• Open Access
    Review

    Physiologically driven nanodrug delivery system for targeted lung cancer treatment

    Shiying Zhang
    Xia Li
    Yang Liu
    Hui Li *
    Zhiyue Zhang *

    Explor Med. 2024;5:280–311 DOI: https://doi.org/10.37349/emed.2024.00221

    Received: November 14, 2023 Accepted: December 18, 2023 Published: April 25, 2024

    Academic Editor: Yun-Long Wu, Xiamen University, China

    This article belongs to the special issue Functional Biomaterials and Tumor Immunotherapy

    Abstract

    Lung cancer remains a leading cause of cancer-related deaths globally, and a significant number of patients are ineligible for surgery, while chemoradiotherapy often shows limited efficacy, a systemic distribution, a low drug concentration at tumor sites, severe side effects, and the emergence of drug resistance. In this context, a nanodrug delivery system (NDDS) has emerged as a promising approach for lung cancer treatment, offering distinct advantages such as targeted delivery, responsiveness to the tumor microenvironment, site-specific release, and enhanced induction of apoptosis in cancer cells, ultimately leading to tumor growth inhibition or even elimination. This review aims to provide an overview of the physiological characteristics of lung cancer, highlight the limitations of conventional treatment methods, and extensively examine recent significant advancements in NDDS utilized for lung cancer therapy. The findings from this review lay the foundation for further development and optimization of NDDSs in the treatment of lung cancer.

    Keywords

    Lung cancer, drug delivery system, nanotechnology, targeted therapy, cancer treatment, active targeting ligand, pharmacology, oncology

    Introduction

    Lung cancer is the leading cause of cancer deaths worldwide, with approximately 2.2 million (11.4%) new cases and 1.8 million (18.0%) deaths reported in 2020 [1]. It is a highly heterogeneous disease with complex clinical effects and a poor prognosis [2]. Due to the lack of preventive measures and early diagnosis methods and the variability of the tumor environment, lung cancer is prone to high metastasis and recurrence rates and multidrug resistance (MDR) [3]. Therefore, improving the therapeutic effect on lung cancer is an urgent problem.

    Histological studies of cancer cells have categorized lung cancer into two types: small cell lung cancer (SCLC) and non-SCLC (NSCLC), with SCLC accounting for approximately 15% and NSCLC accounting for approximately 85% of cases [4]. Although current medical technology can effectively intervene in pulmonary neoplasms to a certain extent through surgical resection, chemotherapy, and radiotherapy, improving the survival rate of lung cancer patients remains challenging [5].

    Surgery is not suitable for all patients, as lung cancer is often diagnosed at an advanced stage and has a poor prognosis and high recurrence rate. Radiotherapy, on the other hand, has strong systemic side effects and can cause skin damage in the radiated area, limiting its therapeutic effect [6]. Chemotherapy, another main treatment for intermediate-stage disease, is hindered by poor water solubility, lack of targeting, and the first-pass effects of most drugs, resulting in low drug concentrations at the tumor site [7, 8]. As a result, current treatments are not highly effective.

    However, due to the significantly higher growth rate of tumor tissues compared to normal tissues, defects between tumor vascular endothelial cells, wider gaps in the vascular wall, poor structural integrity, insufficient lymphatic drainage within the tumor, and a lower blood flow rate, macromolecular-like substances and lipid particles can selectively permeate and be retained in the tumor, which is known as the enhanced permeability and retention (EPR) effect [9]. This phenomenon provides opportunities for developing novel therapeutic approaches that exploit these characteristics.

    Nano-based agents have the ability to overcome biological and chemical barriers within the human body. They can improve the pharmacokinetics and biodistribution of drugs and prevent early inactivation or biodegradation [10]. Additionally, they can be functionalized with modified targeting ligands, antibodies, and peptides to anchor them to receptor-overexpressing structures and specifically bind to cancer cells [11]. This enables active targeting, tumor site accumulation, and responsiveness to the tumor microenvironment (TME) or external stimuli such as temperature, light, and ultrasound, thus influencing these agents’ affinity or binding state, triggering the release of drugs, and promoting accumulation of therapeutic molecules in tumor tissues [12]. Moreover, the use of nano-based agents can reduce accumulation in normal organs, minimize systemic toxicity, and enhance treatment efficacy [13]. The use of nano-based carriers for drug delivery offers several advantages, including a high surface area-to-volume ratio, adjustable thermal, magnetic, optical, and electrical properties, an ability to synthesize diverse shapes and sizes, a high drug loading capacity, and stimulus-response sensitivity for precise spatiotemporal controlled drug release [14]. However, despite promising results in preclinical trials, the clinical translation potential of nano-based agents has not been fully realized due to challenges associated with reproducibility, large-scale manufacturing, and potential toxicological and safety hazards. In this study, we aim to leverage the physiological characteristics of lung cancer to clarify drug targets and behavioral processes in the tumor environment and thus provide insights for the design of more intelligent and successful drug delivery systems.

    Physiological characteristics of lung cancer

    Lung cancer is an intricate and diverse disease marked by the deterioration of lung epithelial cells [15]. Developing a profound understanding of the physiological attributes of the disease is essential to drive research in innovative diagnostic methods, treatments, and prognostic assessments. Targeted regulation of abnormal molecular signals and their subsequent pathways offers a promising avenue for the development of therapeutic interventions. To devise effective strategies for lung cancer treatment, elucidating the molecular pathogenesis, epigenetics, and signaling pathways mediated by pertinent molecules is imperative (Figure 1).

    Typical signaling pathways in lung cancer. Created by Figdraw. The arrows mean the direction of signal transmission, except that before “apoptosis” means induction. AKT: protein kinase B; ALK: anaplastic lymphoma kinase; Apaf: apoptotic protease activating factor; Bcl-2: B cell lymphoma 2; Bak: Bcl-2 homologous antagonist/killer; Bax: Bcl-2-associated X protein; Bcl-XL: Bcl-extra large; EGFR: epidermal growth factor (EGF) receptor; ERK: extracellular signal-regulated kinase; HER: human EGFR receptor; JAK: Janus kinase; MAK: male germ cell-associated kinase; Mcl-1: myeloid cell leukemia 1; MET: mesenchymal-epithelial transition factor; PI3K: phosphatidylinositol 3-kinase; RAF: rapidly accelerated fibrosarcoma; RAS: rat sarcoma; STAT: signal transducer and activator of transcription; VEGFR: vascular endothelial growth factor (VEGF) receptor

    Similar to other tumors, the development of lung cancer involves activation of growth-promoting proteins, such as EGFR, Kirsten RAS oncogene homolog (KRAS), and HER2. Notably, more than 50% of NSCLC patients who smoke and consume alcohol have p53 mutations, and certain compounds have shown potential for reactivating p53 for lung cancer treatment [16].

    Lung cancer biomarkers, which are identified through chromosomal aberrations and mutations, often serve as attractive therapeutic targets. In this section, we will discuss the common variant types and clinical significance of proto-oncogenes, oncogenes, and molecular signaling pathways in lung cancer. Furthermore, we will explore mutated genes, specific ligands, and the tumor metastasis microenvironment associated with lung cancer, aiming to provide insights into the development of nanodrug delivery systems (NDDSs) for lung cancer [17]. By probing into the scientific understanding of lung cancer at various levels, researchers can pave the way for personalized and comprehensive treatment strategies that consider genetic, molecular, histopathological, and clinical factors, which holds promise for improving the diagnosis, treatment, and prognosis of lung cancer.

    Occurrence of lung cancer

    Lung cancer occurs mainly due to genetic mutations, including EGFR oncogenic mutations, ALK and ROS proto-oncogene 1-receptor tyrosine kinase (ROS1) fusion, and v-raf murine sarcoma viral oncogene homolog B1 (BRAF) mutations, which have been approved by the U.S. Food and Drug Administration (FDA) as therapeutic targets for NSCLC. In addition, KRAS mutations, aberrations in MET, which encodes the hepatocyte growth factor (HGF) receptor (HGFR), and mitogen-activated protein kinase kinase 1 (MEK1) overexpression in NSCLC may also be potential targets for the treatment of lung cancer [18].

    EGFR is a member of the Erb-B2 receptor tyrosine kinase 2 (ErbB) family of transmembrane receptor tyrosine kinases (TKs). The ErbB receptor consists of an extracellular ligand binding domain, a transmembrane domain, and an intracellular TK domain [19]. Seven known ligands of EGFR induce different biological effects even in the same cell [20]. EGFR can induce signaling pathways, including the PI3K/AKT/mammalian target of rapamycin (mTOR) pathway, RAS/RAF/MEK/ERK pathway, and JAK/STAT pathway, which are involved in the pathogenesis of various tumors, including NSCLC [21]. Mutations in the EGFR gene affect tumor pathogenesis, including cell proliferation, survival, and differentiation, neovascularization, invasion, and metastasis [22]. Therefore, EGFR inhibitors are typically compounds that directly inhibit TK phosphorylation through physical interaction with ATP or enzyme-substrate binding sites. Third-generation TK inhibitors (TKIs), such as osimertinib, which has shown efficacy against EGFR T790M mutations, have been developed to overcome resistance to previous EGFR inhibitors.

    HER2 is another receptor TK in the ErbB/HER family, which together with EGFR encodes a membrane-bound glycoprotein TK that binds to other ligands to form a heterodimer and activates downstream signaling [23]. HER2 amplifications and mutations are considered to be usually unrelated and clinically distinct driver alterations that can be used to segment lung adenocarcinoma patients for targeted therapy [24, 25], and the development of drugs targeting HER2 remains an urgent clinical need [26]. HER2 aberrations in lung cancer may be resistant to EGFR TKI therapy due to tissue specificity [27]. Currently, trastuzumab/paclitaxel (PTX) combination therapy has promising applications for the treatment of HER2-positive lung cancer.

    The KRAS gene is a member of the RAS proto-oncogene family, which includes neuroblastoma RAS viral oncogene homolog (NRAS) and Harvey RAS viral oncogene homolog (HRAS). G proteins encoded by the KRAS gene play a key role in regulating cell proliferation, differentiation, and growth signal transduction pathways [28]. In normal resting cells, RAS protein binds to guanosine-5’-diphosphate (GDP) in an inactive state; when the upstream growth factor receptor is activated, RAS protein binds to guanosine triphosphate (GTP) and becomes active [29], which causes the protein to lose its intrinsic GTPase activity, preventing GTP from turning into GDP and thus resulting in the unchecked proliferation of the RAS/RAF/MEK/ERK signaling pathway downstream of EGFR [30]. Although KARS mutations are common in lung cancer, the results of clinical trials of relevant targeted drugs have been unsatisfactory, and further research is needed.

    The BRAF gene encodes serine/threonine protein kinase, a downstream effector protein of KRAS [31]. BRAF proteins are RAS-RAF-MEK-activated ERKs that play a critical role in the MAPK/ERK signaling pathway and are involved in the regulation of cell proliferation and growth [32]. The therapeutic regimen of the BRAF inhibitor dabrafenib combined with the downstream MEK inhibitor trametinib for NSCLC with metastatic BRAF V600e mutation has been approved by the FDA and European Medicines Agency (EMA) [33]. Novartis has received approval from the State Drug Administration for a new indication for the treatment of BRAF V600 mutation-positive metastatic NSCLC with the dual-targeted combination of dabrafenib mesylate capsules and trametinib tablets.

    ALK belongs to the insulin receptor superfamily [34], is activated by the family with sequence similarity 72 member A (FAM150A) and FAM150B ligands [35], and is expressed as a fusion gene in various cancers [36]. The ALK fusion protein dimerizes in a ligand-independent manner and activates the PI3K/AKT, RAS/RAF/MAPK1, and JAK/tyrosine aminotransferase (TAT) signaling pathways via the abnormal structure of the ALK TK. The abnormal constitutive activation of ALK TK leads to cell proliferation disorder and participates in the regulation of cell proliferation and apoptosis. To date, three ALK TKIs (crizotinib, ceritinib, and alectinib) have been used clinically for the treatment of ALK-abnormal NSCLC [37]. ALK TKIs often achieve significant tumor regression in patients with NSCLC with ALK rearrangements; however, in most cases, ALK TKI-resistant tumors’ likelihood of reemergence is high within a few years.

    ROS1 encodes a receptor TK containing a large N-terminal extracellular region [38], a 3-hydrophobic one-way transmembrane region, and a C-terminal intracellular TK region [39, 40]. ROS1 regulates cell growth and apoptosis through the MAPK/ERK, PI3K/AKT, JAK/STAT3, and Src homology 2 domain-containing protein tyrosine phosphatase 1/2 (SHP1/2) signaling pathways [41]. All ROS1 fusions retain 14 intact ROS1 kinase structural domains [42]. Bilateral lung adenocarcinomas occur in transgenic mice expressing ezrin (EZR)-ROS1 in the alveolar epithelium [43]. Crizotinib, an ALK/ROS1/MET inhibitor, is an FDA-approved targeted agent for the treatment of advanced ROS1-rearranged NSCLC, and studies have shown that some ALK TKIs have dual inhibitory activity against ALK and ROS1 [44].

    The MET gene encodes HGFR [45], which affects key processes such as cell proliferation, differentiation, motility, and tumor angiogenesis through the PI3K/mTOR, STAT, and MAPK signaling pathways [46, 47] and is closely associated with the development of various cancers [48]. However, MET exon 14 jump mutations and high levels of MET amplification have emerged as potential predictive biomarkers [49, 50]. Capmatinib (INC280) is a potent and selective MET receptor inhibitor that has shown antitumor activity with various MET activations and can cross the blood-brain barrier [51]. Monoclonal antibodies to the MET TKI onartuzumab in combination with the EGFR TKI erlotinib block HGF binding to MET receptors for second and third-line treatment of NSCLC after chemotherapy failure [52].

    Lung cancer metastasis-related molecules

    The main objective of systemic therapy for metastatic NSCLC is to reduce the burden of symptoms and improve the survival rate and quality of life [53]. Many cytotoxic chemotherapy regimens exhibit significant toxicities (e.g., alopecia, nausea), and the role of surgery and radiotherapy in prolonging disease-free survival is being studied for patients with lower metastatic tumor burdens. However, targets for drugs that are essential for tumor cell survival or immune evasion and can be tailored to patients’ tumor characteristics remain to be identified [54].

    CD44 plays an important role in cell adhesion [55] and has a significant negative correlation with transfer potential [56]. The cell surface adhesion receptor CD44 is a positive regulator of programmed cell death ligand 1 (PD-L1) expression in NSCLC cells. CD44 activates PD-L1 transcription by cleavage of the intracytoplasmic domain. In one study, the proportion of PD-L1 tumors in patients with metastatic NSCLC was scored as 1% [57], providing a new rationale for CD44 as a key therapeutic target to inhibit intrinsic PD-L1 tumor function [58]. Effector T cell depletion is caused by decreased phosphorylation of various signaling molecules, such as ERK, Vav, and phospholipase Cγ (PLCγ) [59], which regulate T cell activation and proliferation through nuclear factors that activate T cells, leading to cancer immune evasion and promoting the growth, migration, and invasion of lung cancer cells [60].

    The metastasis of lung cancer to the central nervous system (CNS) is the main clinical obstacle leading to the low 5-year survival rate of advanced diseases. Therapeutic drugs do not easily cross the blood-brain barrier, which substantially limits treatment and results in a poor prognosis. Therefore, the relevant driving factors and molecular mechanisms of tumor metastasis and targeted prevention and treatment require attention. Cytoskeletal proteins and motility-related genes such as metallothionein 2A (MT2A), fascin actin-bundling protein 1 (FSCN1), microtubule-associated protein 7 (MAP7), and chemokine-CXC-motif chemokine ligand 13 (CXCL13) were significantly upregulated in tumor metastases. CXC-motif chemokine receptor 4 (CXCR4) is expressed in 90% of primary tumors and 100% of brain metastases, and CX3C-motif chemokine receptor 1 (CX3CR1) is also a chemokine ligand associated with lung cancer and metastatic spread [61]. Overexpression of N-cadherin and decreased expression of kinesin family member C1 (KIFC1), E-cadherin, and BPTF/FALZ can predict brain metastasis [62]. HGFR expression is observed in approximately 30% of adenocarcinomas, and cellular-MET (c-MET) amplification is observed in approximately 10% of adenocarcinomas [63].

    TME

    The TME consists of extracellular matrix, soluble components (growth factors, chemokines, etc.), and cellular components (tumor cells, fibroblasts, endothelial cells, etc.). Tumor-associated macrophages (TAMs), vascular endothelial cells, and fibroblasts in the tumor matrix secrete growth factors and chemokines to attract and regulate the behavior of stromal cells and immune cells and promote tumor formation and growth. Tumor-derived cytokines such as IL-6, growth arrest specific protein 6 (GAS6), HGF and EGF can promote resistance to targeted therapy through autocrine signaling [64]. The interaction between tumor cells and the TME, such as changing cell adhesion by increasing the expression of cadherin and integrin β1 and reducing the expression of β2-microglobulin, affects the response of tumor cells to targeted therapy. Cytokines secreted by cancer-associated fibroblasts (CAFs) and mesenchymal stem cells (MSCs) in the matrix lead to EGFR TKI resistance by activating CXCR4 and IL-6R [65]. Due to an insufficient oxygen supply, tumor cells are able to metabolize energy only through anaerobic enzymes, resulting in lactic acid accumulation; meanwhile, ion exchange proteins on the tumor cell membrane are constantly transporting intracellular H+ outside the cell to avoid causing acidosis, resulting in a lower pH of the TME and an acidic environment, which will cause apoptosis and release cell debris and chemokines, leading to inflammatory cell infiltration and inflammatory factor secretion. However, the typical characteristics of the TME, such as hypoxia, low pH, and a high bioreduction environment, can be used as the conditions for stimulus-responsive drug release [66]. Chen et al. [67] found that hypoxia can improve the drug resistance of adriamycin in the treatment of NSCLC by inhibiting the expression of MDR-associated protein 1 (MRP1) and P-glycoprotein (P-gp) and enhancing the chemical sensitization of MRP1 and P-gp blockers.

    Tumor angiogenesis and vascular abnormalities are critical to the growth and metastasis of solid tumors [68], and many proteins and small molecules have been demonstrated to be involved in angiogenesis, including VEGF, platelet-derived growth factor (PDGF), and fibroblast growth factor (FGF). Small-molecule TKIs, particularly multitargeted vascular kinase inhibitors, selectively inhibit downstream VEGFR pathway-mediated activation [69]. At present, various small-molecule antiangiogenic TKIs, such as apatinib, anlotinib, and nintedanib, have been used and evaluated in lung cancer clinical studies, as well as FGF receptor (FGFR) with TKIs alone or in combination with chemotherapy [70, 71].

    NSCLC is immunogenic [72], and approximately two-thirds of lung tumor-infiltrating immune cells consist of T and B cells, with the remainder consisting of TAMs, neutrophils, a few dendritic cells (DCs) and natural killer (NK) cells [73], which are involved in antitumor responses [74]. A decrease in NK cells and enrichment of regulatory T (Treg) cells at tumor sites, where CD4+ Th1 cells, activated CD8+ T cells, and even γδ-T cells are frequently involved in type I immune responses, are associated with a good prognosis in lung cancer patients [75]. Immune checkpoint inhibitors (ICIs) have become the treatment of choice for recurrent or metastatic cancers, with the most widely used being anti-cytotoxic T lymphocyte associated protein 4 (CTLA4) and anti-programmed cell death protein-1 (PD1) antibodies [76]. CTLA4 is expressed on tumor-infiltrating T cells and is an inhibitory receptor whose main role is to regulate the degree of early T cell activation [77]. Anti-CTLA4 antibodies block inhibitory signals between antigen-presenting cells and T lymphocytes involving CTLA4 molecules. PD1 is highly expressed in Treg cells, and its main role is to inhibit T cell activation in peripheral tissues and suppress autoimmunity [78, 79]. Targeted immunoregulatory molecules, such as anti-4-1BB, Ox40, inducible T cell costimulator (ICOS), and CD40, block antibodies of lymphocyte-activation gene 3 (LAG3), B7-H4, Tim3, and killer Ig-like receptors (KIRs), and related therapeutic drugs are being developed. Since in vivo immune regulation involves complex regulatory pathways, combination immunotherapy will have a better therapeutic effect [80].

    Ligands targeting tumor-specific receptors

    In the context of cancer treatment, a successful cure often depends on achieving high concentrations of drugs specifically within tumor tissues. Systemic distribution of drugs throughout the body can lead to adverse effects and hinder the desired therapeutic outcomes. To overcome these limitations, nanocarriers have emerged as promising vehicles for drug delivery, offering improved pharmacokinetics and targeted delivery to tumors. Extensive research efforts have focused on designing nanoparticles with multiple functions to address the biological barriers encountered during intravenous administration. When nanoparticles are administered intravenously, they tend to be taken up by resident macrophages of the mononuclear phagocyte system (MPS) [81]. Consequently, significant accumulation of nanoparticles occurs in organs such as the spleen and liver, resulting in nonspecific distribution of nanotherapeutic drugs to healthy organs.

    To overcome these challenges, identifying specific ligands that can selectively target receptors expressed by lung cancer cells is crucial. Active targeting strategies can be employed to achieve site-specific delivery of therapeutic drugs and facilitate their efficient accumulation in lung cancer tissues. By employing nanocarriers with active targeting capabilities, enhancing the specificity and efficiency of drug delivery to lung cancer cells becomes feasible. This approach holds significant potential for improving the therapeutic outcomes of lung cancer treatment while minimizing adverse effects on healthy organs.

    Folate

    The folate receptor (FR) family consists of four members, including FRα, FRβ, FRγ, and FRδ [82]. Silencing FR expression can inhibit the proliferation of cancer cells and promote apoptosis of human NSCLC cells [83]. FRβ has been demonstrated to be overexpressed in M2-polarized TAMs, and FR-positive TAMs are related to the poor prognosis of lung cancer [84]. By targeting FR overexpression of M2 TAMs and tumor cells, the proportion of M2 TAMs was reduced, and the proliferation and metastasis of tumor cells were inhibited, which had a synergistic effect on inhibiting tumors. Tie et al. [85] targeted FRβ-positive TAMs in lung cancer with folic acid-modified liposome complexes, which significantly promoted apoptosis of tumor cells and M2 TAMs. Clinical studies are currently assessing the efficacy of FRα combined with chemotherapy drugs and humanized anti-FRα antibodies [86].

    Transferrin

    The transferrin (Tf) family plays an essential role in transporting ferrum in blood after the formation of ferritin complexes and proliferation, differentiation, and even antioxidation. The expression of Tf receptor (TfR) in tumor cells is approximately 100 times that in normal cells. Tf is a glycoprotein that controls extracellular ferrum levels, reversibly binds multivalent ions, including copper, cobalt, and ruthenium [87], and is widely used as a targeting ligand. Coupling anticancer drugs with Tf can significantly improve selectivity and toxicity and overcome drug resistance, resulting in better therapeutic outcomes [88]. Lu et al. [89] designed self-assembled DNA and Tf to form a dual-targeted ruthenium complex with antitumor and antimetastatic properties that inhibits tumor growth and prevents lung metastasis by acting on Tf/TfR.

    iRGD

    VEGF, αvβ3 integrin, matrix metalloproteinase (MMP), and vascular cell binding molecule-1 (VCAM-1) are overexpressed in tumor vascular endothelial cells. Using the tumor homologous peptide iRGD (CRGDKGPDC) as the binding agent, a peptide-mediated delivery strategy for compounds penetrating into the tumor parenchyma was developed [90]. Combined with iRGD, the sensitivity of tumor imaging agents can be significantly improved, which is helpful for deep tumor penetration in the preparation and delivery of therapeutic molecules to the target site and enhancement of the activity of antitumor drugs. Su et al. [91] designed a dual-targeted combination drug delivery system based on multiwalled carbon nanotubes (MWNTs) for the antiangiogenic treatment of lung cancer, which showed significant tumor growth inhibition in A549 cells and xenograft nude mice.

    β2-adrenergic receptor agonists

    β2-adrenergic receptors (β2-ARs) are highly expressed in bronchial smooth muscle and the lung. Different β2-AR agonists, such as formoterol and salmeterol, have been used as targeted ligands to enhance receptor-mediated delivery to the lungs [92]. β-AR belongs to the G-protein-coupled receptor (GPCR) family and is subdivided into three different subtypes: β1, β2, and β3 [93]. The binding of β2-AR agonists to GPCRs leads to the internalization of receptor/ligand complexes through a process mediated by globulin [94]. Elfinger et al. [95] used the β2-AR agonist clenbuterol (Clen) to improve the gene transfer efficiency of the polyethylenimine (PEI) gene vector in vitro in alveolar epithelial cells and in vivo in the mouse lung, which led to the uptake of clenbuterol-specific cells mainly into alveolar epithelial cells, indicating that the coupling of β2-AR ligands to nonviral gene vectors is a promising method to improve gene delivery to the lungs.

    Tumor necrosis factor-related apoptosis-inducing ligand

    Tumor necrosis factor-related apoptosis-inducing ligand (TRAIL) is an apoptotic target currently under exploration for cancer drug development. TRAIL, a member of the tumor necrosis factor (TNF) receptor superfamily, is a type II transmembrane protein that can induce cancer cell apoptosis [96]. As a stable soluble trimer, this protein selectively induces apoptosis in many cancerous cells but does not induce apoptosis in normal cells. Therefore, TRAIL can be used to selectively target cancer cells and has attracted widespread attention as a potential tumor-specific cancer therapeutic agent [97]. Interferon (IFN) activates Apo2 ligand (Apo2L)/TRAIL transcription through specific regulatory elements in its promoter and may be involved in the activation of NK cells, cytotoxic T lymphocytes, and DCs [98]. The activated antibody against the TRAIL receptor and a soluble truncated TRAIL ligand are in phase I/II clinical trials for cancer treatment [99].

    Bombesin

    Bombesin (BN) receptor (BNR), also known as gastrin-releasing peptide (GRP) receptor, belongs to the GPCR superfamily and has been found to be overexpressed in lung cancer, prostate cancer, breast cancer, and pancreatic cancer [100]. In preclinical studies, some nonradioactively labeled ligand-drug complexes constructed by the combination of BN with camptothecin (CPT), doxorubicin (DOX), PTX, and other chemotherapeutic drugs have successfully improved the selectivity and efficacy of these drugs [101]. A nanostructured lipid carrier (NLC) with DOX and DNA-loaded BN-coupled polyethylene glycol (PEG) stearate (BN-PEG-SA) was synthesized and found to serve as a better carrier for improved cellular targeting and nuclear targeting, and its nanodrug may be a promising active targeting drug/gene therapy system for lung cancer treatment [102].

    N-acetyl-d-glucosamine

    Glucose can be used as a targeted molecule in the drug delivery system to promote drug transport and endocytosis [103]. Glucose transporters (GLUT), such as GLUT-1, are overexpressed in tumors of the brain, colon, liver, and lung [104]. GLUTs targeting various tumors have been successfully used for positron emission tomography, magnetic resonance contrast imaging, and gene targeting. As a glucose receptor targeting ligand, N-acetyl-d-glucosamine (GlcNAc) has good water solubility, which is helpful for enhancing the solubility and internalization of NDDSs. A synthetic polymeric drug conjugate, PEG-DOX conjugate, coupled to GlcNAc as a tumor tissue-targeting ligand, showed significantly enhanced cytotoxicity and stronger internalization and retention in cancer cells [105].

    Hyaluronic acid

    Hyaluronic acid (HA) is a widely distributed extracellular matrix polysaccharide with biocompatible and biodegradable properties that belongs to the glycosaminoglycan family [106]. The linear structure of HA is composed of alternating glucuronic acid (GlcA) and GlcNAc, forming a disaccharide β-d-GlcA-(1→3)-β-d-GlcNAc-(1→4) repeat sequence [107, 108]. HA is the ligand of the overexpressed CD44 receptor in NSCLC cell lines [109]. Mattheolabakis et al. [110] demonstrated that HA-modified polymer nanoparticles (PNPs) can improve the accumulation of DOX in tumor cells and improve antitumor efficiency through their involvement in the regulation of inflammation, tumor development, and healing processes [110]. Hsiao et al. [111] discussed the potential of HA-modified nanoparticles in improving apoptosis, cytotoxicity, and anti-proliferation of A549 lung cancer cells.

    Trophoblast cell-surface antigen 2 antibody

    Human trophoblast cell-surface antigen 2 (Trop2) is a transmembrane glycoprotein with an extracellular EGF-like and thyroglobulin 1 repeat domain [112], which is overexpressed in various solid tumors, including NSCLC [113]. High expression of Trop2 is associated with the growth and proliferation of cancer cells and the low survival rate of patients. Trop2 can recognize specific ligands, such as insulin like growth factor-1 (IGF-1), cyclin D1 [114], and protein kinase C (PKC) [115], participate in intracellular signal transduction pathways, and regulate the cell cycle [116]. Trop2 can directly bind to IGF-1 to weaken the activation of AKT/b-catenin and ERK mediated by IGF-1R signaling. Currently, various novel antibody-drug conjugates have been developed, such as sacituzumab govitecan [117] and datopotamab deruxtecan [118], and the coupling of human anti-Trop2 antibodies with cytotoxic drugs will promote the development of tumor therapeutic targets.

    NDDSs in targeted lung cancer therapy

    The current clinical approach to treating lung cancer primarily involves surgery, radiotherapy, and chemotherapy (Table S1, Table 1). However, due to the lack of effective early detection methods, most lung cancer cases are diagnosed at advanced stages involving local tumor invasion or distant metastases. In cases where surgery is not suitable, chemotherapy becomes an option [119]. Conventional chemotherapy in NSCLC has limitations, including lower efficacy, higher toxicity, increased recurrence rates, and lower five-year survival rates [120]. Precision treatment options for NSCLC currently include a combination of local radiotherapy, targeted therapy, and immunotherapy [121, 122]. Immunotherapy, particularly anti-PD1 antibodies such as nivolumab and pembrolizumab, in combination with platinum-based chemotherapy has shown promising results and has been approved by the FDA to improve survival rates in advanced NSCLC patients [123].

    Clinically approved nanoparticle-based medicines for lung cancer therapy

    NameActive ingredientFormulationClinical phaseFirst approval
    Genexol-PMPTXPolymeric micelleApproved2007
    AbraxanePTXAlbuminApproved2012
    LipusuPTXLiposomeApproved2019
    AGEN1181αCTLA-4PhospholipidI/II phase2020
    PEP02IrinotecanLiposomeII phase2015
    DM-CHOC-PENDNA alkylatorLipoid (chloroethyl-cyclohexyl-nitrosourea)I/II phase2021
    NKTR-214IL-2PEGI/II phase2022
    Display full size

    In the field of oncology, nanotechnology has emerged as a potential solution to enhance cancer diagnosis and treatment. Nanotechnology enables targeted delivery of imaging agents and therapeutic drugs to tumor tissues, offering advantages over conventional systemic drug delivery [124]. NDDSs have garnered extensive attention for their ability to facilitate the development of more effective treatments, reduce systemic toxicity, and improve pharmacokinetics [125]. NDDSs have shown potential in the diagnosis, treatment, and prognosis of lung cancer. NDDSs offer several advantages, including an increased drug loading capacity, improved pharmacokinetic properties, passive or active targeting mechanisms for site-specific delivery, synergistic effects of multiple therapeutic agents, minimized drug resistance, and controlled release to enhance efficacy while reducing toxicity. Passive targeting of tumors is based on the EPR effect, where small nanoparticles can accumulate specifically in tumor sites through tumor vascular leakage and lymphatic drainage [126]. Nanoparticles can also interact with TAMs, enhancing their uptake within tumors [127]. Nanoparticles can be surface modified with high-affinity ligands that bind to receptors overexpressed by cancer cells, enabling targeted delivery to tumor cells and metastatic lesions. Stimulus-responsive crosslinked nanomedicines in the field of cancer treatment and have shown advances in circumventing the drawbacks of conventional drug delivery systems [128] and are classified into three categories based on crosslinking strategies, including built-in, on-surface, and interparticle crosslinking nanomedicines. Due to the stimulus-responsive crosslinkages, stimulus-responsive nanomedicines are capable of maintaining robust stability during systemic circulation. They also respond to particular tumoral conditions to induce a series of dynamic changes, such as changes in size, surface charge, targeting moieties, integrity, and imaging signals. These characteristics allow them to efficiently overcome different biological barriers and substantially improve drug delivery efficiency, tumor-targeting ability, and imaging sensitivities. In addition, micro/nanorobots are propelled by chemical reactions, physical fields, and biological systems and can be manipulated by chemotaxis, remote magnetic guidance, and light [129]. Moreover, self-adaptive nanomaterials, which respond to signal changes emitted from the tumor site, might realize spatiotemporally and quantitatively specific release of drugs. Self-adaptive nanomaterials exhibit self-regulation and self-feedback capabilities, and their properties, such as charge, size, and shape, undergo on-demand transformation in response to specific stimuli. Compared to conventional nanomaterials, self-adaptive nanomaterials successfully decrease the frequency of drug release within normal tissues and maintain drug concentrations in tumor cells for a more extended period, thus promoting rational clinical drug application [130]. Clinical use of nanoparticles has demonstrated improved efficacy and reduced adverse effects through adjustment of the systemic biodistribution of drugs, allowing higher doses of therapeutic agents to reach tumor sites. For example, the FDA-approved albumin-bound PTX drug Abraxane showed superior overall response rates and fewer adverse events than generic PTX in NSCLC patients [131].

    Research progress in nanomaterials for lung cancer treatment involves organic nanomaterials [liposomes, polymers, covalent organic framework (COF)], inorganic nanomaterials [gold, paramagnetic iron oxide, silica, carbon quantum dots, metal organic framework (MOF)], and biomimetic nanomaterials (albumin, biofilm materials, Figure 2). The size and composition of nanoparticles significantly influence their bioavailability in vivo, allowing them to pass through specific barriers while limiting uptake in healthy tissues [132, 133]. In summary, nanotechnology, particularly NDDSs, holds great promise for improving lung cancer treatment outcomes. Ongoing research and development of various nanomaterials contribute to advancing the field of lung cancer therapy.

    Nano-based drug delivery system for lung cancer therapy. Created by BioRender. NPs: nanoparticles

    Liposomes

    Liposomes are spherical bilayer vesicles formed by the self-assembly of cholesterol and phospholipids [134] and are structurally similar to biological membranes and capable of efficiently integrating hydrophobic and hydrophilic drugs into hydrophobic cavities and lipid bilayers. Liposomes as drug carriers have the following advantages: (i) nontoxic, safe, and biocompatible carrier materials; (ii) high drug loading efficiency, good stability, and a prolonged drug half-life; (iii) increased drug uptake by tumor cells; (iv) easy surface modification, which can leverage the difference between cancerous and healthy tissues to increase the residence of liposomes at the target site, enhance targeting and reduce the cytotoxicity of drugs [135]; and (v) stimulus-responsive control of drug release that can be achieved by different mechanisms, such as pH and enzymes, depending on the physicochemical properties of the drug to increase efficacy and reduce toxicity. Since cationic liposomes can self-assemble with nucleic acids, many studies have used cationic liposomes to deliver genes to the lung [136]. The main materials used in the preparation of liposomes are soybean phosphatidylcholine, cholesterol, mannitol, lipoic acid-modified polypeptides, phosphoglycerate mutase, and citraconic anhydride-grafted poly-L-lysine, among others. Zhang et al. [137] developed a liposomal curcumin dry powder inhaler for the treatment of primary lung cancer. The uptake of curcumin liposomes by A549 cells was markedly greater and faster than that of free curcumin. Zhang et al. [138] designed a tumor cell membrane-liposome hybrid bionanoparticle surface modified with a peptide targeting MMP-9 with a negatively charged citric anhydride-grafted poly-L-lysine intermediate layer for pH-triggered charge reversal and codelivery of phosphoglycerate mutase 1 small interfering RNA (PGAM1 siRNA) and docetaxel (DTX) to achieve synergistic drug effects of metabolic therapy and chemotherapy. Hu et al. [139] fused tumor-derived nanovesicles with liposomes with a homologous targeting capability and bionic hybrid nanovesicles loaded with DOX [DOX-loaded biomimetic hybrid nanovesicles (DOX@LINV)], which improved the immunosuppressive TME by infiltrating effector immune cells.

    Drug resistance affects the prognosis and survival of lung cancer patients. The mechanisms associated with MDR in lung cancer include drug inactivation, DNA repair, elevated drug release from transporters such as ATP-binding cassette (ABC) transporter proteins and apoptosis defects [140]. The main protein transporter involved in pump resistance is the MDR-associated protein [141], an efflux pump that reduces intracellular drug levels. Off-pump resistance is mainly due to defective apoptosis, and the Bcl-2 protein is a major player in this mechanism [142, 143]. A dual liposome system consisting of adriamycin and siRNA blocked the MDR of Bcl-2 [144]. Delivery of DOX, Bcl-2, and MDR-associated protein 1 via liposomes markedly reduced MDR by stimulating caspase-dependent apoptotic pathways [145]. Li et al. [146] delivered oxygen and erlotinib through aptamer-modified liposome complexes to reverse hypoxia-induced drug resistance in lung cancer.

    Solid lipid-based nanoparticles

    The use of solid lipid-based nanoparticles (SLBNs) in drug delivery has been studied extensively. The two types of SLBNs are solid lipid nanoparticles (SLNs) and NLCs [147]. SLNs can be understood as nanoemulsions in which the liquid lipid nuclei of the droplets are replaced by solid lipid nuclei, such as triglycerides, glycerides, fatty acids, palmitates, and steroids. They are usually prepared using high-pressure homogenization or microemulsification techniques [148] and have colloidal properties, a suitable size, a high surface area-to-volume ratio, good surface properties, physical stability for controlled drug release and easy diffusion, improved stability in the presence of light, humidity, and chemically unstable drugs, and good biocompatibility. SLNs are mainly composed of the materials squalene, soybean phosphatidylcholine, Tween-80, 1,2-dioleoyloxy-3-trimethylammonium-propane (DOTAP), distearoylphosphatidylethanolamine-PEG2000 (DSPE-PEG2000), stearyl amine, monostearin, and poly-lactide-co-glycolide. More active component spaces of NLCs can be obtained by mixing solid and liquid lipids in different proportions to form different lipid matrices [149]. NLCs show better gas atomization characteristics, better accumulation in the lung, good stability, and biological activity integrity [150]. NLC-based drug delivery systems clinical applications are further advanced by various excipients, such as solid lipids, liquid-phase lipids, and surfactants [151], and the inclusion of various cationic components, target/ligand linkers, and bioactive genetic material [152].

    The combination of EGFR TKIs and other chemotherapy drugs in NSCLC is a feasible strategy to overcome EGFR TKI resistance [153]. Garbuzenko et al. [154] developed NLC systems targeting luteinizing hormone-releasing hormone (LHRH) in prepared NSCLC cells to form inhaled LHRH-NLC-siRNA-PTX nanoparticles and tested them using human lung cancer cells with different sensitivities to gefitinib (EGFR inhibitor) and in situ NSCLC mouse models. Yang et al. [155] prepared a lung inhalation microsphere system for SLNs based on stearic acid-loaded afatinib (AFT) spheres [porous microspheres (pMS)] to form AFT-SLN-PTX-pMS nanoparticles. Due to the large surface area of pMS and the high initial burst release that facilitates the rapid release of PTX, this pulmonary inhalation microsphere system has the advantages of long-lasting performance, a high lung deposition rate, rapid release of PTX, sustained release of AFT, and the synergistic effects of AFT and PTX and can be used for the treatment of EGFR TKI-resistant NSCLC. Soni et al. [156] investigated the construction of gemcitabine (GEM)-loaded mannosylated SLNs with superior in vivo pharmacokinetics for effective intracellular delivery to lung cancer cells with the help of SLNs with slow controlled release and a high drug loading rate to improve drug stability, reduce toxicity, enhance efficacy, and improve pharmacokinetics. Moreover, targeted mannose-based SLNs (M-SLNs) were taken up by lung cancer cells through mannose receptor-mediated endocytosis, enhancing the cytotoxic effect on tumor cells and reducing cytotoxicity to normal cells.

    Metallic and inorganic nanoparticles

    Metallic and inorganic nanoparticle as drug delivery carriers have the advantages of a large surface area-to-volume ratio, surface modification of organic materials, being biological molecules, and no immunogenicity. A variety of metal and inorganic nanoparticles have been used for the treatment of NSCLC, including magnetic nanoparticles (MNPs), MOFs, gold nanoparticle, graphene oxide (GO) nanoparticle, quantum dots, plasma, mesoporous silica (MS) nanoparticles (MSNs), and black phosphorus (BP), among others. Their physical and chemical properties are related to size, shape, and composition. The diverse structures and characteristics of nanoparticles provide scaffolds for drug treatment and imaging [157], enhanced stability and bioavailability of anticancer drugs, and sustained control of drug release rates, enabling efficient and targeted treatment of cancer.

    Metallic nanoparticle as drug carriers have the advantages of a high drug-carrying capacity, functional modifiability, and nonimmunogenicity. Such nanoparticle include Au nanoparticle, Cu nanoparticle, MnO2 nanoparticles, Fe3O4 MNPs, and MOFs, which are used as drug delivery carriers for lung cancer therapy. Particle-based pulmonary drug delivery has considerable potential to provide inhalable therapeutics for local or systemic applications. Designing particles with enhanced aerodynamic properties can improve lung distribution and deposition, thereby enhancing the efficacy of capsule-based inhaled drugs. As nanocarriers for the inhalation of small-molecule drugs and macromolecular drugs, metallic phenol capsules can increase the thickness of the capsule shell by repeatedly depositing thin films on the template, thereby increasing the aerodynamic diameter and accurately controlling the deposition of the capsule shell in a human lung model [158].

    An important therapeutic approach based on metal materials is magnetic nanotherapy, a noninvasive method of tumor ablation based on MNPs acting on their own or with anticancer drugs and external magnetic fields, where drug-laden nanoparticles can be guided into the circulatory system and targeted tissues under the action of an alternating magnetic field [159, 160]. Iron oxide nanoparticles are the most commonly used nanoparticles for MNP therapy due to their degradability, biocompatibility, and superparamagnetic effect [161]. Iron ions can catalyze the generation of free radicals from hydrogen peroxide through the Fenton-type reaction, which can damage mitochondria, lipids, proteins, DNA, and other structures in tumor cells and induce apoptosis [162, 163]. Tseng et al. [164] showed that recombinant adeno-associated virus serotype 2 (AAV2) chemically conjugated with iron oxide nanoparticles (approximately 5 nm) has a remarkable ability to be remotely guided under a magnetic field. Transduction is achieved with microscale precision. Furthermore, a gene for the production of the photosensitive protein KillerRed was introduced into the AAV2 genome to enable photodynamic therapy (PDT) or light-triggered virotherapy. In vivo animal experiments revealed that magnetic guidance of “ironized” AAV2-KillerRed injected through the tail vein in conjunction with PDT significantly decreased tumor growth via apoptosis. Sadhukha et al. [165] investigated the synthesis of inhalable superparamagnetic iron oxide (SPIO) nanoparticle (SPIONs) targeting EGFR, and the results confirmed that magnetic thermotherapy using SPIO-targeted nanoparticles significantly inhibited tumor growth in vivo and that magnetic thermotherapy has strong potential as a treatment modality for NSCLC. Huang et al. [166] combined porous iron oxide nanoagents (PION)-mediated promoting photothermal therapy (PTT) with CRYBG3 long noncoding RNA (lncRNA)-mediated gene therapy, where PIONs were used as a magnetic nanoagent for magnetic resonance imaging (MRI) and photoacoustic imaging (PAI), and a high cancer cell killing effect was observed in vitro and in vivo.

    Functional modification of nanocarriers can enable integration of multiple functions into NDDSs to achieve targeted drug delivery, diagnostic imaging, and combined therapy. Ma et al. [167] used an integrated therapeutic and diagnostic targeting nanoplatform for lung cancer in situ spinal metastases for the first time (Figure 3). Au@MOF was coated with MSNs and connected with the photosensitizer indocyanine green (ICG) to alter the modified targeting peptide dYNH on Au@MOF@MS-ICG. The dYNH peptide was combined with the transmembrane receptor erythropoietin, which produces human hepatocellular A2 (EphA2) [168], and the outer layer was modified with polyacrylic acid (PAA) to enable pH-sensitive codelivery of cisplatin and the alpha-selective PI3K inhibitor BYL719@Au@MOF@MS-ICG, representing an integrated therapeutic nanoplatform combining dual drugs, strong targeting, and photothermal effects. In addition, imaging modalities are expected to serve as an efficient and safe smart drug delivery system for the treatment of in situ metastases.

    Synthesis of BCAMMD modified bivalve nanoparticles (BCAMM) loaded with BYL719-cisplatin. BCAMM: BYL719&cisplatin@Au@MOF@MS-ICG; BCAMMD: dYNH-targeted peptide; Cis: cisplatin; CTAB: hexadecyltrimethylammonium chloride solution; MI: methylimidazole; TEOS: tetraethyl orthosilicate

    Note. Reprinted with permission from “Rationally integrating peptide-induced targeting and multimodal therapies in a dual-shell theranostic platform for orthotopic metastatic spinal tumors” by Ma Y, Chen L, Li X, Hu A, Wang H, Zhou H, et al. Biomaterials. 2021;275:120917 (https://linkinghub.elsevier.com/retrieve/pii/S0142961221002738). © Elsevier 2021.

    Inorganic nonmetallic nanoparticles have unique physicochemical properties and biological effects; for instance, GO nanoparticle can effectively load and deliver antigens, show the potential to activate the immune system, have easily modified surfaces, and are widely used as carriers and immune adjuvants [169]. BP nanoparticle are a new type of phosphorus-source nanoparticle, and BP quantum dots (BPQDs) and BP hydrogels (BPHs) are both common forms BP. Due to their high photothermal conversion efficiency, excellent PTT and PDT and good biocompatibility, BP-based drug delivery systems have received attention and are widely used [170].

    Quantum dot-based fluorescence strategies can improve sensing performance and enhance sensitivity to targets due to their optical and photophysical properties and adjustable emission range [171, 172]. Moreover, integrating organic dyes and quantum dots into peptide substrates provides a well-controlled and scalable strategy for protease-sensing fluorescence resonance energy transfer (FRET) systems. Wu et al. [173] constructed a nitrogen-rich carbon dot (NCD)-mediated DNA nanostructure self-assembly strategy. Due to the excellent photoluminescence and photodynamic properties of NCDs, NCDs can isothermally mediate DNA nanopillar self-assembly in a magnesium-free manner in large temperature and pH ranges and combine with KRAS siRNA for the treatment of KRAS mutant NSCLC. Studies have shown that the combination of NCDs and programmable DNA nanostructures is a powerful strategy for endowing DNA nanostructures with new functions, and the nanoplatform has phototherapy potential.

    MSNs are one of the most widely used inorganic nanoparticles, with the advantages of large pore size, an adjustable particle size and pore diameter, a large specific surface area, a high density of silanol groups on the surface, favorable functional modification, excellent biocompatibility, and can induce reactive oxygen species generation [174], rendering MSNs ideal carriers for adsorbable drug molecules [175, 176]. Zhou et al. [177] prepared a novel injectable thermosensitive hydrogel based on MSNs and thermosensitive hydrogels for the delivery of the oral targeted drug erlotinib to achieve targeted sustained release. The erlotinib-loaded hydrogel composite (ERT@HMSNs/gel) complex showed a longer drug retention time in and around the tumor, enhancing the efficacy of NSCLC. This development also provides ideas for the design and preparation of nanodelivery systems for local antitumor therapy. Cheng et al. [178] developed a d-a-tocopheryl PEG 1000 succinate (TPGS)-functionalized polydopamine-coated MSN drug delivery systems for pH-responsive release of DOX with ideal particle size, drug loading, and drug release characteristics and long cycle advantages. Drug-resistant A549 cells were used to detect the cytotoxicity and cellular uptake of nanoparticle. NDDSs showed outstanding performance in overcoming MDR.

    PNPs

    PNPs are amphiphilic polymers self-assembled by hydrophobic interactions in aqueous solution to form a thermodynamically stable system. Hydrophobic small molecules are trapped within PNPs by covalent bonding or interaction with the hydrophobic core, and hydrophilic drugs can be loaded by physical action or chemical coupling to effectively carry antitumor drugs and remain stable in vivo [179]. PNPs are easy to prepare, have good biocompatibility, low toxicity, structural stability, chemical modification abilities, and multifunctional groups that can bind specific ligands or antibodies [180] to achieve the advantages of targeted drug delivery, and are widely used in drug delivery systems [181, 182], showing strong potential for chemotherapy and gene therapy. CRLX101 is a self-assembled nanoparticle containing PEG-poly(lactic acid)-encapsulated CPT that delivers CPT to cancer cells while significantly reducing systemic exposure and is currently in phase II clinical trials [183]. Tseng et al. [184] reported the exploitation of NSCLC tumor-secreted lactate in designing an acid-degradable nanoparticle containing the acyclic acetal component of oxidized HA for viral release. The virus, lactate oxidase (LOX), and hexanoamide were conjugated with aldehyde-HA through reductive amination (Figure 4) [185]. The lower pH can facilitate virus internalization into cells due to pH-sensitive proteases of the viral capsid. Site-specific delivery was demonstrated by viral transduction in the NSCLC tumor-secreted lactate microenvironment, offering an avenue for improving general or drug-resistant NSCLC treatment outcomes. The exploitation of tumor lactate production in designing a hypoxia-responsive carrier self-assembled from HA conjugated with 6-(2-nitroimidazole)hexylamine for localized release of recombinant AAV2 has also been reported. The carrier is loaded with LOX and is permeable to small molecules such as the lactate that accumulates in a tumor [186]. Wang et al. [185] developed hierarchically responsive nanomedicine (HRNMs) self-assembled via a cyclic Arg-Gly-Asp (RGD) peptide-coupled triblock copolymer, poly(2-(hexamethyleneimino)ethyl methacrylate)-poly(oligo-(ethylene glycol) monomethyl ether methacrylate)-poly reduction-responsive CPT (PC7A-POEG-PssCPT). In circulation, RGD peptides are shielded by the POEG coating, and HRNMs achieve effective tumor aggregation through passive targeting. Upon reaching the tumor site, the acidic microenvironment induces hydrophobic to hydrophilic conversion of PC7A, and RGD peptides are exposed, enhancing tumor retention and intracellularization. Furthermore, HRNMs show effective tumor targeting, potent antitumor effects, and reduced systemic toxicity. Such HRNMs are expected to be used to enhance chemotherapeutic delivery. Zhong et al. [187] designed a series of biodegradable PEG, guanidine-functionalised polycarbonate and polypropylene cross-ester (PEG-PGCx-PDLAy) triblock copolymers as chemotherapeutic agents that self-assemble into micellar nanoparticles against a variety of cancer cell lines, which killed cancer cells through a nonapoptotic mechanism involving significant vacuolization and subsequent membrane disruption without inducing resistance to multiple treatments at sublethal doses of the polymer. Iyer et al. [188] developed glutathione (GSH)-responsive polyurethane nanoparticles loaded with cisplatin, which showed GSH dose-dependent cisplatin release and significantly reduced in vitro survival of A549 lung cancer cells following the action of GSH-responsive polyurethane nanoparticles (GPUs) compared to that achieved with equal concentrations of free cisplatin. In vivo biodistribution studies showed that fluorescently labeled GPUs clustered in lung tumor areas and that tumor suppression was significantly improved following tail vein injection in mice. Wang et al. [189] constructed poly(lactide-co-glycolide)-PEG-FITC-poly(ethylene glycol)-amine-MSC (PLGA-PEG-FITC-MSC) nanoparticles loaded with DTX. MSCs can be used as lung-targeting drug carriers [190] and exhibit naturally high tumor affinity [191], and the MSC/nanoparticles (MSC/NP) system can effectively target drugs to lung tissue. The tumor inhibition efficiency of the MSC/nanoparticles/DTX (MSC/NP/DTX) system was similar to that of nanoparticles/DTX (NP/DTX) but at only 1/8 of the DTX dose. Chen et al. [192] prepared pH and redox dual-responsive methoxy PEG-disulfide bond-poly(beta-amino ester)-PLGA (mPEG-SS-PBAE-PLGA) nanoparticle-loaded platinum-curcumin complexes, which facilitated intracellular release and enhanced synergistic anticancer effects.

    Schematic diagram of programmed administration of HRNMs. (1) HRNMs have high stability in the circulation, (2) due to their nanoparticle size and neutral POEG surface; therefore, they can effectively accumulate, within the tumor through EPR effects. (3) In the TME, the acidic pH leads to charge switching and exposure of RGD peptides by HRNMs, which enhances tumor retention and (4) intracellularization. (5) Thereafter, intracellular GSH will trigger the release of CPT from cancer chemotherapy

    Note. Reprinted with permission from “Hierarchical tumor microenvironment-responsive nanomedicine for programmed delivery of chemotherapeutics” by Wang S, Yu G, Wang Z, Jacobson O, Tian R, Lin LS, et al. Adv Mater. 2018:e1803926 (https://onlinelibrary.wiley.com/doi/10.1002/adma.201803926). © WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 2018.

    Many noninvasive technologies, such as near-infrared (NIR) and ultrasound therapy, have been applied to the treatment of cancer. Compared with lasers, which have a limited penetration depth, ultrasound energy can penetrate deeper tumor tissues in a safe range, which is more suitable for the treatment of lung cancer. Wang et al. [193] proposed a novel NIR light-triggered photothermal polymer containing DAP-F complexed with a reduction-sensitive amphiphilic polymer, P1, to form F-nanoparticles with photothermal effects that can encapsulate Pt(IV) prodrugs via P1 and bind F-nanoparticles to Pt-nanoparticles to construct the final nanosystem F-Pt-nanoparticles, which inhibited DNA repair, effectively overcame cisplatin resistance, and suppressed tumors (Figure 5).

    Schematic illustration of the preparation of the F-Pt-NPs and possible mechanism involved in inhibition of cisplatin resistance under NIR laser irradiation. PDX: patient-derived xenograft

    Note. Reprinted with permission from “A systematic strategy of combinational blow for overcoming cascade drug resistance via NIR-light-triggered hyperthermia” by Wang L, Yu Y, Wei D, Zhang L, Zhang X, Zhang G, et al. Adv Mater. 2021;33:e2100599 (https://onlinelibrary.wiley.com/doi/10.1002/adma.202100599). © Wiley-VCH GmbH 2021.

    Biomimetic nanoparticles

    Biomimetic nanoparticles have the morphology, surface properties, and size of natural constituent structures (e.g., exosomes) that evade clearance by the immune system, an enhanced targeting ability to deliver drugs to target cells or tissues, good biocompatibility, improved therapeutic efficiency, and reduced toxic side effects and are widely studied as drug delivery systems. With hybridization, nucleic acid nanoassemblies, including DNA and RNA nanostructures, can be applied to enhance diagnosis and therapeutic effects [194]. Kuerban et al. [195] obtained outer-membrane vesicles (OMVs) from attenuated Klebsiella pneumonia and prepared DOX-loaded OMVs, which were efficiently transported into A549 cells and presented substantial tumor growth inhibition with favorable tolerability and pharmacokinetic profiles. The appropriate immunogenicity of OMVs enables the recruitment of macrophages to the TME, which might synergize with DOX in vivo. Li et al. [196] found that inhibition of exosome secretion may be an effective strategy to overcome the antagonistic effects of TKIs and chemotherapeutic agents. Zhong et al. [197] constructed three different DNA nanostructures and found that DNA nanostructures could significantly enhance the intracellularization of platinum drugs, thereby increasing their anticancer activity not only against conventional A549 cells but also more importantly against cisplatin-resistant (CisR) cancer cells (A549 CisR), thus effectively inhibiting tumor growth, suggesting that DNA nanostructures are effective carriers for platinum precursor drug delivery to counter chemotherapy resistance.

    Proteins, such as albumin [198] and high-density lipoproteins (HDLs) [199], and silk fibroin nanoparticles (SFNPs) [200] have been used to treat NSCLC, and their amino and carboxylic acid fractions can be chemically modified to actively target cancer cells. Among all types of polypeptide nanoparticles, polycystine 2 (PCys2)-based nanoparticles have drawn increasing attention due to their unique properties. On the one hand, uniform nanogels can be easily obtained through the crosslinking of two active centers during polymerization without an additional self-assembly step. On the other hand, the Cys2-based nanoparticles always showed reduction responsiveness owing to the inherent disulfide bond. With the development of advanced diagnostic and therapeutic technologies, multifunctional PCys2-based nanoparticles were achieved via rational construction of the polymer structure [201]. Elgohary et al. [202] developed inhalable human serum albumin (HSA) nanocomposites for the combined delivery of etoposide (ETP) and berberine (BER) to lung tumors, which reduced toxicity, prevented drug resistance, and enhanced cytotoxicity and internalization in A549 lung cancer cells. Cell and cell membrane-derived nanobiocarriers are being developed to design new drug delivery strategies for the treatment of various malignant tumors, including lung cancer. At present, the sources of this biological nanocarrier mainly include red blood cells, white blood cells, platelets, MSCs, cancer cells, and exosomes. Cell-based nanobiocarriers have good biocompatibility, multimolecular and intrinsic targeting abilities, a long cycle ability, and good host biointegration and can be used for the treatment of various tumors. Cell membrane-derived microparticles (MPs) are important mediators of intercellular information transfer. Guo et al. [203] investigated the therapeutic potential of tumor cell-derived MPs (TMPs) in patients with malignant pleural effusion (MPE) from lung cancer. The safety, immunogenicity, and clinical activity of TMPs-methotrexate (TMPs-MTX) were validated in a human study in 11 patients with MPE in advanced lung cancer, showing a significant reduction in tumor cells and CD163+ macrophages and stimulation of IL-2 release from CD4+ T cells and IFN-γ release from CD8+ T cells. Ye et al. [198] used neutrophil-mediated nanoparticles to promote tumor photothermal therapy by modifying Au nanorods (AuNR) with bovine serum albumin (BSA) coupled to RGD (AuNRBR) and then encapsulated them in neutrophils (AuNRBR/N). Neutrophils can efficiently cross epithelial cells and exhibit enhanced cytotoxicity against Lewis cells under in vitro laser irradiation. In addition, AuNRBR/N showed stronger targeting to tumor tissue than cell-free nanoparticles, and the enhanced tumor-homing efficiency of AuNRBR/N and the release of AuNRBR from neutrophils facilitated further deep tissue dissemination, contributing to tumor growth inhibition and improved survival in PTT.

    Conclusions

    Patients with lung cancer often face low overall survival, progression-free survival, and quality of life. Current interventions for lung cancer primarily target advanced stages of the disease characterized by signs of tumor metastasis and recurrence, which are influenced by genetic mutations, drug resistance, and epithelial-mesenchymal transition (EMT) processes [204]. While new therapies combined with traditional drugs have shown some clinical success, NDDSs have become increasingly important in clinical trials and applications. NDDSs offer the potential to enhance treatment efficacy and reduce side effects by modulating the biodistribution and pharmacokinetics of drugs. However, despite promising results, further research is necessary to develop more effective nanoplatforms. Currently, passive targeting based on the EPR effect is a common approach for tumor uptake of nanobased agents [205]. However, the EPR effect is typically less prominent in humans than in animal models [206]. To address the limited tumor uptake, nanoparticles can be designed to actively target tumor tissue-specific overexpressed proteins using a combination of active and passive targeting strategies, facilitating receptor-mediated endocytosis. This active targeting strategy aims to simultaneously target primary and metastatic tumor cells [207], as metastatic cells may not exhibit strong EPR effects in the early stages.

    However, many remaining challenges require considerable research. While NDDSs have shown promise in overcoming drug resistance in lung cancer treatment, the exact mechanisms by which these systems can effectively deliver drugs to tumor cells and bypass drug resistance mechanisms are still not well understood. Further research is needed to elucidate the specific interactions between nanoparticle-based drug carriers and tumor cells regarding drug resistance. In addition, although many different types of nanoparticles have been used for lung cancer therapy, a consensus on the optimal characteristics of nanoparticle design for effective drug delivery is still lacking. Research is needed to determine the influence of nanoparticle size, shape, surface properties, and targeting ligands on the efficiency of drug delivery to lung tumor cells. Understanding the pharmacokinetics, biodistribution, and clearance of NDDSs is crucial for their successful translation into clinical use. More studies are needed to investigate the fate of nanoparticles in the lung, including their distribution within tumor tissues, metabolism, and excretion pathways. While NDDSs hold great promise, their potential toxicity and safety concerns need to be thoroughly investigated. Systemic administration of nanoparticles for lung cancer therapy may lead to off-target effects and accumulation in healthy tissues, which could have adverse toxicological effects. Further research is needed to assess the safety profiles of different nanoparticle formulations and develop strategies to minimize potential toxicity. Preclinical studies have explored active targeting strategies using EGFR-targeting peptides [165, 208], HA [209], folic acid [210, 211], Tf, and other ligands in lung cancer models. Despite significant advancements in the preclinical development of NDDSs for lung cancer therapy, very few have been successfully translated to clinical use. However, clinical trials involving these active targeting agents in lung cancer patients are still underway. Considering the heterogeneity of tumors, employing multiple ligands targeting different cell surface receptors is preferable [212]. Alternatively, specific targeted agents can be designed based on the individualized driver mutation characteristics of different subtypes of lung cancer considering the TME and controlling drug release to minimize off-target toxicity and enhance precision therapeutic effects. Research gaps exist in terms of the scalability, manufacturing reproducibility, and clinical feasibility of these systems. Further studies are needed to address the challenges related to large-scale production and clinical translation of effective nanodrug delivery platforms for lung cancer treatment. Currently, clinical treatment with nanomedicines in lung cancer patients shows promise. However, further improvements in their structures and the combination of multiple NDDS approaches are necessary to provide multidrug treatment options and achieve optimal NDDSs with superior functional and structural properties, high reproducibility, simple preparation methods, and low costs.

    Abbreviations

    AAV2:

    adeno-associated virus serotype 2

    AFT:

    afatinib

    AKT:

    protein kinase B

    ALK:

    anaplastic lymphoma kinase

    Bcl-2:

    B cell lymphoma 2

    BN:

    bombesin

    BP:

    black phosphorus

    BRAF:

    v-raf murine sarcoma viral oncogene homolog B1

    CPT:

    camptothecin

    CTLA4:

    cytotoxic T lymphocyte associated protein 4

    DOX:

    doxorubicin

    DTX:

    docetaxel

    EGF:

    epidermal growth factor

    EGFR:

    epidermal growth factor receptor

    EPR:

    enhanced permeability and retention

    ErbB:

    Erb-B2 receptor tyrosine kinase 2

    ERK:

    extracellular signal-regulated kinase

    FDA:

    U.S. Food and Drug Administration

    FR:

    folate receptor

    GlcNAc:

    N-acetyl-d-glucosamine

    GLUT:

    glucose transporter

    GPCR:

    G-protein-coupled receptor

    GSH:

    glutathione

    GTP:

    guanosine triphosphate

    HA:

    hyaluronic acid

    HER:

    human epidermal growth factor receptor

    HGF:

    hepatocyte growth factor

    HGFR:

    hepatocyte growth factor receptor

    HRNMs:

    hierarchically responsive nanomedicine

    ICG:

    indocyanine green

    IGF-1:

    insulin like growth factor-1

    JAK:

    Janus kinase

    KRAS:

    Kirsten rat sarcoma virus oncogene homolog

    MDR:

    multidrug resistance

    MEK1:

    mitogen-activated protein kinase kinase 1

    MET:

    mesenchymal-epithelial transition factor

    MNPs:

    magnetic nanoparticles

    MOF:

    metal organic framework

    MS:

    mesoporous silica

    MSCs:

    mesenchymal stem cells

    MSNs:

    mesoporous silica nanoparticles

    NCD:

    nitrogen-rich carbon dot

    NDDS:

    nanodrug delivery system

    NIR:

    near-infrared

    NK:

    natural killer

    NLC:

    nanostructured lipid carrier

    NSCLC:

    non-small cell lung cancer

    OMVs:

    outer-membrane vesicles

    PAA:

    polyacrylic acid

    PC7A:

    peptide-coupled triblock copolymer, poly(2-(hexamethyleneimino)ethyl methacrylate)

    PD-L1:

    programmed cell death ligand 1

    PDT:

    photodynamic therapy

    PEG:

    polyethylene glycol

    PI3K:

    phosphatidylinositol 3-kinase

    pMS:

    porous microspheres

    PNPs:

    polymer nanoparticles

    POEG:

    poly(oligo-(ethylene glycol) monomethyl ether methacrylate)

    PTT:

    promoting photothermal therapy

    PTX:

    paclitaxel

    RAF:

    rapidly accelerated fibrosarcoma

    RAS:

    rat sarcoma

    RGD:

    Arg-Gly-Asp

    ROS1:

    ROS proto-oncogene 1-receptor tyrosine kinase

    SCLC:

    small cell lung cancer

    siRNA:

    small interfering RNA

    SLNs:

    solid lipid nanoparticles

    STAT:

    signal transducer and activator of transcription

    TAMs:

    tumor-associated macrophages

    Tf:

    transferrin

    TKIs:

    tyrosine kinase inhibitors

    TKs:

    tyrosine kinases

    TME:

    tumor microenvironment

    TRAIL:

    tumor necrosis factor-related apoptosis-inducing ligand

    Trop2:

    trophoblast cell-surface antigen 2

    VEGF:

    vascular endothelial growth factor

    VEGFR:

    vascular endothelial growth factor receptor

    β2-ARs:

    β2-adrenergic receptors

    Supplementary materials

    The supplementary material for this article is available at: https://www.explorationpub.com/uploads/Article/file/1001221_sup_1.pdf

    Declarations

    Author contributions

    SZ: Conceptualization, Writing—original draft. XL and YL: Writing—review & editing. HL and ZZ: Supervision.

    Conflicts of interest

    The authors declare no conflicts of interest.

    Ethical approval

    Not applicable.

    Consent to participate

    Not applicable.

    Consent to publication

    Not applicable.

    Availability of data and materials

    Not applicable.

    Funding

    This work was supported by the National Natural Science Foundation of China [82003680, 82111530241], Shandong Provincial Natural Science Foundation [ZR2020QH350, ZR2021QH024, ZR2023YQ065], Open Projects Fund of NMPA Key Laboratory for Technology Research and Evaluation of Drug Products [No. 2022TREDP03], and the Chinese “post-doctoral international exchange program”. The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.

    Copyright

    © The Author(s) 2024.

    References

    Sung H, Ferlay J, Siegel RL, Laversanne M, Soerjomataram I, Jemal A, et al. Global Cancer Statistics 2020: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J Clin. 2021;71:20949. [DOI] [PubMed]
    Rawal S, Patel M. Bio-nanocarriers for lung cancer management: befriending the barriers. Nanomicro Lett. 2021;13:142. [DOI] [PubMed] [PMC]
    Cryer AM, Thorley AJ. Nanotechnology in the diagnosis and treatment of lung cancer. Pharmacol Ther. 2019;198:189205. [DOI] [PubMed]
    Herbst RS, Morgensztern D, Boshoff C. The biology and management of non-small cell lung cancer. Nature. 2018;553:44654. [DOI] [PubMed]
    Sharma P, Mehta M, Dhanjal DS, Kaur S, Gupta G, Singh H, et al. Emerging trends in the novel drug delivery approaches for the treatment of lung cancer. Chem Biol Interact. 2019;309:108720. [DOI] [PubMed]
    Mottaghitalab F, Farokhi M, Fatahi Y, Atyabi F, Dinarvand R. New insights into designing hybrid nanoparticles for lung cancer: diagnosis and treatment. J Control Release. 2019;295:25067. [DOI] [PubMed]
    Ahmad J, Akhter S, Rizwanullah M, Amin S, Rahman M, Ahmad MZ, et al. Nanotechnology-based inhalation treatments for lung cancer: state of the art. Nanotechnol Sci Appl. 2015;8:5566. [DOI] [PubMed] [PMC]
    Cook M, Qorri B, Baskar A, Ziauddin J, Pani L, Yenkanchi S, et al. Small patient datasets reveal genetic drivers of non-small cell lung cancer subtypes using machine learning for hypothesis generation. Explor Med. 2023;4:42840. [DOI]
    Xie W, Liu S, Li G, Xu H, Zhou L. The evolving treatment paradigm of lung cancer in China. Acta Pharm Sin B. 2022;12:15367. [DOI] [PubMed] [PMC]
    Bernabeu E, Cagel M, Lagomarsino E, Moretton M, Chiappetta DA. Paclitaxel: What has been done and the challenges remain ahead. Int J Pharm. 2017;526:47495. [DOI] [PubMed]
    Pelaz B, Alexiou C, Alvarez-Puebla RA, Alves F, Andrews AM, Ashraf S, et al. Diverse applications of nanomedicine. ACS Nano. 2017;11:231381. [DOI] [PubMed] [PMC]
    Patra JK, Das G, Fraceto LF, Campos EVR, Rodriguez-Torres MDP, Acosta-Torres LS, et al. Nano based drug delivery systems: recent developments and future prospects. J Nanobiotechnology. 2018;16:71. [DOI] [PubMed] [PMC]
    Forman HJ, Zhang H. Targeting oxidative stress in disease: promise and limitations of antioxidant therapy. Nat Rev Drug Discov. 2021;20:689709. [DOI] [PubMed] [PMC]
    Sattar R, Shahzad F, Ishaq T, Mukhtar R, Naz A. Nano-drug carriers: a potential approach towards drug delivery methods. ChemistrySelect. 2022;7:e202200884. [DOI]
    Thai AA, Solomon BJ, Sequist LV, Gainor JF, Heist RS. Lung cancer. Lancet. 2021;398:53554. [DOI] [PubMed]
    Larsen JE, Minna JD. Molecular biology of lung cancer: clinical implications. Clin Chest Med. 2011;32:70340. [DOI] [PubMed] [PMC]
    Ahrendt SA, Chow JT, Yang SC, Wu L, Zhang MJ, Jen J, et al. Alcohol consumption and cigarette smoking increase the frequency of p53 mutations in non-small cell lung cancer. Cancer Res. 2000;60:31559. [PubMed]
    Rotow J, Bivona TG. Understanding and targeting resistance mechanisms in NSCLC. Nat Rev Cancer. 2017;17:63758. [DOI] [PubMed]
    Ciardiello F, Tortora G. EGFR antagonists in cancer treatment. N Engl J Med. 2008;358:116074. [DOI] [PubMed]
    Singh B, Carpenter G, Coffey RJ. EGF receptor ligands: recent advances. F1000Res. 2016;5:2270. [DOI] [PubMed] [PMC]
    Ryslik GA, Cheng Y, Cheung KH, Modis Y, Zhao H. Utilizing protein structure to identify non-random somatic mutations. BMC Bioinformatics. 2013;14:190. [DOI] [PubMed] [PMC]
    Marchetti A, Ardizzoni A, Papotti M, Crinò L, Rossi G, Gridelli C, et al. Recommendations for the analysis of ALK gene rearrangements in non-small-cell lung cancer: a consensus of the Italian Association of Medical Oncology and the Italian Society of Pathology and Cytopathology. J Thorac Oncol. 2013;8:3528. [DOI] [PubMed]
    Tomizawa K, Suda K, Onozato R, Kosaka T, Endoh H, Sekido Y, et al. Prognostic and predictive implications of HER2/ERBB2/neu gene mutations in lung cancers. Lung Cancer. 2011;74:13944. [DOI] [PubMed]
    Klotz LV, Courty Y, Lindner M, Petit-Courty A, Stowasser A, Koch I, et al. Comprehensive clinical profiling of the Gauting locoregional lung adenocarcinoma donors. Cancer Med. 2019;8:148699. [DOI] [PubMed] [PMC]
    Li BT, Ross DS, Aisner DL, Chaft JE, Hsu M, Kako SL, et al. HER2 amplification and HER2 mutation are distinct molecular targets in lung cancers. J Thorac Oncol. 2016;11:4149. [DOI] [PubMed] [PMC]
    Liu S, Li S, Hai J, Wang X, Chen T, Quinn MM, et al. Targeting HER2 aberrations in non-small cell lung cancer with osimertinib. Clin Cancer Res. 2018;24:2594604. [DOI] [PubMed] [PMC]
    Liao ZX, Huang KY, Kempson IM, Li HJ, Tseng SJ, Yang PC. Nanomodified strategies to overcome EGFR-tyrosine kinase inhibitors resistance in non-small cell lung cancer. J Control Release. 2020;324:48292. [DOI] [PubMed]
    Downward J. Targeting RAS signalling pathways in cancer therapy. Nat Rev Cancer. 2003;3:1122. [DOI] [PubMed]
    Wen Z, Jiang R, Huang Y, Wen Z, Rui D, Liao X, et al. Inhibition of lung cancer cells and Ras/Raf/MEK/ERK signal transduction by ectonucleoside triphosphate phosphohydrolase-7 (ENTPD7). Respir Res. 2019;20:194. [DOI] [PubMed] [PMC]
    Linardou H, Dahabreh IJ, Kanaloupiti D, Siannis F, Bafaloukos D, Kosmidis P, et al. Assessment of somatic k-RAS mutations as a mechanism associated with resistance to EGFR-targeted agents: a systematic review and meta-analysis of studies in advanced non-small-cell lung cancer and metastatic colorectal cancer. Lancet Oncol. 2008;9:96272. [DOI] [PubMed]
    Davies H, Bignell GR, Cox C, Stephens P, Edkins S, Clegg S, et al. Mutations of the BRAF gene in human cancer. Nature. 2002;417:94954. [DOI] [PubMed]
    Leonetti A, Facchinetti F, Rossi G, Minari R, Conti A, Friboulet L, et al. BRAF in non-small cell lung cancer (NSCLC): pickaxing another brick in the wall. Cancer Treat Rev. 2018;66:8294. [DOI] [PubMed]
    Planchard D, Smit EF, Groen HJM, Mazieres J, Besse B, Helland Å, et al. Dabrafenib plus trametinib in patients with previously untreated BRAFV600E-mutant metastatic non-small-cell lung cancer: an open-label, phase 2 trial. Lancet Oncol. 2017;18:130716. [DOI] [PubMed]
    Morris SW, Kirstein MN, Valentine MB, Dittmer K, Shapiro DN, Look AT, et al. Fusion of a kinase gene, ALK, to a nucleolar protein gene, NPM, in non-Hodgkin’s lymphoma. Science. 1994;263:12814. [DOI] [PubMed]
    Soda M, Choi YL, Enomoto M, Takada S, Yamashita Y, Ishikawa S, et al. Identification of the transforming EML4-ALK fusion gene in non-small-cell lung cancer. Nature. 2007;448:5616. [DOI] [PubMed]
    Guan J, Umapathy G, Yamazaki Y, Wolfstetter G, Mendoza P, Pfeifer K, et al. FAM150A and FAM150B are activating ligands for anaplastic lymphoma kinase. Elife. 2015;4:e09811. [DOI] [PubMed] [PMC]
    Katayama R. Therapeutic strategies and mechanisms of drug resistance in anaplastic lymphoma kinase (ALK)-rearranged lung cancer. Pharmacol Ther. 2017;177:18. [DOI] [PubMed]
    Jiang W, Cai G, Hu P, Wang Y. Personalized medicine of non-gene-specific chemotherapies for non-small cell lung cancer. Acta Pharm Sin B. 2021;11:340616. [DOI] [PubMed] [PMC]
    Satoh H, Yoshida MC, Matsushime H, Shibuya M, Sasaki M. Regional localization of the human c-ros-1 on 6q22 and flt on 13q12. Jpn J Cancer Res. 1987;78:7725. [PubMed]
    Acquaviva J, Wong R, Charest A. The multifaceted roles of the receptor tyrosine kinase ROS in development and cancer. Biochim Biophys Acta. 2009;1795:3752. [DOI] [PubMed]
    Davies KD, Le AT, Theodoro MF, Skokan MC, Aisner DL, Berge EM, et al. Identifying and targeting ROS1 gene fusions in non-small cell lung cancer. Clin Cancer Res. 2012;18:45709. [DOI] [PubMed] [PMC]
    Uguen A, De Braekeleer M. ROS1 fusions in cancer: a review. Future Oncol. 2016;12:191128. [DOI] [PubMed]
    Drilon A, Jenkins C, Iyer S, Schoenfeld A, Keddy C, Davare MA. ROS1-dependent cancers – biology, diagnostics and therapeutics. Nat Rev Clin Oncol. 2021;18:3555. [DOI] [PubMed] [PMC]
    Lin JJ, Shaw AT. Recent advances in targeting ROS1 in lung cancer. J Thorac Oncol. 2017;12:161125. [DOI] [PubMed] [PMC]
    Jänne PA, Engelman JA, Johnson BE. Epidermal growth factor receptor mutations in non-small-cell lung cancer: implications for treatment and tumor biology. J Clin Oncol. 2005;23:322734. [DOI] [PubMed]
    Koch JP, Aebersold DM, Zimmer Y, Medová M. MET targeting: time for a rematch. Oncogene. 2020;39:284562. [DOI] [PubMed]
    Feng J, Lian Z, Xia X, Lu Y, Hu K, Zhang Y, et al. Targeting metabolic vulnerability in mitochondria conquers MEK inhibitor resistance in KRAS-mutant lung cancer. Acta Pharm Sin B. 2023;13:114563. [DOI] [PubMed] [PMC]
    Reungwetwattana T, Liang Y, Zhu V, Ou SI. The race to target MET exon 14 skipping alterations in non-small cell lung cancer: the Why, the How, the Who, the Unknown, and the Inevitable. Lung Cancer. 2017;103:2737. [DOI] [PubMed]
    Schildhaus HU, Schultheis AM, Rüschoff J, Binot E, Merkelbach-Bruse S, Fassunke J, et al. MET amplification status in therapy-naïve adeno- and squamous cell carcinomas of the lung. Clin Cancer Res. 2015;21:90715. [DOI] [PubMed]
    Jenkins RW, Oxnard GR, Elkin S, Sullivan EK, Carter JL, Barbie DA. Response to crizotinib in a patient with lung adenocarcinoma harboring a MET splice site mutation. Clin Lung Cancer. 2015;16:e1014. [DOI] [PubMed] [PMC]
    Baltschukat S, Engstler BS, Huang A, Hao HX, Tam A, Wang HQ, et al. Capmatinib (INC280) is active against models of non-small cell lung cancer and other cancer types with defined mechanisms of MET activation. Clin Cancer Res. 2019;25:316475. [DOI] [PubMed]
    Cooper WA, Lam DC, O’Toole SA, Minna JD. Molecular biology of lung cancer. J Thorac Dis. 2013;5:S47990. [DOI] [PubMed] [PMC]
    Zukin M, Barrios CH, Pereira JR, Ribeiro Rde A, Beato CA, do Nascimento YN, et al. Randomized phase III trial of single-agent pemetrexed versus carboplatin and pemetrexed in patients with advanced non-small-cell lung cancer and Eastern Cooperative Oncology Group performance status of 2. J Clin Oncol. 2013;31:284953. [DOI] [PubMed]
    Arbour KC, Riely GJ. Systemic therapy for locally advanced and metastatic non-small cell lung cancer: a review. JAMA. 2019;322:76474. [DOI] [PubMed]
    Aruffo A, Stamenkovic I, Melnick M, Underhill CB, Seed B. CD44 is the principal cell surface receptor for hyaluronate. Cell. 1990;61:130313. [DOI] [PubMed]
    Nabil G, Alzhrani R, Alsaab HO, Atef M, Sau S, Iyer AK, et al. CD44 targeted nanomaterials for treatment of triple-negative breast cancer. Cancers (Basel). 2021;13:898. [DOI] [PubMed] [PMC]
    Mok TSK, Wu YL, Kudaba I, Kowalski DM, Cho BC, Turna HZ, et al.; KEYNOTE-042 Investigators. Pembrolizumab versus chemotherapy for previously untreated, PD-L1-expressing, locally advanced or metastatic non-small-cell lung cancer (KEYNOTE-042): a randomised, open-label, controlled, phase 3 trial. Lancet. 2019;393:181930. [DOI] [PubMed]
    Kong T, Ahn R, Yang K, Zhu X, Fu Z, Morin G, et al. CD44 promotes PD-L1 expression and its tumor-intrinsic function in breast and lung cancers. Cancer Res. 2020;80:44457. [DOI] [PubMed]
    Shi L, Chen S, Yang L, Li Y. The role of PD-1 and PD-L1 in T-cell immune suppression in patients with hematological malignancies. J Hematol Oncol. 2013;6:74. [DOI] [PubMed] [PMC]
    Hui E, Cheung J, Zhu J, Su X, Taylor MJ, Wallweber HA, et al. T cell costimulatory receptor CD28 is a primary target for PD-1-mediated inhibition. Science. 2017;355:142833. [DOI] [PubMed] [PMC]
    Mauri FA, Pinato DJ, Trivedi P, Sharma R, Shiner RJ. Isogeneic comparison of primary and metastatic lung cancer identifies CX3CR1 as a molecular determinant of site-specific metastatic diffusion. Oncol Rep. 2012;28:64753. [DOI] [PubMed]
    Grinberg-Rashi H, Ofek E, Perelman M, Skarda J, Yaron P, Hajdúch M, et al. The expression of three genes in primary non-small cell lung cancer is associated with metastatic spread to the brain. Clin Cancer Res. 2009;15:175561. [DOI] [PubMed]
    Whitsett TG, Inge LJ, Dhruv HD, Cheung PY, Weiss GJ, Bremner RM, et al. Molecular determinants of lung cancer metastasis to the central nervous system. Transl Lung Cancer Res. 2013;2:27383. [DOI] [PubMed] [PMC]
    Chang H, Sung JH, Moon SU, Kim HS, Kim JW, Lee JS. EGF induced RET inhibitor resistance in CCDC6-RET lung cancer cells. Yonsei Med J. 2017;58:918. [DOI] [PubMed] [PMC]
    Su T, Huang S, Zhang Y, Guo Y, Zhang S, Guan J, et al. miR-7/TGF-β2 axis sustains acidic tumor microenvironment-induced lung cancer metastasis. Acta Pharm Sin B. 2022;12:82137. [DOI] [PubMed] [PMC]
    Kumari R, Sunil D, Ningthoujam RS. Hypoxia-responsive nanoparticle based drug delivery systems in cancer therapy: an up-to-date review. J Control Release. 2020;319:13556. [DOI] [PubMed]
    Chen YL, Yang TY, Chen KC, Wu CL, Hsu SL, Hsueh CM. Hypoxia can impair doxorubicin resistance of non-small cell lung cancer cells by inhibiting MRP1 and P-gp expression and boosting the chemosensitizing effects of MRP1 and P-gp blockers. Cell Oncol (Dordr). 2016;39:41133. [DOI] [PubMed]
    Krzywinska E, Kantari-Mimoun C, Kerdiles Y, Sobecki M, Isagawa T, Gotthardt D, et al. Loss of HIF-1α in natural killer cells inhibits tumour growth by stimulating non-productive angiogenesis. Nat Commun. 2017;8:1597. [DOI] [PubMed] [PMC]
    Li S, Huang L, Sun Y, Bai Y, Yang F, Yu W, et al. Slit2 promotes angiogenic activity via the Robo1-VEGFR2-ERK1/2 pathway in both in vivo and in vitro studies. Invest Ophthalmol Vis Sci. 2015;56:52107. [DOI] [PubMed]
    Chen Y, Ma G, Su C, Wu P, Wang H, Song X, et al. Apatinib reverses alectinib resistance by targeting vascular endothelial growth factor receptor 2 and attenuating the oncogenic signaling pathway in echinoderm microtubule-associated protein-like 4-anaplastic lymphoma kinase fusion gene-positive lung cancer cell lines. Anticancer Drugs. 2018;29:93543. [DOI] [PubMed]
    Paz-Ares L, Hirsh V, Zhang L, de Marinis F, Yang JC, Wakelee HA, et al. Monotherapy administration of sorafenib in patients with non-small cell lung cancer (MISSION) trial: a phase III, multicenter, placebo-controlled trial of sorafenib in patients with relapsed or refractory predominantly nonsquamous non-small-cell lung cancer after 2 or 3 previous treatment regimens. J Thorac Oncol. 2015;10:174553. [DOI] [PubMed]
    Alexandrov LB, Nik-Zainal S, Wedge DC, Aparicio SA, Behjati S, Biankin AV, et al. Signatures of mutational processes in human cancer. Nature. 2013;500:41521. [DOI] [PubMed] [PMC]
    Kargl J, Busch SE, Yang GH, Kim KH, Hanke ML, Metz HE, et al. Neutrophils dominate the immune cell composition in non-small cell lung cancer. Nat Commun. 2017;8:14381. [DOI] [PubMed] [PMC]
    Brambilla E, Le Teuff G, Marguet S, Lantuejoul S, Dunant A, Graziano S, et al. Prognostic effect of tumor lymphocytic infiltration in resectable non-small-cell lung cancer. J Clin Oncol. 2016;34:122330. [DOI] [PubMed] [PMC]
    Bremnes RM, Busund LT, Kilvær TL, Andersen S, Richardsen E, Paulsen EE, et al. The role of tumor-infiltrating lymphocytes in development, progression, and prognosis of non-small cell lung cancer. J Thorac Oncol. 2016;11:789800. [DOI] [PubMed]
    Hodi FS, O’Day SJ, McDermott DF, Weber RW, Sosman JA, Haanen JB, et al. Improved survival with ipilimumab in patients with metastatic melanoma. N Engl J Med. 2010;363:71123. [DOI] [PubMed] [PMC]
    Parikh AR, Szabolcs A, Allen JN, Clark JW, Wo JY, Raabe M, et al. Radiation therapy enhances immunotherapy response in microsatellite stable colorectal and pancreatic adenocarcinoma in a phase II trial. Nat Cancer. 2021;2:112435. [DOI] [PubMed] [PMC]
    Hellmann MD, Ciuleanu TE, Pluzanski A, Lee JS, Otterson GA, Audigier-Valette C, et al. Nivolumab plus ipilimumab in lung cancer with a high tumor mutational burden. N Engl J Med. 2018;378:2093104. [DOI] [PubMed] [PMC]
    Wang SS, Liu W, Ly D, Xu H, Qu L, Zhang L. Tumor-infiltrating B cells: their role and application in anti-tumor immunity in lung cancer. Cell Mol Immunol. 2019;16:618. [DOI] [PubMed] [PMC]
    Spranger S, Koblish HK, Horton B, Scherle PA, Newton R, Gajewski TF. Mechanism of tumor rejection with doublets of CTLA-4, PD-1/PD-L1, or IDO blockade involves restored IL-2 production and proliferation of CD8+ T cells directly within the tumor microenvironment. J Immunother Cancer. 2014;2:3. [DOI] [PubMed] [PMC]
    Zugazagoitia J, Guedes C, Ponce S, Ferrer I, Molina-Pinelo S, Paz-Ares L. Current challenges in cancer treatment. Clin Ther. 2016;38:155166. [DOI] [PubMed]
    Minami K, Hiwatashi K, Ueno S, Sakoda M, Iino S, Okumura H, et al. Prognostic significance of CD68, CD163 and Folate receptor-β positive macrophages in hepatocellular carcinoma. Exp Ther Med. 2018;15:446576. [DOI] [PubMed] [PMC]
    Xu X, Jiang J, Yao L, Ji B. Silencing the FOLR2 gene inhibits cell proliferation and increases apoptosis in the NCI-H1650 non-small cell lung cancer cell line via inhibition of AKT/mammalian target of rapamycin (mTOR)/ribosomal protein S6 kinase 1 (S6K1) signaling. Med Sci Monit. 2018;24:806473. [DOI] [PubMed] [PMC]
    O’Shannessy DJ, Somers EB, Wang LC, Wang H, Hsu R. Expression of folate receptors alpha and beta in normal and cancerous gynecologic tissues: correlation of expression of the beta isoform with macrophage markers. J Ovarian Res. 2015;8:29. [DOI] [PubMed] [PMC]
    Tie Y, Zheng H, He Z, Yang J, Shao B, Liu L, et al. Targeting folate receptor β positive tumor-associated macrophages in lung cancer with a folate-modified liposomal complex. Signal Transduct Target Ther. 2020;5:6. [DOI] [PubMed] [PMC]
    Tian Y, Wu G, Xing JC, Tang J, Zhang Y, Huang ZM, et al. A novel splice variant of folate receptor 4 predominantly expressed in regulatory T cells. BMC Immunol. 2012;13:30. [DOI] [PubMed] [PMC]
    Shamsi A, Ahmed A, Khan MS, Husain FM, Amani S, Bano B. Investigating the interaction of anticancer drug temsirolimus with human transferrin: molecular docking and spectroscopic approach. J Mol Recognit. 2018;31:e2728. [DOI] [PubMed]
    Kaur T, Upadhyay J, Pukale S, Mathur A, Ansari MN. Investigation of trends in the research on transferrin receptor-mediated drug delivery via a bibliometric and thematic analysis. Pharmaceutics. 2022;14:2574. [DOI] [PubMed] [PMC]
    Lu Y, Zhu D, Gui L, Li Y, Wang W, Liu J, et al. A dual-targeting ruthenium nanodrug that inhibits primary tumor growth and lung metastasis via the PARP/ATM pathway. J Nanobiotechnology. 2021;19:115. [DOI] [PubMed] [PMC]
    Sugahara KN, Teesalu T, Karmali PP, Kotamraju VR, Agemy L, Girard OM, et al. Tissue-penetrating delivery of compounds and nanoparticles into tumors. Cancer Cell. 2009;16:51020. [DOI] [PubMed] [PMC]
    Su Y, Hu Y, Wang Y, Xu X, Yuan Y, Li Y, et al. A precision-guided MWNT mediated reawakening the sunk synergy in RAS for anti-angiogenesis lung cancer therapy. Biomaterials. 2017;139:7590. [DOI] [PubMed]
    Szlenk CT, Gc JB, Natesan S. Membrane-facilitated receptor access and binding mechanisms of long-acting β2-adrenergic receptor agonists. Mol Pharmacol. 2021;100:40627. [DOI] [PubMed] [PMC]
    Johnson M. The β-adrenoceptor. Am J Respir Crit Care Med. 1998;158:S14653. [DOI] [PubMed]
    Ferguson SS. Evolving concepts in G protein-coupled receptor endocytosis: the role in receptor desensitization and signaling. Pharmacol Rev. 2001;53:124. [PubMed]
    Elfinger M, Geiger J, Hasenpusch G, Uzgün S, Sieverling N, Aneja MK, et al. Targeting of the β2-adrenoceptor increases nonviral gene delivery to pulmonary epithelial cells in vitro and lungs in vivo. J Control Release. 2009;135:23441. [DOI] [PubMed]
    Iaboni M, Russo V, Fontanella R, Roscigno G, Fiore D, Donnarumma E, et al. Aptamer-miRNA-212 conjugate sensitizes NSCLC cells to TRAIL. Mol Ther Nucleic Acids. 2016;5:e289. [DOI] [PubMed] [PMC]
    Almasan A, Ashkenazi A. Apo2L/TRAIL: apoptosis signaling, biology, and potential for cancer therapy. Cytokine Growth Factor Rev. 2003;14:33748. [DOI] [PubMed]
    Kim I, Byeon HJ, Kim TH, Lee ES, Oh KT, Shin BS, et al. Doxorubicin-loaded porous PLGA microparticles with surface attached TRAIL for the inhalation treatment of metastatic lung cancer. Biomaterials. 2013;34:644453. [DOI] [PubMed]
    Snajdauf M, Havlova K, Vachtenheim J Jr, Ozaniak A, Lischke R, Bartunkova J, et al. The TRAIL in the treatment of human cancer: an update on clinical trials. Front Mol Biosci. 2021;8:628332. [DOI] [PubMed] [PMC]
    Sancho V, Di Florio A, Moody TW, Jensen RT. Bombesin receptor-mediated imaging and cytotoxicity: review and current status. Curr Drug Deliv. 2011;8:79134. [DOI] [PubMed] [PMC]
    Moody TW, Sun LC, Mantey SA, Pradhan T, Mackey LV, Gonzales N, et al. In vitro and in vivo antitumor effects of cytotoxic camptothecin-bombesin conjugates are mediated by specific interaction with cellular bombesin receptors. J Pharmacol Exp Ther. 2006;318:126572. [DOI] [PubMed]
    Du J, Li L. Which one performs better for targeted lung cancer combination therapy: pre- or post-bombesin-decorated nanostructured lipid carriers? Drug Deliv. 2016;23:1799809. [DOI] [PubMed]
    Pawar S, Vavia P. Glucosamine anchored cancer targeted nano-vesicular drug delivery system of doxorubicin. J Drug Target. 2016;24:6879. [DOI] [PubMed]
    Pawar SK, Vavia P. Efficacy interactions of PEG-DOX-N-acetyl glucosamine prodrug conjugate for anticancer therapy. Eur J Pharm Biopharm. 2015;97:45463. [DOI] [PubMed]
    Pawar SK, Badhwar AJ, Kharas F, Khandare JJ, Vavia PR. Design, synthesis and evaluation of N-acetyl glucosamine (NAG)-PEG-doxorubicin targeted conjugates for anticancer delivery. Int J Pharm. 2012;436:18393. [DOI] [PubMed]
    Fuster MM, Esko JD. The sweet and sour of cancer: glycans as novel therapeutic targets. Nat Rev Cancer. 2005;5:52642. [DOI] [PubMed]
    Yang C, Cao M, Liu H, He Y, Xu J, Du Y, et al. The high and low molecular weight forms of hyaluronan have distinct effects on CD44 clustering. J Biol Chem. 2012;287:43094107. [DOI] [PubMed] [PMC]
    Shahriari M, Taghdisi SM, Abnous K, Ramezani M, Alibolandi M. Self-targeted polymersomal co-formulation of doxorubicin, camptothecin and FOXM1 aptamer for efficient treatment of non-small cell lung cancer. J Control Release. 2021;335:36988. [DOI] [PubMed]
    Toole BP. Hyaluronan: from extracellular glue to pericellular cue. Nat Rev Cancer. 2004;4:52839. [DOI] [PubMed]
    Mattheolabakis G, Milane L, Singh A, Amiji MM. Hyaluronic acid targeting of CD44 for cancer therapy: from receptor biology to nanomedicine. J Drug Target. 2015;23:60518. [DOI] [PubMed]
    Hsiao KY, Wu YJ, Liu ZN, Chuang CW, Huang HH, Kuo SM. Anticancer effects of sinulariolide-conjugated hyaluronan nanoparticles on lung adenocarcinoma cells. Molecules. 2016;21:297. [DOI] [PubMed] [PMC]
    Bignotti E, Zanotti L, Calza S, Falchetti M, Lonardi S, Ravaggi A, et al. Trop-2 protein overexpression is an independent marker for predicting disease recurrence in endometrioid endometrial carcinoma. BMC Clin Pathol. 2012;12:22. [DOI] [PubMed] [PMC]
    Fu Y, Jing Y, Gao J, Li Z, Wang H, Cai M, et al. Variation of Trop2 on non-small-cell lung cancer and normal cell membranes revealed by super-resolution fluorescence imaging. Talanta. 2020;207:120312. [DOI] [PubMed]
    Lin JC, Wu YY, Wu JY, Lin TC, Wu CT, Chang YL, et al. TROP2 is epigenetically inactivated and modulates IGF-1R signalling in lung adenocarcinoma. EMBO Mol Med. 2012;4:47285. [DOI] [PubMed] [PMC]
    Cubas R, Li M, Chen C, Yao Q. Trop2: a possible therapeutic target for late stage epithelial carcinomas. Biochim Biophys Acta. 2009;1796:30914. [DOI] [PubMed]
    Guerra E, Trerotola M, Aloisi AL, Tripaldi R, Vacca G, La Sorda R, et al. The Trop-2 signalling network in cancer growth. Oncogene. 2013;32:1594600. [DOI] [PubMed]
    Heist RS, Guarino MJ, Masters G, Purcell WT, Starodub AN, Horn L, et al. Therapy of advanced non-small-cell lung cancer with an SN-38-anti-Trop-2 drug conjugate, sacituzumab govitecan. J Clin Oncol. 2017;35:27907. [DOI] [PubMed]
    Second clinical trial collaboration initiated to evaluate datopotamab deruxtecan in combination with KEYTRUDA® (pembrolizumab) in patients with metastatic non-small cell lung cancer [Internet]. Tokyo: Daiichi Sankyo; c2024 [cited 2021 Oct 25]. Available from: https://daiichisankyo.us/press-releases/-/article/second-clinical-trial-collaboration-initiated-to-evaluate-datopotamab-deruxtecan-in-combination-with-keytruda-pembrolizumab-in-patients-with-metastati
    Thakur C. Chapter 2 - an overview, current challenges of drug resistance, and targeting metastasis associated with lung cancer. In: Kesharwani P, editor. Nanotechnology-based targeted drug delivery systems for lung cancer. Amsterdam: Academic Press; 2019. pp. 21–38.
    Jiang W, Cai G, Hu PC, Wang Y. Personalized medicine in non-small cell lung cancer: a review from a pharmacogenomics perspective. Acta Pharm Sin B. 2018;8:5308. [DOI] [PubMed] [PMC]
    Jablonska PA, Bosch-Barrera J, Serrano D, Valiente M, Calvo A, Aristu J. Challenges and novel opportunities of radiation therapy for brain metastases in non-small cell lung cancer. Cancers (Basel). 2021;13:2141. [DOI] [PubMed] [PMC]
    Hussain S. Nanomedicine for treatment of lung cancer. In: Ahmad A, Gadgeel S, editors. Lung cancer and personalized medicine: novel therapies and clinical management. Cham: Springer; 2016. pp. 137–47.
    Sul J, Blumenthal GM, Jiang X, He K, Keegan P, Pazdur R. FDA approval summary: pembrolizumab for the treatment of patients with metastatic non-small cell lung cancer whose tumors express programmed death-ligand 1. Oncologist. 2016;21:64350. [DOI] [PubMed] [PMC]
    Norouzi M, Yathindranath V, Thliveris JA, Kopec BM, Siahaan TJ, Miller DW. Doxorubicin-loaded iron oxide nanoparticles for glioblastoma therapy: a combinational approach for enhanced delivery of nanoparticles. Sci Rep. 2020;10:11292. [DOI] [PubMed] [PMC]
    Wilson CM, Magnaudeix A, Naves T, Vincent F, Lalloue F, Jauberteau MO. The ins and outs of nanoparticle technology in neurodegenerative diseases and cancer. Curr Drug Metab. 2015;16:60932. [DOI] [PubMed]
    Tzogani K, Penttilä K, Lapveteläinen T, Hemmings R, Koenig J, Freire J, et al. EMA review of daunorubicin and cytarabine encapsulated in liposomes (Vyxeos, CPX-351) for the treatment of adults with newly diagnosed, therapy-related acute myeloid leukemia or acute myeloid leukemia with myelodysplasia-related changes. Oncologist. 2020;25:e141420. [DOI] [PubMed] [PMC]
    Belgiovine C, D’Incalci M, Allavena P, Frapolli R. Tumor-associated macrophages and anti-tumor therapies: complex links. Cell Mol Life Sci. 2016;73:241124. [DOI] [PubMed]
    Xue X, Qu H, Li Y. Stimuli-responsive crosslinked nanomedicine for cancer treatment. Exploration (Beijing). 2022;2:20210134. [DOI] [PubMed] [PMC]
    Sun Z, Hou Y. Intelligent micro/nanorobots for improved tumor therapy. BMEMat. 2023;1:e12012. [DOI]
    Yang G, Liu Y, Chen J, Ding J, Chen X. Self-adaptive nanomaterials for rational drug delivery in cancer therapy. Acc Mater Res. 2022;3:123247. [DOI]
    FDA approves abraxane for late-stage pancreatic cancer [Internet]. New York: ScienceBlog.com; c2024 [cited 2013 Sept 6]. Available from: https://scienceblog.com/66288/fda-approves-abraxane-for-late-stage-pancreatic-cancer/
    Raju GSR, Benton L, Pavitra E, Yu JS. Multifunctional nanoparticles: recent progress in cancer therapeutics. Chem Commun. 2015;51:1324859. [DOI] [PubMed]
    Gindy ME, Prud’homme RK. Multifunctional nanoparticles for imaging, delivery and targeting in cancer therapy. Expert Opin Drug Deliv. 2009;6:86578. [DOI] [PubMed]
    Çağdaş M, Sezer AD, Bucak S. Liposomes as potential drug carrier systems for drug delivery. In: Sezer AD, editor. Application of nanotechnology in drug delivery. London: IntechOpen; 2014.
    Woodman C, Vundu G, George A, Wilson CM. Applications and strategies in nanodiagnosis and nanotherapy in lung cancer. Semin Cancer Biol. 2021;69:34964. [DOI] [PubMed]
    Akbarzadeh A, Samiei M, Davaran S. Magnetic nanoparticles: preparation, physical properties, and applications in biomedicine. Nanoscale Res Lett. 2012;7:144. [DOI] [PubMed] [PMC]
    Zhang T, Chen Y, Ge Y, Hu Y, Li M, Jin Y. Inhalation treatment of primary lung cancer using liposomal curcumin dry powder inhalers. Acta Pharm Sin B. 2018;8:4408. [DOI] [PubMed] [PMC]
    Zhang W, Gong C, Chen Z, Li M, Li Y, Gao J. Tumor microenvironment-activated cancer cell membrane-liposome hybrid nanoparticle-mediated synergistic metabolic therapy and chemotherapy for non-small cell lung cancer. J Nanobiotechnology. 2021;19:339. [DOI] [PubMed] [PMC]
    Hu M, Zhang J, Kong L, Yu Y, Hu Q, Yang T, et al. Immunogenic hybrid nanovesicles of liposomes and tumor-derived nanovesicles for cancer immunochemotherapy. ACS Nano. 2021;15:312338. [DOI] [PubMed]
    Shanker M, Willcutts D, Roth JA, Ramesh R. Drug resistance in lung cancer. Lung Cancer: Targets Ther. 2010;1:2336. [PubMed] [PMC]
    Scagliotti GV, Novello S, Selvaggi G. Multidrug resistance in non-small-cell lung cancer. Ann Oncol. 1999;10:S836. [DOI] [PubMed]
    Kunjachan S, Rychlik B, Storm G, Kiessling F, Lammers T. Multidrug resistance: physiological principles and nanomedical solutions. Adv Drug Deliv Rev. 2013;65:185265. [DOI] [PubMed] [PMC]
    Mamot C, Drummond DC, Hong K, Kirpotin DB, Park JW. Liposome-based approaches to overcome anticancer drug resistance. Drug Resist Updat. 2003;6:2719. [DOI] [PubMed]
    Qu MH, Zeng RF, Fang S, Dai QS, Li HP, Long JT. Liposome-based co-delivery of siRNA and docetaxel for the synergistic treatment of lung cancer. Int J Pharm. 2014;474:11222. [DOI] [PubMed]
    Saad M, Garbuzenko OB, Minko T. Co-delivery of siRNA and an anticancer drug for treatment of multidrug-resistant cancer. Nanomedicine (Lond). 2008;3:76176. [DOI] [PubMed] [PMC]
    Li F, Mei H, Gao Y, Xie X, Nie H, Li T, et al. Co-delivery of oxygen and erlotinib by aptamer-modified liposomal complexes to reverse hypoxia-induced drug resistance in lung cancer. Biomaterials. 2017;145:5671. [DOI] [PubMed]
    Xu W, Lee M. Development and evaluation of lipid nanoparticles for paclitaxel delivery: a comparison between solid lipid nanoparticles and nanostructured lipid carriers. J Pharm Invest. 2015;45:67580. [DOI]
    Freag MS, Elnaggar YS, Abdelmonsif DA, Abdallah OY. Stealth, biocompatible monoolein-based lyotropic liquid crystalline nanoparticles for enhanced aloe-emodin delivery to breast cancer cells: in vitro and in vivo studies. Int J Nanomedicine. 2016;11:4799818. [DOI] [PubMed] [PMC]
    Weber S, Zimmer A, Pardeike J. Solid Lipid Nanoparticles (SLN) and Nanostructured Lipid Carriers (NLC) for pulmonary application: a review of the state of the art. Eur J Pharm Biopharm. 2014;86:722. [DOI] [PubMed]
    Patil TS, Deshpande AS, Deshpande S. Critical review on the analytical methods for the estimation of clofazimine in bulk, biological fluids and pharmaceutical formulations. Crit Rev Anal Chem. 2018;48:492502. [DOI] [PubMed]
    Patil TS, Deshpande AS. Nanostructured lipid carriers-based drug delivery for treating various lung diseases: A State-of-the-Art Review. Int J Pharm. 2018;547:20925. [DOI] [PubMed]
    Patil TS, Deshpande AS, Deshpande S, Shende P. Targeting pulmonary tuberculosis using nanocarrier-based dry powder inhalation: current status and futuristic need. J Drug Target. 2019;27:1227. [DOI] [PubMed]
    Wen M, Xia J, Sun Y, Wang X, Fu X, Zhang Y, et al. Combination of EGFR-TKIs with chemotherapy versus chemotherapy or EGFR-TKIs alone in advanced NSCLC patients with EGFR mutation. Biologics. 2018;12:18390. [DOI] [PubMed] [PMC]
    Garbuzenko OB, Kuzmov A, Taratula O, Pine SR, Minko T. Strategy to enhance lung cancer treatment by five essential elements: inhalation delivery, nanotechnology, tumor-receptor targeting, chemo- and gene therapy. Theranostics. 2019;9:836276. [DOI] [PubMed] [PMC]
    Yang Y, Huang Z, Li J, Mo Z, Huang Y, Ma C, et al. PLGA porous microspheres dry powders for codelivery of afatinib-loaded solid lipid nanoparticles and paclitaxel: novel therapy for EGFR tyrosine kinase inhibitors resistant nonsmall cell lung cancer. Adv Healthc Mater. 2019;8:e1900965. [DOI] [PubMed]
    Soni N, Soni N, Pandey H, Maheshwari R, Kesharwani P, Tekade RK. Augmented delivery of gemcitabine in lung cancer cells exploring mannose anchored solid lipid nanoparticles. J Colloid Interface Sci. 2016;481:10716. [DOI] [PubMed]
    Mattoussi H, Rotello VM. Inorganic nanoparticles in drug delivery. Adv Drug Deliv Rev. 2013;65:6056. [DOI] [PubMed]
    Ju Y, Cortez-Jugo C, Chen J, Wang TY, Mitchell AJ, Tsantikos E, et al. Engineering of nebulized metal-phenolic capsules for controlled pulmonary deposition. Adv Sci (Weinh). 2020;7:1902650. [DOI] [PubMed] [PMC]
    Saadat M, Manshadi MKD, Mohammadi M, Zare MJ, Zarei M, Kamali R, et al. Magnetic particle targeting for diagnosis and therapy of lung cancers. J Control Release. 2020;328:77691. [DOI] [PubMed] [PMC]
    Bao G, Mitragotri S, Tong S. Multifunctional nanoparticles for drug delivery and molecular imaging. Annu Rev Biomed Eng. 2013;15:25382. [DOI] [PubMed] [PMC]
    Orel V, Shevchenko A, Romanov A, Tselepi M, Mitrelias T, Barnes CH, et al. Magnetic properties and antitumor effect of nanocomplexes of iron oxide and doxorubicin. Nanomedicine. 2015;11:4755. [DOI] [PubMed]
    Mahmoudi M, Sahraian MA, Shokrgozar MA, Laurent S. Superparamagnetic iron oxide nanoparticles: promises for diagnosis and treatment of multiple sclerosis. ACS Chem Neurosci. 2011;2:11840. [DOI] [PubMed] [PMC]
    Gaihre B, Khil MS, Kim HY. In vitro anticancer activity of doxorubicin-loaded gelatin-coated magnetic iron oxide nanoparticles. J Microencapsul. 2011;28:28693. [DOI] [PubMed]
    Tseng SJ, Huang KY, Kempson IM, Kao SH, Liu MC, Yang SC, et al. Remote control of light-triggered virotherapy. ACS Nano. 2016;10:1033946. [DOI] [PubMed]
    Sadhukha T, Wiedmann TS, Panyam J. Inhalable magnetic nanoparticles for targeted hyperthermia in lung cancer therapy. Biomaterials. 2013;34:516371. [DOI] [PubMed] [PMC]
    Huang H, Yuan G, Xu Y, Gao Y, Mao Q, Zhang Y, et al. Photoacoustic and magnetic resonance imaging-based gene and photothermal therapy using mesoporous nanoagents. Bioact Mater. 2022;9:15767. [DOI] [PubMed] [PMC]
    Ma Y, Chen L, Li X, Hu A, Wang H, Zhou H, et al. Rationally integrating peptide-induced targeting and multimodal therapies in a dual-shell theranostic platform for orthotopic metastatic spinal tumors. Biomaterials. 2021;275:120917. [DOI] [PubMed]
    Li N, Liu S, Sun M, Chen W, Xu X, Zeng Z, et al. Chimeric antigen receptor-modified T cells redirected to EphA2 for the immunotherapy of non-small cell lung cancer. Transl Oncol. 2018;11:117. [DOI] [PubMed] [PMC]
    Cao W, He L, Cao W, Huang X, Jia K, Dai J. Recent progress of graphene oxide as a potential vaccine carrier and adjuvant. Acta Biomater. 2020;112:1428. [DOI] [PubMed]
    Liu W, Dong A, Wang B, Zhang H. Current advances in black phosphorus-based drug delivery systems for cancer therapy. Adv Sci (Weinh). 2021;8:2003033. [DOI] [PubMed] [PMC]
    Stanisavljevic M, Krizkova S, Vaculovicova M, Kizek R, Adam V. Quantum dots-fluorescence resonance energy transfer-based nanosensors and their application. Biosens Bioelectron. 2015;74:56274. [DOI] [PubMed]
    Hamano N, Murata M, Kawano T, Piao JS, Narahara S, Nakata R, et al. Förster resonance energy transfer-based self-assembled nanoprobe for rapid and sensitive detection of postoperative pancreatic fistula. ACS Appl Mater Interfaces. 2016;8:511423. [DOI] [PubMed]
    Wu D, Li BL, Zhao Q, Liu Q, Wang D, He B, et al. Assembling defined DNA nanostructure with nitrogen-enriched carbon dots for theranostic cancer applications. Small. 2020;16:e1906975. [DOI] [PubMed]
    Saravanakumar G, Kim J, Kim WJ. Reactive-oxygen-species-responsive drug delivery systems: promises and challenges. Adv Sci (Weinh). 2016;4:1600124. [DOI] [PubMed] [PMC]
    Zhang Q, Liu F, Nguyen KT, Ma X, Wang X, Xing B, et al. Multifunctional mesoporous silica nanoparticles for cancer‐targeted and controlled drug delivery. Adv Funct Mater. 2012;22:514456. [DOI]
    Tang F, Li L, Chen D. Mesoporous silica nanoparticles: synthesis, biocompatibility and drug delivery. Adv Mater. 2012;24:150434. [DOI] [PubMed]
    Zhou X, He X, Shi K, Yuan L, Yang Y, Liu Q, et al. Injectable thermosensitive hydrogel containing erlotinib-loaded hollow mesoporous silica nanoparticles as a localized drug delivery system for NSCLC therapy. Adv Sci (Weinh). 2020;7:2001442. [DOI] [PubMed] [PMC]
    Cheng W, Liang C, Xu L, Liu G, Gao N, Tao W, et al. TPGS-functionalized polydopamine-modified mesoporous silica as drug nanocarriers for enhanced lung cancer chemotherapy against multidrug resistance. Small. 2017;13:1700623. [DOI] [PubMed]
    Amreddy N, Babu A, Muralidharan R, Munshi A, Ramesh R. Polymeric nanoparticle-mediated gene delivery for lung cancer treatment. Top Curr Chem (Cham). 2017;375:35. [DOI] [PubMed] [PMC]
    Kim J, Wilson DR, Zamboni CG, Green JJ. Targeted polymeric nanoparticles for cancer gene therapy. J Drug Target. 2015;23:62741. [DOI] [PubMed] [PMC]
    Zhou L, Wang H, Li Y. Stimuli-responsive nanomedicines for overcoming cancer multidrug resistance. Theranostics. 2018;8:105974. [DOI] [PubMed] [PMC]
    Paus C, van der Voort R, Cambi A. Nanomedicine in cancer therapy: promises and hurdles of polymeric nanoparticles. Explor Med. 2021;2:16785. [DOI]
    Hensing TA, Karrison T, Garmey EG, Hennessy MG, Salgia R. Randomized phase II study of IV topotecan versus CRLX101 in the second-line treatment of recurrent extensive-stage small cell lung cancer (ES-SCLC). J Clin Oncol. 2013;31:TPS7610. [DOI]
    Tseng S, Kempson IM, Liao Z, Ho Y, Yang P. An acid degradable, lactate oxidizing nanoparticle formulation for non-small cell lung cancer virotherapy. Nano Today. 2022;46:101582. [DOI]
    Wang S, Yu G, Wang Z, Jacobson O, Tian R, Lin LS, et al. Hierarchical tumor microenvironment-responsive nanomedicine for programmed delivery of chemotherapeutics. Adv Mater. 2018;30:e1803926. [DOI] [PubMed] [PMC]
    Liao ZX, Kempson IM, Hsieh CC, Tseng SJ, Yang PC. Potential therapeutics using tumor-secreted lactate in nonsmall cell lung cancer. Drug Discov Today. 2021;26:250814. [DOI] [PubMed]
    Zhong G, Yang C, Liu S, Zheng Y, Lou W, Teo JY, et al. Polymers with distinctive anticancer mechanism that kills MDR cancer cells and inhibits tumor metastasis. Biomaterials. 2019;199:7687. [DOI] [PubMed]
    Iyer R, Nguyen T, Padanilam D, Xu C, Saha D, Nguyen KT, et al. Glutathione-responsive biodegradable polyurethane nanoparticles for lung cancer treatment. J Control Release. 2020;321:36371. [DOI] [PubMed] [PMC]
    Wang X, Chen H, Zeng X, Guo W, Jin Y, Wang S, et al. Efficient lung cancer-targeted drug delivery via a nanoparticle/MSC system. Acta Pharm Sin B. 2019;9:16776. [DOI] [PubMed] [PMC]
    Kidd S, Spaeth E, Dembinski JL, Dietrich M, Watson K, Klopp A, et al. Direct evidence of mesenchymal stem cell tropism for tumor and wounding microenvironments using in vivo bioluminescent imaging. Stem Cells. 2009;27:261423. [DOI] [PubMed] [PMC]
    Sonabend AM, Ulasov IV, Tyler MA, Rivera AA, Mathis JM, Lesniak MS. Mesenchymal stem cells effectively deliver an oncolytic adenovirus to intracranial glioma. Stem Cells. 2008;26:83141. [DOI] [PubMed]
    Chen Y, Chen C, Zhang X, He C, Zhao P, Li M, et al. Platinum complexes of curcumin delivered by dual-responsive polymeric nanoparticles improve chemotherapeutic efficacy based on the enhanced anti-metastasis activity and reduce side effects. Acta Pharm Sin B. 2020;10:110621. [DOI] [PubMed] [PMC]
    Wang L, Yu Y, Wei D, Zhang L, Zhang X, Zhang G, et al. A systematic strategy of combinational blow for overcoming cascade drug resistance via NIR-light-triggered hyperthermia. Adv Mater. 2021;33:e2100599. [DOI] [PubMed]
    Zhao M, Wang R, Yang K, Jiang Y, Peng Y, Li Y, et al. Nucleic acid nanoassembly-enhanced RNA therapeutics and diagnosis. Acta Pharm Sin B. 2023;13:91641. [DOI] [PubMed] [PMC]
    Kuerban K, Gao X, Zhang H, Liu J, Dong M, Wu L, et al. Doxorubicin-loaded bacterial outer-membrane vesicles exert enhanced anti-tumor efficacy in non-small-cell lung cancer. Acta Pharm Sin B. 2020;10:153448. [DOI] [PubMed] [PMC]
    Li XQ, Liu JT, Fan LL, Liu Y, Cheng L, Wang F, et al. Exosomes derived from gefitinib-treated EGFR-mutant lung cancer cells alter cisplatin sensitivity via up-regulating autophagy. Oncotarget. 2016;7:2458595. [DOI] [PubMed] [PMC]
    Zhong YF, Cheng J, Liu Y, Luo T, Wang Y, Jiang K, et al. DNA nanostructures as Pt(IV) prodrug delivery systems to combat chemoresistance. Small. 2020;16:e2003646. [DOI] [PubMed]
    Ye B, Zhao B, Wang K, Guo Y, Lu Q, Zheng L, et al. Neutrophils mediated multistage nanoparticle delivery for prompting tumor photothermal therapy. J Nanobiotechnology. 2020;18:138. [DOI] [PubMed] [PMC]
    Chen J, Yang J, Ding J. Rational construction of polycystine-based nanoparticles for biomedical applications. J Mater Chem B. 2022;10:717382. [DOI] [PubMed]
    Delk SC, Chattopadhyay A, Escola-Gil JC, Fogelman AM, Reddy ST. Apolipoprotein mimetics in cancer. Semin Cancer Biol. 2021;73:15868. [DOI] [PubMed] [PMC]
    Mottaghitalab F, Kiani M, Farokhi M, Kundu SC, Reis RL, Gholami M, et al. Targeted delivery system based on gemcitabine-loaded silk fibroin nanoparticles for lung cancer therapy. ACS Appl Mater Interfaces. 2017;9:3160011. [DOI] [PubMed]
    Elgohary MM, Helmy MW, Abdelfattah EA, Ragab DM, Mortada SM, Fang JY, et al. Targeting sialic acid residues on lung cancer cells by inhalable boronic acid-decorated albumin nanocomposites for combined chemo/herbal therapy. J Control Release. 2018;285:23043. [DOI] [PubMed]
    Guo M, Wu F, Hu G, Chen L, Xu J, Xu P, et al. Autologous tumor cell-derived microparticle-based targeted chemotherapy in lung cancer patients with malignant pleural effusion. Sci Transl Med. 2019;11:eaat5690. [DOI] [PubMed]
    Ghosh S, Javia A, Shetty S, Bardoliwala D, Maiti K, Banerjee S, et al. Triple negative breast cancer and non-small cell lung cancer: clinical challenges and nano-formulation approaches. J Control Release. 2021;337:2758. [DOI] [PubMed]
    Wilhelm S, Tavares A, Dai Q, Ohta S, Audet J, Dvorak HF, et al. Analysis of nanoparticle delivery to tumours. Nat Rev Mater. 2016;1:16014. [DOI]
    Danhier F. To exploit the tumor microenvironment: Since the EPR effect fails in the clinic, what is the future of nanomedicine? J Control Release. 2016;244:10821. [DOI] [PubMed]
    Pirollo KF, Nemunaitis J, Leung PK, Nunan R, Adams J, Chang EH. Safety and efficacy in advanced solid tumors of a targeted nanocomplex carrying the p53 gene used in combination with docetaxel: a phase 1b study. Mol Ther. 2016;24:1697706. [DOI] [PubMed] [PMC]
    Peng XH, Wang Y, Huang D, Wang Y, Shin HJ, Chen Z, et al. Targeted delivery of cisplatin to lung cancer using ScFvEGFR-heparin-cisplatin nanoparticles. ACS Nano. 2011;5:948093. [DOI] [PubMed] [PMC]
    Ganesh S, Iyer AK, Gattacceca F, Morrissey DV, Amiji MM. In vivo biodistribution of siRNA and cisplatin administered using CD44-targeted hyaluronic acid nanoparticles. J Control Release. 2013;172:699706. [DOI] [PubMed] [PMC]
    Chen Q, Wang X, Chen F, Zhang Q, Dong B, Yang H, et al. Functionalization of upconverted luminescent NaYF4:Yb/Er nanocrystals by folic acid-chitosan conjugates for targeted lung cancer cell imaging. J Mater Chem. 2011;21:76617. [DOI]
    Singh RP, Sharma G, Sonali, Agrawal P, Pandey BL, Koch B, et al. Transferrin receptor targeted PLA-TPGS micelles improved efficacy and safety in docetaxel delivery. Int J Biol Macromol. 2016;83:33544. [DOI] [PubMed]
    Shi J, Kantoff PW, Wooster R, Farokhzad OC. Cancer nanomedicine: progress, challenges and opportunities. Nat Rev Cancer. 2017;17:2037. [DOI] [PubMed] [PMC]