg3 pre-treatment ([129]Figure 2D). The results showed that
   G-Rg3 pre-treatment could effectively reduce the inflammation and
   oxidative stress in the HAPE model.
3.2 Effects of G-Rg3 pre-treatment on oxidative stress and inflammatory
cytokines by HAPE
   HAPE progression is mechanistically linked to acute immune-inflammatory
   cascade activation, characterized by neutrophilic infiltration,
   macrophage aggregation, and mediator-induced alveolar-capillary
   hyperpermeability ([130]Lee et al., 2018; [131]Wang et al., 2022b). Our
   experimental analyses revealed significant upregulation of
   pro-inflammatory IL-1β, IL-6, and TNF-α in BALF during hypobaric
   hypoxia exposure, with prophylactic G-Rg3 administration demonstrating
   potent cytokine suppression efficacy ([132]Figure 3A). Concomitant
   evaluation of pulmonary redox homeostasis identified characteristic
   oxidative imbalance patterns: HAPE-induced samples exhibited heightened
   lipid peroxidation (elevated MDA), compromised antioxidant defenses
   (reduced SOD and GSH), and tissue iron overload - all metabolic
   disturbances effectively normalized through G-Rg3 pre-treatment, as
   evidenced by complete restoration of baseline oxidative stress markers
   and iron homeostasis parameters ([133]Figure 3B).
FIGURE 3.
   [134]Bar graphs labeled A and B show various biochemical parameters. A
   displays IL-1β, IL-6, and TNF-α levels across five groups: Sham, G-Rg3,
   HAPE, G-Rg3-L, and G-Rg3-H. B shows GSH, MDA, SOD, and tissue iron
   levels in the same groups. Significant differences are indicated by
   asterisks and hash marks. Each graph uses different y-axes units
   relevant to the parameter measured.
   [135]Open in a new tab
   Effects of G-Rg3 pre-treatment on oxidative stress and inflammatory
   cytokines by HAPE. (A) G-Rg3-mediated suppression of proinflammatory
   cytokines (IL-1β/IL-6/TNF-α). (B) Antioxidant regulation through
   SOD/GSH/MDA biomarkers. Data reflect mean ± SEM (n = 6 biological
   replicates). One-way ANOVA with Tukey’s post hoc analysis: ^##/### P <
   0.01/0.001 vs. Sham controls; ^*/** P < 0.05/0.01 vs. HAPE group.
3.3 Effects of G-Rg3 pre-treatment on ferroptosis in HAPE
   Given that excessive oxidative stress responses can trigger ferroptosis
   ([136]Djulbegovic and Uversky, 2019), we next investigated the
   involvement of ferroptosis in HAPE pathogenesis. Western blot analysis
   indicated that the expression of anti-ferroptosis proteins, including
   GPX4, Nrf2, HO-1, SLC7A11, FTH1, and FLC, was markedly reduced in model
   tissues, whereas pro-ferroptosis proteins COX2 and TFRC exhibited
   elevated expression ([137]Figure 4A). Notably, pre-treatment with G-Rg3
   reversed these trends in a dose-dependent manner. To further elucidate
   the cellular phenotypes linked to ferroptosis, we conducted TEM
   analysis on lung tissue samples ([138]Figure 4B). Electron microscopic
   observations revealed characteristic ultrastructural alterations
   indicative of ferroptosis in type II alveolar epithelial cells from the
   HAPE group, such as mitochondrial shrinkage, reduced cristae density,
   and outer mitochondrial membrane damage. These changes were mitigated
   following G-Rg3 pre-treatment. Based on these findings, we conclude
   that G-Rg3 pre-administration alleviates HAPE by suppressing
   ferroptosis.
FIGURE 4.
   [139]The image contains two panels labeled A and B. Panel A shows
   Western blot results and corresponding bar graphs for several proteins
   (e.g., GPX4, Nrf2, HO-1) across different treatments (Sham, G-Rg3,
   HAPE). The graphs display protein expression as a fold of Sham, with
   significant differences indicated by asterisks. Panel B presents
   electron microscopy images of tissue samples under various conditions:
   Sham, G-Rg3, HAPE, G-Rg3-L, and G-Rg3-H, at magnifications of x6000 and
   x25000. Red arrows in the images highlight specific structures or
   changes.
   [140]Open in a new tab
   Effects of G-Rg3 pre-treatment on ferroptosis in HAPE. (A) Immunoblot
   quantification of key ferroptotic regulators (GPX4/SLC7A11/FTH1). (B)
   Ultrastructural evidence of mitochondrial pathology in AT2 cells (Red
   arrows: cristae disruption; scale bars: 2 μm [×6,000], 500 nm
   [×25,000]). Data represent mean ± SEM (n = 6 biological replicates).
   One-way ANOVA with Tukey’s post hoc analysis: ^#/##/### P <
   0.05/0.01/0.001 vs. Sham; ^*/**/*** P < 0.05/0.01/0.001 vs. HAPE group.
3.4 Network pharmacology prediction and molecular docking analysis of G-Rg3
pre-treatment in HAPE
   Our multi-modal investigation combining network pharmacology and
   computational modeling revealed G-Rg3’s mechanistic actions against
   HAPE. Bioinformatic interrogation identified 1,285 HAPE-associated
   targets and 248 G-Rg3-related genes, with 52 shared candidates forming
   the core interaction network ([141]Figures 5A–C). Protein interactome
   mapping prioritized five hub genes (TNF, IL6, AKT1, IL1B, ESR1) through
   topological centrality analysis ([142]Figure 5D). Functional enrichment
   demonstrated these mediators coordinate apoptotic regulation
   (biological process), granular secretory mechanisms (cellular
   component), and nuclear receptor activation (molecular function), with
   pathway analysis implicating PI3K/AKT signaling and endocrine
   resistance as primary therapeutic targets ( [143]Figures 5E,F).
FIGURE 5.
   [144]Panel A shows a Venn diagram comparing Ginsenoside Rg3 and HAPE,
   with an overlap of 54 entities. Panel B features a complex network
   diagram of interacting proteins. Panel C displays a network diagram
   centered around Ginsenoside Rg3 with connected genes. Panel D is a
   dense gene interaction network. Panel E presents a bar graph detailing
   various biological processes and their significance, while Panel F is a
   dot plot illustrating KEGG pathway enrichment, with color and size
   indicating statistical significance and gene count.
   [145]Open in a new tab
   Network pharmacology prediction of G-Rg3 pre-treatment in HAPE. (A)
   Pharmacological target convergence between G-Rg3 and HAPE-associated
   genes. (B) Protein-protein interaction (PPI) network of hub targets,
   with node size/color intensity reflecting interaction centrality. (C)
   Tripartite network mapping G-Rg3-target-disease-pathway
   interrelationships. (D) High-fidelity subnetwork of critical protein
   interactions, highlighting topological significance through node degree
   gradation. (E) GO functional annotation categorizing targets into
   biological processes, molecular functions, and cellular components. (F)
   KEGG pathway enrichment analysis of core therapeutic targets.
   Further molecular docking and dynamics simulations revealed the
   interaction between Ginsenoside_Rg3 and PI3K. [146]Figure 6A
   illustrates the binding mode of the PI3K_Ginsenoside_Rg3 complex,
   demonstrating eight hydrogen bonds (including VAL-803 and LYS-807) and
   hydrophobic interactions within the active pocket, with a docking
   energy of −8.48 kcal/mol ([147]Table 1). The 100 ns molecular dynamics
   simulations confirmed binding stability, showing low RMSD (<2 Å) and
   RMSF (<2 Å) values ([148]Figures 6B,C) ([149]Chen et al., 2024).
   MM-GBSA calculations yielded a binding energy of −34.10 ± 3.10 kcal/mol
   ([150]Table 2), primarily driven by van der Waals and electrostatic
   interactions. Ten key residues (including TRP-812 and ILE-963)
   contributed significantly to binding ([151]Figure 6D). Sustained
   hydrogen bonding (4 bonds/frame on average, [152]Figure 6E) validated
   the high-affinity binding potential of this interaction.
FIGURE 6.
   [153]Five panels show molecular dynamics results. Panel A depicts a
   protein-ligand interaction with highlighted amino acids. Panel B is a
   graph of ligand RMSD over time, measuring conformational stability.
   Panel C illustrates protein RMSF across residues, indicating
   flexibility. Panel D is a bar chart showing binding free energy changes
   by residue. Panel E displays the number of hydrogen bonds over time,
   examining stability fluctuations.
   [154]Open in a new tab
   Molecular docking analysis of G-Rg3 pre-treatment in HAPE. (A)
   Molecular docking pose visualization: Gold stick = G-Rg3; Cyan cartoon
   = PI3K backbone; Blue dashes = hydrogen bonds; Yellow surfaces =
   hydrophobic interfaces. (B) Conformational stability assessment via
   RMSD. (C) Per-residue flexibility mapping (RMSF) with mobile loop
   regions. (D) Binding energy hotspot identification through MM-PBSA
   decomposition (Top 10). (E) Time-dependent hydrogen bond population
   analysis.
TABLE 1.
   Scores for the binding affinity of the complexes.
   Target_name   Ligand_name   Docking_score (kcal/mol)
    3ml9-PI3K  Ginsenoside_Rg3          −8.484
   [155]Open in a new tab
TABLE 2.
   MM/GBSA predictions for binding free energies and energy components
   (kcal/mol).
   System name PI3K/Ginsenoside_Rg3
    ΔE [vdw]      −58.96 ± 2.20
   ΔE [elec]      −38.11 ± 2.68
     ΔG[GB]        71.67 ± 5.68
     ΔG[SA]        −8.70 ± 0.68
    ΔG[bind]      −34.10 ± 3.10
   [156]Open in a new tab
   ΔE[vdW]: van der Waals energy.
   ΔE[elec]: electrostatic energy.
   ΔG[GB]: electrostatic contribution to solvation.
   ΔG[SA]: non-polar contribution to solvation.
   ΔG[bind]: binding free energy.
   According to the findings from network pharmacology and docking
   analysis, G-Rg3 might mitigate HAPE and ferroptosis by stimulating the
   PI3K/AKT signaling pathway via strong affinity binding with PI3K, which
   was needed additional experimental validation.
3.5 G-Rg3 ameliorates HAPE by inhibiting ferroptosis through activation of
the PI3K/AKT pathway
   To confirm the predictions of network pharmacology through biochemical
   approaches, we conducted Western blot and immunofluorescence analyses.
   The Western blot results showed that in mice exposed to high-altitude
   hypobaric hypoxia-induced HAPE, the phosphorylation levels of PI3K and
   AKT were substantially decreased. Nevertheless, G-Rg3 pre-treatment
   markedly inhibited the reduction in phosphorylation levels of PI3K and
   AKT ([157]Figure 7A). Immunofluorescence analysis further indicated
   that in the HAPE group, the co-localization between PI3K activation and
   GPX4 antioxidant enzyme expression was inhibited. In contrast,
   prophylactic treatment with G-Rg3 successfully restored the expression
   of these two biomarkers ([158]Figure 7B). Taken together, these
   findings suggest that G-Rg3 can effectively activate the PI3K/AKT
   signaling pathway.
FIGURE 7.
   [159]Panel A shows Western blot results for proteins p-PI3K, PI3K,
   p-AKT, AKT, and β-actin with bar graphs comparing their expression.
   Panel B presents immunofluorescence images stained with DAPI, p-PI3K,
   and GPX4 across different conditions: Sham, G-Rg3, HAPE, G-Rg3-L, and
   G-Rg3-H. Each row represents a different stain, culminating in a merged
   image.
   [160]Open in a new tab
   G-Rg3 ameliorates HAPE by inhibiting ferroptosis through activation of
   the PI3K/AKT pathway. (A) Immunoblot quantification of PI3K, p-PI3K,
   AKT, and p-AKT. (B) Confocal microscopy of merged p-PI3K (green)/GPX4
   (red) fluorescence signals (scale bar: 20 μm, ×400). Data represent
   mean ± SEM (n = 6 biological replicates). One-way ANOVA with Tukey’s
   post hoc analysis: ^##/### P < 0.01/0.001 vs. Sham; ^*/*** P <
   0.05/0.001 vs. HAPE.
3.6 G-Rg3 pre-treatment inhibition of ferroptosis was dependent on the
PI3K/AKT pathway
   To determine whether the inhibitory action of G-Rg3 on ferroptosis in a
   HAPE mouse model is dependent on the activation of the PI3K/AKT
   signaling pathway, this study utilized the specific PI3K inhibitor
   LY294002. Western blotting analysis demonstrated G-Rg3’s capacity to
   enhance PI3K/AKT phosphorylation, an effect abrogated by concurrent
   LY294002 administration ([161]Figure 8A). Subsequent ferroptotic
   biomarker profiling revealed G-Rg3’s dual regulatory function:
   coordinated upregulation of GPX4, Nrf2, HO-1, SLC7A11, and ferritin
   complexes (FTH1/FLC) concurrent with suppression of TFRC/COX2
   expression ([162]Figure 8B). Crucially, the inhibitor LY294002 negated
   these modulatory effects. Complementary validation through
   immunofluorescence and transmission electron microscopy corroborated
   the molecular analyses, demonstrating restored subcellular architecture
   and antioxidant enzyme localization ([163]Figures 8C,D). These findings
   mechanistically establish PI3K/AKT signaling as the principal conduit
   mediating G-Rg3’s anti-ferroptotic efficacy in HAPE pathophysiology.
FIGURE 8.
   [164]The figure consists of four panels (A, B, C, and D). Panel A shows
   Western blot results for proteins like p-PI3K, PI3K, p-AKT, and AKT,
   and corresponding bar graphs comparing expression levels under
   different treatments involving G-Rg3 and LY294002 in a HAPE context.
   Panel B displays Western blots for proteins such as GPX4, Nrf2, HO-1,
   and others, with bar graphs illustrating their expression levels under
   various conditions. Panel C contains fluorescent microscopy images for
   DAPI, p-PI3K, GPX4, and a merged view under different treatment
   conditions. Panel D features electron microscopy images at
   magnifications ×6000 and ×25000 comparing cellular ultrastructure
   across treatments.
   [165]Open in a new tab
   G-Rg3 pre-treatment inhibition of ferroptosis was dependent on the
   PI3K/AKT pathway. (A) Immunoblot quantification of PI3K, p-PI3K, AKT,
   and p-AKT. (B) Ferroptosis biomarker analysis. (C) Confocal microscopy
   of p-PI3K (green, Alexa Fluor 488) and GPX4 (red, Cy3) co-localization
   in alveolar epithelia (scale bar: 20 μm; ×400). (D) Ultrastructural
   ferroptosis markers: mitochondrial cristae dissolution (red arrows) in
   type II alveolar epithelial cells (scale bars: 2 μm [×6,000]; 500 nm
   [×25,000]). Data represent mean ± SEM (n = 6 biological replicates).
   One-way ANOVA with Tukey’s post hoc analysis: ^*/**/*** P < 0.05/0.001
   vs. HAPE, ^$/$$ P < 0.05/0.01 vs. G-Rg3-H group.
3.7 The protective effect of G-Rg3 pre-treatment on HAPE was mediated through
the PI3K/AKT signaling pathway
   Histopathological evaluation of LY294002-treated lung tissues revealed
   pronounced structural damage characterized by inflammatory
   infiltration, alveolar wall thickening, and intra-alveolar congestion
   ([166]Figure 9A). These pathological changes were corroborated through
   semi-quantitative scoring and edema quantification ([167]Figures 9B,C).
   G-Rg3 pre-treatment notably mitigated these alterations, restoring
   near-normal pulmonary architecture. Concurrent analysis demonstrated
   LY294002-induced upregulation of inflammatory mediators and oxidative
   stress markers, effects substantially attenuated by G-Rg3
   administration ([168]Figures 9D,E). This integrated analysis confirms
   PI3K/AKT signaling modulation as the central mechanism underlying
   G-Rg3’s therapeutic efficacy against HAPE-associated pulmonary injury.
FIGURE 9.
   [169]Histological images and bar graphs illustrate the effects of
   treatments on lung tissue in different conditions. Panel A shows
   stained lung tissue sections at two magnifications (200x and 400x)
   under four conditions: HAPE, G-Rg3-H, LY294002, and G-Rg3-H+LY294002.
   Panels B to E present bar graphs comparing lung injury scores,
   wet-to-dry ratios, cytokine levels (IL-1, IL-6, TNF-α), and oxidative
   stress markers (GSH, MDA, SOD, tissue iron) across the conditions, with
   statistical annotations indicating significance levels.
   [170]Open in a new tab
   The protective effect of G-Rg3 pre-treatment on HAPE was mediated
   through the PI3K/AKT signaling pathway. (A) H&E-stained lung sections
   (upper: ×200, 50 μm scale; lower: ×400, 20 μm scale). (B)
   Semi-quantitative histopathological assessment of HAPE. (C) Pulmonary
   edema quantification (W/D ratio). (D) Analysis of the effect of
   LY294002 on inflammatory markers (IL-1β, IL-6, and TNF-α) in BALF. (E)
   Evaluation of LY294002’s influence on antioxidant enzymes (GSH, SOD)
   and oxidative stress markers (MDA) as well as tissue iron levels in
   lung tissues. Data represent mean ± SEM (n = 6 biological replicates).
   One-way ANOVA with Tukey’s post hoc analysis: ^*/**/*** P < 0.05/0.001
   vs. HAPE, ^$/$$/$$$ P < 0.05/0.01 < 0.001 vs. G-Rg3-H group.
4 Discussion
   Recent years have witnessed increased human activity in high-altitude
   regions, amplifying clinical demands for managing hypoxia-related
   cardiopulmonary disorders. Particularly concerning is the prevalence of
   HAPE. Our study employed an HAPE mouse model to systematically evaluate
   the therapeutic potential of G-Rg3. Pretreated mice exhibited markedly
   attenuated HAPE-associated pathological features, concurrent with
   normalized HIF-1α overexpression and robust
   anti-inflammatory/antioxidant responses. Further mechanistic
   investigations identified ferroptosis suppression as a critical
   component of G-Rg3’s action. Central to this protective effect is the
   metabolite’s activation of PI3K/AKT signaling, as evidenced by
   integrated network pharmacology analyses, computational modeling, and
   targeted pathway inhibition experiments. Investigations reveal that the
   traditional botanical metabolite G-Rg3 holds promise for high-altitude
   HAPE treatment through its multi-action regulatory effects. By
   targeting acute pathological changes and alleviating hypoxia-induced
   secondary injuries, it offers significant pharmacological insights
   supporting the clinical translation of plant-based therapies.
   The complex physiological demands imposed by high-altitude hypoxia
   present significant challenges. A key mechanism driving the development
   of HAPE involves the deregulation of HIF-1α and VEGF ([171]Wang et al.,
   2022b). Under conditions of reduced pressure and oxygen availability,
   cells initiate an adaptive response by stabilizing HIF-1α, which
   subsequently triggers elevated VEGF expression. Both factors are
   crucial in the progression of HAPE. Studies have shown that hypoxic
   conditions amplify the activity of HIF-1α, positioning it as the
   principal regulator of genes responsive to hypoxia—genes that are vital
   for cellular survival. In pathological scenarios, the heightened
   induction of VEGF, mediated by HIF-1α, fosters irregular angiogenesis
   and accelerates disease progression. Our experimental data corroborate
   these findings, indicating that increased levels of HIF-1α and VEGF
   intensify the pathological processes associated with pulmonary edema.
   Importantly, this investigation expands on earlier studies ([172]Ahmmed
   et al., 2019; [173]Lv et al., 2023) by illustrating the suppressive
   impact of G-Rg3 on the expression levels of HIF-1α and VEGF in HAPE
   models.
   Simultaneously, pro-inflammatory cytokines such as TNF-α, IL-1β, and
   IL-6 have been identified as pivotal factors influencing the dynamic
   changes in white blood cells during HAPE, which play critical roles in
   regulating the proliferation, migration, and differentiation of immune
   cells ([174]Tanaka et al., 2014) In mouse models, the systemic levels
   of these cytokines were significantly elevated following HAPE
   induction. These observations align with previous reports associating
   cytokine storms with inflammatory lung injury ([175]Arya et al., 2013).
   The ability of G-Rg3 to modulate these pathways underscored its dual
   therapeutic potential, illustrating its influence on the inflammatory
   cascade within the pathophysiology of HAPE.
   Hypobaric hypoxia in HAPE induces mitochondrial dysfunction and
   excessive ROS generation (superoxide, hydrogen peroxide), driving
   oxidative damage and depleting antioxidant reserves ([176]Irarrázaval
   et al., 2017). his oxidative cascade triggers membrane lipid
   peroxidation, quantified by elevated MDA levels ([177]Weismann et al.,
   2011), while concurrently suppressing SOD activity which is a critical
   antioxidant enzyme ([178]Wang et al., 2018)). The hypoxia-induced
   inhibition of system Xc^− (SLC7A11/SLC3A2) further diminishes GSH
   synthesis through cysteine deprivation, impairing both ROS
   neutralization and GPX4-mediated ferroptosis prevention ([179]Endale et
   al., 2023; [180]Mailloux et al., 2013). Our findings reveal a
   pathological feedback loop in HAPE: diminished SOD activity facilitates
   superoxide accumulation, heightened MDA concentrations signify
   accelerated lipid peroxidation, and GSH depletion undermines
   antioxidant mechanisms, all of which contribute to pulmonary edema.
   Importantly, pre-treatment with G-Rg3 can counteract these disruptions,
   restore SOD function and GSH levels in lung tissue, and lower MDA,
   effectively interrupting this disease progression pathway.
   Emerging evidence identifies oxidative stress-mediated iron metabolism
   dysregulation and ferroptosis as pivotal mechanisms in HAPE. The
   pathogenic triad of ROS accumulation, lipid peroxidation, and iron
   overload which established contributors to pulmonary disorders, is
   exacerbated under hypoxic conditions ([181]Wang et al., 2024).
   Mechanistically, hypoxia inhibits Nrf2, suppressing key iron regulators
   (SLC7A11, HO-1, FTH1) and weakening antioxidant defenses ([182]Loboda
   et al., 2016; [183]Su et al., 2025; [184]Zhang S. et al., 2025). This
   suppression triggers two detrimental pathways: 1) impaired cystine
   transport through System Xc^− disrupts glutathione synthesis and GPX4
   antioxidant activity; 2) reduced HO-1/FTH1 expression enhances free
   iron toxicity via Fenton reactions ([185]Basiouny et al., 2025; [186]Fu
   et al., 2022; [187]Qian et al., 2024). Concurrent elevation of
   ferroptosis markers TFRC and COX2 further drives pathological iron
   uptake and inflammatory lipid damage ([188]Liu M. J. et al., 2025;
   [189]Zhuang et al., 2024). Our experimental data validate this
   mechanistic model in HAPE progression. Diseased mice exhibited
   characteristic ferroptosis signatures: depressed anti-ferroptosis
   proteins (GPX4, Nrf2, HO-1, SLC7A11, FTH1/FLC) alongside elevated TFRC,
   COX2, and tissue iron. Crucially, G-Rg3 treatment ameliorated these
   imbalances, reactivating ferroptosis-suppressing pathways and restoring
   iron homeostasis. This dual antioxidant-iron regulatory capability
   positions G-Rg3 as a promising therapeutic agent targeting the
   oxidative-ferroptosis axis in HAPE.
   The PI3K/AKT signaling axis ([190]Liu et al., 2025c) functions as a
   master regulator of cellular homeostasis, coordinating processes from
   metabolic regulation to survival pathways ([191]Huang et al., 2022).
   Mechanistically, PI3K initiates signaling by phosphorylating membrane
   phosphatidylinositol to generate PIP3, which facilitates AKT membrane
   recruitment and subsequent activation through phosphorylation events.
   Activated AKT then orchestrates downstream biological responses through
   its effector network ([192]Vanhaesebroeck et al., 2010). This pathway
   exhibits dual ferroptosis-inhibitory capacities: (1) SREBP1-mediated
   lipogenesis enhancement reduces membrane phospholipid peroxidation
   susceptibility ([193]Zheng et al., 2024); (2) AKT-driven Nrf2
   phosphorylation (Ser40) promotes KEAP1 dissociation, enabling nuclear
   translocation to boost antioxidant defenses (glutathione synthesis) and
   iron regulatory capacity ([194]Cheng et al., 2023). Our integrated
   pharmacological investigation reveals G-Rg3’s therapeutic mechanism
   through PI3K/AKT modulation. Network pharmacology prioritized this
   pathway as the prime HAPE intervention target, corroborated by
   computational modeling showing stable G-Rg3-PI3K binding through
   hydrogen bonds and hydrophobic interactions. Immunoblot analyses
   demonstrated G-Rg3’s capacity to restore depressed PI3K/AKT
   phosphorylation in HAPE lungs. Crucially, co-treatment with LY294002
   (PI3K inhibitor) abolished G-Rg3’s protective effects against both
   inflammatory cytokines and ferroptosis markers (GPX4/HO-1
   downregulation), conclusively validating pathway centrality.
   This study identified G-Rg3 as a selective PI3K/AKT activator that can
   alleviate ferroptosis in high-altitude pulmonary edema (HAPE) by
   coordinating the regulation of lipid metabolism, iron homeostasis, and
   oxidative defense. Four limitations need to be addressed when
   establishing the therapeutic framework of G-Rg3: (1) The limitations of
   the acute hypobaric model require gradient altitude studies that
   simulate the progression of human HAPE; (2) Further in vitro studies
   are needed to clarify the mechanism by which G-Rg3 prevents HAPE; (3)
   Orthogonal validation through gene editing, co-immunoprecipitation
   target identification, and surface plasmon resonance is necessary for
   the confirmation of the PI3K-based pathway; (4) GLP-compliant
   toxicological analyses (chronic toxicity, tissue-specific
   pharmacokinetics) are crucial for clinical translation. These further
   studies will provide a more solid theoretical basis for G-Rg3-based
   intervention strategies for HAPE.
5 Conclusion
   This research highlights that the active constituent of ginseng, G-Rg3,
   can successfully mitigate acute high-altitude pulmonary edema triggered
   by hypobaric hypoxia. This effect is achieved through the activation of
   the PI3K/AKT signaling pathway and the cooperative modulation of
   inflammatory responses, oxidative stress, and ferroptosis ([195]Figure
   10). Importantly, the therapeutic timing for G-Rg3 includes preventive
   administration up to 72 h prior to ascent, with optimal
   efficacy-to-dose performance observed at a dosage of 30 mg/kg. Beyond
   counteracting the HIF-driven inflammatory-oxidative cascade, focusing
   on the PI3K/AKT regulatory axis offers a molecular foundation for
   designing combination therapies. These underlying mechanisms indicate
   that G-Rg3 could serve both as a standalone treatment in high-altitude
   emergency care and leverage its distinctive iron homeostasis regulation
   capabilities for managing chronic high-altitude illnesses or systemic
   inflammation-related hypoxic conditions. This provides a comprehensive
   approach for advancing traditional medicinal plants into modern
   applications.
FIGURE 10.
   [196]Illustration of cellular mechanisms influenced by hypoxia, showing
   the role of G-Rg3 and pulmonary edema. Arrows indicate pathways from
   hypoxia through cytokines, oxidative stress, and ferroptosis. Molecular
   interactions involve PI3K/Akt, COX2, ROS, and various proteins like
   FTH1 and SLC7A11. The diagram includes mitochondria-related processes
   like the TCA cycle and Fenton reaction, highlighting oxidative stress
   and cellular responses.
   [197]Open in a new tab
   Mechanistic Framework of G-Rg3-Mediated Pulmonary Protection in Murine
   HAPE Models. Pharmacological preconditioning with G-Rg3 (30 mg/kg, i.
   p.) ameliorates hypoxia-induced pulmonary edema through coordinated
   ferroptosis-PI3K/AKT axis modulation.
Funding Statement
   The author(s) declare that financial support was received for the
   research and/or publication of this article. This study was financially
   supported by National Natural Science Foundation of China (82205043),
   China Association of Chinese Medicine Qiushi Project (2023-QNQS-11),
   Natural Science Foundation of Sichuan Province (2024NSFSC1865,
   2025ZNSFSC1851), Yunnan Provincial Science and Technology Department
   (202301AZ070001-119, C012018005, Y0120180018), Sichuan Cadre Health
   Care Research Project (2024-503), Research Project of Sichuan
   Provincial Administration of Traditional Chinese Medicine (2023MS555),
   Scientific Research Project of Health Commission of Chengdu
   (WXLH202403114, WXLH202402022) and “Xinglin Scholars” Discipline Talent
   Research Enhancement Program Postdoctoral Special Project
   (QJRC2023011).
Data availability statement
   The original contributions presented in the study are included in the
   article/supplementary material, further inquiries can be directed to
   the corresponding authors.
Ethics statement
   The animal study was approved by the Medical Ethics Committee at
   Chengdu University of Traditional Chinese Medicine (Approval number:
   2022-18). The animal experiments followed 3R (reduction, replacement,
   and refinement) principles, where all measures are taken to ensure
   minimal discomfort and minimal suffering. The study was conducted in
   accordance with the local legislation and institutional requirements.
Author contributions
   YH: Writing – original draft. YW: Writing – original draft. HD: Writing
   – original draft. DH: Writing – original draft. NJ: Writing – original
   draft. ZS: Writing – original draft. ZW: Writing – review and editing.
   MW: Writing – review and editing. TZ: Writing – review and editing.
Conflict of interest
   The authors declare that the research was conducted in the absence of
   any commercial or financial relationships that could be construed as a
   potential conflict of interest.
Generative AI statement
   The author(s) declare that no Generative AI was used in the creation of
   this manuscript.
Publisher’s note
   All claims expressed in this article are solely those of the authors
   and do not necessarily represent those of their affiliated
   organizations, or those of the publisher, the editors and the
   reviewers. Any product that may be evaluated in this article, or claim
   that may be made by its manufacturer, is not guaranteed or endorsed by
   the publisher.
References