Abstract

Carcinoma of the lungs is among the most menacing forms of malignancy and has a poor prognosis, with a low overall survival rate due to delayed detection and ineffectiveness of conventional therapy. Therefore, drug delivery strategies that may overcome undesired damage to healthy cells, boost therapeutic efficacy, and act as imaging tools are currently gaining much attention. Advances in material science have resulted in unique nanoscale-based theranostic agents, which provide renewed hope for patients suffering from lung cancer. Nanotechnology has vastly modified and upgraded the existing techniques, focusing primarily on increasing bioavailability and stability of anti-cancer drugs. Nanocarrier-based imaging systems as theranostic tools in the treatment of lung carcinoma have proven to possess considerable benefits, such as early detection and targeted therapeutic delivery for effectively treating lung cancer. Several variants of nano-drug delivery agents have been successfully studied for therapeutic applications, such as liposomes, dendrimers, polymeric nanoparticles, nanoemulsions, carbon nanotubes, gold nanoparticles, magnetic nanoparticles, solid lipid nanoparticles, hydrogels, and micelles. In this Review, we present a comprehensive outline on the various types of overexpressed receptors in lung cancer, as well as the various targeting approaches of nanoparticles.
1. Introduction
Cancer has become one of the paramount and most serious health concerns globally, affecting people in both developing and developed countries.1,2 Lung cancer is one of the deadliest cancers, accounting for a high mortality rate, and is the most frequent cause of tumor-related fatalities. It is essentially categorized into primary and secondary cancers. Based on the histological assessment, lung carcinoma is subdivided into non-small-cell lung carcinoma (NSCLC), small-cell lung carcinoma (SCLC), mesothelioma, sarcoma, and carcinoid. SCLC and NSCLC are the most commonly occurring, i.e., they represent around 90% of all forms of lung cancers, while the incidence of other types has been rare.3,4 NSCLC is further subclassified into four types: adenocarcinoma, squamous cell carcinoma, large-cell undifferentiated carcinoma, and Pancoast tumor.5,6 SCLC is an aggressively growing tumor and is categorized into small-cell and ''oat cell'' cancer combined.7 Smoking habits, hereditary factors, urbanization, and environmental factors (like exposure to arsenic, toxins, and asbestos) are the primary causative factors contributing to the pathogenesis of lung cancer.8−12 As per the statistics released by the World Health Organization (WHO), in 2020 alone there were around 2.21 million diagnosed cases of lung cancers, along with 180 million deaths.13
Conventional strategies such as chemotherapy, radiotherapy, and surgical resection, either separately or in combination, have been the mainstay of the current treatment regimens.14,15 However, severe adverse reactions are well reported with such traditional treatment methods. Moreover, the damage caused by radiotherapy to the surrounding healthy cells is evident, and thus it should not be the preferred treatment in patients who already have severely compromised pulmonary systems, as this may lead to loss in the functionality of the lungs.16,17 Radiotherapy has also been used in combination with chemotherapy and/or surgery lately.18 The main principle on which chemotherapy works is by hindering the synthesis of DNA (deoxyribonucleic acid) and mitosis, which eventually leads to the death of rapidly spreading cancerous cells. As the chemotherapeutic agents are non-selective, they tend to impart undesirable adverse side effects to the surrounding normal cells, which is one of the major reasons behind the high mortality rate in tumor patients. The drug regimen is, moreover, administered in specific cycles with regular breaks in between, which is essential for the recovery of non-selectively damaged cells. But these breaks may eventually restore even the carcinoma cells which were supposed to be weakened or killed by the therapy. Furthermore, this could also lead to myelosuppression, which requires extending the duration of the break and/or even discontinuation of the agents.19 In addition, non-targeted cytotoxicity and the prevalence of highly structured physical, physiological, and enzymatic barriers make targeted drug delivery extremely difficult, which restricts drug distribution to the target location.20 Moreover, the advent of multi-drug resistance (MDR) substantially reduces treatment effectiveness for the majority of malignancies.21 As a result, nanotechnology has emerged as a widely recognized and promising approach in the battle against cancer.
Nanoparticles (NPs) are colloidal particles with a diameter of less than 100 nm, which are combined with a therapeutic agent that is usually encapsulated inside the particle matrix, which is either adsorbed or conjugated on the surface by structural changes to enhance therapeutic stability and selective targeting. Nanoparticles are effective theranostic agents due to their dimensional resemblance to biomolecules, high surface-to-volume ratio, and potential for surface modification.22 Theranostic agents are tools that include cancer cell imaging and treatment by incorporating numerous methods that allow for a comprehensive diagnosis, molecular imaging, and a specific treatment regimen all at the same time. Nanoparticles possess a sub-micron size, which allows for deep tissue infiltration, penetration through epithelial fenestrations, and usually effective uptake by targeted cells, which elevates the bioavailability of therapeutic components.23,24 By fabricating the particle polymer properties, it is possible to maximize the amount and rate of drug release.
Nanotheranostics provide significant advantages over conventional therapeutic, diagnostic, and pharmaceutical approaches (Table 1). As there is a probability of surface modification25 and a tendency for both passive and active drug targeting,26 nanoparticles are intriguing agents which can be used in extremely precise in vitro and in vivo imaging and as targeted treatment sensors.27,28 The nanotechnology research in this field encompasses a wide variety of component systems that modulate the speed and direct the drug to be released at a specific site.11 Quantum dots, liposomes, polymeric nanoparticles, inorganic nanoparticles, magnetic nanoparticles, dendrimers, and gold nanoparticles are among the most investigated and promising systems.
Table 1. Various Nanoparticles with Their Theranostic Properties.

Different synthetic and natural compounds have been synthesized and isolated and evaluated against lung cancer cell lines, but still their mechanism of action needs to be explored.29−33 To overcome the challenges associated with conventional chemotherapy, the development of novel drug delivery systems which can either passively or actively target carcinoma cells is urgently required.34 Given the larger surface area of the alveolar region, drug delivery to the lungs has provided numerous promising opportunities for enhancing treatment efficacy. Lack of mechanistic insights into the underlying cellular mechanisms behind tumoral heterogeneity has drawn considerable attention to develop the relevant treatment approaches and therapeutic strategies. Because therapeutic responses are generally influenced by the emergence of resistant subpopulations and variations in molecular phenotypes, hence understanding tumor heterogeneity is therapeutically beneficial.35 For the past two decades, the rise of nano-based delivery systems has created immense opportunities in clinical therapeutics and enhanced bioavailability by demonstrating desirable properties such as the prolonged retention of therapeutics in the lungs, consequently reducing the amount of dose administered and improving patient compliance.36 This comprehensive review also discusses the applications of novel drug delivery vehicles that employ nanomolecules in the treatment and diagnosis of lung cancer.
2. Overexpressed Receptors
2.1. Epidermal Growth Factor Receptor (EGFR)
Overexpression of transmembrane protein EGFR plays a crucial role in the advancement of carcinoma. It consists of an intracellular fragment of tyrosine kinase activity and an extracellular section that includes the ligand-binding area responsible for the regulation of tumor growth (invasion, angiogenesis, metastasis, and proliferation of cells).37 Conformational modifications arise when ligands bind to the attaching area. The binding of ligands and activity of tyrosine kinase further leads to autophosphorylation that triggers alterations in the signaling pathways.38,39 These tyrosine kinase inhibitors are more effective for altered EGFR in cancer.40 Mutations arising due to the overactivation of EGFR alter the signal transduction pathways. Attachment of ligands present on the outer side of the cell further leads to cell multiplication and anti-apoptosis.41,42
2.2. Growth Hormone Receptor (GHR)
GHR is a single-pass transmembrane receptor with at least one cytokine receptor homology domain (CRHD) in its extracellular portion. The CRHD has two fibronectin III (FNIII)-like folds. GHR overexpression has been identified in A549 NSCLC and other related cancers.43,44
2.3. Folate Receptors
The folate receptors demonstrate a greater binding affinity for folic acid.45 It comprises four different forms: Folate Receptor Alpha (FRA), Folate Receptor Beta (FRB), Folate Receptor Gamma (FRG), and Folate Receptor Delta (FRD). FRA is a glycosylphosphatidylinositol-anchored glycoprotein present on the cell surface which is referred as FOLR-1 or folate binding protein (FBP) and assists in the delivery of 5-methyltetrahydrofolate (5-MTF) (an active form of folate).46−48 Overexpression of the FRA has been observed in lung-like solid tumor types and direct or indirect pathway of absorption of folate proves to be advantageous for the growth of cancer cells.49−51 Several studies have reported on the elevated levels of overexpressed FRA in NSCLC.52,53 The U.S. FDA (United States Food and Drug Administration) has approved various techniques that target FRA.54
2.4. Vascular Endothelial Growth Factor Receptors (VEGFR)
The VEGFRs were first recognized as a vascular permeability factor that trigger vascular leakage when exuded by tumor cells.55,56 FLT-1 (VEGFR1), FLK-1 (VEGFR2), and FLT-4 (VEGFR3) are three major types of tyrosine kinase receptors that exclusively bind to mammalian VEGFR. These receptors feature seven extracellular immunoglobulin-like domains which get stimulated by ligand-mediated dimerization.57 These receptors are frequently overexpressed in NSCLC and possess an acknowledged significance in angiogenesis, proliferation, and metastasis of tumor cells.58,59
2.5. Luteinizing Hormone-Releasing Hormone Receptor (LHRHR)
Overexpressed LHRHR found in the cytoplasmic membrane of lung cancer cells could be utilized for targeting the cancer cells without causing any harm to healthy natural cells.60 Several studies have shown successful targeted screening and delivery of antitumor drugs to the cancerous cells by employing LHRH peptides, by boosting their concentration in the tumor microenvironment and reducing undesirable interactions with the surrounding healthy cells.61
2.6. Fibroblast Growth Factor Receptor (FGFR)
The extracellular component of FGFR is made up of immunoglobulin resembling domains that possess greater affinity for FGF ligand, consist of a transmembrane region, and possess tyrosine kinase activity.62 This family of receptors is classified into four groups FGFR1 to FGFR4. Unusual FGF/FGFR signaling, alterations or amplification of genes (oncogenic) may contribute to tumorigenesis and therapy resistance in various cancers such as malignant melanoma and solid tumors.63,64 Elevated expression of FGFR1 is reported in NSCLC with varying ratios in squamous cell carcinomas and lung adenocarcinomas.65,66
2.7. Cluster of Differentiation 44 (CD44)
CD44 is a membrane glycoprotein receptor that plays an important role by binding to hyaluronic acid particularly to help normal and cancer stem cells attach, differentiate, perform homing, and migrate.67 CD44 overexpression was found in pneumocytes (type II) and squamous metaplasia, in lung cancer.68 The cluster of differentiation 44v6 has been linked to lymph node metastases in NSCLC patients.69−71
2.8. Integrins
Integrins belong to transmembrane heterodimeric glycoprotein family that consists of non-covalently linked alpha and beta subunits.72 They are dependent on the orientation patterns exhibited in the middle of 18 α and 8 β subunits, and there are 24 distinct integrin receptors.73 Integrins are detected in almost 82% of all NSCLC patients, regardless of differentiation type or degree. On the other hand, just 13% of SCLCs demonstrated an overexpressed α3β1.74 Downregulated levels of integrin expression have been related to SCLC, as well as the disease’s high severity and potential to spread.75
2.9. Mer or Axl as Receptor Tyrosine Kinase (RTK)
RTKs are members of the group of transmembrane proteins, with a large amount of variation in the extracellular area and a common intracellular tyrosine kinase domain. Axl is a key member of the RTK family TYRO3-AXL-MER (TAM). AXL overexpression was observed in NSCLC ADC (antibody–drug conjugates).76 The TAM family’s second most significant member, MER, was identified from a chicken retrovirus RLP30. Mer overexpression was reported in T-cell acute lung carcinoma.77,78 Overexpression of these receptors has been linked to chemotherapeutic treatment resistance, tumor growth, and advancement in the cancer metastatic stage.79
2.10. Interleukin-22 Receptor
The cytokine interleukin-22 (IL-22) is a member of the IL-10 cytokine family. Expression is shown by macrophages, alveolar cells, and T cells in the lungs and operates entirely on IL-22-R1 (IL-22-receptor 1).80 Zhang with his team (2008)349 found IL-22 and IL-22R1 upregulation in cancerous tissue and serum of patients, and since then it has been shown that it protects cancerous cell lines from serum deprivation led cell death triggered by therapeutics.
2.11. Adenosine Receptors (AR)
A1, A2A, A2B, and A3 are the members of G-protein-coupled adenosine receptors. Of these, it is reported that A3 has been highly expressed in the lung tissues,81 although the overexpression varies when compared to that of other ARs that are elevated at various degrees in the lung.82 The Cancer Genome Anatomy Project (CGAP) reported that the expression of A3AR was higher in malignant tissues as compared to normal tissues of the same patient.83
2.12. Chemotactic Cytokine CXCR4 Receptor
Chemotactic cytokine receptors have 7 transmembrane-spanning helixes and belong to the class of G-protein-coupled receptors (GPCR). They bind to a chemotactic cytokine ligand and participate in the migration of a number of cell types.84 CXC and CC are the two primary types of chemokines. The first two cysteine residues in CXC are separated by an amino acid, whereas it is adjacently placed in chemotactic cytokines.85 CXCR4 is a chemotactic cytokine receptor that is solely activated through CXCL12.86 This pair is responsible for the spread of NSCLC cancer. According to Saintigny and Burger (2012), elevation of CXCR4/CXCL12 linkage was related to a later stage of illness and a worsened diagnosis.87 The receptor linkage was reported in approximately 80% of non-small-cell lung carcinoma tissue sections collected from cancer patients in stages IA to IIB.88
2.13. Bombesin Receptor
The GPCR family of human bombesin receptors has three subcategories: the BB1 receptor, which links neuromedin (NMB) with strong potential, the BB2 receptor, which aids in binding GRP (gastrin-releasing peptide), and the BRS-3 (bombesin receptor subtype-3).89 Due to its strong structural similarity with the other two bombesin receptors, BRS-3 is categorized as a member of the bombesin receptor family. BB1 and BB2 receptors have been found to be overexpressed in most types of lung carcinoma, and they essentially contribute in enhancing the growth of malignant cells, which frequently function by autocrine mechanisms.90 When BRS-3 gets activated, it leads to the elevation of intracellular calcium and phospholipase C, further leading to activating PKC (protein kinase C) and MAPK (mitogen-activated protein kinase pathway).91 Hence, BRS-3 consequently phosphorylates ERK (extracellular signal-regulated kinase) tyrosine and expresses Elk-1 (ETS like-1 protein) and c-Fos, which is the reason behind excessive growth of lung tumor cells.89
2.14. PETA-3/CD151
Tetraspanins (Tspan) consist of the following transmembrane domain proteins: (i) N-cytoplasmic domain, (ii) C-cytoplasmic domain, and (iii) two extracellular domains.92 CD151 (cluster of differentiation 151), CD9, and Tspan12 are three of the 33 human tetraspanin proteins that have shown significance in tumorigenesis.93 Clinical investigations have found a strong link, associating the expression of CD151 protein with the sufferers of NSCLC, suggesting that it might be an efficient diagnostic tool for NSCLC, especially in patients of adenocarcinoma.94
2.15. Sigma (σ) Receptors
σ1 (sigma-1) and σ2 (sigma-2) are membrane-bound protein sigma receptors, each with its own pharmacological profile. Six out of 15 human adenocarcinoma samples and 12 out of 15 human SCLC samples have shown considerable overexpression of σ2 receptors. Various studies conducted on σ2 have demonstrated that it is a promising bioindicator which has the potential to be exploited as a new chemotherapeutic target.95,96
2.16. Anaplastic Lymphoma Kinase (ALK) Receptors
The ALK gene is responsible for encoding the human enzyme ALK receptor, also known as CD26. This receptor belongs to the superfamily of insulin receptors and exhibits an enhanced conformance to leukocyte tyrosine kinase. Genetic arrangement had earlier led to the initial identification of ALK receptor in anaplastic large cell lymphomas,97 following which, it has been reconfigured, mutated, or amplified in a series of cancers including neuroblastoma, lymphoma, and NSCLC.98
2.17. ROS Proto-oncogene 1, Tyrosine Kinase Receptor
The ROS1 (c-ros oncogene 1) gene is a receptor tyrosine kinase that structurally resembles the ALK protein. After the first report on ROS1 rearrangements in NSCLC was published,99 consequently altered ROS1 was identified in patients as a distinct type of molecule.100
3. Targeting Approach of Nanocarrier
The interactions between nanomaterials and biological systems must be thoroughly investigated in order to develop an efficient nanoparticulate system for delivery of drugs. Three phases must be fulfilled before the nanoparticle-drug delivery system reaches the tumor. They initially circulate through the body and are substantially ingested by macrophages from the reticuloendothelial system (RES).101 Subsequently, they proceed to permeate and deliver drugs to the tumor. Eventually, they accumulate at the tumor site where they exert pharmacological effects. However, before NPs reach the malignant cells, they are more inclined to interact with immune cells (macrophages and neutrophils) in the RES (liver and spleen) via blood circulation and are eliminated by urinary clearance.102 Perpetuating the blood circulation is required to reduce RES deposition, however; a prolonged vascular circulation of NPs enhances the likelihood of immunological interactions with the RES. Thus, specific surface alteration can considerably boost the absorption and dispersion of smaller NPs in tumor tissues. Drugs that are hydrophobic are more soluble when they are combined with polyethylene glycol (PEG). It is advantageous to modify the nanoparticle surface of hydrophobic drugs by amalgamating it with PEG because it elevates solubility, prolongs circulation period, and targets specified tumors through an enhanced permeation and retention (EPR) effect.103 Two generations of PEG conjugation approaches are used: the first generation, also known as PEGylation by randomization, and the second generation, also known as the targeted approach. Specific targeted drug delivery can be achieved by PEGylation, through active and passive routes. Numerous nanoparticles have been synthesized with an improved PK (pharmacokinetic) profile via PEGylation technique. Since such PEGylated NPs can significantly evade RES observation than the control nanoparticles, they are frequently referred to as “stealth” nanoparticles.103 Strategies involved in the development of target-specific small molecules are respectively modified through inserting a nanocarrier inside the surface-functionalized novel nanosystem or by surface modification of the novel carrier (see Figure 1). Localized drug delivery of nanosystem for lung cancer therapy has fundamental advantages, but an undiscriminating release of anti-cancer drugs can be lethal for non-cancerous cells and the lung parenchyma. Targeted drug delivery systems provide a selective release of the drug directly into the cell. These modified novel delivery compositions are required to reduce cell toxicity. In addition, they also help in increasing the retention time and efficacy of the drug. These novel drug deliveries of small molecules play an important role in lung cancer treatment.
Figure 1.
Nanoparticles as an efficient carrier in the therapeutics of lung tumor.
Strategies are generally classified into three categories: (1) receptor-based targeting, (2) passive targeting, and (3) stimuli-responsive system-based targeting.
3.1. Receptor-Based Targeting
Modification of the drug nanocarrier surface results in the delivery of the drug which is specific to the tumor cell by receptor-mediated endocytosis. Folic acid, transferrin, albumin, hyaluronic acid, RGD (arginine-glycine-aspartic acid), and herceptin have been utilized for modifying the surface of nanocarriers and now have improved the tumor-targeting efficiency. Here are some receptor-based targeting approaches which are contributed by many researchers for providing targeted delivery of novel nanocarrier in terms of permeation, cell uptake, and tumor internalization
EGFR Receptor-Based Targeting: Herceptin could be used as an option for receptor-based targeting. Herceptin identifies HER2 receptors overexpressed in breast cancer tumor cells. Paclitaxel (PTX) and docetaxel (DTX) nanocarriers modified with herceptin were employed to avoid drug loss before it reaches the targeted cells and organs. Herceptin-modified nanocarriers could be prepared by adsorption reactions which are carried out overnight with moderate shaking at room temperature. Such modifications have enhanced cellular uptake, improved the release of drugs, and have demonstrated cytotoxic activity.104
Transferrin/Folic Acid Receptor-Based Targeting: Paclitaxel nanocarrier was reportedly inserted with apo-transferrin to provide a selective target delivery. Another example of paclitaxel is the modification of the paclitaxel nanocarrier superficially with transferrin surface protein which had previously provided higher tumor inhibition. Selective tumor-targeting peptide, XQ1(HAIYPRHGGGF), binds with transferrin receptors and thus improves the antineoplastic action of camptothecin nanocarriers on HeLa and A549 cell lines.105
Vascular Endothelial Growth Factor Receptor-Based Targeting: Oral treatment of bovine lactoferrin (BLF) reduced VEGF overexpression in lung cancer cells, which is linked to vascularization and angiogenesis, as per immunohistochemistry examination. These mechanisms serve as a crucial contributor to the tumor development, including tumor progression, penetration, and metastasis. BLF could block angiogenesis, which makes it useful for treating the carcinoma efficiently.106
Luteinizing Hormone-Releasing Receptor-Based Targeting: Tartula et al. discovered that luteinizing hormone-releasing receptor fabricated nanostructured lipid carriers (LHRH-NLC) had success in delivering paclitaxel to the lungs via inhalation. The results showed that Luteinizing hormone-releasing receptor conjugated nanostructured lipid paclitaxel had 16 times better apoptosis than the standard drug and was selectively localized in tumor sites compared to non-targeted NLC nanoparticles.107
CD44 Receptor-Based Targeting: Hyaluronic acid-coated lapatinib nanocarriers are used in active targeting of the CD44 receptor which is overexpressed by the cancer cells.108 Another type of a nanocarrier of paclitaxel was combined with hyaluronic acid that binds to CD44, which is a cell surface receptor upregulated in some other types of cancer (gastric or breast).109 This intercellular adhesion molecule binds with a receptor for hyaluronic acid-mediated motility for tumor-selective targeting.
Integrin Adhesion Molecule αvβ3-Based Targeting: RGD engineered nanocrystals can attach to αvβ3, an integrin attachment protein seen on tumor cells, and then could be incorporated via receptor-mediated endocytosis. To achieve αvβ3-based targeting, RGD-modified NCs (nanocapsules), such as RGD-modified PTX NC,110 may be synthesized via the Schiff base reaction.
Sigma-1 (σ1) Receptor-Based Targeting: Ligands include [18F]FTC-146, [18F]fluspidine, donezepil,111 while for sigma2, it includes [11C]RHM-1, [18F]ISO-1, siramesine, WC-26, DTG (1, 3-ditolyl guanidine), and SV119.112 Drugs that modulate ion channels and antiproliferative processes, have shown anti-cancer efficacy when acting on the sigma1 and sigma-2 receptors. As a result, the medication acting on receptors makes it an effective potential chemotherapeutic tool in treating lung carcinoma cell lines.113
ROS Proto-oncogene 1 Receptor-Based Targeting: Crizotinib was first discovered to have anti-cancer efficacy in individuals with advanced type NSCLC who had a ROS1 fusion.114 Other examples that inhibit the ROS1 receptor (c-ros oncogene 1) are ROS1 Ceritinib (LDK378),115 Lorlatinib116 DS-6051ba117 and Entrectinib118
Other Endocytosis: Targeted drug delivery could be done by the endocytosis pathway such as caveolar or clathrin-mediated endocytosis. Modified paclitaxel nanocarriers are prepared by albumin absorbed paclitaxel nanocarrier and triphenyl phosphonium cation.
3.2. Passive Targeting
Passive targeting approaches enhance the infiltration rate in circulatory vessels in the cancer microenvironment by EPR effect as compared to the usual healthy condition. Enhancing the permeability and retention generally increases 20% to 30% targeting ability as compared to healthy cells.119,120 Tumors generally grow in unfavorable conditions such as hypoxia or inflammation and rapidly growing cancerous tumors block the new vessels or blood vessels which are present in the surrounding environment. This selectively enhanced permeation of nanosystems might help the tumor stroma and furthermore such enhancement in the permeation in newly generated vessels could improve the retention time. However, drainage of lymph could remove such small molecules and may also decrease the retention time of the drug. Therefore, the regulation of lymphatic drainage is necessary. Physico-chemical characteristics including charge, particle size, and surface-related chemistry are known to improve the EPR effect-based passive targeting to the tumor. It is challenging for a drug molecule which possesses a certain degree of therapeutic value to lodge itself effectively in all target cells.121 Uneven targeting of tumor cells is the basic limitation that could be overcome by producing more target-specific novel nanocarriers that efficiently link with tumor cells following the irruption process. NDDSs (novel drug delivery systems) containing small molecules improve their pharmacokinetic profile by increasing the retention time and reducing unwanted adverse effects that occur due to the target specificity of the drug.
3.3. Stimuli-Responsive System-Based Targeting
A stimuli-responsive system-based targeting could provide effective and selective therapies. Tumor cells have many key variations in their intracellular cytosolic levels and extracellular tumor microenvironments such as hypoxia, temperature, pH, and glutathione levels. Release of the drug from polymers on exposure to a stimulus specific to a tumor cell microenvironment or providing various types of polymers from the surroundings might be a viable strategy for the preparation of such nanocarrier inserted drugs.
pH-Responsive Systems: Over the years, the lower pH of the tumor microenvironment (4.5–6.5 pH) and lysosomes (5.5–6.8 pH) have been explored to directly release of a drug to the extracellular tumor microenvironment or intracellular cytosol, mitochondria, and nucleus. Liang et al.122 used a pH-responsive peptide and mannitol-based spray-dried powder inhalation system to improve nucleic acid transfection and ensure stability in the prevalence of broncho-alveolar lavage media, delivering payloads to tumor areas with lower pH values.
Temperature-Responsive Systems: As hyperthermia is one of the characteristics of tumorous tissues, modifying the medication release rate in a temperature-dependent way at a specific location has proven highly beneficial in cancer therapy. Moreover, the surface of nanocrystals may be modified using thermo-responsive polymers. To encapsulate 5-Fluorouracil, Wu et al.123 employed cellulose nanocrystal grafted thermo-responsive PNIPAM [poly(N-isopropylacrylamide)] brushes. When the temperature was elevated from 25 °C to 37–40 °C, PNIPAM was shown to produce a multiple-fold increase in drug release from the system. The thermo-responsive release method can also be utilized with stimulus-guided methods, in which the medication is released in response to an external signal.
Enzyme-Responsive Systems: Since specific enzymes are present at the tumor site, it is possible to employ them to selectively release medicines from enzyme-responsive systems. Tagami et al.124 described a doxorubicin-loaded hybrid liposomal system made up of phospholipids and poloxamer 188 that was sensitive to the PLA2 (phospholipase A2) enzyme expressed in cancer cells. The presence of the PLA2 enzyme resulted in an 8-fold increase in the release of drugs. Enzyme-responsive nanocarriers can be utilized to encapsulate medication nanocrystals and transport them via the pulmonary route for cancer therapy at a local level.
Ligand-Attached Stimuli-Targeted Systems: Pulmonary delivered receptor-based targeting mixed with stimuli-responsive system-based drug delivery can be utilized to improve drug intracellular localization in lung cancer cells. Tseng et al.125 created gelatin nanoparticle-based aerosols with EGFR receptor targeting for pulmonary administration in mice. The aerosolized method improved the intracellular accumulation of the drug substance in lung cancer cells. Such systems could be exploited to deliver additional treatments such as drug nanocrystals through the lungs. Similarly, the extracellular release of the payload was inhibited when the intracellular concentration of siRNA (small interfering RNA) was enhanced with higher stability by a siRNA-loaded dextran nanogel covered with pulmonary surfactant surface and adorned with folate.
Redox-Responsive Systems: Redox-sensitive systems offer great potential and offer several advantages in cancer therapy. Such stimuli-responsive methods have been proved to be effective for delivering anti-cancer medicines to cancer cells selectively. Tumors, or more particularly, cancer cells, have an increased amount of intracellular ROS, which can be exploited to deliver pulmonary nanoparticle systems to cancer cells. A liposome-based nanosystem comprising of a redox-sensitive cationic oligopeptide lipid with proton sponge effect, soya PC (phosphatidylcholine), and cholesterol was developed for the co-delivery of paclitaxel and anti-survival siRNA in a study done by Chen and co-workers.126 The synthesized liposomes were quick to release the drug and siRNA into the cytosol, through endo-lysosomal leakage, which consequently resulted in a reduced survival expression and elevated apoptosis induction.
Magnetic Field-Responsive Systems: Using magnetic fields as an external trigger, site-specific delivery of therapeutic substances to the cytosol or extracellularly localized malignancies in the lungs have also been described. Park et al.127 used PEG-polyethylenimine-DOPA (dopamine) to synthesize a magnetite nanocrystal cluster that exhibited magnetically induced intracellular transport of loaded siRNA into cancer cells.
Multi-Stimuli-Responsive Systems: Multiple multi-responsive delivery methodologies for the pulmonary delivery of biomolecules to cancer cells have been established previously. Qu et al.128 presented a dual redox and pH-responsive system dependent on a GSH (glutathione) sensitive disulfide linker and a pH-sensitive moiety, which specifically released the drug in the tumor microenvironment with elevated anti-tumor activity.
4. Nanocarrier-Mediated Drug Delivery Systems
Figure 2 illustrates several types of nanocarriers which are being employed in lung cancer therapy. In addition, due to the regulated drug release pattern and improved in vivo efficacy of encapsulated drugs, their acceptance is now well recognized. Various types of natural, semi-synthetic, and synthetic forms of nanomaterials are now utilized as drug delivery vectors.129
Figure 2.

Various nanocarrier-mediated drug delivery systems.
4.1. Liposomes
Liposomes are gaining much significance lately as a potent anti-cancer drug delivery tool. Liposomal research has expanded in the past decade because of its excellent biocompatibility, resulting in a plethora of newly formed formulations like: archaeosomes, virosomes, temperature-responsive liposomes, and cationic liposomes.130 Liposomes are nanocarrier tools, synthesized from natural or synthetic lipids and occur as single or multiple bilayer forms.131 Banham developed liposome-derived phospholipid vesicles in 1965, and they were quickly identified as prospective medication carriers.132 These are classified as multilamellar vesicles (Figure 3) made of several concentric phospholipid bilayers (1.0–5.0 μm), as well as big (500–1000 Å) and small (300–500 Å) unilamellar vesicles.133 Several drugs which were approved lung cancer therapy are being continuously developed into liposomal formulations. Etoposide (ETP), doxorubicin (DOX), paclitaxel (PTX), irinotecan (IRI), erlotinib, docetaxel (DTX), vinorelbine (VNB), cisplatin (CPPD), and epirubicin were developed as liposomal formulations. In addition, the antimetastatic efficacy of DOX conjugated liposome and tretinoin have been studied as chemotherapeutic agents for cancers that have spread to the lungs (melanoma, breast cancer).134−136 The importance of liposomes is associated with its enhanced efficiency and less invasive drug delivery to the bronchial tissues which is a consequence of their prolonged release characteristic and the increased probability to target-specific administration of the therapeutic agents.137 Remote-loading or gradient methods have been used to load hydrophilic pharmaceuticals such as DOX, which is a hydrophilic weak base molecule. Liposomal-based formulation serves as a solubilizing matrix for hydrophobic, water-insoluble compounds like PTX, as well as a means for lowering medication-related toxicity.138
Figure 3.
A unilamellar liposome nanoparticle. Lipophilic drugs are contained between the amphiphilic bilayer, while hydrophilic drugs are confined in the central core.
Stealth liposomes or PEGylated liposomes are synthesized through modifying the lipid bilayer surface and by incorporating the synthetic PEG into the mixture. When insertion of cholesterol into the lipoidal layer is executed, attaching the liposomal membrane with PEG allows the liposome to circulate in the blood system for longer periods which results in its diminished absorption by the phagocytic system.139 Therapeutic activity is initiated, when the drug molecule passes through the cellular membrane to intracellular sites of action. As a result, ligand-targeted liposomes were proposed to improve the selectivity of liposome-mediated administration. This technique has gained considerable attention for utilization in lung cancer-based treatment and research. In lung cancer, the main targeting strategies include (i) specific targeting of the tumor microenvironment (TME), (ii) overexpressed receptor-mediated targeting, and (iii) targeting the organelle in demand. Attaching a target-specific ligand with DSPE-MPEG (1,2-distearoyl-sn-glycero-3-phosphoethanolamine-N-[maleimide(polyethylene glycol)] integrates the ligand with the membrane during the generation of liposome. Sometimes the ligand gets infiltrated into the lipoplasmic surface. Cetuximab140 possesses significant inhibitory activity on EGFR and EphA2 (Ephrin A-family) receptors. This ligand was linked to the maleimide group of DSPE-MPEG through the surface of liposomal bilayer. DSPE-PEG was coupled with small molecules such as arginylglycylaspartic acid (which targets overexpressed integrin receptors) and anisamide (which targets overexpressed sigma receptors). A CD44 receptor-targeting hyaluronic acid ligand was reported to be coupled with DOPE (1,2-dioleoyl-sn-glycero-3- phosphoethanolamine).141 Khatri et al.142 demonstrated that RGD grafted liposomes carrying siRNA were more efficiently absorbed by NSCLC cells from a target-specific and multipurpose-based epirubicin liposomal formulation.143 The formulated multipurpose epirubicin liposomes were found to have better tumor targeting efficacy, a longer survival time, and a greater anti-cancer effect than non-targeted liposomes liposomes.144,145 Octreotide is a tumor marker that may be linked onto the liposomal membrane by selectively binding with the overexpressed somatostatin receptors.146,147
There are several types of liposomal formulations in various phases of clinical trials to treat NSCLC (Table 2).148 Several preclinical investigations have shown that targeted and non-targeted liposomal drug delivery methods could improve biodistribution and therapeutic efficacy.149 Doxil was the first FDA-approved (U.S. Food and Drug Administration) nanodrug in 1995. Utilization of PEGylated nanoliposomes, aided in extending the drug circulation time and preventing RES.150 Doxil and topotecan have entered phase I clinical trials as potential treatment modules in SCLC. However, no information from this study is available.151 The chemotherapeutic drug CPPD is the most widely utilized drug in lung cancer treatment regimens. SPI-077 (Alza Corporation) was the first liposomal CPPD to be developed, containing the same lipid makeup as Doxil. Animal studies and preclinical trials revealed promising efficacy in NSCLC treatment and a potential delay in proliferation and growth of tumor cells when compared to CPPD. According to the data provided by phase I/II trials, large doses of liposome-based CPPD when administered are safe and have less toxicity in patients with NSCLC. The potency of lipoplatin in NSCLC was compared in two clinical trials which were randomly conducted (phase II and phase III). Lipoplatin conjugated gemcitabine or lipoplatin conjugated paclitaxel were observed to be as effective as cisplatin conjugated gemcitabine or cisplatin conjugated paclitaxel. Additionally, these formulations were proven to be substantially less toxic.152 In NSCLC patients (stage III and stage IV), stimuvex exhibited satisfactory results in the initial stage trials.153 This was the earliest cancer vaccine which entered multiple advanced phase III clinical studies around the world (START, INSPIRE, STRIDE).
Table 2. Liposomal-Based Formulations under Clinical Trialsa.
| formulation | application | clinical trial phase |
|---|---|---|
| BLP25 liposome vaccine | squamous non-small-cell lung cancer | II |
| liposomal lurtotecan | small-cell lung cancer | II |
| cholesterol-fus1 liposome complex | lung cancer | I |
| irinotecan liposome injection | small-cell lung cancer | II/III |
| paclitaxel liposome | stage IIIA non-small-cell lung cancer | IV |
| stage IIIB non-small-cell lung cancer | ||
| adenocarcinoma of the lung | ||
| liposomal cisplatin (lipoplatin) | non-small-cell lung cancer | III |
Data obtained from U.S. National Institutes of Health Web site (http://clinicaltrials.gov/) on October 4, 2021.
The necessity for consistency in biological fluids is one of the concerns with liposomes, when employed as drug delivery vehicles. Drug molecules leak into natural tissues, as a result, producing undesired effects. In conclusion, future research on liposomal conjugated drug delivery could be designed through functionally versatile platforms by administering target-specific theranostic agents for cancerous cells. Liposomes have also shown magnificent biodegradability and biocompatibility for lung cancer as theranostics. Furthermore, liposomes have an advantage over other nanoparticles since they can retain a large number of therapeutic substances and are easy to develop and utilize for long-term drug delivery. For morphological evaluation using MRI, contrasting compounds such as gadolinium (Gd) are often utilized.154 Although liposomes conjugated with Gd have been proven to have no toxicity or adverse effects in normal cells, these complexes might be used as a promising theranostic tool for cancer therapy if loaded with antitumor agents and coupled with particular ligands of interest. Cheng et al.155 employed the EGFR binding interactions of a novel peptide GE11 in doxorubicin-encapsulated liposomes to evaluate size range, zeta potential, drug encapsulation efficiency, and shape. In A549 cytotoxicity, it was reported that a 10% GE11 density was optimum. They discovered that GE-11 altered liposomes accumulated and retained 2.2-fold more than unmodified liposomes using a near-infrared (NIR) fluorescence imaging system.155 Additionally, they will be capable enough for triggering targeted delivery at specific locations where an outcome is required. Furthermore, surmounting the therapeutic resistance will be the objective in modified and improved liposome-based drugs. Advancements in treating lung cancer through therapy, could be achieved by implementing the suggestions discussed, which also aim at minimizing the side effects caused on healthy cells and tissues.
4.2. Polymeric Nanoparticles
Different polymers are employed to encapsulate drugs which constitutes a class of nanoparticles known as polymeric nanoparticles (Figure 4). Such polymeric NPs are employed to target-specific tumor cells or tissues by linking them to the ligand and increasing the affinity and stability of the drug.103 Polymeric nanoparticles are widely demanded for both active and passive strategies of targeting malignant cells. There are several polymers that can be used in the synthesis of polymer nanoparticles. Poly(lactide-co-glycolide) (PLGA), poly(lactic acid) (PLA), and a few additional polymers [poly(glycolic acid) (PGA), poly(ε-caprolactone) (PCL), poly(acrylic acid) , etc.] are frequently used polymers in nanoparticles. PLA is a polyester created from lactic acid which is biodegradable. The copolymer of lactic acid and glycolic acid, known as polylactic-co-glycolic acid (PLGA), is both biodegradable and biocompatible. The U.S. FDA and EMA (European Medicines Agency) have validated PLGA as a drug delivery vehicle for parenteral route of administration.156 The use of polymeric nanoparticles in the treatment of lung cancer demonstrates encouraging outcomes. Taxanes-loaded PEG-PLA (polyethylene-glycol-modified polylactic acid) nanoparticles have greatly increased the efficacy of chemotherapy and radiation treatment in vitro and in vivo (A549 lung tumor xenograft model).157 By co-entrapping paclitaxel and cisplatin within PEG-PLA block copolymers, Kim et al. created nanoparticles for the treating lung cancer,158 which have entered clinical trial phase II under the name Genexol-PM for advanced NSCLC.
Figure 4.

Structural representation of polymeric nanoparticles.
Wang and his team developed a prodrug-based nano-drug delivery system where prodrugs of baicalein (BCL) and paclitaxel (PTX) were used.159 These NPs were created by nanoprecipitation, and the innermost core was made of PLGA polymer conjugated with BCL and PTX. The MTT (3-(4,5-dimethyl-2-thiazolyl)-2,5-diphenyltetrazolium bromide) assay was implemented to conduct in vitro cytotoxicity tests. In A549 cells, PTX-BCL NPs showed higher cytotoxic effects (P < 0.05) compared to other NP formulations or free-drug solutions. Additionally, PTX-BCL solution demonstrated lower cytotoxicity than PTX-NPs and BCL NPs. Thus, the exponential usage of polymeric nanoparticles is producing chemotherapeutics for tumor treatment which are advanced, revolutionary formulations with least harm to healthy cells and which can be optimally delivered to cancerous tissues.
4.3. Dendrimers
Dendrimers are intensely branched, highly biocompatible, uniformly structured, and complex molecules (having a diameter of 2–10 nm) of spherical or globular appearance. The innermost core of the dendrimer has a tree resembling orientation which has numerous extensions attached to it. It exists in two different forms: high-molecular-weight (hyperbranched, dendronized, and brush polymers) and low-molecular-weight dendrimers (monodispersed and highly symmetrical polymers).160 The unique properties exhibited by dendrimers are their multivalent nature, their excessively branched core, their uniform globular shape, and their distinguishable molecular weight. These promising features include them in the list of intriguing new drug delivery scaffolds.161 Their tendency to precisely modify or adjust their size, form, and surface characteristics, make them a typical and adaptable nanocarrier system for the targeted delivery of chemotherapeutics.162 Dendrimers are made up of three parts: a central core acting as inducer, various interior branches arranged in layered form, and multiple surface end groups for covalent conjugation. The internal central core moiety has reacting sites, with branches extending out from it. The first-generation monomers (G1) are those linked to the center (G0), and surrounding it, a pair of second-generation monomers (G2) are observed. Arrangement of the next generations will be followed by a similar approach (Figure 5).163 Their wide-ranging nanosized particles, instantaneous synthesis, tendency to conjugate, and their ability to demonstrate numerous copies of lethal functional groups necessary for rearrangement, make them appealing platforms for drug delivery.36
Figure 5.

Demonstrated dendrimer generation.
A concise list of dendrimer-based formulations for treating lung carcinoma and their outcomes has been summarized in Table 3. Therapeutic small molecules could be encapsulated in the innermost core through either hydrogen/chemical linkage or lipophilic interaction.162 To improve bioavailability and target-specific delivery, methods like acylating or PEGylating the active terminal surface and fabrication of dendrimers with receptor proteins could be done, respectively. PEGylated dendrimers are a type of dendrimer which have attracted the interest of formulation scientists because of their prolonged distribution duration in the circulatory system, reduced accumulation in various body organs, and minimal toxicity. Incorporating poly(ethylene oxide) (PEO) with PEG aids in synthesizing PEG-dendrimers.163 PEGylated dendrimer nanoparticles exhibited promising potential as an aerosol-inhaled drug delivery method when studied in a related investigation.171 Furthermore, doxorubicin-conjugated PEGylated dendrimers when administered through inhalation routes resulted in sustained duration of residence of drug in carcinoma tissues and efficiently lowered carcinoma load in rats.172 Commercially available dendrimers are primarily divided into two types: poly(amidoamine) (PAMAM) and poly(propyleneimine) (PPI).174 The former is found to be used more preferably for delivery of chemotherapeutic and other peptides which must be delivered efficiently and effectively to specified sites.176 These PAMAM dendrimers are hydrophilic, biocompatible, and nonimmunogenic. They have the capacity for increasing the bioavailability of chemical entities while reducing their frequency of administration which supports their usage as a medication delivery technique.177 A study was conducted to synthesize a nano drug delivery system of doxorubicin conjugated generation-4 succinamic acid (G4SA) PAMAM dendrimers and to evaluate them in an in vitro lung carcinoma 3D model (Scheme 1). It was observed that in comparison with free doxorubicin, dendrimer-based doxorubicin formulation exhibited three times greater permeation rate with 3.1-fold increased rate of aggregation of drug in the innermost core region of co-cultured spheroids. PPI dendrimers could also be incorporated with the drug moiety by diversified techniques which makes them essential for evaluation in biomedical applications.178 PPI dendrimers exhibit internal hydrophobic properties as a result of the presence of branched alkyl chains. Furthermore, because of the linkage of amides with branched alkyl chains, these PPI dendrimers have a more lipophilic core as compared to PAMAM dendrimers.179,180 The toxicity of PPI dendritic delivery, like that of PAMAM dendrimers, can be minimized by surface-acetylation or other comparable methods.181
Table 3. Dendrimer-Based Approaches for Effective Drug Delivery in Lung Cancer.
| drug | type of dendrimer | formulation | observations | ref |
|---|---|---|---|---|
| cisplatin | PAMAM [poly(amidoamine)]dendrimer | cisplatin-conjugated dendrimer (NCI-H460 cells) | Through sonication or centrifugation, cisplatin load-bearing capacity of dendrimer was increased. | (164) |
| Drug was released at targeted sites with minimum loss. | ||||
| Due to the pH of tumor the drug release rate was higher. | ||||
| Conjugated dendrimer was found to be biocompatible and inhibited proliferation activity in tumor cells. | ||||
| gold NPs | folic acid-conjugated PAMAM dendrimer | folic acid-decorated gold-conjugated PAMAM dendrimer (SPC-A1 cells) | After treatment with Au DENPs-FA (folic acid-modified dendrimer-entrapped gold NPs), the picture of SPC-A1 (human lung adenocarcinoma cell line) cells became clearer. | (165) |
| Biocompatibility observed but without any changes in viability of cells. | ||||
| doxorubin (DOX) | PAMAM | doxorubicin-conjugated dendrimer (melanoma (B16–F10) lung metastasis mouse (male C57BL/6) model) | Better stability shown with respect to extracellular pH. | (166) |
| Better chemotherapy performance shown instead of I.V. (intravenous). | ||||
| More nodules in lungs were reduced when DOX-conjugated dendrimers were administered instead of free DOX. | ||||
| paclitaxel | peptide-conjugated dendrimer | PTX-conjugated dendrimer (L132 and 293T cells) | Better encapsulation efficiency (95%) was reported with 25% of loading capacity. | (167) |
| Sustained release pattern demonstrated higher antiproliferative effect on 293T and L132 cells as compared to free PTX. | ||||
| EndoNt | PAMAM | CTC (circulating tumor cells) detection viability in SCLC (PolySia-positive SCLC cells) | Small-cell lung cancer cells and EndoNt (non-catalytic endosialidase)-immobilized surface have good binding stability. | (168) |
| In small-cell lung cancer cells under flow, EndoNt capturing efficiency tends to be higher than that of aEpCAM (epithelial cell adhesion molecule). | ||||
| doxorubicin | PEGylated-PLL (poly l-lysine) dendrimers | sustained and controlled exposure of therapeutic | Doxorubicin shows prolonged and sustained action. | (169) |
| After I.T. (intratracheal) administration of dendrimer-incorporated DOX, lung-related toxicity concerns were lower when compared to those with an equivalent dose of DOX solution. | ||||
| gemcitabine | mannose-conjugated PPI [poly(propyleneimine)] dendrimer | mannose-conjugated PPI dendrimer incorporated with gemcitabine (A-549 cell line in albino rats) | Drug loading profile was enhanced. | (170) |
| Sustained release with better cytotoxicity against A-549 cell lines. | ||||
| This formulation tends to be a potential drug-delivery tool against lung cancer. |
Scheme 1. Synthesis of G4SA-GFLG-DOX (Generation 4 Succinamic Acid PAMAM Dendrimer Conjugated with Doxorubicin).
DMF: dimethylformamide, EDC: 1-ethyl-3-(3-dimethylaminopropyl) carbodiimide, and NHS: N-hydroxysuccinimide.
Dendrimer-integrated photo-medicinal platforms can be applied as molecular imaging contrast agents to facilitate detection of cancer precisely by coupling the optical features of photoactive agents with the qualities of other distinct components. Luong et al.173 used a folic acid (FA)-conjugated PAMAM dendrimer with a superparamagnetic iron oxide (SPIO) core to successfully deliver the hydrophobic anti-tumor medication 3,4-difluorobenzylidene curcumin (CDF). Furthermore, incorporating CDF in a FA-PAMAM dendrimer resulted in higher targeting ability, anti-cancer activity, and MRI contrast-enhancing capabilities. Chen et al.175 demonstrated that a dendrimer complexed with a nanocomposite of Au nanoparticles, cyanine dye (Cy5.5), and Gd(III) could be used as a tri-modal imaging agent for CT (Au nanoparticle), fluorescence imaging (Cy5.5), and MRI (Gd(III)) to impart improved spatial and density resolutions with high specificity. Developing nanomedicines will surely benefit from such integrated multimodal theranostic strategies.
4.4. Nanoemulsions
Administration of chemotherapeutics in the form of nanoemulsions might influence intratumoral chemotherapeutic accumulation, as nanoemulsion delivery employs various pathways for enhanced bioavailability.182 Nanoemulsions are droplets of water in oil (w/o) or oil in water (o/w) having an average radius of 10–100 nm which are translucent or transparent. These heterogeneous dispersions exhibit a stable thermodynamic profile. They are formed by dispersing aqueous phase in oil phase, as both these phases are immiscible with each other, the dispersion is maintained by reducing the interfacial tension with the help of different surfactants and co-surfactants.183 The distinctive nanoemulsion architecture is administered with hydrophobic or lipophilic chemotherapeutics for creating hydrophilic and lipophilic conditions. Such a design of nanoemulsion aids in reducing liver bypass, inhibits drug decomposition in abnormal environments, eliminates P-glycoprotein outflow, promotes mucosal penetration, and therefore improves the chemotherapeutic systemic availability.184 Furthermore, decreased stomach emptying and enhanced drug solubility due to a large expansion in the interfacial area is made possible by producing nanoemulsion droplets which may be linked to improved bioavailability.185 Subsequently, the functionalization of 2-hydroxypropyl-β-cyclodextrin (HP-β-CD), lipoid E 80 (agents which elevates permeation) enhance the bioavailability of embedded chemotherapeutics by temporarily opening cellular tight junctions.185,186 Improved bioavailability of chemotherapeutics in the tumor microenvironment (TME) depends, whether the targeting is active or passive when nanoemulsion is administered intravenously.182,184 Furthermore, nanoemulsions are advantageous, as they possess enhanced therapeutic absorption due to increased surface area and hence improved drug bioavailability. An additional feature is that it shields the drug from UV (ultraviolet) radiation and oxidative degradation, which improves the drug’s stability in the formulation.187 Nanoemulsions also have the advantage of being resistant to bacteria, fungus, and viruses.188 Furthermore, its customizable droplet size has the potential to ameliorate the infiltration tendency; for instance, decreased skin irritability was observed when the mean droplet size was smaller, as the small-sized globule easily infiltered into the skin through hair follicles and pores (where the surrounding healthy tissues were undisturbed).187 Doxorubicin is an anti-cancer substance which is used to treat a variety of malignancies. On the other hand, it has hazardous effects on healthy tissues because of its broader dispersing characteristic feature and a shorter half-life. A pH-responsive lapidated nanoemulsion containing doxorubicin prodrug (DNE) was developed which possesses a significant biocompatibility. The bioluminescent findings from the experiment on the mouse model showed a substantial decrease by DNE in distant lung metastases.189 Kim and Park created a NSCLC-targeting nanoemulsion having hyaluronic acid-conjugated paclitaxel.190 While creating the nanoemulsion of conjugated paclitaxel, the formulation was subsequently chelated with the targeting moiety, hyaluronic acid, using a microfluidizer at high pressure. The experiment was carried out on nude mice which had overexpressed CD-44 in NCI-H460 cell tumor xenograft. Cancer progression was more effectively inhibited due to the astounding encapsulation efficiency (>100%).190−192 Li et al. also reported on a nanoemulsion of docetaxel,193 a cancer-fighting drug from the same family. Researchers had created an o/w emulsion with the use of medium-chain triglyceride (MCT) composed of the given constituents: (i) poloxamer 188, (ii) oleic acid, and (iii) egg lecithin. In nanoemulsion and commercial formulation groups, histological examination of tumor development indicated necroptosis of cancer cells (by formation of a capsule along the fibrous connective tissues).193 Due to the absence of breaching of the fibrous connective tissue barrier, malignancy of cancer cells was prevented, and the cancerous cells did not metastasize to other body parts.194 The production and effectiveness of curcuminoid (derived via Curcuma longa) nanoemulsion were observed against cancerous cell lines.195 The curcuminoids and nanoemulsion of curcuminoids were treated with two distinct cell lines: H460 cells (large cell lung cancer) and A549 (adenocarcinoma of lung). The curcuminoid containing nanoemulsion affected the H460 cancerous cells by overexpressing the P21 gene (cyclin-dependent kinase inhibitor) which was followed by reduced cyclin-dependent kinase 1 (CDK1) and cyclin B, which eventually led to the cessation of cell cycle in G2/M phase. Moreover, an elevation in the P53 gene expression of A549 cancerous cells was indirectly responsible for an overexpression of p21 gene which ultimately lowered CDK1 and increased cyclin B levels, halting the cell cycle in the G2/M phase.195 The o/w carvacrol encapsulated nanoemulsion was developed (by ultrasonication), as a delivery mechanism for carvacrol on a nanoemulsion platform. CANE or Carvacrol nanoemulsion tested against human A549 cancer cell lines (in vitro) and against a nude mouse xenograft model (in vivo), resulted in a dose-dependent cytotoxic response in the A549 cancerous cell lines.196 Inhalable nanostructured lipid particles were formulated by Jyoti et al. to administer 9-bromonoscapine, a tubulin-binding alkaloid that inhibits cell growth and induces death in NSCLC cells. The quick incorporation of the formulated nanoemulsion system through a specific mechanism (endocytosis) may be ascribed to the uneven external surface instead of uniform surface of lipid containing nanoemulsions, which may justify the increased apoptosis.197 Tanshinone conjugated nanoemulsion was synthesized by scientists for testing the ability of nanoemulsion to inhibit the growth of A549 cells. By ultrasonication, Tween 80, Capryol 90, and ethanol were used to create nanoemulsion. The produced formulation had an enhanced storage stability, homogeneity as well as improved encapsulation. Furthermore, as compared to tanshinone extract, the nanoemulsion exhibited substantial tumor cell suppression or death. Tanshinones can indeed halt the G0/G1 phase of cell cycle by overexpressing p53 and p21 genes, PJNK (Phospho-c-jun NH2 terminal kinase) while decreasing the expression of cyclin E1, cyclinD1, and CDK2.198 Nanoemulsions are drawing much interest in the biomedical application of lung cancer therapy and R&D. However, numerous obstacles need to be addressed in order to ensure the safe and stable administration of nanoemulsions. The major requirements for developing nanoemulsion formulation involves employing the excipients for providing a stable, effective, and target-specific profile by preventing the occurrence of undesirable adverse effects and confirming the biological outcome of the drug.
4.5. Micelles
Micelles are colloidal nanoparticles (Figure 6) which are formed in an aqueous environment when synthetic amphiphilic copolymer or surfactant is greater than a specific concentration, which is essentially the CMC (critical micelle concentration).199,200
Figure 6.
Structural representation of micelles.
Smaller size (diameter less than 100 nm) of the polymeric micelles is one of the reasons behind selecting them as an ideal drug delivery tool as they actively escape RES and renal exclusion. They even allow themselves to passively increase the penetration of the endothelial barrier in the area prevailing with the tumor. Polymeric micelles show great potential as hydrophobic drug delivery vehicles.36 Rapid self-assembling of amphiphilic block copolymers in aqueous environment due to hydrophobic or lipophilic interactions is responsible for designing the distinctive zones which form the structural configuration of micelles. Micelles consist of a hydrophobic center and a hydrophilic barrier which functions by preventing protein linkage, particle agglomeration, and opsonin linkage. These get decomposed in the circulatory pathways, before entering the targeted region.201 Liu et al. (2012) covalently functionalized β-CD (β cyclodextrin) fabricated micelles with folic acid (FA), doxorubicin (DOX), and contrast media (DOTA-Gd) which showed decreased cell viability (Scheme 2).202
Scheme 2. Schematic Representation of DOX, FA, and DOTA-Gd-Conjugated β-CD-Based Micelles.

Kataoka et al. demonstrated effective entrapment of doxorubicin (DOX) by poly(ethylene glycol)-poly(-aspartic acid) block copolymer [PEG-PAsp (DOX)] and the formulation is termed as NK911 which is currently being assessed in phase-II clinical trials.36 Unimeric micelles are more appropriate for drug delivery than diblock (hydrophilic–hydrophobic) or triblock (hydrophilic–hydrophobic–hydrophilic) particles due to their lower molecular mass of polymeric residues after micelle disintegration.201 Micelles are effective nanocarriers for the delivery of cytotoxic hydrophobic substances like antineoplastic drugs. These medications get enmeshed in the innermost central region, resulting with increased aqueous solubility, reduced toxic profile, tumor cell-specific aggregation and drug resistivity reversal. Furthermore, prolonged blood circulation time could be achieved with hydrophilic shell by inhibiting identification through reticuloendothelial barrier and preventing alveolar uptake. As a result of the tighter particle size characterization and extended plasma circulation (in vivo), they specifically aggregate around tumorous sites, which eventually improves therapeutic efficiency and bioavailability of weakly water-soluble drugs.203 The biodegradability of therapeutic delivery systems is an essential feature which makes this approach a preferable option for formulating anti-cancer and ocular drugs. Drug integration into micelles enhanced circulatory residence time, lowered tumorigenesis, and reduced the toxicity significantly. Various polymeric micelles were conjugated with the purpose of treating lung cancer and were evaluated through preclinical studies which are summarized in Table 4. Distributions based on conjugation with polymeric micelles are preferred because of their inherently adaptable structures. The capacity to influence their surface permits delivery of theranostic chemicals to particular regions, increasing the efficacy, susceptibility, and selectivity of therapeutic and diagnostic approaches. They may also be readily manipulated to increase theranostic solubility and biocompatibility. Micelles conjugated with targeting agents (alpha fetoprotein (AFP) antibodies), surface incorporated with MRI imaging tools (Gd ions), and antitumor drug (paclitaxel) (TGPM) showed growth inhibitory potential, and MRI revealed the deposition of TGPM at the target tissue of H22 tumor containing mice due to AFP (alpha-fetoprotein) overexpression.204 Polymeric poly(ethylene glycol)-phosphatidylethanolamine micelles incorporated with paclitaxel and coupled with SPIO nanoparticles imaging agents displayed concurrent MRI and antitumor action.205
Table 4. Micelle-Based Formulations under Preclinical Trialsa.
| drug | conjugation with polymer | phase |
|---|---|---|
| cisplatin | PEG-b-poly(glutamic acid) | phase Ib/II |
| paclitaxel | PEG-b-PLA copolymer | approved in South Korea (2007) and Europe (2013), Phase II (FDA, USA) |
| SN-38 | PEG-b-poly(glutamic acid) | phase II |
| DACHPt | PEG-b-poly(glutamic acid) | phase I |
| epirubicin | PEG-b-poly | phase I (aspartic acid hydrazone) |
Data gathered from US National Institutes of Health Web site (http://clinicaltrials.gov/).
4.6. Carbon Nanotubes (CNTs)
CNTs are lipophilic hollow tubes composed of carbon atoms with diameters ranging from 4 to 100 mm that vary depending on how graphene molecules are arranged.206 CNTs have the appearance of a single rolled-up sheet of graphene (Figure 7) which is related to the family of fullerene (third allotrope of carbon family).207 There are two forms in which nanotubes are categorized: (i) having a single wall and (ii) having multiple walls, having an outer diameter of 4–20 Å and 20–1000 Å, respectively. The former can be made when just a single graphene sheet is rolled into a cylinder-like shape, whereas the latter is created when several concentric layers of graphene are rolled to form a cylindrical appearance. Since the past few decades, their applications have been investigated to deliver chemotherapeutic substances because of their vast ability to develop new and better delivery systems. However, their negative health effects are a huge concern.208 Specific biological and physicochemical properties, such as hollow monolith formation, enhanced mechanical efficiency, higher length and diameter ratio, increased surface area, ultra-massive molecular weight, nanoneedle shape, elevated thermal and electrical conductivity, and the ability to alter the surface, make them a promising drug delivery vehicle.209,210 They can pass through the cellular membrane via endocytosis and then enter the cell because of their needle-shaped features.11 Drug substances could be injected into the inner chamber of CNTs or bonded to the surface, either covalently or non-covalently. Furthermore, the anchoring of various ligands to the surface of CNTs allows for the targeted therapy of a chemotherapeutic drug to a desired location. Certain chemotherapy drugs, such as doxorubicin, paclitaxel, and cisplatin, have been delivered using CNTs.211 Three common ways for manufacturing CNTs include laser ablation, thermally enhanced chemical vapor deposition (CVD), and arc discharge.214 CNTs are a carbon only moiety with a unique chemical structure which has hundredfold strength when compared to steel while weighing only one-sixth of it. They also have an outstanding thermal conductance. It is therefore, considered as a fascinating nanocarrier, which is capable for identification of cancer cells as well as delivering medicines or smaller chemotherapeutics to the carcinomic cells.
Figure 7.

Structural representation of a carbon nanotube.
CNTs are fundamentally soluble in neither aqueous nor organic medium. They cause severe toxic effects in biological fluids are a massive concern which needs to be addressed.215 Chemical variation enhances biocompatibility, lowers toxicity, turns them into water-soluble nanocarriers, and makes them serum-stable substances.216 The first anti-cancer medication to be coupled with single-walled carbon nanotubes (SWCNTs) to target epidermal growth factor receptors was Cisplatin. The findings revealed that the drug was more effective against squamous cancer cells with high EGFR expression. Furthermore, due to its nano size, it demonstrated improved efficacy in preventing tumorigenesis instead of targeting it passively.217In vitro and preclinical studies on docetaxel-loaded CNTs in PC3 and S180 (murine sarcoma) cell lines have shown that they are more impactful than free docetaxel.218 In another study, the efficacy of PEGylated nanotubes incorporated with a chemotherapeutic agent was studied with regard to its mitochondrial apoptosis, cytotoxic profile in NSCLC, its cellular uptake and dynamics. The authors attempted to see that the compound was mostly received by clathrin-mediated endocytosis, which would help the A549 cell line generate early endosomes. In addition, the cytotoxicity studies revealed that PEGylated conjugates induced mitochondrial death at a higher degree than the unconjugated moiety.219 A549 cancer cell lines were targeted by administering doxorubicin hydrochloride via nanotube/gold hybrids. Non-covalent interactions between aromatic rings (π–π) of doxorubicin and CNTs, resulted in the adsorption of large amounts of doxorubicin on both internal and external surfaces of the nanotubes. Findings from cellular experiments in vitro indicated efficient delivery and administration of doxorubicin via nanoparticles.220 Poloxamer, a hydrophilic non-ionic surfactant, was utilized for stabilizing supramolecular nanotubes containing doxorubicin complexes. For assessing the drug payload capacity of multi-walled carbon nanotubes (MWCNTs), fluorescence spectrophotometry was employed. It was observed that the concentration of CNTs has a significant impact on the intensity of doxorubicin fluorescence.221 Additionally, a transmission micrograph may be employed to ensure that CNTs are released efficiently. In contrast to doxorubicin–pluronic complexes and simple doxorubicin, cytotoxic assessment on MCF-7 cell lines revealed significant toxicity.222 According to Arya et al., paclitaxel conjugated single walled nanotubes resulted in an upregulated rate of apoptosis in cancerous tissues. Furthermore, they concluded that synergistic or additional effects (ROS-associated) produced due to CNTs were mostly responsible for the increase in paclitaxel’s cytotoxic impact in lung cancer.223 Sobhani et al. used fabricated CNTs to study the cytotoxicity of paclitaxel. The drug was attached to the surface of MWCNTs using a poly(citric acid) spacer that lowers the nanotube hydrophobicity. Based on their findings, it was concluded that paclitaxel loaded on functionalized CNTs had a greater level of cytotoxicity due to elevated cell infiltration when assessed in comparison with free paclitaxel drug.224
CNT linked nanomedicine imparts ways to combine imaging and pharmacological action as a “theranostic” tool in biomedicine. Specifically, CNTs can be combined with other nano-based formulations from various sources to obtain synergetic, imaging and therapeutic effects. These theranostic blended nanoparticles might be applied to identify, visualize, and treat numerous kinds of cancers. Furthermore, unique conjugated nano-platforms can aid in cancer therapy optimization and chemotherapeutic toxicity reduction. Mashal et al. synthesized tissue-mimicking conjugates with varying proportions of SWCNTs and assessed their dielectric characteristics and microwave heating response.212 In CNT-blended mixes, the researchers noticed a linear temperature rise. Al Faraj et al. synthesized doxorubicin-encapsulated SWCNTs for targeting and monitoring metastasized tumors.213 Free-breathing MRI and bioluminescence imaging were used to observe the patients. The magnets were implanted in the lungs of tumor-bearing mice at the metastatic tumor locations. The study depicted that amalgamation of non-invasive MRI (to locate tumor areas sensitively) with magnets that promote the magnetic targeting of CNTs improved overall treatment outcome. CNTs may be integrated with a variety of materials, including metals, polymers, and mesoporous silica, for application in theranostics for cancer treatment.
4.7. Gold Nanoparticles (AuNPs)
It is well known that drugs, genetic materials, proteins, and tiny molecules can all be delivered using AuNPs, which have been recently identified as a potential delivery vehicle for transporting and releasing drugs into various types of cells. Several enticing characteristics are responsible for making them an excellent tool for delivering chemotherapeutics. To begin with, the innermost section (Figure 8) is safe, biocompatible, and inert, which is the primary reason behind using these particles as an initiating point for building carriers.
Figure 8.

Structural representation of gold nanoparticles.
AuNPs are found in various sizes ranging from 5 nm to 200 nm in diameter, which may also be synthesized by regulated dispersion.225 In 1951, Turkevitch described a simple, methodical bottom-up approach for synthesizing small gold molecules. Reduction of sodium tetrachloroaurate ions with trisodium citrate, resulted in the formation of nanosized dispersed particles (ca. 20 nm). The technique has then been improvised progressively over time. Several chemical and physical techniques now synthesize AuNPs with controlled particle size and shape.226 Essentially, drug delivery vehicles must consider both size and dispersion. Finally, AuNPs’ highly adjustable and multivalent surface configurations allow for conjugation (covalently or non-covalently) with a variety of drug substances or biomacromolecules on their surfaces. Moreover, the therapeutic applications of the noble metal, gold have been widely explored, including its application in sensitized diagnosis, detection, and classification of lung tumor types.
The flexibility with which AuNPs may be fabricated is one of their most appealing features. AuNPs are beneficial in both active and passive targeting because of their ability to modify the surface. Due to high water solubility, the drug, methotrexate (MTX) generally has a poor tumor retention capacity, which is responsible for its delayed or weaker therapeutic effect in cancer patients. A mouse model with Lewis lung carcinoma when administered with MTX encapsulated AuNPs, demonstrated a significant accumulation in cancerous tissues which eventually improved the treatment effectiveness.228 Covalent binding of hydrophobic paclitaxel to AuNPs via DNA linker was demonstrated by Zhang et al. (2011) (Scheme 3),229 where they reported enhanced in vitro cytotoxicity against cancer cells. Gold nanoclusters with fluorescent tagged antibodies on their surfaces have been created with the purpose of specifically targeting overexpressed EGFR tumor cells.230 This configuration imparts the pathway for gold-based nanoparticles to be applied in both magnetic resonance imaging (MRI) and treatment. Currently, photodynamic therapy (PDT) has been widely employed for AuNPs to transport purpurin-18-N-methyl-d-glucamine (Pu-18-NMGA), a water-soluble PDT drug, to A549 cell lines.231 The photodynamic activity of Pu-18-NMGA-AuNPs was greater than that of free Pu-18-NMGA.
Scheme 3. Synthesis Scheme for Paclitaxel Incorporated with DNA-Linked Gold Nanoparticles.
The potential of utilizing gold nanocarriers as medication and biomolecule transporting tools, originates from the versatility of the AuNP monolayer platform.225 Researchers described a new AuNP carrier system with polyethylene glycol (coating), heptakis-(6-deoxy-6-mercapto)-block-cyclodextrin (chemotherapeutic), and anti-EGFR monoclonal antibodies (targeted ligand). Biocompatibility of PEG derivatives is responsible for protecting nanocarriers and the drug encapsulated within it from enzymatic decomposition. Because of the tendency to defend pharmaceuticals from enzymatic, physical, or chemical destruction, cyclodextrins (CDs) are one of the most preferred drug delivery tools. B-Lapachone entrapped within the core cavity of β-CD is recognized for use in the treatment of various carcinomas (like breast, lung, and prostate).232
AuNPs have been widely used as cancer theranostics because of their ease of synthesis and functionalization, excellent biocompatibility, and multifunctional theranostic features. Knights et al.227 recently published a study that showed the size-dependent effects of gold nanorods (AuNRs) on photoacoustic (PA) imaging response and pulsed-wave photothermal therapeutic (PW-PPTT) efficacy, that is critical for AuNRs clinical translation. Additionally, Nanospectra has developed a silica-gold nanoshell stabilized by PEG for the photothermal therapy to the solid tumors using an NIR light source.233
4.8. Magnetic Nanoparticles (MNPs)
MNPs have extensive pharmaceutical and biomedical applications, mainly in MRI, tailored drug/gene therapy, and cell transfection (magnetofection). The appended ligands influence the magnetic field around the nanocarriers and aid in targeted drug delivery either actively or passively. Application of various MNPs has been approved by the FDA when they are used simultaneously with conventional chemotherapy regimen, resulting in a significant outcome in cancer treatment.234 Nanoparticles of gadolinium, and platinum are utilized for chemotherapy and other imaging strategies because of their aggregating tendency near the tumor location.235 MNPs comprise superparamagnetic material with an average size of more than 25 nm.236 MNPs that are non-biodegradable are covered with a substance that enables the magnetic core to leach thereby aiding in their elimination from the body. Magnetic hyperthermia is a safe lung cancer treatment that uses heat to destroy tumor tissue. The exposure of these MNPs to an external magnetic field results in the induction of thermal effects that eventually cause slower cell death when the temperature is above 42 °C, but instant death when temperature is above 45 °C.238 Similarly, the exposure of MNPs to alternating current results in the generation of sublethal heat, thereby exacerbating the tissue injury. In a mouse model of NSCLC-based hyperthermic destruction, Sadhukha et al.239 investigated the efficiency of tumor-targeted SPIO nanoparticles. The SPIO small molecules targeted the overexpressed EGFR which improved drug retention in tumors and reduced cancer progression substantially. Magnetic core and a surrounding functional covering, together constitute MNPs. Elements including iron (Fe), cobalt (Co), gold (Au), and nickel (Ni) make up the innermost core which offers it magnetic characteristics. The surface coat, on the other hand, inhibits aggregation and restricts magnetic core interaction with other particles.240 This surrounding coat contains a biologically effective therapeutic drug or a ligand. The first experiment with MNPs that can detect lung cancer micrometastasis was carried out by Wang and co-workers.241 Fabrication with the pan-cytokeratin epithelial tumor cell marker effectively identifies tumor cells in the blood circulation from patients with lung cancer. Drug resistance is subsequently surmounted using MNPs. Using cisplatin-loaded MNPs, an A549 cancer cell xenograft model which is resistant toward cisplatin, can be chemosensitized. Several molecular biologists then concluded that tumors treated with cisplatin-loaded MNPs had a substantial decrease in the concentration of lung resistance-related proteins and elevated cisplatin cytotoxicity.242 Iron oxide and doxorubicin exhibited magnetic and paramagnetic characters respectively, because of which, they tend to have greater cytotoxic activity.243 To target EGFR, gold-coated iron oxide nanoparticles were incorporated with 225-antibody, which caused DNA deterioration and cell cycle termination in the M/G2-phase.244 Another study245 demonstrated the cytotoxicity through C-225 (Cetuximab) conjugated nanoparticles on NSCLC. Cisplatin (a hydrophobic pharmaceutical agent), necessitates the use of a specific carrier to enhance its anti-cancer action. As a result, a functionalized Fe3O4 nanoparticle coupled with a PEG-PLGA copolymer has been constructed, demonstrating enhanced cisplatin anti-cancer efficacy in lung cancer.246 Nanoparticles of lower sizes are well known to be inhaled forming agglomerates in the lungs.247 MNPs are now encapsulated in micro-sized particles to effectively overcome this challenge. The most frequent materials utilized to make magnetic nanocomposite particles are iron oxide and d-mannitol nanoparticles. This method circumvents therapeutic molecules from diffusing through the bloodstream and transports the drug directly to the specific target location.248 Iron oxide nanoparticles (IONPs) coupled with gold have been reported to increase nanocarrier bioavailability and are deemed safe for treating lung malignancies.249 The transport of MNPs to the lungs could be beneficial and worth investigating further as a promising diagnostic tool or a medicine delivery method. In theranostic nanoparticle synthesis, MNPs, particularly IONPs consisting of magnetite (Fe3O4) or hematite, have been regarded important nanomaterials (Fe2O3) They are used in MRI contrast, medication administration, controlled and prolonged release, and hyperthermia therapy. Even human clinical trials are now being conducted on several of them.237
4.9. Solid Lipid Nanoparticles (SLNs)
Another type of vehicle for drug and gene delivery are SLNs. SLNs have a myriad of benefits, including the ability to encapsulate hydrophobic drugs more easily, reducing their negative consequences on the GI (gastro-intestinal) tract, and preserving sensitive drugs from an acidic pH (Figure 9).250
Figure 9.

Effects shown by solid lipid nanoparticle-encapsulated drugs.
These particles were first developed in the early 1990s. Many lipophilic or hydrophobic drugs were cytotoxic, so SLNs provided an essential foundation for such medicines to be entrapped. SLNs are colloidal nanocarriers for dissolved and dispersed therapeutics (size ranging from 50 nm to 1 μm) with a highly lipophilic lipid matrix comprising of solid lipids such as monoglycerides, diglycerides, triglycerides, waxes, fatty alcohols, free fatty acids, and steroids.252 SLNs have a solid lipid content of 0.1–30% (w/w), having the tendency to easily dissolve in aqueous media. Generally, 0.5–5% of surfactants are helpful to stabilize the SLNs.254 In terms of stability, enhanced biocompatibility, enhanced drug payload, and convenient production on commercial scale, SLNs outperform their lipid counterparts. SEM (scanning electron microscopy) and TEM (transmission electron microscopy) evaluation have indicated that SLNs have a spherical form with a size range of 50 to 1000 nm.255 A regulated release profile, sustained release characteristics, with a quicker in vivo decomposition are all considered benefits of SLN-based drugs when compared to particles composed of PLA or PLGA. Furthermore, as compared to particles produced from certain polymeric materials, SLNs were shown to have better tolerability in the lungs which contributes significantly to their safety profile.256 As a result, this system is suggested for the administration of pulmonary drugs, in a solution or as a dry powder, without causing inflammation.257 SLNs combine the properties of fat-emulsion carriers, liposome carriers, and polymeric nanoparticles, making them excellent for targeted drug delivery.258
This strategy has demonstrated enormous potential for targeted utilization (particularly intracellular) by manipulating gene expression and cellular communication. Currently, ultra-solid lipid nanoparticles are employed to target the tumor location.259 Anti-tumor drugs such as doxorubicin, etoposide, and idarubicin260 are currently administered using these nanocarriers which increase their bioavailability. Furthermore, SLNs254 are considered as good and efficient gene carriers. Cationic SLNs have demonstrated excellent transfection of p53 gene loaded with chemotherapeutic agents at targeted sites.261 Trimyristin (high-melting triglyceride) was used to synthesize novel SLNs of DCX (docetaxel). This strategy has resulted in elevated levels of DCX in tumor tissues while lowering them in other organs such as the heart, kidney, liver, and lungs. For optimizing the efficacy of SLNs, usually one or more glycerides (having high melting point) are used for synthesis.262 Choi et al.263 transfected SLN-carrier p53 into p53 null lung carcinoma cell line H1299. When compared to commercially available Lipofectin, the authors of this study effectively mapped overexpressed p53 protein which ensured that this system of drug delivery could be employed in genetic transfection with adequate drug payload.
According to a recent study, synergistic effects were observed when amalgamation of paclitaxel and Bcl-2 siRNA were incorporated into SLN.264 Yuan and colleagues observed that drug-loaded targeted SLNs had a greater proportion of drug uptake in A549 cells, as compared to SLNs incorporated with drug. Hydrophilicity of PEG facilitates an initial burst release, and it was further reported that drug-loaded PEGylated SLNs released more drugs in comparison to target-specific SLNs.265 When doxorubicin containing SLNs were administered through IV route in mice it was observed that the drug concentration in the lungs was greater than when they were administered with the solution of doxorubicin.256 According to Leiva et al., in vitro anti-cancer activity was considerably improved by paclitaxel-loaded tripalmitin SLN formulation as it facilitated prolonged or sustained release of drug and reduced IC50 of paclitaxel which resulted in greater cytotoxicity in lung and breast cancer tissues.266
Theranostic applicability, biocompatibility, and biodegradability of created nanomedicine have all been demonstrated in previous research studies, as well as the precise in vivo screening and drug administration utilizing SLNs. The findings showed that the created SLN theranostic nanoformulation has the potential to make a significant contribution in the field of oncology and diagnosis at the same time. Paclitaxel and siRNA were used as the anti-cancer agent. Paclitaxel and QDs (quantum dots) were distributed throughout the lipid core, whereas anionic siRNA was electrostatically coupled to the external cationic interface. In lung cancer, the integration of the dual therapeutic drug paclitaxel and siRNA loaded in SLN was efficiently deposited and showed synergistic anti-cancer action. Significantly, QD fluorescence in SLN allowed researchers to measure increased in vivo cellular uptake of the drug on-site while reducing off-site cancer cell uptake and thus its theranostic action was assessed.251,253
4.10. Hydrogels
Hydrogels are polymeric three-dimensional mesh, which has the property of retaining a significant amount of water within their fibers.267 It is a binary polymer solution system which has a cross-linker. Hydrogels have gained much attention from the last two to three decades because they possess various unique characteristics which are useful in biomedical applications. P-HEMA, or poly(2-hydroxyethyl methacrylate), was the earliest designed hydrogel to be produced by DuPont scientists in 1936. But it was not until 1960 that Wichterle and Lim realized the relevance of P-HEMA hydrogels as good candidates for contact focal point purposes.268 Based on their cross-linking process, hydrogels may be divided into physical and chemical hydrogels using different production procedures.269,270 Cross-linking of hydrogels could be done either chemically (ionic, covalent, and atomic) or physically (hydrogen bonding). Chemical-based approaches, such as thermoresponsive, Michael addition, self-assembling hydrogel, and pH-responsive strategies, are therefore recommended for creating hydrogels of various strengths and controlled drug release.
Hydrogels (sol–gel–sol) are generally functionalized with polymers (PEG-PCL-PEG/DDP, PECE/DDP) possessing a polymeric micellar combination of paclitaxel and cisplatin. At room temperature, hydrogels exist in solution form, but when it attains body temperature it gets transformed into a gel. The principal mechanism of this gel form by which it treats cancer, is to form a drug depot through which the medication keeps diffusing gradually. Cisplatin diffuses from the thermosensitive polymer and the gradual release of micellar paclitaxel occurs through the gel state to the nearby cancerous tissues. Hydrogel formulations inhibit tumorigenesis while also prolonging the organism’s lifespan. Lung cancer may be potentially treated with hydrogels having two-drug combinations.271
For the treatment of NSCLC, scientists have designed a hydrogel-based drug delivery system for intravenous administration.272 Targeting the cancerous tissues was achieved by two methods. The first stage is passive targeting, followed by IV treatment where GMPs (glycomacropeptides) selectively concentrate in the lungs. Passive targeting yields a 10-times enhancement in anti-cancer medication efficiency with a ten-time decreased peak systemic drug levels, according to preliminary results. The second method is active targeting where two different types of nanoparticles are suggested. Initially, ligands on the NP surfaces are designed for targeting cancerous nodules specifically. The secondary NP group is likewise incorporated with cell surface ligands, but instead of selectively delivering drug contents within the cancer cell, these NPs are intended to firmly attach to cancer cell surface receptors and persist there to block the metastatic signaling cascade. The NPs contained in the GMPs disperse out and approach cancerous cells, once the GMPs have accumulated passively in the lung, resulting in a high level of targeted selectivity.273
Following cellular absorption of nanocarriers, expansion of the polymeric nanoparticles is observed in moderately acidic medium because it has evolved to a hydrophilic structure from a hydrophobic structure. This produces nanosized hydrogels, which exhibit intracellular activity and drug depot traits. These nanoparticles are promptly absorbed by A549 cell lines (88.3 ± 0.8% uptake in 24 h), which consequently, expands for releasing the entrapped moieties, and may efficiently deliver the chemotherapeutics in vitro. Paclitaxel, an antitumor drug, has been found to have a 3- to 4-fold affinity for enlarged nanoparticles, permitting them to function as intracellular drug depots which facilitates aggregation of the drug locally.275
Incorporating SWCNTs with peptide hydrogels may widen tissue engineering applications and may provide hydrogels a broader application-oriented perspective for cell biomaterials. EFK8 is a self-assembling peptide with enhanced mechanical properties. Since phenylalanine is prevalent, it confers stronger hydrophobic interactions. The inclusion of SWCNTs to the EFK8 hydrogel (while maintaining its modulus) yields higher motility in cells comparable to firmer hydrogels, emphasizing the relevance of cell scaffold interactions in metastasis.276
In A549 cell line, Khatun et al.274 reported versatile cancer theranostic uses of graphene-doxorubicin conjugated in hyaluronic acid nanogel. This nanocomposite has been employed for thermal chemotherapy applications, as well as real-time and non-invasive optical imaging and controlled drug release.274 Carbon dots (CDs) are predominantly used in theranostics and are implanted into the chitosan-containing hydrogel, where the hydrogel serves as a matrix for embedding CDs via non-covalent conjugation. Elimination of cancerous cells is also facilitated by the anti-cancer drug, 5-fluorouracil (5-FU) which is implanted in the hydrogel through hydrophobic interactions and further integrated with CDs to form 5FU@CDs-HY, which has anti-cancer action. The moiety also has properties of a screening agent and is employed for visualization during cellular uptake in cancer cells.277 The silk nanoparticles of cisplatin were developed by Kim et al. Spray-drying was used to create a composition including silk fibers and cisplatin with or without cross-linkage. Different release profiles result from the presence or absence of a cross-linking agent in the formulation. Without the use of a cross-linker, cisplatin silk nanoparticles showed significant cytotoxicity. The addition of acid and/or ions cause(s) hydrogel sol–gel transitions in silk fibroin nanoparticles.278,279 Silk strands are more durable and help to keep nanoparticles steady throughout the storage period and in vivo investigations. Therefore, anti-cancer medicines are delivered to the lungs using silk fiber nanoparticles, which have greater cytotoxicity and bioavailability.
The anti-cancer drug docetaxel (DTX) and the triblock copolymer hydrogel are employed in the formulation of PLGA-PEG-PLGA triblock copolymer, which does not influence the solubility of DTX but reduces its toxicity concerns. On A-549 lung tumor-bearing BALB/cA mice, a triblock copolymer hydrogel formulation was assessed, and it was found to improve the pharmacokinetic profile over 3 weeks in a sustained release pattern.280
4.11. Quantum Dots (QDs)
QDs are nanosized colloidal particles with atom-like characteristics that have recently been synthesized via nanofabrication (Figure 10).281 These QDs are thought to be a potential treatment option for lung cancer. Modifying the surface of QDs makes it more soluble and biocompatible which converts it to a superior fluorescent probe.282 The core of the material is composed of a crystalline metalloid which is shielded by a cap or shell which enhances the overall bioavailability.283
Figure 10.

Structural representation of quantum dots.
Large absorption spectra, high photobleaching, and photostability are few of the unique properties exhibited by QDs.284 Because of the restricted wavelength of the emission spectrum, this nanoparticle system displays redundant fluorescence and excitation cycles.285 These characteristics have made QDs a trendy topic, particularly for clinical imaging of diseased tissue specimens from patients who had cancer, for identifying the type and phase of disease, directing therapy regimen, and predicting prognosis.286,293 QDs might potentially be utilized to investigate cell differentiation and development throughout embryogenesis.287 In general, QDs are made up of elements from groups II–VI/III–V. Cadmium selenide (CdSe), cadmium telluride (CdTe), zinc sulfide (ZnS), and zinc selenide (ZnSe) are among the elements in groups II–IV. Gallium nitride (GaN), gallium arsenide (GaAs), indium phosphide (InP) and indium arsenide (InAs) are among the elements in group IIIV.288
When using QDs in biological research investigations, solubility in aqueous medium is a frequent concern for various forms of QDs. Apart from directly synthesizing QDs in aqueous solutions, layering with polymer and exchanging the ligands have efficiently overcome solubility issues.289 Lipid-coated QDs demonstrate promising outcomes toward tumor cells.291 An amalgamated treatment strategy was studied and constructed by incorporating several anti-cancer agents (paclitaxel, doxorubicin, and carboplatin) and Bcl-2-targeted siRNA into a novel QD nanocarrier. The findings suggested that the QD-based delivery method resulted in 3- to 4-fold improvement in cytotoxicity for A549 cell lines when compared to free drug therapy. Furthermore, effectiveness of the regimen was improved, revealing that QD-based multifunctional delivery systems could be successfully employed in treating lung cancer when used in combination therapy.
Quantum dots are noted for their high fluorescence emission, which might be valuable for diagnostic and theranostic applications; unfortunately, the heavy metals in their composition could cause severe toxicity, necessitating the employment of alternative imaging strategies. Since tumor cells tend to overexpress biotin, biotinylated graphene QDs might be a viable option for cancer theranostics. Aside from being less cytotoxic than heavy metal QDs, graphene QDs have ideal surface tenability characteristics that provide specific changes and may therefore be developed for a myriad of theranostic purposes.292 Wu and colleagues described carbon dots functionalized with positively charged polyethylenimine (PEI) and negatively charged siRNA molecules, as well as folate as the tumor-specific moiety for being applied as theranostic agent. Bioluminescence (the dots demonstrate absorbance and emission at 360 and 400 nm, respectively) could be used to track the accumulation of nanomaterials, and the authors were able to report a decline in lung tumor volume in vivo.290
4.12. Inhalational
The pulmonary route is a more significant, direct, and appealing delivery strategy when compared to conventional methodologies of drug delivery (oral or parenteral). The local approach of delivery of drugs with the help of nanocarriers is to avoid the first-pass metabolism or to overcome the need for injections. Direct therapeutic delivery to the lungs significantly contributes to the local treatment of asthma, cystic fibrosis, and COPD (chronic obstructive pulmonary disease) and advanced treatment of cancer and diabetes. Direct pulmonary targeted delivery has more absorption efficiency compared to conventional administration, and therefore reduces systemic toxicity. Local toxicity could occur due to the deposition of both drugs and excipients. Traditional inhalation methods were not able to achieve specific targets which necessitates the need of developing novel delivery methods.294 Furthermore, reduced thickness, high surface area, and intense vascularization of the alveoli makes the pulmonary delivery an excellent delivery approach. The prolonged release of drugs, strong tendency for association with drugs, and target specificity are some of the reasons scientists are attracted to explore inhalable nanoparticles. As they can withstand high nebulization forces, they can easily be formulated into aerosols. The duration of the administered drug in the respiratory tract could be enhanced by evading mucociliary transport and phagocytosis by lungs.295
Polymeric nanocarriers: Drug delivery through nanotechnology in the treatment of cancer has significantly progressed over the past decades. Polymeric nanocarriers play a vital role in inhalable drug and gene delivery. Doxorubicin released by isobutyl cyanoacrylate polymeric nanoparticles showed that polymeric nanocarrier caused cytotoxicity (secondary) triggered via alveolar phagocytes which provides a superior treatment for lung cancer. After initial phagocytosis, exocytosis of segments of nanoparticles occurs when alveolar phagocytes start imitating the action of Kupffer cells which causes the death of surrounding tumorous cells. Murine macrophages of alveoli were treated with nanocarriers and then were introduced to H460 cells.294 Free doxorubicin did not produce any secondary cytotoxic effects when it was introduced to macrophages, but in the presence of nanocarriers containing doxorubicin, macrophages demonstrated optimum cytotoxicity after 8 and 24 h. The effects of nanoparticle treatment resulted in the formation of interferon gamma and macrophage inflammatory protein which initiated cellular immunity and inflammation, thus augmenting tumor rejection.
Lipid Nanocarriers: Many in vivo and in vitro studies have revealed the therapeutic efficacy of 9-nitrocamptothecin (9NC) via an inhalable route instead of oral or i.v. or i.m. routes for treating lung cancer. Formulation of 9NC conjugated liposomes with dilauroylphosphatidylcholine [MMAD (mass-mean aerodynamic diameter) 1.2–1.6 μm] inhibited B16 melanoma pulmonary metastases in mice after 16–21 days of treatment. Interestingly, the quantity of inhaled dose of 9NC was 3 to 20 times lower when compared with intragastric, intravenous or intramuscular dosage forms but showed a similar effect.296 Preparation of solid lipid nanocarrier by amalgamating glycerol and various esters of behenic acid showed greater stability during nebulization time with better inhalable fraction instead of epirubicin (EPI) solution. Inhalable EPI-solid lipid nanocarriers have a greater concentration of drug within the cancer-affected area. In another study with solid lipid nanocarriers, inhalable paclitaxel-solid lipid nanocarriers showed better effect and lesser side effects compared to i.v. formulation.297 Inhalable celecoxib entrapped lipid nanocarriers demonstrated controlled release of celecoxib. In comparison with celecoxib solution, nebulized celecoxib-encapsulated nanostructure showed 4 times better retention time in the lungs.298 Nebulizing the drugs enhances the homogeneity and stability of lipid nanocapsules. Paclitaxel–lipid nanocapsules and paclitaxel showed similar cytotoxic effects against lung cancer but lipid nanocapsules influence the stability of the drug in the lungs.299
Hybrid Lipid–Polymer Nanoparticles: The innermost core of poly glutamic acid provides viscosity to lipid polymer nanoparticles, which leads to prolonged release of 5-FU when compared to polymeric microspheres and liposomes. Drug release rate of 5-FU is affected through diameter of nanoparticle and thickness of lipid shell.300 Enhanced release rate of 5-FU was observed (70% to 85%) when the thickness of the shell was reduced from 300 to 100 nm which also influenced the overall size reduction from 1000 to 600 nm. Drug disintegration was affected by the thickness of the shell because it regulated the amount of water which could enter inside the LNPs (lipid NPs). The initial 5-FU concentration in the lungs was 1000 times higher than in the systemic circulation after inhaling LNPs, indicating a larger deposition fraction. Another study was conducted where Paclitaxel-loaded hybrid PEG5000–1,2-distearoyl-phosphatidylethanolamine (PEG5000-DSPE) micelles were administered to deeper alveolar tissues via nebulization.301 The core moiety of micelles was secured from phagocytosis by PEG chain through macrophages which led to its increased residence in lung tissues. When inhalable micelles conjugated with paclitaxel were administered, 45 times greater AUC (area under the curve) was observed in lungs of rat as compared to administration through intravenous route, which further resulted in significant enhancement in the concentration of drug in the kidneys, liver, and spleen (because of instant absorption through RES macrophages). Overall prolonged release of drug was observed from the core (12 h) with reduced enzymatic degradation of DSPE when compared to free drug which promptly enters the systemic pathway.
Inorganic Nanocarriers: Pulmonary application of metal NPs are limited because of their adverse effects of causing toxicity in lung tissues. Toxicity caused by metal NPs could be regulated by influencing the concentration of inhaled metal, particle size of aerosol, and time of exposure with lung tissues which could enhance their use as a theranostic agent (diagnosis and therapy). Employing nanomagnetic aerosols for targeted delivery of chemotherapeutic agents to various affected regions of lungs through “nanomagnetosols” has an impact on enhancing lung deposition of drugs at tumorous sites while decreasing the toxicity toward healthy tissues.302 Furthermore, concentration of SPIONs aerosols could be directed greatly toward the central airway instead of the entire lung by regulating the applied magnetic field. When plasmid DNA loaded SPIONs aerosol were administered to mice in the presence or absence of magnetic field, it was observed that the right lung which was influenced with application of magnetic field had a 2-times greater concentration of plasmid DNA as compared to the left lung where magnetic field was absent.303 In a study conducted to prevent tumorigenesis, PLGA coated Fe3O4 nanocarriers were loaded with aerosolized quercetin.304 Coating avoided agglomeration of particles by improving the size distribution of nanoparticles. In a culture media of cells from human lung epithelium, PLGA-coated MNPs did not show cytotoxicity. But PLGA-MNPs encapsulated with drugs demonstrated cytotoxicity by decreasing the quantity of viable A549 cancerous cell lines.
Actively Targeted Inhalable Nanoparticles: The particle size of inhaled aerosol is critical for precise targeting of tumorous cells which are located in various regions of the lung. Accumulation of aerosol particles with size of 5 μm–10 μm is observed in larger airways and oropharynx while particles with smaller size of 1 μm–5 μm deposits in alveoli and smaller airways.305 An 8- to 9-fold inclination was observed in the quantity of internalized drug, when administered via nanosized particles (100 nm–150 nm) as compared to microsized particles (3 μm–5 μm).306 Precisely controlled modification of aerosol particle size is required as it will directly affect the internalization of drug into tumorous tissues and accumulation of drug in lungs.57,307−309
5. Diagnostic Agents Used for Lung Cancer
The advantages of timely diagnosis of carcinoma have sparked a lot of scientific interest, since it significantly improves the rate of survival. Early diagnosis of tumors could aid in the total cure of the disease. As a result, the demand for methods that could identify pulmonary nodules in their earlier stages of growth has increased. Early diagnosis is related to a considerably improved survival pattern, highlighting the need for correct LC staging for reasonable treatment regimen selection for maximum anti-cancer effectiveness. Sputum cytology, chest radiography (CXR), bronchial biopsy, polymerase chain reaction (PCR), fluorescence bronchoscopy, and computed tomography (CT scan) are all widely applied techniques to diagnose LC and to determine the severity and extent of the disease (also known as staging). In the case of inadequate or restricted scope of different traditional diagnostic methods for early-stage tumor identification, NPs have plenty of promise in the clinical environment since they can precisely target the diseased region. In diagnostic radiography, the problem of “missing lung cancer” has also been a subject of concern. These missed detections in LC imaging result in the development of early-stage illness to advanced-stage disease.310 Nanotechnology is progressing toward the development of novel chemicals for cancer screening, staging, and treatment regimens, resulting in the intriguing potential for biomedical imaging.311 Particles at the nanoscale size have unique physical and chemical properties that allow detecting tools to be built with increased signal intensity, enumeration, magnification, contrast, and dispersion.312 Nanomedicines have therefore evolved progressively as a game-changing phenomenon for developing anti-cancer drugs. By regulating nanocarriers they may carry screening agents and chemotherapeutic agents for improved imaging and treatment effectiveness.313−315
Metallic and Non-metallic Nanocarriers: Noble metal nanoparticles have sparked interest because of their admirable thermal, electrical, and optical characteristics. To begin with, the core is devoid of toxicity, possesses high biocompatibility, and is inert, which makes it a great moiety to start when building a carrier.
Polymeric Nanocarriers: Because of its biocompatibility, biodegradability, noncytotoxicity, and variety of physicochemical characteristics, these have piqued importance in the field of oncology for screening as well as in therapeutic strategies.
Biologically Derived Nanocarriers: These nanocarriers have risen to prominence because of their related biocompatibility, simplicity of morphological changes, and notable tendency of fabricating the surface with multiple active drugs, therefore they are being used in medical procedures.
Table 5 summarizes the many types of nanodiagnostic devices employed for identifying early stages of lung cancer with pulmonary metastases.
Table 5. Nano-imaging Diagnostic Methodologies.
| strategy | formulation | diagnosis | observations | ref |
|---|---|---|---|---|
| Metallic and Non-metallic NPs | ||||
| IONPs | conjugated iron-oxide/silica nanocomposite | improvement in imaging | In the instance of FA-overexpressing lung cancer cells, improved imaging findings were revealed. | (316) |
| conjugated superparamagnetic iron oxide NPs | MRI | Through targeted administration and improved T2 relaxivity rates, peptide functionalization boosted lung cancer detection. | (317) | |
| immune SPIO-NPs | overexpression of CD44v6 receptors in metastasized lung cancer cells, leading to more accurate detection | Targeted delivery of anti-CD44v6 Ab-attached SPIO-NPs to CD44v6 overexpressing cancer cells improves contrast screening of metastasized lung cancer. | (318) | |
| QD/iron-oxide-based nanocomposite | αv-integrin overexpressing lung cancer cells: targeted administration and improved imaging | Targeted administration of iRGD-attached SPIO-NPs boosted optical and MR diagnostic findings in αv-integrin-overexpressing lung cancer cells. | (319) | |
| ultra-small conjugated SPIO-NPs | lung cancer screening with radionucleotide and optical imaging probe combined US-SPIO-NPs | The use of integrated MR-PET (positron emission tomography) optical imaging resulted in optimal imaging of lung cancers. | (320) | |
| functionalized NIR dye-loaded SPIO-NPs | boosted imaging | In the condition of FA-overexpressed lung cancer cells, improved optical imaging and contrast findings were observed. | (321) | |
| AuNPs | dendrimer-coupled AuNPs (Au DENPs) | X-ray-based CT scan | In vivo imaging of a lung tumor that is both precise and biocompatible. | (322−324) |
| aptamer-coupled AuNPs | adherence of aptamer | The presence of cancer cells was indicated by a change in color in the colorimetric assay. | (325) | |
| Au-conjugated triangular nanoplates | CT scan based on dual X-ray | Dual CT/PA (computed tomography/pulmonary angiography) screening aids in the determination of the most effective treatment period and gives real-time visual assistance. | (326) | |
| AuNP-conjugated electrochemical immunodetectors | enolase-alpha was discovered in patient blood serum | Elevated enolase-alpha levels were linked to a higher risk of lung cancer. | (327) | |
| MMP2-conjugated Au nanoclusters (matrix metalloproteinase 2) | fluorescence in near-IR region | Due to the significant buildup of clusters inside tumor cells by passive and active targeting, NIR fluorescence-based identification of lung tumors is effective. | (328) | |
| conjugated AuNP interaction | electrochemical probes with hydrophilic characteristics tested on tumor cell surfaces | Lung cancer cells are quantified by evaluating their contact angle and electrochemical characteristics. | (329) | |
| AuNPs and iodine NPs-based dual CT scan | X-ray-based CT scan | In vivo imaging of lung cancer and its accompanying vasculature. | (330) | |
| hollow Au nanospheres (HAuNs) | carcinoembryonicantigen (CEA) detected | CEA levels greater than 5 ng/mL were linked to a higher risk of lung cancer. | (331) | |
| silica-based NPs | nanospheres of SiO2 silver-based support shells | single miRNAs detected in lung cancer cells using fluorescent metal nanoshells as molecular diagnostic agents | Lung cancer incidence was confirmed by elevated photostability of NPs and increased emission intensity linked with the genetic signal. | (332) |
| SMART SiO2 nanoshells | targeted distribution of encapsulated cobalt-ferrite NPs | The overexpression of αvβ3-integrin, tenascin-C, and nucleolin in lung cancer cells resulted in improved optical imaging and contrast. | (333) | |
| carbon-based NPs | graphene-AgNP-conjugated 3D bio-nanosensor | CYFRA21-1 DNA biomarker detected | In preclinical and clinical conditions, the cancer-based nucleic acid may be detected with a sensitivity of 1 × 10–14 M. | (334) |
| SWCNT/AgNP-based nanohybrids | single miRNAs linked to lung cancer revealed | By detecting the related miRNA biomarker, the nanohybrids were able to accurately diagnose lung cancer cells. | (335) | |
| manganese oxide NPs | fabricated manganese dioxide (MnO2) NPs | MR contrast imaging applied for NSCLC diagnosis | In the instance of NSLC diagnosis, efficient T1/T2-weighted MR contrast imaging findings with increased transverse relaxivity were seen. | (336) |
| Mn3O4-based NPs | overexpressed MMP2 and LHRH-receptor | In metastatic lung cancer cells, targeted administration of MMP2-cleavable peptide and LHRH-targeting peptide-attached Mn3O4-NPs resulted in improved MR imaging outcomes. | (337) | |
| Polymeric NPs | ||||
| albumin NPs | hematoporphyrin-conjugated albumin NPs | diagnosis of metastatic lung cancer in vitro and in vivo | Based on the results of confocal microscopy and scintigraphic imaging, it was shown that cancer cells could be imaged effectively. | (338) |
| chitosan NPs | functionalized glycol-chitosan NPs | in vivo imaging | Time-dependent NIRF imaging demonstrated effective imaging of lung cancer by targeted uptake of Cy5.5-labeled gCNPs. | (339) |
| gelatin NPs | fabricated gelatin NPs | in vitro and in vivo diagnosis | Confocal microscopy and in vivo imaging provided effective imaging of lung cancer by targeted uptake of FITC-EGFR-labeled gelatin NPs. | (340, 341) |
| Biological Nanocarriers | ||||
| viral NPs | bioengineered T4-ViNPs | early lung adenocarcinoma detected | Owing to the rapid absorption of Alexa Fluor/Cy3-conjugated viral NPs by cells, the imaging findings are quite sharp. | (342) |
6. Drug Delivery Changes
Over the past decade, drug delivery strategies based on nanoparticle systems have seen tremendous growth and development. Nanocarriers are most likely three-dimensional constructs of multiple components arranged in the proper spatial arrangement for their operations. Minor modifications in the method or composition might have a deleterious impact on the complex superposition of components. Synthesizing adequately sized nanoparticles for delivering an efficacious payload of therapeutic agent with the capability of targeting the specific site has always made the technique troublesome. In the biological surroundings, various risk factors such as indistinct shape, improper size characterization, inadequate biocompatibility, and unsuitable surface topology are observed. Because of the variations in physical, biological, and chemical profile of the nanoparticles, developing an appropriate and relevant delivery system which could administer medication to the lungs is considered to be arduous. Numerous challenges have to be surmounted by these nanosized systems when they reach clinical and preclinical stages of drug discovery, which include systemic clearance, immunological reactivity, tendency to overcome biological barriers, and target specificity. Certain characteristics must be considered while developing an appropriate nanocarrier system to be used for clinical purposes: Absolute knowledge about critical components with their respective interactions, capability for exhibiting target-specific action by overcoming biological barriers, identifying significant features and their impact on performance, and capacity for replication within manufacturing duration, efficient production in sterile form, convenience for storage and administration and high stability.343 To achieve successful delivery of chemotherapeutics to the alveolar region several considerations must be met which include formation of definite sized particles for better tumor retention, eluding proteolysis, and overcoming mucus and epithelial absorption barriers (which may affect absorption of drug). The determination of physicochemical parameters is crucial to identify interaction between particles in a biological media, tendency to aggregate, adsorption of the protein surface of nanoparticles, and nanoparticle intracellular trafficking. If any significant change occurs in the parameters, it may affect the therapeutic efficiency (where insufficient drug is delivered) or may cause undesirable side effects which could be toxic. Intravenous injection of particles greater than 500 nm is not recommended since these particles are quickly removed from systemic circulation.344 Meticulously controlled steps should be carried out for fabrication of polymeric nanoparticles, to obtain a homogeneous size of less than 200 nm. Using standard extrusion techniques, the particle size composition could be modified.345 When in vivo studies are conducted, the fate of the nanocarrier is determined by surface charge. The zeta potential of nanoparticles has a significant impact on particle–particle interactions and aggregation tendencies. Positively charged nanoparticles have a stronger attraction to the body’s negatively charged cellular membranes. However, careful examination of these nanoparticle systems is required before testing in an in vivo system. In general, multi-functionalization enhances the complexity of the nanoparticles. Although metallic and inorganic nanoparticles have shown potential anti-cancer effects, they also possess the risk of causing damage to normal cells. For example, amorphous silica nanoparticles could trigger inflammatory responses in the targeted site.346 Titanium dioxide is biologically significant and is used in multiple biomedical purposes but its toxicity to the normal cells is a concerning issue which needs to be resolved.347 There’s a need to improve immunotherapy’s adaptability, productivity, and persistence to stimulate the anti-cancer immune system.348 To ensure safety, individual assessment of various nanocarrier conjugated systems should be performed (by in vivo and in vitro toxicity studies) before they are allowed for clinical use.
7. Future Perspectives
Novel drug delivery systems (NDDSs) are being explored for a broad array of applications. They have shown potential for enhancing the safety, effectiveness, and accuracy of therapeutic regimens. These systems surmount the fundamental concerns with conventional regimens, which include absence of targeted delivery, lower therapeutic efficiency, undesirable adverse effects, and drug resistivity, while also outperforming their predecessors in terms of early metastatic detection. The ability of nanoparticles to distribute medication exclusively to specific targeted tissues via peptide, PEGylation, or receptor ligand linkage delivers an option that creates an optimum local concentration of drug with lower dosages and lower cytotoxicity than topical administration of the free drug. Multiple investigations have reported that embedding chemotherapeutics in nanocarriers reduces cytotoxicity compared to administering them as free drug solutions. With the existing advancements in nano-based systems, they have been hailed as one of the potential strategies for the treatment of lung cancer. However, there are still several unmet needs that must be addressed before they can be formalized. Drug resistance can be caused by insufficient drug release from nanocarriers. Resistance offered by multiple drugs is likely to be resolved using techniques such as enhanced multifunctionalized nanocarriers for targeted drug delivery to specific tumor, endosomal disruption for immediate cytoplasmic release of drug, and providing combination treatment via multiple chemotherapeutics. Further, as we progress from microsized to nanosized particles, there is a decline in size but, inclination in surface area and particle number. Because nanoparticles have a larger surface area, they have more chemical reactivity, which makes it difficult to predict how they would react under varied situations. Nanoparticles have enhanced chemical reactivity which produces reactive oxygen species (ROS), that can trigger inflammatory responses, DNA deterioration, and oxidative stress, eventually causing cytotoxicity. However, only a few liposomes and nanoparticle-based compositions have received FDA approval so far. Liposomes and lipid-based nanoparticles are prone to oxidative breakdown, and have the tendency to accumulate, leading to reduced therapeutic efficacy and are therefore less stable when compared with other nanocarriers. Even though most lung malignancies are detected at an advanced stage, a suitable delivery approach via an appropriate nanocarrier which can release the drug to peripheral and central lung tissues, along with circulatory pathways, are needed for targeting local and distant metastatic malignant nodules. Additional advancements are needed to facilitate the regulated release of chemotherapeutic drugs through stimuli sensitive carriers. Cancer theranostics includes both diagnosis and target-specific therapy delivered through multi-purpose nanocarriers which include screening and chemotherapeutic chemicals, which might become crucial in the future to regulate cancer therapy. It has the potential for detecting cells with tumors and then eliminating them with minimal adverse effects, and provide instant in vivo screening technologies, and demonstrate treatment benefits. As a result, future studies should focus on identifying efficacious nanoparticles that are highly stable, have enhanced deagglomeration potential, and improved infiltrating efficacy. Varying the physical (mass median diameter and porosity) and chemical characteristics of nanoparticles (amphiphilic particles, target ligands, and stable polymers) might help preserve stability and prevent agglomeration during storage. To have a synergistic boosting impact on local and systemic distribution, a drug can be embedded into several nanoparticles, such as liposomes, dendrimers, and micelles.
In addition to the challenges of designing and synthesizing nanocarriers, researchers must also consider the lack of standards in the evaluation of nanomedicines, which involves manufacturing procedures, functional testing, and safety protocols. Therapies including nanocarriers must fundamentally circumvent identical challenges that a newly synthesized drug possesses, which includes: ideal composition and characteristics of compounds, repeatable manufacturing techniques, establishment of analytical techniques to adequately characterize the compound, preferable pharmacological evaluation with toxicity tests, and demonstration of safety and effectiveness in pre-clinical and clinical studies. A single active ingredient is generally found in standard medicines. On the other hand, various active compounds are embedded in nanoparticles which make them complicated. As a result of such a complexity, conventional pharmacokinetic, bioequivalence, and safety inspections must be updated. To strategically tackle the development of new nanoproducts obtained from novel methodologies with ease of drug delivery, regulatory authorities must establish an intensive set of tests with shortened clearance procedures. Substantial investigation in this sector is expected to seek new perspectives and replace the standard dose method with NDDSs in the future, therefore improving health care delivery.
8. Conclusion
Lung cancer is one of the most common cancers with a high mortality rate. Early identification and better treatment strategies are highly essential to elevate the efficiency of treatment of carcinoma. Theranostic agents use a distribution platform based on nanotechnology, to deliver both therapeutic and imaging compounds to the desired location of the body. Because of unique physical characteristics possessed by nanostructures, they frequently have the capacity to resolve solubility and stability problems of chemotherapeutic drugs by surface modification. Nanotechnology has limitless potential and novel applications are constantly being investigated as it is a discipline that has recently spawned novel methods for lung cancer treatment. Researchers have used several forms of nanocarriers to overcome the shortcomings of traditional treatment by improving active or passive target specificity and reducing cytotoxicity toward healthy cells. Aiming the regimen toward overexpressed targets greatly enhances lung cancer treatment results. The greater surface area of nanoparticles contributes to the larger therapeutic payload as well. Liposomes, SLNs, micelles, dendrimers, MNPs, CNTs, QDs, hydrogels, and nanoemulsions, polymeric NPs are some of the nanocarriers which are formulated to be utilized as theranostics. The quality of nanocarriers is significantly crucial for transferring the drug to the tumor site as it directly affects the infiltration of drug and accumulation of drug in the cancerous tissue. The capacity to identify and quantify cancer at its initial stage with higher resolution and predicting the response of cancer therapy will also depend on the theranostic agent. These significant advancements in lung cancer therapy, as well as the creation of different superior novel drug delivery systems, have helped to boost the morale of patients undergoing cancer treatment.
Acknowledgments
All the authors are thankful to their respective affiliated institutions for providing research facilities to carry out this work.
The authors declare no competing financial interest.
References
- Fitzmaurice C.; Dicker D.; Pain A.; Hamavid H.; Moradi-Lakeh M.; MacIntyre M. F.; Allen C.; Hansen G.; Woodbrook R.; Wolfe C.; et al. The global burden of cancer 2013. JAMA Oncol 2015, 1, 505–527. 10.1001/jamaoncol.2015.0735. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kumar S.; Sharma A. K.; Lalhlenmawia H.; Kumar D. Natural Compounds Targeting Major Signaling Pathways in Lung Cancer. Targeting Cellular Signalling Pathways in Lung Diseases 2021, 821–846. 10.1007/978-981-33-6827-9_37. [DOI] [Google Scholar]
- Litzky L. A. Pulmonary sarcomatous tumors. Arch. Pathol. Lab. Med. 2008, 132, 1104–1117. 10.5858/2008-132-1104-PST. [DOI] [PubMed] [Google Scholar]
- Zorzetto M.; Ferrari S.; Saracino L.; Inghilleri S.; Stella G. M. MET genetic lesions in non-small-cell lung cancer: pharmacological and clinical implications. Transl. Lung Cancer Res. 2012, 1, 194. 10.3978/j.issn.2218-6751.2012.09.03. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pikor L. A.; Ramnarine V. R.; Lam S.; Lam W. L. Genetic alterations defining NSCLC subtypes and their therapeutic implications. Lung cancer 2013, 82, 179–189. 10.1016/j.lungcan.2013.07.025. [DOI] [PubMed] [Google Scholar]
- Panagopoulos N.; Leivaditis V.; Koletsis E.; Prokakis C.; Alexopoulos P.; Baltayiannis N.; Hatzimichalis A.; Tsakiridis K.; Zarogoulidis P.; Zarogoulidis K.; et al. Pancoast tumors: characteristics and preoperative assessment. J. Thorac. Dis. 2014, 6, S108. 10.3978/j.issn.2072-1439.2013.12.29. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sher T.; Dy G. K.; Adjei A. A. Small cell lung cancer. Mayo Clin. Proc. 2008, 83, 355–367. 10.4065/83.3.355. [DOI] [PubMed] [Google Scholar]
- Field R. W.; Withers B. L. Occupational and environmental causes of lung cancer. Clin. Chest Med. 2012, 33, 681–703. 10.1016/j.ccm.2012.07.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dresler C. The changing epidemic of lung cancer and occupational and environmental risk factors. Thorac. Surg. Clin. 2013, 23, 113–122. 10.1016/j.thorsurg.2013.01.015. [DOI] [PubMed] [Google Scholar]
- Wogan G. N.; Hecht S. S.; Felton J. S.; Conney A. H.; Loeb L. A. Environmental and chemical carcinogenesis. Semin. Cancer Biol. 2004, 14, 473–486. 10.1016/j.semcancer.2004.06.010. [DOI] [PubMed] [Google Scholar]
- Pérez-Herrero E.; Fernández-Medarde A. Advanced targeted therapies in cancer: Drug nanocarriers, the future of chemotherapy. Eur. J. Pharm. Biopharm. 2015, 93, 52–79. 10.1016/j.ejpb.2015.03.018. [DOI] [PubMed] [Google Scholar]
- Jiang X. Q.; Mei X. D.; Feng D. Air pollution and chronic airway diseases: what should people know and do?. J. Thorac. Dis. 2016, 8, E31. 10.3978/j.issn.2072-1439.2015.11.50. [DOI] [PMC free article] [PubMed] [Google Scholar]
- https://www.who.int/news-room/fact-sheets/detail/cancer
- Huang C. Y.; Ju D. T.; Chang C. F.; Reddy P. M.; Velmurugan B. K. A review on the effects of current chemotherapy drugs and natural agents in treating non-small cell lung cancer. Biomedicine 2017, 7, 23. 10.1051/bmdcn/2017070423. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Regassa H.; Sourirajan A.; Kumar V.; Pandey S.; Kumar D.; Dev K. A review of medicinal plants of the Himalayas with anti-proliferative activity for the treatment of various cancers. Cancers 2022, 14, 3898. 10.3390/cancers14163898. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hirsch F. R.; Suda K.; Wiens J.; Bunn P. A. Jr New and emerging targeted treatments in advanced non-small-cell lung cancer. Lancet Oncol 2016, 388, 1012–1024. 10.1016/S0140-6736(16)31473-8. [DOI] [PubMed] [Google Scholar]
- Yadav R.; Das J.; Lalhlenmawia H.; Tonk R. K.; Singh L.; Kumar D.. Targeting cancer using phytoconstituents-based drug delivery. Advanced Drug Delivery Systems in the Management of Cancer; Academic Press, 2021; pp 499–508. [Google Scholar]
- Liauw S. L.; Connell P. P.; Weichselbaum R. R. New paradigms and future challenges in radiation oncology: an update of biological targets and technology. Sci. Transl. Med. 2013, 5, 173sr2. 10.1126/scitranslmed.3005148. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Smith R. E. Trends in recommendations for myelosuppressive chemotherapy for the treatment of solid tumors. J. Natl. Compr. Cancer Network 2006, 4, 649–658. 10.6004/jnccn.2006.0056. [DOI] [PubMed] [Google Scholar]
- Nguyen K. T. Targeted nanoparticles for cancer therapy: promises and challenges. J. Nanomed. Nanotechnol. 2011, 2, 103e. 10.4172/2157-7439.1000103e. [DOI] [Google Scholar]
- Callaghan R.; Luk F.; Bebawy M. Inhibition of the multidrug resistance P-glycoprotein: time for a change of strategy?. Drug Metab. Dispos. 2014, 42, 623–31. 10.1124/dmd.113.056176. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sanvicens N.; Marco M. P. Multifunctional nanoparticles - properties and prospects for their use in human medicine. Trends Biotechnol 2008, 26, 425–33. 10.1016/j.tibtech.2008.04.005. [DOI] [PubMed] [Google Scholar]
- Prokop A.; Davidson J. M. Nanovehicular intracellular delivery systems. J. Pharm. Sci. Exp. Pharmacol. 2008, 97, 3518–90. 10.1002/jps.21270. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dua K.; Shukla S. D.; de Jesus Andreoli Pinto T.; Hansbro P. M. Nanotechnology: advancing the translational respiratory research. Interv. Med. Appl. Sci. 2017, 9, 39–41. 10.1556/1646.9.2017.1.02. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Singh R.; Lillard J. W. Nanoparticle-based targeted drug delivery. Exp. Mol. Pathol. 2009, 86, 215–23. 10.1016/j.yexmp.2008.12.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jiménez A.; Peltzer M.; Ruseckaite R.. Poly (lactic acid) science and technology: processing, properties, additives and applications; Royal Society of Chemistry, 2014. [Google Scholar]
- Elsabahy M.; Heo G. S.; Lim S. M.; Sun G.; Wooley K. L. Polymeric nanostructures for imaging and therapy. Chem. Rev. 2015, 115, 10967–1011. 10.1021/acs.chemrev.5b00135. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Charrueau C.; Zandanel C. Drug delivery by polymer nanoparticles: the challenge of controlled release and evaluation. Polymer Nanoparticles for Nanomedicines 2016, 439–503. 10.1007/978-3-319-41421-8_14. [DOI] [Google Scholar]
- Devi J.; Kumar S.; Kumar D.; Jindal D. K.; Poornachandra Y. Synthesis, characterization, in vitro antimicrobial and cytotoxic evaluation of Co (II), Ni (II), Cu (II) and Zn (II) complexes derived from bidentate hydrazones. Res. Chem. Intermed. 2022, 48, 423–455. 10.1007/s11164-021-04602-8. [DOI] [Google Scholar]
- Bisht D.; Kumar D.; Kumar D.; Dua K.; Chellappan D. K. Phytochemistry and pharmacological activity of the genus artemisia. Arch. Pharmacal Res. 2021, 44, 439–474. 10.1007/s12272-021-01328-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kumar D.; Harshavardhan S. J.; Chirumarry S.; Poornachandra Y.; Jang K.; Ganesh Kumar C.; Yoon Y. J.; Zhao B. X.; Miao J. Y.; et al. Design, Synthesis In Vitro Anticancer Activity and Docking Studies of (−)-Catechin Derivatives. Bull. Korean Chem. Soc. 2015, 36, 564–570. 10.1002/bkcs.10108. [DOI] [Google Scholar]
- Devi J.; Yadav J.; Lal K.; Kumar N.; Paul A. K.; Kumar D.; Dutta P. P.; Jindal D. K. Design, synthesis, crystal structure, molecular docking studies of some diorganotin (IV) complexes derived from the piperonylic hydrazide Schiff base ligands as cytotoxic agents. J. Mol. Struct. 2021, 1232, 129992. 10.1016/j.molstruc.2021.129992. [DOI] [Google Scholar]
- Singh G.; Kumar D.; Dash A. K. Pinus gerardiana Wallichex. D. Don. -A review. Phytomedicine Plus 2021, 1, 100024. 10.1016/j.phyplu.2021.100024. [DOI] [Google Scholar]
- Attri A.; Thakur D.; Kaur T.; Sensale S.; Peng Z.; Kumar D.; Singh R. P. Nanoparticles Incorporating a Fluorescence Turn-on Reporter for Real-Time Drug Release Monitoring, a Chemoenhancer and a Stealth Agent: Poseidon’s Trident against Cancer?. Mol. Pharmaceutics 2021, 18, 124–147. 10.1021/acs.molpharmaceut.0c00730. [DOI] [PubMed] [Google Scholar]
- Hinohara K.; Polyak K. Intratumoral heterogeneity: more than just mutations. Trends Cell Biol. 2019, 29, 569–579. 10.1016/j.tcb.2019.03.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sung J. C.; Pulliam B. L.; Edwards D. A. Nanoparticles for drug delivery to the lungs. Trends Biotechnol 2007, 25, 563–570. 10.1016/j.tibtech.2007.09.005. [DOI] [PubMed] [Google Scholar]
- Bethune G.; Bethune D.; Ridgway N.; Xu Z. Epidermal growth factor receptor (EGFR) in lung cancer: an overview and update. J. Thorac. Dis. 2010, 2, 48. [PMC free article] [PubMed] [Google Scholar]
- Inamura K.; Ninomiya H.; Ishikawa Y.; Matsubara O. Is the epidermal growth factor receptor status in lung cancers reflected in clinicopathologic features?. Arch. Pathol. Lab. Med. 2010, 134, 66–72. 10.5858/2008-0586-RAR1.1. [DOI] [PubMed] [Google Scholar]
- Agustoni F.; Suda K.; Yu H.; Ren S.; Rivard C. J.; Ellison K.; Caldwell C. Jr; Rozeboom L.; Brovsky K.; Hirsch F. R. EGFR-directed monoclonal antibodies in combination with chemotherapy for treatment of non-small-cell lung cancer: an updated review of clinical trials and new perspectives in biomarkers analysis. Cancer Treat. Rev. 2019, 72, 15–27. 10.1016/j.ctrv.2018.08.002. [DOI] [PubMed] [Google Scholar]
- Gupta R.; Dastane A. M.; Forozan F.; Riley-Portuguez A.; Chung F.; Lopategui J.; Marchevsky A. M. Evaluation of EGFR abnormalities in patients with pulmonary adenocarcinoma: the need to test neoplasms with more than one method. Mod. Pathol. 2009, 22, 128–133. 10.1038/modpathol.2008.182. [DOI] [PubMed] [Google Scholar]
- Pao W.; Chmielecki J. Rational, biologically based treatment of EGFR-mutant non-small-cell lung cancer. Nat. Rev. Cancer 2010, 10, 760–774. 10.1038/nrc2947. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mitsudomi T.; Kosaka T.; Yatabe Y. Biological and clinical implications of EGFR mutations in lung cancer. Int. J. Clin. Oncol. 2006, 11, 190–198. 10.1007/s10147-006-0583-4. [DOI] [PubMed] [Google Scholar]
- Waters M. J. The growth hormone receptor. Growth Hormone & IGF Res. 2016, 28, 6–10. 10.1016/j.ghir.2015.06.001. [DOI] [PubMed] [Google Scholar]
- Cheng C. Y.; Perevedentseva E.; Tu J. S.; Chung P. H.; Cheng C. L.; Liu K. K.; Chao J. I.; Chen P. H.; Chang C. C. Direct and in vitro observation of growth hormone receptor molecules in A549 human lung epithelial cells by nanodiamond labeling. Appl. Phys. Lett. 2007, 90, 163903. 10.1063/1.2727557. [DOI] [Google Scholar]
- Matherly L. H.; Goldman D. L. Membrane transport of folates. Vitam. Horm 2003, 66, 405–457. 10.1016/s0083-6729(03)01012-4. [DOI] [PubMed] [Google Scholar]
- Elnakat H.; Ratnam M. Distribution, functionality and gene regulation of folate receptor isoforms: implications in targeted therapy. Adv. Drug Delivery Rev. 2004, 56, 1067–1084. 10.1016/j.addr.2004.01.001. [DOI] [PubMed] [Google Scholar]
- Salazar M. D. A.; Ratnam M. The folate receptor: what does it promise in tissue-targeted therapeutics?. Cancer Metastasis Rev. 2007, 26, 141–152. 10.1007/s10555-007-9048-0. [DOI] [PubMed] [Google Scholar]
- Kelemen L. E. The role of folate receptor α in cancer development, progression and treatment: cause, consequence or innocent bystander?. Int. J. Cancer 2006, 119, 243–250. 10.1002/ijc.21712. [DOI] [PubMed] [Google Scholar]
- Shi H.; Guo J.; Li C.; Wang Z. A current review of folate receptor alpha as a potential tumor target in non-small-cell lung cancer. Drug Des., Dev. Ther. 2015, 9, 4989. 10.2147/DDDT.S90670. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Boogerd L. S.; Boonstra M. C.; Beck A. J.; Charehbili A.; Hoogstins C. E.; Prevoo H. A.; Singhal S.; Low P. S.; van de Velde C. J.; Vahrmeijer A. L. Concordance of folate receptor-α expression between biopsy, primary tumor and metastasis in breast cancer and lung cancer patients. Oncotarget 2016, 7, 17442. 10.18632/oncotarget.7856. [DOI] [PMC free article] [PubMed] [Google Scholar]
- O’Shannessy D. J.; Somers E. B.; Maltzman J.; Smale R.; Fu Y. S. Folate receptor alpha (FRA) expression in breast cancer: identification of a new molecular subtype and association with triple negative disease. Springerplus 2012, 1, 22. 10.1186/2193-1801-1-22. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cagle P. T.; Zhai Q. J.; Murphy L.; Low P. S. Folate receptor in adenocarcinoma and squamous cell carcinoma of the lung: potential target for folate-linked therapeutic agents. Arch. Pathol. Lab. Med. 2013, 137, 241–244. 10.5858/arpa.2012-0176-OA. [DOI] [PubMed] [Google Scholar]
- O’Shannessy D. J.; Yu G.; Smale R.; Fu Y. S.; Singhal S.; Thiel R. P.; Somers E. B.; Vachani A. Folate receptor alpha expression in lung cancer: diagnostic and prognostic significance. Oncotarget 2012, 3, 414. 10.18632/oncotarget.519. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Christoph D. C.; Asuncion B. R.; Hassan B.; Tran C.; Maltzman J. D.; O’Shannessy D. J.; Wynes M. W.; Gauler T. C.; Wohlschlaeger J.; Hoiczyk M.; et al. Significance of folate receptor alpha and thymidylate synthase protein expression in patients with non–small-cell lung Cancer treated with Pemetrexed. J. Thorac. Oncol. 2013, 8, 19–30. 10.1097/JTO.0b013e31827628ff. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Holmes K.; Roberts O. L.; Thomas A. M.; Cross M. J. Vascular endothelial growth factor receptor-2: structure, function, intracellular signalling and therapeutic inhibition. Cell. Signalling 2007, 19, 2003–2012. 10.1016/j.cellsig.2007.05.013. [DOI] [PubMed] [Google Scholar]
- Blakely C.; Jahan T. Emerging antiangiogenic therapies for non-small-cell lung cancer. Expert Rev. Anticancer Ther 2011, 11, 1607–1618. 10.1586/era.11.146. [DOI] [PubMed] [Google Scholar]
- Stuttfeld E.; Ballmer-Hofer K. Structure and function of VEGF receptors. IUBMB Life 2009, 61, 915–922. 10.1002/iub.234. [DOI] [PubMed] [Google Scholar]
- Devery A. M.; Wadekar R.; Bokobza S. M.; Weber A. M.; Jiang Y.; Ryan A. J. Vascular endothelial growth factor directly stimulates tumour cell proliferation in non-small cell lung cancer. Int. J. Oncol. 2015, 47, 849–856. 10.3892/ijo.2015.3082. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shibuya M. Vascular endothelial growth factor (VEGF) and its receptor (VEGFR) signaling in angiogenesis: a crucial target for anti-and pro-angiogenic therapies. Genes Cancer 2011, 2, 1097–1105. 10.1177/1947601911423031. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kuzmov A.; Minko T. Nanotechnology approaches for inhalation treatment of lung diseases. J. Controlled Release 2015, 219, 500–518. 10.1016/j.jconrel.2015.07.024. [DOI] [PubMed] [Google Scholar]
- Li X.; Taratula O.; Taratula O.; Schumann C.; Minko T. LHRH-targeted drug delivery systems for cancer therapy. Mini-Rev. Med. Chem. 2017, 17, 258–267. 10.2174/1389557516666161013111155. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ahmad I.; Iwata T.; Leung H. Y. Mechanisms of FGFR-mediated carcinogenesis. Biochim. Biophys. Acta, Mol. Cell Res. 2012, 1823, 850–860. 10.1016/j.bbamcr.2012.01.004. [DOI] [PubMed] [Google Scholar]
- Brooks A. N.; Kilgour E.; Smith P. D. Molecular Pathways: Fibroblast Growth Factor Signaling: A New Therapeutic Opportunity in Cancer FGF/FGFR Signaling in Cancer. Clin. Cancer Res. 2012, 18, 1855–1862. 10.1158/1078-0432.CCR-11-0699. [DOI] [PubMed] [Google Scholar]
- Dieci M. V.; Arnedos M.; Andre F.; Soria J. C. Fibroblast growth factor receptor inhibitors as a cancer treatment: from a biologic rationale to medical perspectives. Cancer Discovery 2013, 3, 264–279. 10.1158/2159-8290.CD-12-0362. [DOI] [PubMed] [Google Scholar]
- Johnson D. E.; Williams L. T. Structural and functional diversity in the FGF receptor multigene family. Adv. Cancer Res. 1992, 60, 1–41. 10.1016/S0065-230X(08)60821-0. [DOI] [PubMed] [Google Scholar]
- Dutt A.; Ramos A. H.; Hammerman P. S.; Mermel C.; Cho J.; Sharifnia T.; Chande A.; Tanaka K. E.; Stransky N.; Greulich H.; et al. Inhibitor-sensitive FGFR1 amplification in human non-small cell lung cancer. PloS One 2011, 6, 20351. 10.1371/journal.pone.0020351. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Templeton A. K.; Miyamoto S.; Babu A.; Munshi A.; Ramesh R. Cancer stem cells: progress and challenges in lung cancer. Stem Cell Invest 2014, 1, 9. 10.3978/j.issn.2306-9759.2014.03.06. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Penno M. B.; August J. T.; Baylin S. B.; Mabry M.; Linnoila R. I.; Lee V. S.; Croteau D.; Yang X. L.; Rosada C. Expression of CD44 in human lung tumors. Cancer Res. 1994, 54, 1381–1387. [PubMed] [Google Scholar]
- Miyoshi T.; Kondo K.; Hino N.; Uyama T.; Monden Y. The expression of the CD44 variant exon 6 is associated with lymph node metastasis in non-small cell lung cancer. Clin. Cancer Res. 1997, 3, 1289–1297. [PubMed] [Google Scholar]
- Wang F. L.; Wei L. X. Expression of CD44 variant exon 6 in lung cancers. Zhongguo yi xue ke xue Yuan xue bao. Acta Acad. Med. Sin. 2001, 23, 401–402. [PubMed] [Google Scholar]
- Zhao S.; He J. L.; Qiu Z. X.; Chen N. Y.; Luo Z.; Chen B. J.; Li W. M. Prognostic value of CD44 variant exon 6 expression in non-small cell lung cancer: a meta-analysis. Asian Pac. J. Cancer Prev. 2014, 15, 6761–6766. 10.7314/APJCP.2014.15.16.6761. [DOI] [PubMed] [Google Scholar]
- Hynes R. O. Integrins: bidirectional, allosteric signaling machines. Cell 2002, 110, 673–687. 10.1016/S0092-8674(02)00971-6. [DOI] [PubMed] [Google Scholar]
- Barczyk M.; Carracedo S.; Gullberg D. Integrins. Cell Tissue Res. 2010, 339, 269–280. 10.1007/s00441-009-0834-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bartolazzi A.; Cerboni C.; Flamini G.; Bigotti A.; Lauriola L.; Natali P. G. Expression of α3β1 integrin receptor and its ligands in human lung tumors. Int. J. Cancer 1995, 64, 248–252. 10.1002/ijc.2910640407. [DOI] [PubMed] [Google Scholar]
- Barr L. F.; Campbell S. E.; Bochner B. S.; Dang C. V. Association of the decreased expression of α3β1 integrin with the altered cell: environmental interactions and enhanced soft agar cloning ability of c-myc-overexpressing small cell lung cancer cells. Cancer Res. 1998, 58, 5537–5545. [PubMed] [Google Scholar]
- Rankin E. B.; Giaccia A. J. The receptor tyrosine kinase AXL in cancer progression. Cancers 2016, 8, 103. 10.3390/cancers8110103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Verma A.; Warner S. L.; Vankayalapati H.; Bearss D. J.; Sharma S. Targeting Axl and Mer Kinases in CancerAxl and Mer in Cancer. Mol. Cancer Ther. 2011, 10, 1763–1773. 10.1158/1535-7163.MCT-11-0116. [DOI] [PubMed] [Google Scholar]
- Graham D. K.; Bowman G. W.; Dawson T. L.; Stanford W. L.; Earp H. S.; Snodgrass H. R. Cloning and developmental expression analysis of the murine c-mer tyrosine kinase. Oncogene 1995, 10, 2349–2360. [PubMed] [Google Scholar]
- Schoumacher M.; Burbridge M. Key roles of AXL and MER receptor tyrosine kinases in resistance to multiple anticancer therapies. Curr. Oncol. Rep. 2017, 19, 19. 10.1007/s11912-017-0579-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Guillon A.; Gueugnon F.; Mavridis K.; Dalloneau E.; Jouan Y.; Diot P.; Heuzé-Vourc’h N.; Courty Y.; Si-Tahar M. Interleukin-22 receptor is overexpressed in non-small-cell lung cancer and portends a poor prognosis. Eur. Respir. J. 2016, 47, 1277–1280. 10.1183/13993003.01580-2015. [DOI] [PubMed] [Google Scholar]
- Sajjadi F. G.; Firestein G. S. cDNA cloning and sequence analysis of the human A3 adenosine receptor. Biochim. Biophys. Acta, Mol. Cell Res. 1993, 1179, 105–107. 10.1016/0167-4889(93)90077-3. [DOI] [PubMed] [Google Scholar]
- Salvatore C. A.; Jacobson M. A.; Taylor H. E.; Linden J.; Johnson R. G. Molecular cloning and characterization of the human A3 adenosine receptor. Proc. Natl. Acad. Sci. U. S. A. 1993, 90, 10365–10369. 10.1073/pnas.90.21.10365. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fishman P.; Bar-Yehuda S.; Synowitz M.; Powell J. D.; Klotz K. N.; Gessi S.; Borea P. A. Adenosine receptors and cancer. Adenosine Receptors in Health and Disease 2009, 193, 399–441. 10.1007/978-3-540-89615-9_14. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zlotnik A.; Burkhardt A. M.; Homey B. Homeostatic chemokine receptors and organ-specific metastasis. Nat. Rev. Immunol. 2011, 11, 597–606. 10.1038/nri3049. [DOI] [PubMed] [Google Scholar]
- Zlotnik A.; Yoshie O. Chemokines: a new classification system and their role in immunity. Immunity 2000, 12, 121–127. 10.1016/S1074-7613(00)80165-X. [DOI] [PubMed] [Google Scholar]
- Wu B.; Chien E. Y.; Mol C. D.; Fenalti G.; Liu W.; Katritch V.; Abagyan R.; Brooun A.; Wells P.; Bi F. C.; et al. Structures of the CXCR4 chemokine GPCR with small-molecule and cyclic peptide antagonists. Science 2010, 330, 1066–1071. 10.1126/science.1194396. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Saintigny P.; Burger J. A. Recent advances in non-small cell lung cancer biology and clinical management. Disc. Med. 2012, 13, 287–297. [PubMed] [Google Scholar]
- Wald O.; Izhar U.; Amir G.; Avniel S.; Bar-Shavit Y.; Wald H.; Weiss I. D.; Galun E.; Peled A. CD4+ CXCR4highCD69+ T cells accumulate in lung adenocarcinoma. J. Immunol. 2006, 177, 6983–6990. 10.4049/jimmunol.177.10.6983. [DOI] [PubMed] [Google Scholar]
- Moody T. W.; Sancho V.; Di Florio A.; Nuche-Berenguer B.; Mantey S.; Jensen R. T. Bombesin receptor subtype-3 agonists stimulate the growth of lung cancer cells and increase EGF receptor tyrosine phosphorylation. Peptides 2011, 32, 1677–1684. 10.1016/j.peptides.2011.06.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jensen R. T.; Battey J. F.; Spindel E. R.; Benya R. V. International Union of Pharmacology. LXVIII. Mammalian bombesin receptors: nomenclature, distribution, pharmacology, signaling, and functions in normal and disease states. Pharmacol. Rev. 2008, 60, 1–42. 10.1124/pr.107.07108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Moreno P.; Mantey S. A.; Lee S. H.; Ramos-Álvarez I.; Moody T. W.; Jensen R. T. A possible new target in lung-cancer cells: The orphan receptor, bombesin receptor subtype-3. Peptides 2018, 101, 213–226. 10.1016/j.peptides.2018.01.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Orlowski E.; Chand R.; Yip J.; Wong C.; Goschnick M. W.; Wright M. D.; Ashman L. K.; Jackson D. E. A platelet tetraspanin superfamily member, CD151, is required for regulation of thrombus growth and stability in vivo. J. Thromb. Haemostasis 2009, 7, 2074–2084. 10.1111/j.1538-7836.2009.03612.x. [DOI] [PubMed] [Google Scholar]
- Wang H. X.; Li Q.; Sharma C.; Knoblich K.; Hemler M. E. Tetraspanin protein contributions to cancer. Biochem. Soc. Trans. 2011, 39, 547–552. 10.1042/BST0390547. [DOI] [PubMed] [Google Scholar]
- Kwon M. J.; Seo J.; Kim Y. J.; Kwon M. J.; Choi J. Y.; Kim T. E.; Lee D. H.; Park S.; Shin Y. K.; Han J.; et al. Prognostic significance of CD151 overexpression in non-small cell lung cancer. Lung Cancer 2013, 81, 109–116. 10.1016/j.lungcan.2013.03.014. [DOI] [PubMed] [Google Scholar]
- John C. S.; Bowen W. D.; Varma V. M.; McAfee J. G.; Moody T. W. Sigma receptors are expressed in human non-small cell lung carcinoma. Life Sciences 1995, 56, 2385–2392. 10.1016/0024-3205(95)00232-U. [DOI] [PubMed] [Google Scholar]
- Moody T. W.; Leyton J.; John C. Sigma ligands inhibit the growth of small cell lung cancer cells. Life Sciences 2000, 66, 1979–1986. 10.1016/S0024-3205(00)00523-3. [DOI] [PubMed] [Google Scholar]
- Palmer R. H.; Vernersson E.; Grabbe C.; Hallberg B. Anaplastic lymphoma kinase: signalling in development and disease. Biochem. J. 2009, 420, 345–361. 10.1042/BJ20090387. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Van den Eynden J.; Umapathy G.; Ashouri A.; Cervantes-Madrid D.; Szydzik J.; Ruuth K.; Koster J.; Larsson E.; Guan J.; Palmer R. H.; et al. Phosphoproteome and gene expression profiling of ALK inhibition in neuroblastoma cell lines reveals conserved oncogenic pathways. Sci. Signaling 2018, 11, 5680. 10.1126/scisignal.aar5680. [DOI] [PubMed] [Google Scholar]
- Lin J. J.; Shaw A. T. Recent advances in targeting ROS1 in lung cancer. J. Thorac. Oncol. 2017, 12, 1611–1625. 10.1016/j.jtho.2017.08.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bergethon K.; Shaw A. T.; Ou S. H. I.; Katayama R.; Lovly C. M.; McDonald N. T.; Massion P. P.; Siwak-Tapp C.; Gonzalez A.; Fang R.; et al. ROS1 rearrangements define a unique molecular class of lung cancers. J. Clin. Oncol. 2012, 30, 863. 10.1200/JCO.2011.35.6345. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang M.; Gao S.; Yang D.; Fang Y.; Lin X.; Jin X.; Liu Y.; Liu X.; Su K.; Shi K. Influencing factors and strategies of enhancing nanoparticles into tumors in vivo. Acta Pharm. Sin. B 2021, 11, 2265–85. 10.1016/j.apsb.2021.03.033. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kim S. W.; Khang D. Multiple cues on the physiochemical, mesenchymal, and intracellular trafficking interactions with nanocarriers to maximize tumor target efficiency. Int. J. Nanomed. 2015, 10, 3989. 10.2147/IJN.S83951.. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shukla R.; Handa M.; Lokesh S. B.; Ruwali M.; Kohli K.; Kesharwani P. Conclusion and future prospective of polymeric nanoparticles for cancer therapy. Polymeric Nanoparticles as a Promising Tool for Anti-cancer Therapeutics 2019, 389–408. 10.1016/B978-0-12-816963-6.00018-2. [DOI] [Google Scholar]
- Choi J. S.; Park J. S. Surface modification of docetaxel nanocrystals with HER2 antibody to enhance cell growth inhibition in breast cancer cells. Colloids Surf., B 2017, 159, 139–150. 10.1016/j.colsurfb.2017.07.064. [DOI] [PubMed] [Google Scholar]
- Zhan H.; Jagtiani T.; Liang J. F. A new targeted delivery approach by functionalizing drug nanocrystals through polydopamine coating. Eur. J. Pharm. Biopharm. 2017, 114, 221–229. 10.1016/j.ejpb.2017.01.020. [DOI] [PubMed] [Google Scholar]
- Tung Y. T.; Chen H. L.; Yen C. C.; Lee P. Y.; Tsai H. C.; Lin M. F.; Chen C. M. Bovine lactoferrin inhibits lung cancer growth through suppression of both inflammation and expression of vascular endothelial growth factor. J. Dairy Sci. 2013, 96, 2095–2106. 10.3168/jds.2012-6153. [DOI] [PubMed] [Google Scholar]
- Taratula O.; Kuzmov A.; Shah M.; Garbuzenko O. B.; Minko T. Nanostructured lipid carriers as multifunctional nanomedicine platform for pulmonary co-delivery of anticancer drugs and siRNA. J. Controlled Release 2013, 171, 349–357. 10.1016/j.jconrel.2013.04.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Agrawal S.; Dwivedi M.; Ahmad H.; Chadchan S. B.; Arya A.; Sikandar R.; Kaushik S.; Mitra K.; Jha R. K.; Dwivedi A. K. CD44 targeting hyaluronic acid coated lapatinib nanocrystals foster the efficacy against triple-negative breast cancer. Nanomed.: Nanotechnol. Biol. Med. 2018, 14, 327–337. 10.1016/j.nano.2017.10.010. [DOI] [PubMed] [Google Scholar]
- Sohn J. S.; Yoon D. S.; Sohn J. Y.; Park J. S.; Choi J. S. Development and evaluation of targeting ligands surface modified paclitaxel nanocrystals. Mater. Sci. Eng. 2017, 72, 228–237. 10.1016/j.msec.2016.11.065. [DOI] [PubMed] [Google Scholar]
- Huang Z. G.; Lv F. M.; Wang J.; Cao S. J.; Liu Z. P.; Liu Y.; Lu W. Y. RGD-modified PEGylated paclitaxel nanocrystals with enhanced stability and tumor-targeting capability. Int. J. Pharm. 2019, 556, 217–225. 10.1016/j.ijpharm.2018.12.023. [DOI] [PubMed] [Google Scholar]
- Georgiadis M. O.; Karoutzou O.; Foscolos A. S.; Papanastasiou I. Sigma receptor (σR) ligands with antiproliferative and anticancer activity. Molecules 2017, 22, 1408. 10.3390/molecules22091408. [DOI] [PMC free article] [PubMed] [Google Scholar]
- van Waarde A.; Rybczynska A. A.; Ramakrishnan N. K.; Ishiwata K.; Elsinga P. H.; Dierckx R. A. Potential applications for sigma receptor ligands in cancer diagnosis and therapy. Biochim. Biophys. Acta, Biophys. Incl. Photosynth. 2015, 1848, 2703–2714. 10.1016/j.bbamem.2014.08.022. [DOI] [PubMed] [Google Scholar]
- Aydar E.; Palmer C. P.; Djamgoz M. B. Sigma receptors and cancer: possible involvement of ion channels. Cancer Res. 2004, 64, 5029–5035. 10.1158/0008-5472.CAN-03-2329. [DOI] [PubMed] [Google Scholar]
- Sehgal K.; Patell R.; Rangachari D.; Costa D. B. Targeting ROS1 rearrangements in non-small cell lung cancer with crizotinib and other kinase inhibitors. Transl. Cancer Res. 2018, 7, S779. 10.21037/tcr.2018.08.11. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mehta P.; Kadam S.; Pawar A.; Bothiraja C. Dendrimers for pulmonary delivery: current perspectives and future challenges. New J. Chem. 2019, 43, 8396–8409. 10.1039/C9NJ01591D. [DOI] [Google Scholar]
- Akamine T.; Toyokawa G.; Tagawa T.; Yamazaki K.; Seto T.; Takeo S.; Mori M. Lorlatinib for the treatment of patients with non-small cell lung cancer. Drugs Today (Barc) 2019, 55, 107–16. 10.1358/dot.2019.55.2.2927983. [DOI] [PubMed] [Google Scholar]
- Katayama R.; Gong B. O.; Togashi N.; Miyamoto M.; Kiga M.; Iwasaki S.; Kamai Y.; Tominaga Y.; Takeda Y.; Kagoshima Y.; et al. The new-generation selective ROS1/NTRK inhibitor DS-6051b overcomes crizotinib resistant ROS1-G2032R mutation in preclinical models. Nat. Commun. 2019, 10, 3604. 10.1038/s41467-019-11496-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Facchinetti F.; Friboulet L. Profile of entrectinib and its potential in the treatment of ROS1-positive NSCLC: evidence to date. Lung Cancer: Targets Ther 2019, 10, 87. 10.2147/LCTT.S190786. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Albanese A.; Tang P. S.; Chan W. C. The effect of nanoparticle size, shape, and surface chemistry on biological systems. Annu. Rev. Biomed. Eng. 2012, 14, 1–16. 10.1146/annurev-bioeng-071811-150124. [DOI] [PubMed] [Google Scholar]
- Attia M. F.; Anton N.; Wallyn J.; Omran Z.; Vandamme T. F. An overview of active and passive targeting strategies to improve the nanocarriers efficiency to tumour sites. J. Pharm. Pharmacol 2019, 71, 1185–1198. 10.1111/jphp.13098. [DOI] [PubMed] [Google Scholar]
- Huynh N. T.; Roger E.; Lautram N.; Benoît J. P.; Passirani C. The rise and rise of stealth nanocarriers for cancer therapy: passive versus active targeting. Nanomedicine 2010, 5, 1415–1433. 10.2217/nnm.10.113. [DOI] [PubMed] [Google Scholar]
- Lombardo D.; Kiselev M. A.; Caccamo M. T. Smart nanoparticles for drug delivery application: development of versatile nanocarrier platforms in biotechnology and nanomedicine. J. Nanomater. 2019, 2019, 1–26. 10.1155/2019/3702518. [DOI] [Google Scholar]
- Lu X. Y.; Wu D. C.; Li Z. J.; Cheng G. Q.. Polymer nanoparticles. In Progress in Molecular Biology and Translational Science; Villaverde A., Ed.; Academic Press, 2011; pp 299–323. [DOI] [PubMed] [Google Scholar]
- Casalini T.; Rossi F.; Castrovinci A.; Perale G. A perspective on polylactic acid-based polymers use for nanoparticles synthesis and applications. Front. Bioeng. Biotechnol 2019, 7, 259. 10.3389/fbioe.2019.00259. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lu Y.; Low P. S. Folate-mediated delivery of macromolecular anticancer therapeutic agents. Adv. Drug Delivery Rev. 2002, 54, 675–693. 10.1016/S0169-409X(02)00042-X. [DOI] [PubMed] [Google Scholar]
- Marasini N.; Haque S.; Kaminskas L. M. Polymer-drug conjugates as inhalable drug delivery systems: A review. Curr. Opin. Colloid Interface Sci. 2017, 31, 18–29. 10.1016/j.cocis.2017.06.003. [DOI] [Google Scholar]
- Harush-Frenkel O.; Debotton N.; Benita S.; Altschuler Y. Targeting of nanoparticles to the clathrin-mediated endocytic pathway. Biochem. Biophys. Res. Commun. 2007, 353, 26–32. 10.1016/j.bbrc.2006.11.135. [DOI] [PubMed] [Google Scholar]
- Vasir J. K.; Labhasetwar V. Quantification of the force of nanoparticle-cell membrane interactions and its influence on intracellular trafficking of nanoparticles. Biomaterials 2008, 29, 4244–4252. 10.1016/j.biomaterials.2008.07.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kohli A. G.; Kierstead P. H.; Venditto V. J.; Walsh C. L.; Szoka F. C. Designer lipids for drug delivery: from heads to tails. J. Controlled Release 2014, 190, 274–287. 10.1016/j.jconrel.2014.04.047. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Simoes S.; Filipe A.; Faneca H.; Mano M.; Penacho N.; Duzgunes N.; Pedroso de Lima M. Cationic liposomes for gene delivery. Expert Opin. Drug Delivery 2005, 2, 237–254. 10.1517/17425247.2.2.237. [DOI] [PubMed] [Google Scholar]
- Bangham A. D. Liposomes: The Babraham connection. Chem. Phys. Lipids 1993, 64, 275–285. 10.1016/0009-3084(93)90071-A. [DOI] [PubMed] [Google Scholar]
- Kaskowitz L.; Graham M. V.; Emami B.; Halverson K. J.; Rush C. Radiation therapy alone for stage I non-small cell lung cancer. Int. J. Radiat. Oncol., Biol., Phys. 1993, 27, 517–523. 10.1016/0360-3016(93)90374-5. [DOI] [PubMed] [Google Scholar]
- Kommareddy S.; Amiji M. Targeted drug delivery to tumor cells using colloidal carriers. Cell. Drug Delivery 2004, 181–215. 10.1007/978-1-59259-745-1_11. [DOI] [Google Scholar]
- Zhang Y.; Li A.; Wang Z.; Han Z.; He J.; Ma J. Antimetastatic activities of pegylated liposomal doxorubicin in a murine metastatic lung cancer model. J. Drug Targeting 2008, 16, 679–687. 10.1080/10611860802201340. [DOI] [PubMed] [Google Scholar]
- Suzuki S.; Kawakami S.; Chansri N.; Yamashita F.; Hashida M. Inhibition of pulmonary metastasis in mice by all-trans retinoic acid incorporated in cationic liposomes. J. Controlled Release 2006, 116, 58–63. 10.1016/j.jconrel.2006.08.025. [DOI] [PubMed] [Google Scholar]
- Siddikuzzaman; Grace V. B. Anti-metastatic study of liposome-encapsulated all trans retinoic acid (ATRA) in B16F10 melanoma cells-implanted C57BL/6 mice. Cancer Invest 2014, 32, 507–517. 10.3109/07357907.2014.964408. [DOI] [PubMed] [Google Scholar]
- Liu C.; Shi J.; Dai Q.; Yin X.; Zhang X.; Zheng A. In-vitro and in-vivo evaluation of ciprofloxacin liposomes for pulmonary administration. Drug Dev. Ind. Pharm. 2015, 41, 272–278. 10.3109/03639045.2013.858740. [DOI] [PubMed] [Google Scholar]
- Zhou J.; Zhao W. Y.; Ma X.; Ju R. J.; Li X. Y.; Li N.; Sun M. G.; Shi J. F.; Zhang C. X.; Lu W. L. The anticancer efficacy of paclitaxel liposomes modified with mitochondrial targeting conjugate in resistant lung cancer. Biomaterials 2013, 34, 3626–3638. 10.1016/j.biomaterials.2013.01.078. [DOI] [PubMed] [Google Scholar]
- Immordino M. L.; Dosio F.; Cattel L. Stealth liposomes: review of the basic science, rationale, and clinical applications, existing and potential. Int. J. Nanomed. 2006, 1, 297. [PMC free article] [PubMed] [Google Scholar]
- Lee Y. K.; Lee T. S.; Song I. H.; Jeong H. Y.; Kang S. J.; Kim M. W.; Ryu S. H.; Jung I. H.; Kim J. S.; Park Y. S. Inhibition of pulmonary cancer progression by epidermal growth factor receptor-targeted transfection with Bcl-2 and survivin siRNAs. Cancer Gene Ther. 2015, 22, 335–343. 10.1038/cgt.2015.18. [DOI] [PubMed] [Google Scholar]
- Nascimento T. L.; Hillaireau H.; Vergnaud J.; Rivano M.; Deloménie C.; Courilleau D.; Arpicco S.; Suk J. S.; Hanes J.; Fattal E. Hyaluronic acid-conjugated lipoplexes for targeted delivery of siRNA in a murine metastatic lung cancer model. Int. J. Pharm. 2016, 514, 103–111. 10.1016/j.ijpharm.2016.06.125. [DOI] [PubMed] [Google Scholar]
- Khatri N.; Baradia D.; Vhora I.; Rathi M.; Misra A. cRGD grafted liposomes containing inorganic nano-precipitate complexed siRNA for intracellular delivery in cancer cells. J. Controlled Release 2014, 182, 45–57. 10.1016/j.jconrel.2014.03.003. [DOI] [PubMed] [Google Scholar]
- Song X. L.; Ju R. J.; Xiao Y.; Wang X.; Liu S.; Fu M.; Liu J. J.; Gu L. Y.; Li X. T.; Cheng L. Application of multifunctional targeting epirubicin liposomes in the treatment of non-small-cell lung cancer. Int. J. Nanomed. 2017, 12, 7433. 10.2147/IJN.S141787. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Allen T. M.; Cullis P. R. Liposomal drug delivery systems: from concept to clinical applications. Adv. Drug Delivery Rev. 2013, 65, 36–48. 10.1016/j.addr.2012.09.037. [DOI] [PubMed] [Google Scholar]
- Bazak R.; Houri M.; El Achy S.; Kamel S.; Refaat T. Cancer active targeting by nanoparticles: a comprehensive review of literature. J. Cancer Res. Clin. Oncol. 2015, 141, 769–784. 10.1007/s00432-014-1767-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Su Z.; Shi Y.; Xiao Y.; Sun M.; Ping Q.; Zong L.; Li S.; Niu J.; Huang A.; You W.; et al. Effect of octreotide surface density on receptor-mediated endocytosis in vitro and anticancer efficacy of modified nanocarrier in vivo after optimization. Int. J. Pharm. 2013, 447, 281–292. 10.1016/j.ijpharm.2013.01.068. [DOI] [PubMed] [Google Scholar]
- Zhang H. Y.; Xu W. Q.; Wang Y. W.; Omari-Siaw E.; Wang Y.; Zheng Y. Y.; Cao X.; Tong S. S.; Yu J. N.; Xu X. M. Tumor targeted delivery of octreotide-periplogenin conjugate: synthesis, in vitro and in vivo evaluation. Int. J. Pharm. 2016, 502, 98–106. 10.1016/j.ijpharm.2016.02.024. [DOI] [PubMed] [Google Scholar]
- http://clinicaltrials.gov/
- Bibi S.; Lattmann E.; Mohammed A. R.; Perrie Y. Trigger release liposome systems: local and remote controlled delivery?. J. Microencapsulation 2012, 29, 262–276. 10.3109/02652048.2011.646330. [DOI] [PubMed] [Google Scholar]
- Barenholz Y. C. Doxil®—The first FDA-approved nano-drug: Lessons learned. J. Controlled Release 2012, 160, 117–134. 10.1016/j.jconrel.2012.03.020. [DOI] [PubMed] [Google Scholar]
- ClinicalTrials.gov. Doxil Topotecan Doublet Cancer Study. https://clinicaltrials.gov/ct2/show/NCT00252889.
- Stathopoulos G. P.; Antoniou D.; Dimitroulis J.; Michalopoulou P.; Bastas A.; Marosis K.; Stathopoulos J.; Provata A.; Yiamboudakis P.; Veldekis D.; et al. Liposomal cisplatin combined with paclitaxel versus cisplatin and paclitaxel in non-small-cell lung cancer: a randomized phase III multicenter trial. Ann. Oncol. 2010, 21, 2227–2232. 10.1093/annonc/mdq234. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Palmer M.; Parker J.; Modi S.; Butts C.; Smylie M.; Meikle A.; Kehoe M.; MacLean G.; Longenecker M. Phase I study of the BLP25 (MUC1 Peptide) liposomal vaccine for active specific immunotherapy in stage IIIB/IV non–small-cell lung cancer. Clin. Lung Cancer 2001, 3, 49–57. 10.3816/CLC.2001.n.018. [DOI] [PubMed] [Google Scholar]
- Petersen G. H.; Alzghari S. K.; Chee W.; Sankari S. S.; La-Beck N. M. Meta-analysis of clinical and preclinical studies comparing the anticancer efficacy of liposomal versus conventional non-liposomal doxorubicin. J. Controlled Release 2016, 232, 255–264. 10.1016/j.jconrel.2016.04.028. [DOI] [PubMed] [Google Scholar]
- Cheng L.; Huang F. Z.; Cheng L. F.; Zhu Y. Q.; Hu Q.; Li L.; Wei L.; Chen D. W. GE11-modified liposomes for non-small cell lung cancer targeting: preparation, ex vitro and in vivo evaluation. Int. J. Nanomed. 2014, 9, 921. 10.2147/IJN.S53310. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tosi G.; Bortot B.; Ruozi B.; Dolcetta D.; Vandelli M. A.; Forni F.; Severini G. M. Potential use of polymeric nanoparticles for drug delivery across the blood-brain barrier. Curr. Med. Chem. 2013, 20, 2212–25. 10.2174/0929867311320170006. [DOI] [PubMed] [Google Scholar]
- Jung J.; Park S. J.; Chung H.; Kang H. W.; Lee S. W.; Seo M. H.; Park H. J.; Song S. Y.; Jeong S. Y.; Choi E. K. Polymeric nanoparticles containing taxanes enhance chemoradiotherapeutic efficacy in non-small cell lung cancer. Int. J. Radiat. Oncol., Biol., Phys. 2012, 84, 77–83. 10.1016/j.ijrobp.2012.02.030. [DOI] [PubMed] [Google Scholar]
- Kim D. W.; Kim S. Y.; Kim H. K.; Kim S. W.; Shin S. W.; Kim J. S.; Park K.; Lee M. Y.; Heo D. S. Multicenter phase II trial of Genexol-PM, a novel Cremophor-free, polymeric micelle formulation of paclitaxel, with cisplatin in patients with advanced non-small-cell lung cancer. Ann. Oncol. 2007, 18, 2009–14. 10.1093/annonc/mdm374. [DOI] [PubMed] [Google Scholar]
- Wang W.; Xi M.; Duan X.; Wang Y.; Kong F. Delivery of baicalein and paclitaxel using selfassembled nanoparticles: synergistic antitumor effect in vitro and in vivo. Int. J. Nanomed. 2015, 10, 3737–3750. 10.2147/IJN.S80297. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sajja H. K.; East M. P.; Mao H.; Wang Y. A.; Nie S.; Yang L. Development of multifunctional nanoparticles for targeted drug delivery and noninvasive imaging of therapeutic effect. Curr. Drug Discovery Technol. 2009, 6, 43–51. 10.2174/157016309787581066. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gillies E. R.; Frechet J. M. Dendrimers and dendritic polymers in drug delivery. Drug Discovery Today 2005, 10, 35–43. 10.1016/S1359-6446(04)03276-3. [DOI] [PubMed] [Google Scholar]
- ud Din F.; Aman W.; Ullah I.; Qureshi O. S.; Mustapha O.; Shafique S.; Zeb A. Effective use of nanocarriers as drug delivery systems for the treatment of selected tumors. Int. J. Nanomed. 2017, 12, 7291. 10.2147/IJN.S146315. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Heath J. R.; Davis M. E. Nanotechnology and cancer. Annu. Rev. Med. 2008, 59, 251. 10.1146/annurev.med.59.061506.185523. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang J. A.; Xuan T.; Parmar M.; Ma L.; Ugwu S.; Ali S.; Ahmad I. Development and characterization of a novel liposome-based formulation of SN-38. Int. J. Pharm. 2004, 270, 93–107. 10.1016/j.ijpharm.2003.10.015. [DOI] [PubMed] [Google Scholar]
- Harasym T. O.; Tardi P. G.; Harasym N. L.; Harvie P.; Johnstone S. A.; Mayer L. D. Increased preclinical efficacy of irinotecan and floxuridine coencapsulated inside liposomes is associated with tumor delivery of synergistic drug ratios. Oncol. Res. 2006, 16, 361–374. 10.3727/000000006783980937. [DOI] [PubMed] [Google Scholar]
- Boulikas T. Clinical overview on Lipoplatin: a successful liposomal formulation of cisplatin. Expert Opin. Invest. Drugs 2009, 18, 1197–1218. 10.1517/13543780903114168. [DOI] [PubMed] [Google Scholar]
- Trang P.; Wiggins J. F.; Daige C. L.; Cho C.; Omotola M.; Brown D.; Weidhaas J. B.; Bader A. G.; Slack F. J. Systemic delivery of tumor suppressor microRNA mimics using a neutral lipid emulsion inhibits lung tumors in mice. Mol. Ther. 2011, 19, 1116–1122. 10.1038/mt.2011.48. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wei Y.; Liang J.; Zheng X.; Pi C.; Liu H.; Yang H.; Zou Y.; Ye Y.; Zhao L. Lung-targeting drug delivery system of baicalin-loaded nanoliposomes: development, biodistribution in rabbits, and pharmacodynamics in nude mice bearing orthotopic human lung cancer. Int. J. Nanomed. 2017, 12, 251. 10.2147/IJN.S119895. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Koshkina N. V.; Waldrep J. C.; Roberts L. E.; Golunski E.; Melton S.; Knight V. Paclitaxel liposome aerosol treatment induces inhibition of pulmonary metastases in murine renal carcinoma model. Clin. Cancer Res. 2001, 7, 3258–3262. [PubMed] [Google Scholar]
- Lin C. C.; Yeh H. H.; Huang W. L.; Yan J. J.; Lai W. W.; Su W. P.; Chen H. H.; Su W. C. Metformin enhances cisplatin cytotoxicity by suppressing signal transducer and activator of transcription–3 activity independently of the liver kinase B1–AMP-activated protein kinase pathway. Am. J. Respir. Cell Mol. Biol. 2013, 49, 241–250. 10.1165/rcmb.2012-0244OC. [DOI] [PubMed] [Google Scholar]
- Ryan G. M.; Kaminskas L. M.; Kelly B. D.; Owen D. J.; McIntosh M. P.; Porter C. J. Pulmonary administration of PEGylated polylysine dendrimers: absorption from the lung versus retention within the lung is highly size-dependent. Mol. Pharmaceutics 2013, 10, 2986–2995. 10.1021/mp400091n. [DOI] [PubMed] [Google Scholar]
- Kaminskas L. M.; McLeod V. M.; Ryan G. M.; Kelly B. D.; Haynes J. M.; Williamson M.; Thienthong N.; Owen D. J.; Porter C. J. Pulmonary administration of a doxorubicin-conjugated dendrimer enhances drug exposure to lung metastases and improves cancer therapy. J. Controlled Release 2014, 183, 18–26. 10.1016/j.jconrel.2014.03.012. [DOI] [PubMed] [Google Scholar]
- Luong D.; Sau S.; Kesharwani P.; Iyer A. K. Polyvalent folate-dendrimer-coated iron oxide theranostic nanoparticles for simultaneous magnetic resonance imaging and precise cancer cell targeting. Biomacromolecules 2017, 18, 1197–1209. 10.1021/acs.biomac.6b01885. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Singh J.; Jain K.; Mehra N. K.; Jain N. K. Dendrimers in anticancer drug delivery: mechanism of interaction of drug and dendrimers. Artif. Cells, Nanomed., Biotechnol. 2016, 44, 1626–1634. 10.3109/21691401.2015.1129625. [DOI] [PubMed] [Google Scholar]
- Chen J.; Sun Y.; Chen Q.; Wang L.; Wang S.; Tang Y.; Shi X.; Wang H. Multifunctional gold nanocomposites designed for targeted CT/MR/optical trimodal imaging of human non-small cell lung cancer cells. Nanoscale 2016, 8, 13568–13573. 10.1039/C6NR03143A. [DOI] [PubMed] [Google Scholar]
- Qin W.; Yang K.; Tang H.; Tan L.; Xie Q.; Ma M.; Zhang Y.; Yao S. Improved GFP gene transfection mediated by polyamidoamine dendrimer-functionalized multi-walled carbon nanotubes with high biocompatibility. Colloids Surf., B 2011, 84, 206–213. 10.1016/j.colsurfb.2011.01.001. [DOI] [PubMed] [Google Scholar]
- Palmerston Mendes L.; Pan J.; Torchilin V. P. Dendrimers as nanocarriers for nucleic acid and drug delivery in cancer therapy. Molecules 2017, 22, 1401. 10.3390/molecules22091401. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gorain B.; Choudhury H.; Pandey M.; Amin M. C. I. M.; Singh B.; Gupta U.; Kesharwani P. Dendrimers as effective carriers for the treatment of brain tumor. Nanotechnol.-Based Targeted Drug Delivery Syst. Brain Tumors 2018, 267–305. 10.1016/B978-0-12-812218-1.00010-5. [DOI] [Google Scholar]
- Shao N.; Su Y.; Hu J.; Zhang J.; Zhang H.; Cheng Y. Comparison of generation 3 polyamidoamine dendrimer and generation 4 polypropylenimine dendrimer on drug loading, complex structure, release behavior, and cytotoxicity. Int. J. Nanomed. 2011, 6, 3361. 10.2147/IJN.S27028. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Richter-Egger D. L.; Tesfai A.; Tucker S. A. Spectroscopic investigations of poly (propyleneimine) dendrimers using the solvatochromic probe phenol blue and comparisons to poly (amidoamine) dendrimers. Anal. Chem. 2001, 73, 5743–5751. 10.1021/ac0155355. [DOI] [PubMed] [Google Scholar]
- Wang F.; Cai X.; Su Y.; Hu J.; Wu Q.; Zhang H.; Xiao J.; Cheng Y. Reducing cytotoxicity while improving anti-cancer drug loading capacity of polypropylenimine dendrimers by surface acetylation. Acta Biomater. 2012, 8, 4304–4313. 10.1016/j.actbio.2012.07.031. [DOI] [PubMed] [Google Scholar]
- Mahato R. Nanoemulsion as targeted drug delivery system for cancer therapeutics. J. Pharm. Sci. Pharmacol. 2017, 3, 83–97. 10.1166/jpsp.2017.1082. [DOI] [Google Scholar]
- Meghani N.; Patel P.; Kansara K.; Ranjan S.; Dasgupta N.; Ramalingam C.; Kumar A. Formulation of vitamin D encapsulated cinnamon oil nanoemulsion: its potential anti-cancerous activity in human alveolar carcinoma cells. Colloids Surf., B 2018, 166, 349–357. 10.1016/j.colsurfb.2018.03.041. [DOI] [PubMed] [Google Scholar]
- Choudhury H.; Gorain B.; Chatterjee B.; Mandal U. K.; Sengupta P.; Tekade R. K. Pharmacokinetic and pharmacodynamic features of nanoemulsion following oral, intravenous, topical and nasal route. Curr. Pharm. Des. 2017, 23, 2504–2531. 10.2174/1381612822666161201143600. [DOI] [PubMed] [Google Scholar]
- Khani S.; Keyhanfar F.; Amani A. Design and evaluation of oral nanoemulsion drug delivery system of mebudipine. Drug delivery 2016, 23, 2035–2043. 10.3109/10717544.2015.1088597. [DOI] [PubMed] [Google Scholar]
- Sun D.; Wei X.; Xue X.; Fang Z.; Ren M.; Lou H.; Zhang X. Enhanced oral absorption and therapeutic effect of acetylpuerarin based on D-α-tocopheryl polyethylene glycol 1000 succinate nanoemulsions. Int. J. Nanomed. 2014, 9, 3413. 10.2147/IJN.S63777. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Singh Y.; Meher J. G.; Raval K.; Khan F. A.; Chaurasia M.; Jain N. K.; Chourasia M. K. Nanoemulsion: Concepts, development and applications in drug delivery. J. Controlled Release 2017, 252, 28–49. 10.1016/j.jconrel.2017.03.008. [DOI] [PubMed] [Google Scholar]
- Hwang Y. Y.; Ramalingam K.; Bienek D. R.; Lee V.; You T.; Alvarez R. Antimicrobial activity of nanoemulsion in combination with cetylpyridinium chloride in multidrug-resistant Acinetobacter baumannii. Antimicrob. Agents Chemother. 2013, 57, 3568–3575. 10.1128/AAC.02109-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
- dos Santos Câmara A. L.; Nagel G.; Tschiche H. R.; Cardador C. M.; Muehlmann L. A.; de Oliveira D. M.; Alvim P. Q.; Azevedo R. B.; Calderón M.; Figueiro Longo J. P. Acid-sensitive lipidated doxorubicin prodrug entrapped in nanoemulsion impairs lung tumor metastasis in a breast cancer model. Nanomedicine 2017, 12, 1751–1765. 10.2217/nnm-2017-0091. [DOI] [PubMed] [Google Scholar]
- Kim J. E.; Park Y. J. Improved antitumor efficacy of hyaluronic acid-complexed paclitaxel nanoemulsions in treating non-small cell lung cancer. Biomol. Ther. 2017, 25, 411. 10.4062/biomolther.2016.261. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wu J.; Zhang J.; Deng C.; Meng F.; Cheng R.; Zhong Z. Robust, responsive, and targeted PLGA anticancer nanomedicines by combination of reductively cleavable surfactant and covalent hyaluronic acid coating. ACS Appl. Mater. Interfaces 2017, 9, 3985–3994. 10.1021/acsami.6b15105. [DOI] [PubMed] [Google Scholar]
- Wang G.; Gao S.; Tian R.; Miller-Kleinhenz J.; Qin Z.; Liu T.; Li L.; Zhang F.; Ma Q.; Zhu L. Theranostic hyaluronic acid–iron micellar nanoparticles for magnetic-field-enhanced in vivo cancer chemotherapy. ChemMedChem 2018, 13, 78–86. 10.1002/cmdc.201700515. [DOI] [PubMed] [Google Scholar]
- Li X.; Du L.; Wang C.; Liu Y.; Mei X.; Jin Y. Highly efficient and lowly toxic docetaxel nanoemulsions for intravenous injection to animals. Die Pharmazie- Int. J. Pharm. Sci. 2011, 66, 479–483. [PubMed] [Google Scholar]
- Yamada Y.; Kinoshita I.; Kenichi K.; Yamamoto H.; Iwasaki T.; Otsuka H.; Yoshimoto M.; Ishihara S.; Toda Y.; Kuma Y.; et al. Histopathological and genetic review of phosphaturic mesenchymal tumours, mixed connective tissue variant. Histopathology 2018, 72, 460–471. 10.1111/his.13377. [DOI] [PubMed] [Google Scholar]
- Chang H. B.; Chen B. H. Inhibition of lung cancer cells A549 and H460 by curcuminoid extracts and nanoemulsions prepared from Curcuma longa Linnaeus. Int. J. Nanomed. 2015, 10, 5059. 10.2147/IJN.S87225. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Khan I.; Bahuguna A.; Kumar P.; Bajpai V. K.; Kang S. C. In vitro and in vivo antitumor potential of carvacrol nanoemulsion against human lung adenocarcinoma A549 cells via mitochondrial mediated apoptosis. Sci. Rep. 2018, 8, 144. 10.1038/s41598-017-18644-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jyoti K.; Kaur K.; Pandey R. S.; Jain U. K.; Chandra R.; Madan J. Inhalable nanostructured lipid particles of 9-bromo-noscapine, a tubulin-binding cytotoxic agent: in vitro and in vivo studies. J. Colloid Interface Sci. 2015, 445, 219–230. 10.1016/j.jcis.2014.12.092. [DOI] [PubMed] [Google Scholar]
- Lee W. D.; Liang Y. J.; Chen B. H. Effects of tanshinone nanoemulsion and extract on inhibition of lung cancer cells A549. Nanotechnology 2016, 27, 495101. 10.1088/0957-4484/27/49/495101. [DOI] [PubMed] [Google Scholar]
- Torchilin V. P. Micellar nanocarriers: pharmaceutical perspectives. Pharm. Res. 2006, 24, 1–16. 10.1007/s11095-006-9132-0. [DOI] [PubMed] [Google Scholar]
- Zhu Y.; Liao L. Applications of nanoparticles for anticancer drug delivery: a review. J. Nanosci. Nanotechnol. 2015, 15, 4753–4773. 10.1166/jnn.2015.10298. [DOI] [PubMed] [Google Scholar]
- Batrakova E. V.; Bronich T. K.; Vetro J. A.; Kabanov A. V. Polymer micelles as drug carriers. Nanopart. Drug Carriers 2006, 57–93. 10.1142/9781860949074_0005. [DOI] [Google Scholar]
- Liu T.; Li X.; Qian Y.; Hu X.; Liu S. Multifunctional pH-disintegrable micellar nanoparticles of asymmetrically functionalized β-cyclodextrin-based star copolymer covalently conjugated with doxorubicin and DOTA-Gd moieties. Biomaterials 2012, 33, 2521–2531. 10.1016/j.biomaterials.2011.12.013. [DOI] [PubMed] [Google Scholar]
- Biswas S.; Kumari P.; Lakhani P. M.; Ghosh B. Recent advances in polymeric micelles for anti-cancer drug delivery. Eur. J. Pharm. Sci. 2016, 83, 184–202. 10.1016/j.ejps.2015.12.031. [DOI] [PubMed] [Google Scholar]
- Liu Y.; Li J.; Liu F.; Zhang L.; Feng L.; Yu D.; Zhang N. Theranostic polymeric micelles for the diagnosis and treatment of hepatocellular carcinoma. J. Biomed. Nanotechnol. 2015, 11, 613–622. 10.1166/jbn.2015.1945. [DOI] [PubMed] [Google Scholar]
- Upponi J. R.; Jerajani K.; Nagesha D. K.; Kulkarni P.; Sridhar S.; Ferris C.; Torchilin V. P. Polymeric micelles: theranostic co-delivery system for poorly water-soluble drugs and contrast agents. Biomaterials 2018, 170, 26–36. 10.1016/j.biomaterials.2018.03.054. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Saifuddin N.; Raziah A. Z.; Junizah A. R. Carbon nanotubes: a review on structure and their interaction with proteins. J. Chem. 2013, 2013, 676815. 10.1155/2013/676815. [DOI] [Google Scholar]
- Bianco A. Carbon nanotubes for the delivery of therapeutic molecules. Expert Opin. Drug Delivery 2004, 1, 57–65. 10.1517/17425247.1.1.57. [DOI] [PubMed] [Google Scholar]
- Luanpitpong S.; Wang L.; Rojanasakul Y. The effects of carbon nanotubes on lung and dermal cellular behaviors. Nanomedicine 2014, 9, 895–912. 10.2217/nnm.14.42. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Madani S. Y.; Naderi N.; Dissanayake O.; Tan A.; Seifalian A. M. A new era of cancer treatment: carbon nanotubes as drug delivery tools. Int. J. Nanomed. 2011, 6, 2963. 10.2147/IJN.S16923. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ng C. M.; Loh H. S.; Muthoosamy K.; Sridewi N.; Manickam S. Conjugation of insulin onto the sidewalls of single-walled carbon nanotubes through functionalization and diimide-activated amidation. Int. J. Nanomed. 2016, 11, 1607. 10.2147/IJN.S98726. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lay C. L.; Liu H. Q.; Tan H. R.; Liu Y. Delivery of paclitaxel by physically loading onto poly (ethylene glycol) (PEG)-graftcarbon nanotubes for potent cancer therapeutics. Nanotechnology 2010, 21, 065101. 10.1088/0957-4484/21/6/065101. [DOI] [PubMed] [Google Scholar]
- Mashal A.; Sitharaman B.; Li X.; Avti P. K.; Sahakian A. V.; Booske J. H.; Hagness S. C. Toward carbon-nanotube-based theranostic agents for microwave detection and treatment of breast cancer: Enhanced dielectric and heating response of tissue-mimicking materials. IEEE Trans. Biomed. Eng. 2010, 57, 1831–1834. 10.1109/TBME.2010.2042597. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Al Faraj A.; Shaik A. S.; Halwani R.; Alfuraih A. Magnetic targeting and delivery of drug-loaded SWCNTs theranostic nanoprobes to lung metastasis in breast cancer animal model: noninvasive monitoring using magnetic resonance imaging. Mol. Imaging Biol. 2016, 18, 315–324. 10.1007/s11307-015-0902-0. [DOI] [PubMed] [Google Scholar]
- Ji S. R.; Liu C.; Zhang B.; Yang F.; Xu J.; Long J.; Jin C.; Fu D. L.; Ni Q. X.; Yu X. J. Carbon nanotubes in cancer diagnosis and therapy. Biochim. Biophys. Acta, Rev. Cancer 2010, 1806, 29–35. 10.1016/j.bbcan.2010.02.004. [DOI] [PubMed] [Google Scholar]
- Wu H. C.; Chang X.; Liu L.; Zhao F.; Zhao Y. Chemistry of carbon nanotubes in biomedical applications. J. Mater. Chem. 2010, 20, 1036–1052. 10.1039/B911099M. [DOI] [Google Scholar]
- Vardharajula S.; Ali S. Z.; Tiwari P. M.; Eroğlu E.; Vig K.; Dennis V. A.; Singh S. R. Functionalized carbon nanotubes: biomedical applications. Int. J. Nanomed. 2012, 7, 5361. 10.2147/IJN.S35832. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bhirde A. A.; Patel V.; Gavard J.; Zhang G.; Sousa A. A.; Masedunskas A.; Leapman R. D.; Weigert R.; Gutkind J. S.; Rusling J. F. Targeted killing of cancer cells in vivo and in vitro with EGF-directed carbon nanotube-based drug delivery. ACS Nano 2009, 3, 307–316. 10.1021/nn800551s. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang L.; Zhang; Zhang; Shi; Zhang; Zhang Z.; Wang L.; Li Synergistic enhancement of cancer therapy using a combination of docetaxel and photothermal ablation induced by single-walled carbon nanotubes. Int. J. Nanomed. 2011, 6, 2641. 10.2147/IJN.S24167. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kim S. W.; Lee Y. K.; Lee J. Y.; Hong J. H.; Khang D. PEGylated anticancer-carbon nanotubes complex targeting mitochondria of lung cancer cells. Nanotechnology 2017, 28, 465102. 10.1088/1361-6528/aa8c31. [DOI] [PubMed] [Google Scholar]
- Minati L.; Antonini V.; Dalla Serra M.; Speranza G. Multifunctional branched gold–carbon nanotube hybrid for cell imaging and drug delivery. Langmuir 2012, 28, 15900–15906. 10.1021/la303298u. [DOI] [PubMed] [Google Scholar]
- Ciofani G.; Raffa V.; Pensabene V.; Menciassi A.; Dario P. Dispersion of multi-walled carbon nanotubes in aqueous pluronic F127 solutions for biological applications. Fullerenes, Nanotubes, Carbon Nanostruct 2009, 17, 11–25. 10.1080/15363830802515840. [DOI] [Google Scholar]
- Ali-Boucetta H.; Al-Jamal K. T.; McCarthy D.; Prato M.; Bianco A.; Kostarelos K. Multiwalled carbon nanotube–doxorubicin supramolecular complexes for cancer therapeutics. Chem. Commun. 2008, 4, 459–461. 10.1039/B712350G. [DOI] [PubMed] [Google Scholar]
- Arya N.; Arora A.; Vasu K. S.; Sood A. K.; Katti D. S. Combination of single walled carbon nanotubes/graphene oxide with paclitaxel: a reactive oxygen species mediated synergism for treatment of lung cancer. Nanoscale 2013, 5, 2818–2829. 10.1039/c3nr33190c. [DOI] [PubMed] [Google Scholar]
- Sobhani Z.; Dinarvand R.; Atyabi F.; Ghahremani M.; Adeli M. Increased paclitaxel cytotoxicity against cancer cell lines using a novel functionalized carbon nanotube. Int. J. Nanomed. 2011, 6, 705. 10.2147/IJN.S17336. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rana S.; Bajaj A.; Mout R.; Rotello V. M. Monolayer coated gold nanoparticles for delivery applications. Adv. Drug Delivery Rev. 2012, 64, 200–216. 10.1016/j.addr.2011.08.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Peng G.; Trock E.; Haick H.. Composites of carbon nanotubes and non-polymeric materials for diagnosing lung cancer via breath samples. Organic Field-Effect Transistors VII and Organic Semiconductors in Sensors and Bioelectronics; 2008; pp 132–140. [Google Scholar]
- Knights O. B.; McLaughlan J. R. Gold nanorods for light-based lung cancer theranostics. Int. J. Mol. Sci. 2018, 19, 3318. 10.3390/ijms19113318. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chen Y. H.; Tsai C. Y.; Huang P. Y.; Chang M. Y.; Cheng P. C.; Chou C. H.; Chen D. H.; Wang C. R.; Shiau A. L.; Wu C. L. Methotrexate conjugated to gold nanoparticles inhibits tumor growth in a syngeneic lung tumor model. Mol. Pharmaceutics 2007, 4, 713–722. 10.1021/mp060132k. [DOI] [PubMed] [Google Scholar]
- Zhang X. Q.; Xu X.; Lam R.; Giljohann D.; Ho D.; Mirkin C. A. Strategy for increasing drug solubility and efficacy through covalent attachment to polyvalent DNA–nanoparticle conjugates. ACS Nano 2011, 5, 6962–6970. 10.1021/nn201446c. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ma L. L.; Tam J. O.; Willsey B. W.; Rigdon D.; Ramesh R.; Sokolov K.; Johnston K. P. Selective targeting of antibody conjugated multifunctional nanoclusters (nanoroses) to epidermal growth factor receptors in cancer cells. Langmuir 2011, 27, 7681–7690. 10.1021/la200659z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lkhagvadulam B.; Kim J. H.; Yoon I.; Shim Y. K. Size-dependent photodynamic activity of gold nanoparticles conjugate of water soluble purpurin-18-N-methyl-D-glucamine. BioMed. Res. Int. 2013, 2013, 720579. 10.1155/2013/720579. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Park C.; Youn H.; Kim H.; Noh T.; Kook Y. H.; Oh E. T.; Park H. J.; Kim C. Cyclodextrin-covered gold nanoparticles for targeted delivery of an anti-cancer drug. J. Mater. Chem. 2009, 19, 2310–2315. 10.1039/b816209c. [DOI] [Google Scholar]
- Anselmo A. C.; Mitragotri S. Nanoparticles in the clinic. Bioeng. Transl. Med. 2016, 1, 10–29. 10.1002/btm2.10003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bao G.; Mitragotri S.; Tong S. Multifunctional nanoparticles for drug delivery and molecular imaging. Annu. Rev. Biomed. Eng. 2013, 15, 253. 10.1146/annurev-bioeng-071812-152409. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu Y.; Zhang P.; Li F.; Jin X.; Li J.; Chen W.; Li Q. Metal-based nanoenhancers for future radiotherapy: radiosensitizing and synergistic effects on tumor cells. Theranostics 2018, 8, 1824. 10.7150/thno.22172. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Akbarzadeh A.; Samiei M.; Davaran S. Magnetic nanoparticles: preparation, physical properties, and applications in biomedicine. Nanoscale Res. Lett. 2012, 7, 144. 10.1186/1556-276X-7-144. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ganapathe L. S.; Mohamed M. A.; Mohamad Yunus R.; Berhanuddin D. D. Magnetite (Fe3O4) nanoparticles in biomedical application: From synthesis to surface functionalisation. Magnetochemistry 2020, 6, 68. 10.3390/magnetochemistry6040068. [DOI] [Google Scholar]
- Revia R. A.; Zhang M. Magnetite nanoparticles for cancer diagnosis, treatment, and treatment monitoring: recent advances. Mater. Today 2016, 19, 157–168. 10.1016/j.mattod.2015.08.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sadhukha T.; Wiedmann T. S.; Panyam J. Inhalable magnetic nanoparticles for targeted hyperthermia in lung cancer therapy. Biomaterials 2013, 34, 5163–5171. 10.1016/j.biomaterials.2013.03.061. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Issa B.; Obaidat I. M.; Albiss B. A.; Haik Y. Magnetic nanoparticles: surface effects and properties related to biomedicine applications. Int. J. Mol. Sci. 2013, 14, 21266–21305. 10.3390/ijms141121266. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang Y.; Zhang Y.; Du Z.; Wu M.; Zhang G. Detection of micrometastases in lung cancer with magnetic nanoparticles and quantum dots. Int. J. Nanomed. 2012, 7, 2315. 10.2147/IJN.S30593. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li K.; Chen B.; Xu L.; Feng J.; Xia G.; Cheng J.; Wang J.; Gao F.; Wang X. Reversal of multidrug resistance by cisplatin-loaded magnetic Fe3O4 nanoparticles in A549/DDP lung cancer cells in vitro and in vivo. Int. J. Nanomed. 2013, 8, 1867. 10.2147/IJN.S43752. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gaihre B.; Khil M. S.; Kim H. Y. In vitro anticancer activity of doxorubicin-loaded gelatin-coated magnetic iron oxide nanoparticles. J. Microencapsulation 2011, 28, 286–293. 10.3109/02652048.2011.559286. [DOI] [PubMed] [Google Scholar]
- Kuroda S.; Tam J.; Roth J. A.; Sokolov K.; Ramesh R. EGFR-targeted plasmonic magnetic nanoparticles suppress lung tumor growth by abrogating G2/M cell-cycle arrest and inducing DNA damage. Int. J. Nanomed. 2014, 9, 3825. 10.2147/IJN.S65990. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yokoyama T.; Tam J.; Kuroda S.; Scott A. W.; Aaron J.; Larson T.; Shanker M.; Correa A. M.; Kondo S.; Roth J. A.; et al. EGFR-targeted hybrid plasmonic magnetic nanoparticles synergistically induce autophagy and apoptosis in non-small cell lung cancer cells. PloS one 2011, 6, e25507. 10.1371/journal.pone.0025507. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nejati-Koshki K.; Mesgari M.; Ebrahimi E.; Abbasalizadeh F.; Fekri Aval S.; Khandaghi A. A.; Abasi M.; Akbarzadeh A. Synthesis and in vitro study of cisplatin-loaded Fe3O4 nanoparticles modified with PLGA-PEG6000 copolymers in treatment of lung cancer. J. Microencapsulation 2014, 31, 815–823. 10.3109/02652048.2014.940011. [DOI] [PubMed] [Google Scholar]
- Iyer R.; Hsia C.; Nguyen K. Nano-therapeutics for the lung: state-of-the-art and future perspectives. Curr. Pharm. Des. 2015, 21, 5233–5244. 10.2174/1381612821666150923095742. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Stocke N. A.; Meenach S. A.; Arnold S. M.; Mansour H. M.; Hilt J. Z. Formulation and characterization of inhalable magnetic nanocomposite microparticles (MnMs) for targeted pulmonary delivery via spray drying. Int. J. Pharm. 2015, 479, 320–328. 10.1016/j.ijpharm.2014.12.050. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Orel V.; Shevchenko A.; Romanov A.; Tselepi M.; Mitrelias T.; Barnes C. H.; Burlaka A.; Lukin S.; Shchepotin I. Magnetic properties and antitumor effect of nanocomplexes of iron oxide and doxorubicin. Nanomed.: Nanotechnol. Biol. Med. 2015, 11, 47–55. 10.1016/j.nano.2014.07.007. [DOI] [PubMed] [Google Scholar]
- Liu D.; Liu Z.; Wang L.; Zhang C.; Zhang N. Nanostructured lipid carriers as novel carrier for parenteral delivery of docetaxel. Colloids Surf., B 2011, 85, 262–9. 10.1016/j.colsurfb.2011.02.038. [DOI] [PubMed] [Google Scholar]
- Mussi S. V.; Torchilin V. P. Recent trends in the use of lipidic nanoparticles as pharmaceutical carriers for cancer therapy and diagnostics. J. Mater. Chem. B 2013, 1, 5201–5209. 10.1039/c3tb20990c. [DOI] [PubMed] [Google Scholar]
- Müller R. H.; Mäder K.; Gohla S. Solid lipid nanoparticles (SLN) for controlled drug delivery–a review of the state of the art. Eur. J. Pharm. Biopharm 2000, 50, 161–177. 10.1016/S0939-6411(00)00087-4. [DOI] [PubMed] [Google Scholar]
- Bentolila L. A.; Ebenstein Y.; Weiss S. Quantum dots for in vivo small-animal imaging. J. Nucl. Med. 2009, 50, 493–496. 10.2967/jnumed.108.053561. [DOI] [PMC free article] [PubMed] [Google Scholar]
- De Jesus M. B.; Zuhorn I. S. Solid lipid nanoparticles as nucleic acid delivery system: Properties and molecular mechanisms. J. Controlled Release 2015, 201, 1–13. 10.1016/j.jconrel.2015.01.010. [DOI] [PubMed] [Google Scholar]
- Madan J. R.; Khude P. A.; Dua K. Development and evaluation of solid lipid nanoparticles of Mometasone furoate for topical delivery. Int. J. Pharm. Invest. 2014, 4, 60. 10.4103/2230-973X.133047. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Davaran S.; Akbarzadeh A.; Nejati-Koshki K.; Alimohammadi S.; Ghamari M. F.; Soghrati M. M.; Rezaei A.; Khandaghi A. A. In vitro studies of NIPAAM-MAA-VP copolymer-coated magnetic nanoparticles for controlled anticancer drug release. J. Encapsulation Adsorpt. Sci. 2013, 3, 108–115. [Google Scholar]
- Mukherjee S.; Ray S.; Thakur R. S. Solid lipid nanoparticles: a modern formulation approach in drug delivery system. Indian J. Pharm. Sci. 2009, 71, 349. 10.4103/0250-474X.57282. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Andreozzi E.; Seo J. W.; Ferrara K.; Louie A. Novel method to label solid lipid nanoparticles with 64cu for positron emission tomography imaging. Bioconjugate Chem. 2011, 22, 808–818. 10.1021/bc100478k. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lim E. K.; Kim T.; Paik S.; Haam S.; Huh Y. M.; Lee K. Nanomaterials for theranostics: recent advances and future challenges. Chem. Rev. 2015, 115, 327–394. 10.1021/cr300213b. [DOI] [PubMed] [Google Scholar]
- Sukumar U. K.; Bhushan B.; Dubey P.; Matai I.; Sachdev A.; Packirisamy G. Emerging applications of nanoparticles for lung cancer diagnosis and therapy. Int. Nano Lett. 2013, 3, 45. 10.1186/2228-5326-3-45. [DOI] [Google Scholar]
- Choi S. H.; Jin S. E.; Lee M. K.; Lim S. J.; Park J. S.; Kim B. G.; Ahn W. S.; Kim C. K. Novel cationic solid lipid nanoparticles enhanced p53 gene transfer to lung cancer cells. Eur. J. Pharm. Biopharm. 2008, 68, 545–554. 10.1016/j.ejpb.2007.07.011. [DOI] [PubMed] [Google Scholar]
- Naguib Y. W.; Rodriguez B. L.; Li X.; Hursting S. D.; Williams R. O. III; Cui Z. Solid lipid nanoparticle formulations of docetaxel prepared with high melting point triglycerides: in vitro and in vivo evaluation. Mol. Pharmaceutics 2014, 11, 1239–1249. 10.1021/mp4006968. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Choi S. H.; Jin S. E.; Lee M. K.; Lim S. J.; Park J. S.; Kim B. G.; Ahn W. S.; Kim C. K. Novel cationic solid lipid nanoparticles enhanced p53 gene transfer to lung cancer cells. Eur. J. Pharm. Biopharm. 2008, 68, 545–554. 10.1016/j.ejpb.2007.07.011. [DOI] [PubMed] [Google Scholar]
- Bae K. H.; Lee J. Y.; Lee S. H.; Park T. G.; Nam Y. S. Optically traceable solid lipid nanoparticles loaded with siRNA and paclitaxel for synergistic chemotherapy with in situ imaging. Adv. Healthcare Mater. 2013, 2, 576–584. 10.1002/adhm.201200338. [DOI] [PubMed] [Google Scholar]
- Yuan H.; Miao J.; Du Y. Z.; You J.; Hu F. Q.; Zeng S. Cellular uptake of solid lipid nanoparticles and cytotoxicity of encapsulated paclitaxel in A549 cancer cells. Int. J. Pharm. 2008, 348, 137–145. 10.1016/j.ijpharm.2007.07.012. [DOI] [PubMed] [Google Scholar]
- Leiva M. C.; Ortiz R.; Contreras-Cáceres R.; Perazzoli G.; Mayevych I.; López-Romero J. M.; Sarabia F.; Baeyens J. M.; Melguizo C.; Prados J. Tripalmitin nanoparticle formulations significantly enhance paclitaxel antitumor activity against breast and lung cancer cells in vitro. Sci. Rep. 2017, 7, 13506. 10.1038/s41598-017-13816-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sudhakar C. K.; Upadhyay N.; Jain A.; Verma A.; Charyulu R. N.; Jain S. Hydrogels—promising candidates for tissue engineering. Nanotechnol. Appl. Tiss. Eng. 2015, 77–94. [Google Scholar]
- Wichterle O.; Lim D. Hydrophilic gels for biological use. Nature 1960, 185, 117–118. 10.1038/185117a0. [DOI] [Google Scholar]
- Chung H. J.; Park T. G. Self-assembled and nanostructured hydrogels for drug delivery and tissue engineering. Nano Today 2009, 4, 429–437. 10.1016/j.nantod.2009.08.008. [DOI] [Google Scholar]
- Slaughter B. V.; Khurshid S. S.; Fisher O. Z.; Khademhosseini A.; Peppas N. A. Hydrogels in regenerative medicine. Adv. Mater. 2009, 21, 3307–3329. 10.1002/adma.200802106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wu Z.; Zou X.; Yang L.; Lin S.; Fan J.; Yang B.; Sun X.; Wan Q.; Chen Y.; Fu S. Thermosensitive hydrogel used in dual drug delivery system with paclitaxel-loaded micelles for in situ treatment of lung cancer. Colloids Surf., B 2014, 122, 90–98. 10.1016/j.colsurfb.2014.06.052. [DOI] [PubMed] [Google Scholar]
- Zubris K. A. V.; Colson Y. L.; Grinstaff M. W. Hydrogels as intracellular depots for drug delivery. Mol. Pharmaceutics 2012, 9, 196–200. 10.1021/mp200367s. [DOI] [PubMed] [Google Scholar]
- Patrick J. S.Multifunctional PEG Hydrogel Nano/Microparticles for Targeted Treatment of NSCLC; 2018. [Google Scholar]
- Chao P.; Deshmukh M.; Kutscher H. L.; Gao D.; Rajan S. S.; Hu P.; Laskin D. L.; Stein S.; Sinko P. J. Pulmonary targeting microparticulate camptothecin delivery system: anti-cancer evaluation in a rat orthotopic lung cancer model. Anti-Cancer Drugs 2010, 21, 65. 10.1097/CAD.0b013e328332a322. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yao J.; Feng J.; Chen J. External-stimuli responsive systems for cancer theranostic. Asian J. Pharm. Sci. 2016, 11, 585–595. 10.1016/j.ajps.2016.06.001. [DOI] [Google Scholar]
- Sheikholeslam M.; Wheeler S. D.; Duke K. G.; Marsden M.; Pritzker M.; Chen P. Peptide and peptide-carbon nanotube hydrogels as scaffolds for tissue & 3D tumor engineering. Acta Biomater. 2018, 69, 107–119. 10.1016/j.actbio.2017.12.012. [DOI] [PubMed] [Google Scholar]
- Sachdev A.; Matai I.; Gopinath P. Carbon dots incorporated polymeric hydrogels as multifunctional platform for imaging and induction of apoptosis in lung cancer cells. Colloids Surf., B 2016, 141, 242–252. 10.1016/j.colsurfb.2016.01.043. [DOI] [PubMed] [Google Scholar]
- Matsumoto A.; Chen J.; Collette A. L.; Kim U. J.; Altman G. H.; Cebe P.; Kaplan D. L. Mechanisms of silk fibroin sol– gel transitions. J. Phys. Chem. B 2006, 110, 21630–21638. 10.1021/jp056350v. [DOI] [PubMed] [Google Scholar]
- Yucel T.; Cebe P.; Kaplan D. L. Vortex-induced injectable silk fibroin hydrogels. Biophys. J. 2009, 97, 2044–2050. 10.1016/j.bpj.2009.07.028. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gao Y.; Ren F.; Ding B.; Sun N.; Liu X.; Ding X.; Gao S. A thermo-sensitive PLGA-PEG-PLGA hydrogel for sustained release of docetaxel. J. Drug Targeting 2011, 19, 516–527. 10.3109/1061186X.2010.519031. [DOI] [PubMed] [Google Scholar]
- Milton F. P.; Govan J.; Mukhina M. V.; Gun’ko Y. K. The chiral nano-world: chiroptically active quantum nanostructures. Nanoscale Horiz 2016, 1, 14–26. 10.1039/C5NH00072F. [DOI] [PubMed] [Google Scholar]
- Chinen A. B.; Guan C. M.; Ferrer J. R.; Barnaby S. N.; Merkel T. J.; Mirkin C. A. Nanoparticle probes for the detection of cancer biomarkers, cells, and tissues by fluorescence. Chem. Rev. 2015, 115, 10530–10574. 10.1021/acs.chemrev.5b00321. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cuenca A. G.; Jiang H.; Hochwald S. N.; Delano M.; Cance W. G.; Grobmyer S. R. Emerging implications of nanotechnology on cancer diagnostics and therapeutics. Cancer 2006, 107, 459–466. 10.1002/cncr.22035. [DOI] [PubMed] [Google Scholar]
- Zrazhevskiy P.; Sena M.; Gao X. Designing multifunctional quantum dots for bioimaging, detection, and drug delivery. Chem. Soc. Rev. 2010, 39, 4326–4354. 10.1039/b915139g. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rosenthal S. J.; Chang J. C.; Kovtun O.; McBride J. R.; Tomlinson I. D. Biocompatible quantum dots for biological applications. Chem. Biol. 2011, 18, 10–24. 10.1016/j.chembiol.2010.11.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dubertret B.; Skourides P.; Norris D. J.; Noireaux V.; Brivanlou A. H.; Libchaber A. In vivo imaging of quantum dots encapsulated in phospholipid micelles. Science 2002, 298, 1759–1762. 10.1126/science.1077194. [DOI] [PubMed] [Google Scholar]
- Ghasemi Y.; Peymani P.; Afifi S. Quantum dot: magic nanoparticle for imaging, detection and targeting. Acta. Biomed. 2009, 80, 156–165. [PubMed] [Google Scholar]
- Lewinski N.; Colvin V.; Drezek R. Cytotoxicity of nanoparticles. Small 2008, 4, 26–49. 10.1002/smll.200700595. [DOI] [PubMed] [Google Scholar]
- Jin S.; Hu Y.; Gu Z.; Liu L.; Wu H. C. Application of quantum dots in biological imaging. J. Nanomater. 2011, 2011, 834139. 10.1155/2011/834139. [DOI] [Google Scholar]
- Lütje S.; Slavik R.; Fendler W.; Herrmann K.; Eiber M. PSMA ligands in prostate cancer–Probe optimization and theranostic applications. Methods 2017, 130, 42–50. 10.1016/j.ymeth.2017.06.026. [DOI] [PubMed] [Google Scholar]
- Hu L.; Zhang C.; Zeng G.; Chen G.; Wan J.; Guo Z.; Wu H.; Yu Z.; Zhou Y.; Liu J. Metal-based quantum dots: synthesis, surface modification, transport and fate in aquatic environments and toxicity to microorganisms. RSC Adv. 2016, 6, 78595–78610. 10.1039/C6RA13016J. [DOI] [Google Scholar]
- Ko N. R.; Nafiujjaman M.; Lee J. S.; Lim H. N.; Lee Y. K.; Kwon I. K. Graphene quantum dot-based theranostic agents for active targeting of breast cancer. RSC Adv. 2017, 7, 11420–11427. 10.1039/C6RA25949A. [DOI] [Google Scholar]
- Kumar A.; Singh K. R.; Ghate M. D.; Lalhlenmawia H.; Kumar D.; Singh J. Bioinspired quantum dots for cancer therapy: A mini-review. Mater. Lett. 2022, 313, 131742. 10.1016/j.matlet.2022.131742. [DOI] [Google Scholar]
- Al-Hallak K. M.; Azarmi S.; Anwar-Mohamed A.; Roa W. H.; Löbenberg R. Secondary cytotoxicity mediated by alveolar macrophages: a contribution to the total efficacy of nanoparticles in lung cancer therapy?. Eur. J. Pharm. Biopharm. 2010, 76, 112–119. 10.1016/j.ejpb.2010.05.002. [DOI] [PubMed] [Google Scholar]
- Beck-Broichsitter M.; Merkel O. M.; Kissel T. Controlled pulmonary drug and gene delivery using polymeric nano-carriers. J. Controlled Release 2012, 161, 214–224. 10.1016/j.jconrel.2011.12.004. [DOI] [PubMed] [Google Scholar]
- Knight V.; Kleinerman E. S.; Waldrep J. C.; Giovanella B. C.; Gilbert B. E.; Koshkina N. V. 9-Nitrocamptothecin liposome aerosol treatment of human cancer subcutaneous xenografts and pulmonary cancer metastases in mice. Ann. N.Y. Acad. Sci. 2000, 922, 151–163. 10.1111/j.1749-6632.2000.tb07033.x. [DOI] [PubMed] [Google Scholar]
- Hu L.; Jia Y. 2010. Preparation and characterization of solid lipid nanoparticles loaded with epirubicin for pulmonary delivery. Die Pharmazie- Int. J. Pharm. 2010, 65, 585–587. [PubMed] [Google Scholar]
- Patlolla R. R.; Chougule M.; Patel A. R.; Jackson T.; Tata P. N.; Singh M. Formulation, characterization and pulmonary deposition of nebulized celecoxib encapsulated nanostructured lipid carriers. J. Controlled Release 2010, 144, 233–241. 10.1016/j.jconrel.2010.02.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chana J.; Forbes B.; Jones S. A. Triggered-release nanocapsules for drug delivery to the lungs. Nanomed.: Nanotechnol. Biol. Med. 2015, 11, 89–97. 10.1016/j.nano.2014.07.012. [DOI] [PubMed] [Google Scholar]
- Hitzman C. J.; Elmquist W. F.; Wiedmann T. S. Development of a respirable, sustained release microcarrier for 5-fluorouracil II: In vitro and in vivo optimization of lipid coated nanoparticles. J. Pharm. Sci. 2006, 95, 1127–1143. 10.1002/jps.20590. [DOI] [PubMed] [Google Scholar]
- Gill K. K.; Nazzal S.; Kaddoumi A. Paclitaxel loaded PEG5000–DSPE micelles as pulmonary delivery platform: formulation characterization, tissue distribution, plasma pharmacokinetics, and toxicological evaluation. Eur. J. Pharm. Biopharm. 2011, 79, 276–284. 10.1016/j.ejpb.2011.04.017. [DOI] [PubMed] [Google Scholar]
- Dames P.; Gleich B.; Flemmer A.; Hajek K.; Seidl N.; Wiekhorst F.; Eberbeck D.; Bittmann I.; Bergemann C.; Weyh T.; et al. Targeted delivery of magnetic aerosol droplets to the lung. Nat. Nanotechnol. 2007, 2, 495–499. 10.1038/nnano.2007.217. [DOI] [PubMed] [Google Scholar]
- Mykhaylyk O.; Dudchenko N.; Dudchenko A. Doxorubicin magnetic conjugate targeting upon intravenous injection into mice: High gradient magnetic field inhibits the clearance of nanoparticles from the blood. J. Magn. Magn. Mater. 2005, 293, 473–482. 10.1016/j.jmmm.2005.01.063. [DOI] [Google Scholar]
- Verma N. K.; Crosbie-Staunton K.; Satti A.; Gallagher S.; Ryan K. B.; Doody T.; McAtamney C.; MacLoughlin R.; Galvin P.; Burke C. S.; et al. Magnetic core-shell nanoparticles for drug delivery by nebulization. J. Nanobiotechnol. 2013, 11, 1–12. 10.1186/1477-3155-11-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gagnadoux F.; Hureaux J.; Vecellio L.; Urban T.; Le Pape A.; Valo I.; Montharu J.; Leblond V.; Boisdron-Celle M.; Lerondel S.; et al. Aerosolized chemotherapy. J. Aerosol Med. Pulm. Drug Delivery 2008, 21, 61–70. 10.1089/jamp.2007.0656. [DOI] [PubMed] [Google Scholar]
- Gratton S. E.; Ropp P. A.; Pohlhaus P. D.; Luft J. C.; Madden V. J.; Napier M. E.; DeSimone J. M. The effect of particle design on cellular internalization pathways. Proc. Natl. Acad. Sci. U. S. A. 2008, 105, 11613–11618. 10.1073/pnas.0801763105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Minko T.; Dharap S. S.; Pakunlu R. I.; Wang Y. Molecular targeting of drug delivery systems to cancer. Curr. Drug Targets 2004, 5, 389–406. 10.2174/1389450043345443. [DOI] [PubMed] [Google Scholar]
- Saad M.; Garbuzenko O. B.; Ber E.; Chandna P.; Khandare J. J.; Pozharov V. P.; Minko T. Receptor targeted polymers, dendrimers, liposomes: which nanocarrier is the most efficient for tumor-specific treatment and imaging?. J. Controlled Release 2008, 130, 107–114. 10.1016/j.jconrel.2008.05.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shen J.; Chelvam V.; Cresswell G.; Low P. S. Use of folate-conjugated imaging agents to target alternatively activated macrophages in a murine model of asthma. Mol. Pharmaceutics 2013, 10, 1918–1927. 10.1021/mp3006962. [DOI] [PubMed] [Google Scholar]
- Shah P. K.; Austin J. H.; White C. S.; Patel P.; Haramati L. B.; Pearson G. D.; Shiau M. C.; Berkmen Y. M. Missed non–small cell lung cancer: radiographic findings of potentially resectable lesions evident only in retrospect. Radiology 2003, 226, 235–241. 10.1148/radiol.2261011924. [DOI] [PubMed] [Google Scholar]
- Weissleder R.; Pittet M. J. Imaging in the era of molecular oncology. Nature 2008, 452, 580–589. 10.1038/nature06917. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chapman S.; Dobrovolskaia M.; Farahani K.; Goodwin A.; Joshi A.; Lee H.; Meade T.; Pomper M.; Ptak K.; Rao J.; et al. Nanoparticles for cancer imaging: The good, the bad, and the promise. Nano Today 2013, 8, 454–460. 10.1016/j.nantod.2013.06.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Safdar M. H.; Hussain Z.; Abourehab M. A.; Hasan H.; Afzal S.; Thu H. E. New developments and clinical transition of hyaluronic acid-based nanotherapeutics for treatment of cancer: reversing multidrug resistance, tumour-specific targetability and improved anticancer efficacy. Artif. Cells, Nanomed., Biotechnol. 2018, 46, 1967–1980. 10.1080/21691401.2017.1397001. [DOI] [PubMed] [Google Scholar]
- Arshad R.; Kiani M. H.; Rahdar A.; Sargazi S.; Barani M.; Shojaei S.; Bilal M.; Kumar D.; Pandey S. Nano-Based Theranostic Platforms for Breast Cancer: A Review of Latest Advancements. Bioengineering 2022, 9, 320. 10.3390/bioengineering9070320. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ali I.; Rahis-Uddin; Salim K.; Rather M. A.; Wani W. A.; Haque A. Advances in nano drugs for cancer chemotherapy. Curr. Cancer Drug Targets 2011, 11, 135–146. 10.2174/156800911794328493. [DOI] [PubMed] [Google Scholar]
- Chen D.; Jiang M.; Li N.; Gu H.; Xu Q.; Ge J.; Xia X.; Lu J. Modification of magnetic silica/iron oxide nanocomposites with fluorescent polymethacrylic acid for cancer targeting and drug delivery. J. Mater. Chem. 2010, 20, 6422–6429. 10.1039/c0jm00320d. [DOI] [Google Scholar]
- Huang G.; Zhang C.; Li S.; Khemtong C.; Yang S. G.; Tian R.; Minna J. D.; Brown K. C.; Gao J. A novel strategy for surface modification of superparamagnetic iron oxide nanoparticles for lung cancer imaging. J. Mater. Chem. 2009, 19, 6367–6372. 10.1039/b902358e. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wan X.; Song Y.; Song N.; Li J.; Yang L.; Li Y.; Tan H. The preliminary study of immune superparamagnetic iron oxide nanoparticles for the detection of lung cancer in magnetic resonance imaging. Carbohydr. Res. 2016, 419, 33–40. 10.1016/j.carres.2015.11.003. [DOI] [PubMed] [Google Scholar]
- Nafiujjaman M.; Khan H. A.; Lee Y. K. Peptide-influenced graphene quantum dots on iron oxide nanoparticles for dual imaging of lung cancer cells. J. Nanosci. Nanotechnol. 2017, 17, 1704–1711. 10.1166/jnn.2017.12825. [DOI] [Google Scholar]
- Key J.; Leary J. F. Nanoparticles for multimodal in vivo imaging in nanomedicine. Int. J. Nanomed. 2014, 9, 711. 10.2147/IJN.S53717. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yoo M. K.; Park I. K.; Lim H. T.; Lee S. J.; Jiang H. L.; Kim Y. K.; Choi Y. J.; Cho M. H.; Cho C. S. Folate–PEG–superparamagnetic iron oxide nanoparticles for lung cancer imaging. Acta Biomater 2012, 8, 3005–3013. 10.1016/j.actbio.2012.04.029. [DOI] [PubMed] [Google Scholar]
- Guo R.; Wang H.; Peng C.; Shen M.; Pan M.; Cao X.; Zhang G.; Shi X. X-ray attenuation property of dendrimer-entrapped gold nanoparticles. J. Phys. Chem. C 2010, 114, 50–56. 10.1021/jp9078986. [DOI] [Google Scholar]
- Wang H.; Zheng L.; Peng C.; Guo R.; Shen M.; Shi X.; Zhang G. Computed tomography imaging of cancer cells using acetylated dendrimer-entrapped gold nanoparticles. Biomaterials 2011, 32, 2979–2988. 10.1016/j.biomaterials.2011.01.001. [DOI] [PubMed] [Google Scholar]
- Wang H.; Zheng L.; Peng C.; Shen M.; Shi X.; Zhang G. Folic acid-modified dendrimer-entrapped gold nanoparticles as nanoprobes for targeted CT imaging of human lung adencarcinoma. Biomaterials 2013, 34, 470–480. 10.1016/j.biomaterials.2012.09.054. [DOI] [PubMed] [Google Scholar]
- Medley C. D.; Smith J. E.; Tang Z.; Wu Y.; Bamrungsap S.; Tan W. Gold nanoparticle-based colorimetric assay for the direct detection of cancerous cells. Anal. Chem. 2008, 80, 1067–1072. 10.1021/ac702037y. [DOI] [PubMed] [Google Scholar]
- Zhao Y.; Liu W.; Tian Y.; Yang Z.; Wang X.; Zhang Y.; Tang Y.; Zhao S.; Wang C.; Liu Y.; et al. Anti-EGFR peptide-conjugated triangular gold nanoplates for computed tomography/photoacoustic imaging-guided photothermal therapy of non-small cell lung cancer. ACS Appl. Mater. Interfaces 2018, 10, 16992–17003. 10.1021/acsami.7b19013. [DOI] [PubMed] [Google Scholar]
- Ho J. A. A.; Chang H. C.; Shih N. Y.; Wu L. C.; Chang Y. F.; Chen C. C.; Chou C. Diagnostic detection of human lung cancer-associated antigen using a gold nanoparticle-based electrochemical immunosensor. Anal. Chem. 2010, 82, 5944–5950. 10.1021/ac1001959. [DOI] [PubMed] [Google Scholar]
- Xia F.; Hou W.; Zhang C.; Zhi X.; Cheng J.; Jesús M.; Song J.; Cui D. pH-responsive gold nanoclusters-based nanoprobes for lung cancer targeted near-infrared fluorescence imaging and chemo-photodynamic therapy. Acta Biomater 2018, 68, 308–319. 10.1016/j.actbio.2017.12.034. [DOI] [PubMed] [Google Scholar]
- He F.; Shen Q.; Jiang H.; Zhou J.; Cheng J.; Guo D.; Li Q.; Wang X.; Fu D.; Chen B. Rapid identification and high sensitive detection of cancer cells on the gold nanoparticle interface by combined contact angle and electrochemical measurements. Talanta 2009, 77, 1009–1014. 10.1016/j.talanta.2008.07.063. [DOI] [PubMed] [Google Scholar]
- Ashton J. R.; Clark D. P.; Moding E. J.; Ghaghada K.; Kirsch D. G.; West J. L.; Badea C. T. Dual-energy micro-CT functional imaging of primary lung cancer in mice using gold and iodine nanoparticle contrast agents: a validation study. PloS One 2014, 9, 88129. 10.1371/journal.pone.0088129. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chon H.; Lee S.; Son S. W.; Oh C. H.; Choo J. Highly sensitive immunoassay of lung cancer marker carcinoembryonic antigen using surface-enhanced Raman scattering of hollow gold nanospheres. Anal. Chem. 2009, 81, 3029–3034. 10.1021/ac802722c. [DOI] [PubMed] [Google Scholar]
- Zhang J.; Fu Y.; Mei Y.; Jiang F.; Lakowicz J. R. Fluorescent metal nanoshell probe to detect single miRNA in lung cancer cell. Anal. Chem. 2010, 82, 4464–4471. 10.1021/ac100241f. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ko H. Y.; Choi K. J.; Lee C. H.; Kim S. A multimodal nanoparticle-based cancer imaging probe simultaneously targeting nucleolin, integrin αvβ3 and tenascin-C proteins. Biomaterials 2011, 32, 1130–1138. 10.1016/j.biomaterials.2010.10.034. [DOI] [PubMed] [Google Scholar]
- Chen M.; Wang Y.; Su H.; Mao L.; Jiang X.; Zhang T.; Dai X. Three-dimensional electrochemical DNA biosensor based on 3D graphene-Ag nanoparticles for sensitive detection of CYFRA21–1 in non-small cell lung cancer. Sens. Actuators, B 2018, 255, 2910–2918. 10.1016/j.snb.2017.09.111. [DOI] [Google Scholar]
- Asadzadeh-Firouzabadi A.; Zare H. R. Preparation and application of AgNPs/SWCNTs nanohybrid as an electroactive label for sensitive detection of miRNA related to lung cancer. Sens. Actuators, B 2018, 260, 824–831. 10.1016/j.snb.2017.12.195. [DOI] [Google Scholar]
- Cho M. H.; Choi E. S.; Kim S.; Goh S. H.; Choi Y. Redox-responsive manganese dioxide nanoparticles for enhanced MR imaging and radiotherapy of lung cancer. Front. Chem. 2017, 5, 109. 10.3389/fchem.2017.00109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Savla R.; Garbuzenko O. B.; Chen S.; Rodriguez-Rodriguez L.; Minko T. Tumor-targeted responsive nanoparticle-based systems for magnetic resonance imaging and therapy. Pharm. Res. 2014, 31, 3487–3502. 10.1007/s11095-014-1436-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yang S. G.; Chang J. E.; Shin B.; Park S.; Na K.; Shim C. K. 99mTc-hematoporphyrin linked albumin nanoparticles for lung cancer targeted photodynamic therapy and imaging. J. Mater. Chem. 2010, 20, 9042–9046. 10.1039/c0jm01544j. [DOI] [Google Scholar]
- Na J. H.; Koo H.; Lee S.; Min K. H.; Park K.; Yoo H.; Lee S. H.; Park J. H.; Kwon I. C.; Jeong S. Y.; et al. Real-time and non-invasive optical imaging of tumor-targeting glycol chitosan nanoparticles in various tumor models. Biomaterials 2011, 32, 5252–5261. 10.1016/j.biomaterials.2011.03.076. [DOI] [PubMed] [Google Scholar]
- Tseng C. L.; Wang T. W.; Dong G. C.; Wu S. Y. H.; Young T. H.; Shieh M. J.; Lou P. J.; Lin F. H. Development of gelatin nanoparticles with biotinylated EGF conjugation for lung cancer targeting. Biomaterials 2007, 28, 3996–4005. 10.1016/j.biomaterials.2007.05.006. [DOI] [PubMed] [Google Scholar]
- Tseng C. L.; Wu S. Y. H.; Wang W. H.; Peng C. L.; Lin F. H.; Lin C. C.; Young T. H.; Shieh M. J. Targeting efficiency and biodistribution of biotinylated-EGF-conjugated gelatin nanoparticles administered via aerosol delivery in nude mice with lung cancer. Biomaterials 2008, 29, 3014–3022. 10.1016/j.biomaterials.2008.03.033. [DOI] [PubMed] [Google Scholar]
- Robertson K. L.; Soto C. M.; Archer M. J.; Odoemene O.; Liu J. L. Engineered T4 viral nanoparticles for cellular imaging and flow cytometry. Bioconjugate Chem. 2011, 22, 595–604. 10.1021/bc100365j. [DOI] [PubMed] [Google Scholar]
- Desai N. Challenges in development of nanoparticle-based therapeutics. AAPS J. 2012, 14, 282–295. 10.1208/s12248-012-9339-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Babu A.; Templeton A. K.; Munshi A.; Ramesh R. Nanoparticle-based drug delivery for therapy of lung cancer: progress and challenges. J. Nanomater. 2013, 2013, 863951. 10.1155/2013/863951. [DOI] [Google Scholar]
- Kumar A.; Sharipov M.; Turaev A.; Azizov S.; Azizov I.; Makhado E.; Rahdar A.; Kumar D.; Pandey S. Polymer-Based Hybrid Nanoarchitectures for Cancer Therapy Applications. Polymers 2022, 14, 3027. 10.3390/polym14153027. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hubbs A. F.; Mercer R. R.; Benkovic S. A.; Harkema J.; Sriram K.; Schwegler-Berry D.; Goravanahally M. P.; Nurkiewicz T. R.; Castranova V.; Sargent L. M. Nanotoxicology—A pathologist’s perspective. Toxicol. Pathol. 2011, 39, 301–324. 10.1177/0192623310390705. [DOI] [PMC free article] [PubMed] [Google Scholar]
- McCarthy J.; Inkielewicz-Stępniak I.; Corbalan J. J.; Radomski M. W. Mechanisms of toxicity of amorphous silica nanoparticles on human lung submucosal cells in vitro: protective effects of fisetin. Chem. Res. Toxicol. 2012, 25, 2227–2235. 10.1021/tx3002884. [DOI] [PubMed] [Google Scholar]
- Sharma A.; Tonk R.; Shekhar R.; Dohare S.; Kumar D. Need to focus on inhibitory activity of benzimidazole analogues against indolamine 2,3-dioxygenase-1 (IDO-1). EXCLI J. 2022, 21, 904–905. 10.17179/excli2022-4988. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang W.; Chen Y.; Wei H.; Zheng C.; Sun R.; Zhang J.; Tian Z. Antiapoptotic activity of autocrine interleukin-22 and therapeutic effects of interleukin-22-small interfering RNA on human lung cancer xenografts. Clin. Cancer Res. 2008, 14, 6432–6439. 10.1158/1078-0432.CCR-07-4401. [DOI] [PubMed] [Google Scholar]





