Next Article in Journal
A New Glucosyl Flavone with Inhibitory Activity of Cancer Cell Viability and Other Bioactive Constituents from the Traditional Kurdish Plant Plantago loeflingii L.
Previous Article in Journal
Illuminating the Tiny World: A Navigation Guide for Proper Raman Studies on Microorganisms
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Adjuvant Novel Nanocarrier-Based Targeted Therapy for Lung Cancer

1
School of Pharmaceutical and Population Health Informatics (SoPPHI), DIT University, Dehradun 248009, India
2
Department of Clinical Laboratory Sciences, College of Applied Medical Sciences, King Khalid University, Abha 62521, Saudi Arabia
3
Department of Pharmaceutical Chemistry, College of Pharmacy, Prince Sattam bin Abdulaziz University, Al-Kharj 11942, Saudi Arabia
4
Smart-Health Initiative (SHI) and Red Sea Research Center (RSRC), Division of Biological and Environmental Sciences and Engineering (BESE), King Abdullah University of Science and Technology (KAUST), Thuwal 23955, Saudi Arabia
5
Core Labs, King Abdullah University of Science and Technology (KAUST), Thuwal 23955, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Molecules 2024, 29(5), 1076; https://doi.org/10.3390/molecules29051076
Submission received: 25 May 2023 / Revised: 19 July 2023 / Accepted: 25 July 2023 / Published: 29 February 2024

Abstract

:
Lung cancer has the lowest survival rate due to its late-stage diagnosis, poor prognosis, and intra-tumoral heterogeneity. These factors decrease the effectiveness of treatment. They release chemokines and cytokines from the tumor microenvironment (TME). To improve the effectiveness of treatment, researchers emphasize personalized adjuvant therapies along with conventional ones. Targeted chemotherapeutic drug delivery systems and specific pathway-blocking agents using nanocarriers are a few of them. This study explored the nanocarrier roles and strategies to improve the treatment profile’s effectiveness by striving for TME. A biofunctionalized nanocarrier stimulates biosystem interaction, cellular uptake, immune system escape, and vascular changes for penetration into the TME. Inorganic metal compounds scavenge reactive oxygen species (ROS) through their photothermal effect. Stroma, hypoxia, pH, and immunity-modulating agents conjugated or modified nanocarriers co-administered with pathway-blocking or condition-modulating agents can regulate extracellular matrix (ECM), Cancer-associated fibroblasts (CAF),Tyro3, Axl, and Mertk receptors (TAM) regulation, regulatory T-cell (Treg) inhibition, and myeloid-derived suppressor cells (MDSC) inhibition. Again, biomimetic conjugation or the surface modification of nanocarriers using ligands can enhance active targeting efficacy by bypassing the TME. A carrier system with biofunctionalized inorganic metal compounds and organic compound complex-loaded drugs is convenient for NSCLC-targeted therapy.

Graphical Abstract

1. Introduction

According to the World Health Organization (WHO), lung cancer was the second-highest diagnosis (11.4%) and the leading cause of death (18%) among all forms of cancer in 2020 [1]. The 5-year survival rate of lung cancer patients worldwide was 19% from 2010–2014. A few countries, such as Japan (33%), Israel (27%), and the Republic of Korea (25%), had a higher survival rate [2]. The prime reasons for the low survival rate are late-stage diagnosis, lack of awareness, socioeconomic conditions, environmental contamination, and the metastatic and intra-tumoral heterogeneity nature of the tumor [3,4,5,6].
The common etiological factors for lung cancer development are tobacco smoking (which causes 80% of cases in the United States and other countries), occupational asbestos exposure (5–10% globally), cannabis or marijuana smoking (4% in the USA in 2002–2014), radon exposure (10% in the Western World), air pollution, group 1 carcinogen arsenic exposure, inflammation and cellular damage during respiratory infection, chronic obstructive pulmonary disease-related inflammation and scarring, and family history of lung cancer [7,8,9,10,11,12,13,14,15]. Other associated increased risk factors for lung cancer are systemic sclerosis patients, smoker breast cancer survivors, HIV-infected patients with idiopathic pulmonary fibrosis, certain fibrotic pneumoconioses patients, and lung cancer survivors.
The etiological agents such as radon, smoke, and cannabis have free radicals, reactive oxygen species, and reactive electrophiles [16,17,18]. Depending on the dose, dimension, durability, and surface reactivity that react with nitrogen and oxygen atom lesions in the DNA, modifying nucleotides to distort the base pairing leads to incorrect nucleotide incorporation during replication [19,20,21,22,23,24]. Cell repair mechanisms can repair DNA damage. But the escape portions change the coding of the DNA. Repeated exposure to etiological factors leads to a series of genomic changes such as copy number variations (CNVs), single-nucleotide variations (SNVs), and insertions/deletions (INDELs) of exomes in the autosomal chromosome that lead to a permanent change in the sequence and that start from the primary tumor, followed by metastasis via circulating tumor cells [25,26,27]. Genetic mutations affect protein synthesis, disrupt cell cycle progression, and promote carcinogenesis. Circulatory tumor cells for metastatic cancer and the genomics of the tumor cells for the non-invasive type help in diagnosis and prognosis. Circulating tumor cell analysis is helpful for the prediction of disease progression, the survivability of patients, and personalized therapy because cell-free DNA fragments are found in peripheral blood [24,25,26,27,28,29,30]. Lung cancer occurs through either one or a combination of the etiological factors that affect and cause mutation of protooncogenes, tumor suppressor genes, DNA repair gene dysfunction, apoptotic mechanism erosion, limitless telomere replication, sustained angiogenesis, increment of invasion, metastasis, and escape from immunity [30,31,32,33,34,35].
Histologically, lung cancer is classified into non-small cell lung cancer (NSCLC, 85%), and small cell lung cancer (SCLC, 13%) [36]. Further, NSCLC subdivides into lung adenocarcinoma (40%), squamous cell carcinoma (25–30%), and large cell carcinoma (5–10%) [34,35,37]. In 2015, the WHO modified the classification of lung cancer based on immunohistochemistry, genetic studies for personalized treatment strategies, and small biopsy and cytologic samples [36,38,39]. This new classification objective is to overcome drug resistance, intracellular accumulation, metastasis, invasion, side effects, and toxicity, and develop a more personalized novel treatment regime [40]. The current treatment regime depends upon the stage of cancer progression, the health of the patients, and affordability at the time of diagnosis. The different treatment methods are surgery (wedge resection, segmental resection, lobectomy, and pneumonectomy), radiation therapy, chemotherapy, stereotactic body radiotherapy, targeted drug therapy, immunotherapy, and palliative care.
Surgery remains the top priority for multimodality cancer patients with advanced stages (III and IV). It is better suited for stage I and II lung cancer because surgery eliminates it before it has spread to lymph nodes or distant sites. Neo-adjuvant therapy is commonly used before surgery to reduce tumor size. The common side effects of lung cancer surgery are severe chest pain, blood clots, bronchopleural fistula, excessive bleeding, collapsed lungs, depression, difficulty breathing, fatigue, loss of appetite, gastrointestinal problems, heart problems, and a sore mouth [41,42,43].
Radiation therapy acts as both a primary and adjuvant therapy for cancer patients by damaging the DNA of the cancerous cell in a dose-dependent manner. It can generate antitumor T cells by activating the stimulator of the interferon genes (STING) pathway to produce interferon-1. Interferon-1 can deliver the DNA fragment to dendritic cells to produce antitumor T cells. Radiation therapy also helps in the trafficking of chemokines by homing T cells to the TME and modulating immunity [44,45,46]. The adverse effects of radiation therapies include damage to epithelial surfaces, intestinal discomfort, inflammation, infertility, fibrosis, epilation, lymphedema, polyneuropathy, and dryness [47].
Chemotherapy is another prominent therapy to control the growth of cancer cells. It can be used before and after surgery in NSCLC patients and with targeted or radiation therapy in the late stage of cancer. Excessive toxicity makes it controversial regarding the effective use of chemotherapeutic agents in lung cancer treatment. Chemotherapeutic agents can damage the DNA or RNA of cancer cells to inhibit their reproduction. The common adverse effects of chemotherapy are nausea, vomiting, sore mouth, weight change, and hair loss [48,49].
Stereotactic body radiotherapy (SBRT) is a neo-adjuvant or adjuvant therapy for operable patients with co-existing interstitial lung disease. It is an alternative therapy for non-operable early-stage localized NSCLC patients. Using target delineation, motion management, image guidance, and dose optimization in treatment planning is possible with this technique. The reported toxicity level for this treatment regime is lower. The adverse effects of SBRT are shortness of breath, chest wall pain, urinary irritation, nausea, and vomiting [50,51,52].
Targeting therapy is designed to alter the specific abnormalities in the cancer cells and their microenvironment. This therapy acts as an adjuvant in the early as well as late stages of the disease’s progression. It involves targeting specific genes or proteins using a drug-loaded carrier system to deliver them to a projected site. A modification of the carrier system enhances the efficacy of the drug at the targeted site. The limitations of conventional therapy can be overcome by using targeted drug delivery systems. It may cause site-specific nano-toxicity and minimal toxicity to surrounding cells. Optimization of targeted drug delivery is one of the biggest challenges [40,53,54,55,56].
Immunotherapy is a treatment regime to identify optimal neoantigen candidates and inhibit the functions of tumor cells. It produces higher effectivity when tumor-neoantigen reaches the dendritic cell and is recognized by the T-cell receptor. Then, it circulates through the periphery, overcomes the immunosuppression of the tumor cell, causes stimulation within the tumor cell, and finally releases the potent molecule to kill the tumor cells. This treatment regime can be used as adjuvant therapy after surgery with chemotherapy, radiation, hormones, and targeted therapy. Fatigue, cough, nausea, itching, skin rash, constipation, diarrhea, and joint pain are a few common adverse effects of immunotherapy [57,58].
Palliative care is an adjuvant therapy to enhance the quality of life for patients with severe diseases such as lung cancer. It focuses on the physical, psychological, spiritual, and practical burdens after a disease diagnosis. It continues with surgery, radiotherapy, and immunotherapy [59].
Though the advancement of the treatment regime impacted the treatment profile, late-stage diagnosis (metastasis stage) creates a burden [3,4,5]. So, the emphasis has increased on chemotherapy and pathway-blocking agents through targeted drug delivery systems for advanced-stage patients [60].
Nanocarriers, a colloidal preparation with a higher number of pores, can be used for the diagnosis and delivery of targeted drugs, nucleic acids, proteins, and diagnostic agents at the desired rate and time to the targeted site through passive, active targeting, pH, and temperature specificity to block pathways and reduce systemic drug toxicity [61,62,63]. As a result of NPs’ small size, tailored surfaces, improved solubility, and multifunctionality, NPs provide superior stability, solubility, and bioavailability. It delivers the magnetic, thermal, electrical, and optical forms of active pharmaceutical ingredients used as targeted radiational, chemotherapeutic, gene therapeutic, immunotherapeutic, and combinational agents to treatment sites through the EPR effects. Depending upon the types, nature, and intention of the use of drugs, they are encapsulated, entrapped, dissolved, or absorbed in nanocarriers. Nanocarriers, a circulating cargo, can enhance the circulation lifetime, permeability, and retention of active pharmaceutical ingredients [64,65]. Viral vector nanocarriers can deliver nucleic acid therapies [66]. In short, a nanocarrier is a system that can control, manipulate, and fabricate micron-sized structures and devices.
Optimization of the physical properties of NPs facilitates the delivery of drugs at a specific rate and time to the desired sites. In addition to protecting the active medicament from premature degradation, nanocarriers control and improve drug distribution through intracellular accumulation, penetration, and shelf life [67,68]. Nanobiocarriers are bioactive or targeting vectors or ligands that deliver the active pharmaceutical moiety to mimic and control unnecessary cellular extravasations, growth, and cellular events. Additionally, nanobiocarriers enable the delivery of drugs with optimal biocompatibility, biointeraction, safety, and efficacy [69].

2. Drug Delivery Constraints in Lung Cancer Management

The self-defense mechanism of the respiratory tract impacts drug delivery and absorption on the lung surface through mechanical, chemical, and immunological barriers [69]. A pulmonary drug delivery method can overcome biological barriers by avoiding the cough reflex, using aerosol, interacting with airway surface liquid and mucus to reach cellular targets, overcoming systemic absorption, degradation, or clearance of the active molecule to reach intracellular targets, and maintaining drug concentrations within therapeutic windows until the next administration [70].
However, in lung cancer, the biological barriers act differently, and tumor heterogeneity also adds instruments [69]. Different attempts have been made to solve these problems. The control of tumor endothelia and sprouting angiogenesis are two such attempts. Generally, sprouting angiogenesis is caused by the vascular endothelial growth factor’s (VEGF) action on its receptor. So, selecting a drug that can access VEGF can enhance the vascular permeability of chemotherapeutics [71]. Another method is to control vascular co-option. Vascular co-option is a process of tumor cell interaction and exploitation with normal tissue vasculature by migrating through the host tissue’s blood vessels in a VEGF non-dependent manner. Clinical tumor-coopted vessels can be identified through specific functional biomarkers or molecular markers [71]. The other ways to overcome the biological barriers are to control vasculogenic mimicry, intussusceptions, and vasculogenesis. Using high intestinal fluid pressure in the tumor neovasculature may be another approach to overcome the barriers because negative or low transcapillary pressure gradients of normal tissue enable outer flow into its tissues. By passing the heterogenic biological barriers of cancer, paracellular transport across the endothelium can transport chemotherapeutic fluids by maintaining concentration gradients. Furthermore, transcellular routes allow molecules to travel through endothelial cells and are an alternative strategy for cancer drug delivery that needs to be considered by current therapeutics. Again, normalization of vessels, vascular promotion, and tumor-specific peptide incorporation with nanocarriers can improve passive drug delivery into tumors. Finally, active transvascular delivery using caveolae pumping may be another way to overcome biological barriers [71].
Targeted therapy is the formulation approach to overcome the lung surface barriers to the targeted site. The lung-targeted therapy enhances the safety and efficacy of the drugs. Examples include topical targeted lung therapy, which acts rapidly in low doses with no systemic side effects. A systemic targeted lung therapy bypasses the gastrointestinal tract and has a better pharmacokinetic profile [69]. The relationship between the active drug’s physicochemical properties and its biological functions affects the development of targeted therapy and treatment profiles [72]. The particle size distribution of active pharmaceutical ingredients, or nanocarriers, is an example. Particles of larger size deposit more in the lung’s central airways, whereas fine particles deposit more in the peripheral airways [73]. The geometrical shape of the nanocarriers affects drug delivery to the neovasculature. Geometrical shapes influence dissolution/diffusion rates and the release of drugs [74]. It also impacts the circulation, margination, adhesion, and internalization of the nanocarriers and active pharmaceutical ingredients. Generally, cylindrical and spherical dispersed drugs have a lower dissolution/diffusion ratio, and the solubility/drug loading ratio is lower, which enhances drug release. A drug’s release pattern may vary based on its geometric shape. Again, surface charges are the distribution of the drug carriers. In addition to surface charges and geometrical shapes, surface modifications can improve drug delivery effectiveness in nanocarriers [52,53]. A modified cuboidal cyclodextrin metal–organic framework is one example of delivering targeted drugs to injured blood vessels [75]. In addition, comparing the ratio between the metalloproteinase and its inhibitor in the targeted site and consequently choosing the type of NP drug delivery system can enhance the vascularization of the NPs [76,77,78]. The development of the targeted drug delivery system after the physicochemical properties are modulated according to biological functions can optimize the delivery of the active drug to the lung cancer site.
The mononuclear phagocytes of the immune system reduce the reach of nanotherapeutics through opsonization and sequestration processes. It occurs in a protein corona around NPs using the opsonization and sequestration processes. The formation of protein corona depends upon the size and surface chemistry of the NPs. After protein corona formation, it absorbs the NPs, internalizes them, fuses them to the lysosomes, and reduces their specificity [65,78,79,80,81,82,83].
In this study of nanocarrier-based targeting drug delivery to overcome the TME barriers, we have found that particle size and active targeting using receptor-based bioconjugating agents play roles in bypassing TME to enhance the targeting efficacy of the loaded drug. Different studies on lung cancer have found that inflammatory mediators are overexpressed, especially IL-6 [84]. A high-affinity protein can block it. So, nanocarriers biofunctionalized with proteins such as RGD can be useful for targeting cancer cells, and the results of a few studies also complement them [85]. Another study found that folic acid deficiency promotes IL-6/JAK-1/pSTAT3 interactions in astrocytes after ischemia-reperfusion. So, folic acid-biofunctionalized nanocarriers may be another approach to improving targeting precision [86]. Further, the PEGylation of NPs reduces the interaction with serum proteins and enhances the stability of the nanocarrier in the reticuloendothelial system [87].
Functionalizing the NPs according to the targeted profile through the intravenous route can decrease the circulation’s lifetime. Using self-peptide, biomimetic particle coating, and conformation-changing coating molecules on the surface of the NPs can overcome this biological barrier [65,88,89,90].
Cancer cells chisel their TME using different factors, such as the release of chemokines and cytokines. These secretions reprogrammed the environment for further tumor growth and disease progression. NPs can passively and limitedly reach the TME through the EPR effect. The tumor heterogeneity acts as a barrier for drug delivery to the TME through uncontrolled vascular events, resistance produced by the stroma, hypoxia, pH, and immune reshaping. For stable drug delivery to the targeted TME, there is a need for favorable vascular network events, regulation of stromal activities, or manipulation of hypoxia, pH, and immunity. In a heterogenic TME, the incremental demand for nutrients increases growth factors and forms leaky vessels. It increases the angiogenesis of tumor cells. It also enhances the interstitial fluid pressure through the leaky vessels and decreases blood flow to the site. So, the drug-loaded NPs cannot reach and accumulate in the targeted space [91,92,93,94,95,96,97,98,99,100,101]. In lung cancer treatment, multiple drug resistances decrease the effectiveness of the treatment regime. A combination of medications for respiratory tract disease changes the compliance rate of the drugs. The modulation of the TME using a single drug therapy with multiple targeting strategies can overcome these issues [92,102,103,104]. A few strategies to optimize drug delivery to the TME are active targeting, TME modulation, and TME-responsive targeted drug delivery [94,101,105].
Nanocarriers also have potential risks and downsides. Nanocarriers can trigger an immune response, leading to inflammation or hypersensitivity reactions. This immune response can be more pronounced if the nanocarriers are derived from foreign materials or have surface properties. Mitigation strategies include using biocompatible and biodegradable materials for nanocarriers or modifying the surface of the nanocarriers to minimize immune recognition [106]. Nanocarriers are designed for targeted delivery but can accumulate in non-target tissues or organs. This can result in off-target effects and potential toxicity. Preclinical studies and careful formulation design can help minimize this risk by optimizing the specificity and stability of the nanocarriers [107]. Nanocarriers may exhibit inherent toxicity if not adequately eliminated from the body. Rigorous toxicity evaluations and optimization of nanocarrier properties, such as size, surface charge, and composition, can help mitigate this risk. Nanocarriers can also experience drug leakage or premature release of the therapeutic payload before reaching the target site. This can result in suboptimal drug concentrations at the intended site and reduce its efficacy. Strategies such as improved encapsulation techniques, surface modifications, or utilizing stimuli-responsive nanocarriers can help minimize premature drug release. Targeted therapy using nanocarriers can be affected by drug resistance mechanisms and the heterogeneity of lung cancer tumors. Combining nanocarrier-based therapy with other treatment modalities or developing strategies to address drug resistance can help overcome this limitation [106,107].

3. Nucleic Acid Role in Lung Cancer Management

Cancer is an acquired disease of genetic alteration. Nucleic acids have a promising treatment profile for cancer. This genetic alteration can be improved using the delivery of DNA and other nucleic acids to control the genetic expression profile of target cells. The delivery of nucleic acid to the targeted cell is challenging due to its instability, off-target effects, and traversal of biological barriers [108].The delivery of nucleic acids to the targeted site can be achieved using a nucleic acid cargo or nanocarrier as the nucleic acid vehicle. A nanocarrier charge can deliver DNA or mRNA to overexpress a gene, small interfering RNAs or microRNAs to knock down a gene, or nucleic acids to trigger pattern-recognition receptors to stimulate the immune system [109]. A plasmid containing both a promoter and the gene of interest is used to treat DNA overexpression by bypassing the plasma membrane and nuclear envelope. After reaching the nucleus, it exports and transcribes into mRNA, which is translated into the desired protein in the cytoplasm. Single-stranded mRNA can also be used for the same purpose, but it is less stable and has a lesser chance of undesired insertion into the genome, such as plasmid DNA, to cause mutagenesis [108,109,110]. However, RNAi can interrupt mRNA translation to decrease target gene expression, and this problem can be solved using short-length dsRNA such as siRNA. Although the sequence of nucleic acids can have functional impacts on biological targets, many physical and chemical considerations are not heavily dependent on the nucleic acid sequence encapsulated in a nanocarrier for delivery. So, the chemical and physical properties of the nucleic acid should be considered [108,109,110].
Co-delivery of multiple nucleic acids of the same type but with different sequences in a single delivery vehicle follows the same design principles, necessitating changes to nanocarrier design to deliver distinct cellular and subcellular locations. Again, tumor heterogeneity and MDR cause multiple therapeutic agents to target different cellular pathways. However, the multitargeted nucleic acid cargo can cause intrinsic toxicity and virus immunogenicity to prevent repetitive administration [108,109]. The challenges of nucleic acid cargo are the physical and chemical properties’ similarity and the overlap of extracellular and intracellular trafficking routes. As nucleic acids possess a negative charge in their structure, generally positively charged polymers can be used to prepare NPs. Cationic polymers such as poly(l-lysine), polyethyleneimine, polyamidoamine, poly(beta-amino ester), and cationic lipids are used [109]. Again, the size and physical properties of the nucleic acid impact its loading on nanocarriers. Further, surface modification of the nucleic acid NP improves its cellular uptake at the targeted site. Commonly used nanocarriers for nucleic acid delivery are liposomes, SLN, polymeric, gold, mesoporous silica, and iron oxide NPs [111].

4. Strategies to Overcome the Tumor Microenvironment

4.1. Vascular Remodulation

The vascular network events can be modified using either disrupting agents or normalizing them (Figure 1 and Figure 2). Disrupting vascular events differs from anti-angiogenesis strategies, and when used with chemo, radio, and angiogenesis-inhibiting therapeutic agents, it enhances treatment efficacy (Figure 1). Recently, researchers reported that the co-administration of combretastatin A4 (CA4) NPs with doxorubicin, CA4 NPs with imiquimod, and nanocomposite hydrogel antitumor therapy with near-infrared radiation enhances the treatment efficacy [105,112,113,114,115,116]. Vascular normalization is another approach to re-modulate vascular events that enhances the effectiveness of chemotherapy, radiation therapy, and immunotherapy by reducing tumor invasion and metastasis (Table 1). In addition, the normalization process targets endothelial cell metabolism, microRNA, and extracellular matrix and balances proangiogenic and antiangiogenic factors [117,118]. Anti-VEGF-receptor-2 antibody DC101 modulates NPs, and nitric oxide deliveries with nanocarriers are a few reported approaches that cannormalize vascular events [112,118,119,120,121].

4.2. Stromal Regulation

In cancer, stromal cells lose their tumor-suppressing abilities and promote tumor growth, invasion, and metastasis. Stroma regulates ECM synthesis, degradation, and signaling pathways (Figure 3) [122]. So, different attempts have been made to enhance stromal regulation.
Researchers reported that metelimumab (a transforming growth factor-β ligand-blocking antibody) conjugate NPs and prolyl-4-hydroxylase inhibitor (which inhibits collagen synthesis in vascular smooth muscle cells) conjugate NPs enhance drug effectivity through ECM synthesis through stromal regulation [123,124,125,126,127,128]. Further research has indicated that volociximab inhibits angiogenesis by preventing integrins from binding to fibronectin (Table 1). Volociximab can be more effective when combined with other tumor-mimicking drugs based on its ability to modify ECM signaling [123,129]. Evidence suggests hyaluronidase, collagenase, and putrescine inhibition can destroy the ECM [122,130,131]. The preparation of an artificial extracellular matrix (AECM) based on laminin (LN)-mimic peptide hydrogel-fabricated NPs loaded with a drug can enhance effectivity and mimic the ECM [92,132,133]. Another factor that changes the stroma at the invasion front is carcinoma-associated fibroblasts (CAFs). This problem can be solved by either directly disrupting or reducing CAF activity. Applying the disruption strategy, researchers reported that artemisinin (which inhibits vimentin expression) as a capping agent for an NP loaded with therapeutics treats the invasion process. In general, vimentin expression increases the migration and invasion of cancer cells. Further, fibroblast activation protein was overexpressed in the stroma. N-(4-quinolinoyl)-Gly-(2-cyanopyrrolidine) capping NPs (which inhibit FAP overexpression) with the active drug may regulate stroma [134,135,136,137,138]. Reprogramming the CAFs can delay the disease’s progression. CAFs act as either immune suppressive or supportive agents (Table 1). Angiogenic receptor blockers (ARBs) latently inhibit CAF activity, and ARB nanoconjugates can enhance immune-supportive activity [139,140].

4.3. Hypoxia Manipulation

Hypoxia is the normal state in the TME that changes the metabolic pathways. It activates hypoxia-induced factors to regulate energy demand through anaerobic glycolysis. Anaerobic glycolysis increases lactate production, and H+ ions lead to acidosis in the TME [92,96,100,101]. It decreases drug compliance and increases resistance and angiogenesis. Hypoxia can be manipulated through an elevation or decrease in oxygen consumption or by using hypoxia-activated prodrugs [92]. According to the literature, oxygen supply elevates with theranostic upconversion nanoprobe MnO2 NPs. The tumor cell produces excessive amounts of H2O2 and lactic acid. Theranostic MnO2 reacts with acidic H2O2 to produce Mn2+ and enhance O2 production [141,142]. Encapsulating photothermal therapy with electron-transporting chain-inhibiting agents using NPs reduces oxygen consumption [140]. Further, hypoxia-activated prodrug (HAP) activates spontaneous electron oxidoreductases (Table 1). HAP agents combined with targeted therapy and checkpoint blockers increase the influx into the hypoxic zone [143,144,145].

4.4. pH Manipulation

A variation in the external to internal pH of tumor cells causes disturbances in biological functions, such as proliferation, migration, and aggression. pH can be manipulated using small-molecule drugs, acidity-neutralizing inhibitors, or pH-regulating enzymes [146]. One reported mechanism for pH manipulation is acid neutralization using sodium potassium citrate, which increases the blood HCO3− level in the oral dose and neutralizes the TME pH [147,148]. Proton pump inhibitors and carbonic anhydrase IX/XII can be used to inhibit the enzymes that manipulate pH [149,150]. In a study, researchers found that pH-dependent dendritic polyglycerol-co-polycaprolactone-derived polymer NPs loaded with gemcitabine are stable at pH 7.4 and can release the drug in a time-dependent manner with improved cellular uptake in the desmoplastic stroma of pancreatic cancer [151].

4.5. Immunity Modulation

Tumor-associated macrophages and regulatory T cells of the TME suppress and escape the immune system. As the TME hinders the trafficking of CD8+ cells, the immune system cannot inhibit the tumor. They decrease nano-therapeutic compliance and increase nano-therapeutic resistance [92,102,103,104,105]. Regulatory T-cell (Treg) inhibition, Tyro3, Axl, and Mertk receptor (TAM) regulation, and myeloid-derived suppressor cells (MDSC) inhibition are a few strategies to overcome immunity. TAM overexpression increases cell survival and decreases apoptotic signaling. Again, TAM down-streaming promotes metastasis via migration and invasion. Polarization of macrophage M1 to M2 by TAM has an immunosuppressive effect. M2 releases immune-suppressive cytokines. M2 blocking agents and M1 reprogramming agents can regulate this immunity suppression. Small-molecule tyrosine inhibitors and TAM receptor-targeted ligands can regulate it [152,153,154,155]. Treg cells regulate T-cell immune responses to maintain cell homeostasis. Adversely, in TME, Treg cells decrease the entry of T cells. Transforming growth factor-β (TGF-β) inhibitors and anti-PD-L1 antibodies reduce the TGF-β signal to promote T-cell infiltration into the TME (Table 1). The cytotoxic T-lymphocyte-associated protein-4 (CTLA-4) antibodies remove the Treg cells and can enhance T-cells’ functions [156]. MDSC induces immune suppression by inhibiting T-cell, NK-cell, and macrophage functions. Targeting and inhibiting phosphatidylinositol 3-kinase (PI3K)δ, PD-L1, CTLA-4, and multi-kinase MDSC can control the same [157,158]. According to a clinical update by the US Patent Office, researchers have found that anti-CTLA-4 antibodies antagonize CTLA-4 to control T-cell functions for controlling immunity [159]. Recently reported research has found that a small-regularity self-replicating RNA intradermal vaccine against SARS-CoV-2 elicits predominantly cellular immunity because of T-cell induction [160].In another study, researchers found that pembrolizumab, as a monotherapy, enhances disease regression with minimal toxicity. Moreover, no researcher has reported a targeted drug delivery system for immunotherapy until now [161,162].

4.6. Active Targeting

Surface and biomimetic modification of NPs improves the active targeting strategy. As a form of active or ligand-mediated targeting, NPs ligation affinity on their surface ensures that they are retained and taken up by the targeted cells. Ligands (antibodies, proteins, peptides, nucleic acids, sugars, and small molecules such as vitamins) are selected to target surface molecules or receptors overexpressed in diseased organs, tissues, cells, or subcellular domains to benefit from active targeting. Activelytargeted materials must be specific to their targets. This approach aims to enhance NP–cell interactions and drug internalization without affecting overall biodistribution ortoxicity to normal cells [163]. These ligands can recognize and bind to specific receptors or markers on the target cells, facilitating selective uptake and enhancing the targeting efficacy of the NPs along with their specificity, stability, and interaction with target cells or tissues [163]. This strategy bypasses TME and synergizes both active and passive targeting. Cancer cell membrane-coated NPs can carry antigens and drugs to the target [164,165]. Protein-conjugated NPs can leverage both active targeting (targeting ligand-mediated) and passive targeting (based on NP properties, such as size and surface charge) mechanisms. This synergy can enhance targeting efficacy and improve therapeutic outcomes [163,164,165]. Depending onthe choice of protein, conjugation method, and NP characteristics, the targeting of protein-conjugated NPs can be affected. The selection of appropriate proteins and optimization of conjugation strategies should be based on understanding the target cells, tissues, and disease context. Researchers found that folate discs enhanced permeability and photothermal efficacy. So, surface modification with folate can be a better therapeutic approach [153,166]. Recently, researchers have found that folic acid-conjugated chitosan NPs loaded with 5-fluorouracilshow higher cytotoxicity than chitosan 5-fluorouraciland enhance the targeting of tumor cells [167]. Research and development of protein-conjugated NPs are ongoing, aiming to improve their targeting efficacy, therapeutic potential, and clinical translation. In lung cancer patients, researchers have found that protein-functionalized lipid hybrid NPs in response to transferrin exhibit improved therapeutic efficacy when loaded with cis-platin and docetaxel [168]. In another study, researchers reported that cyclic RGD-functionalized cyclodextrin metal–organic NPs loaded with doxorubicin enhanced the drug’s efficacy by 4–5 times. As a result, it exhibits transferrin-dependent targeting that reduces off-targeting [85]. One recent study reported that PLGA NPs functionalized with RGD control the loaded drug delivery and enhance its efficacy [169].

4.7. Tumor Environment Responsive Drug Delivery

The stimuli of tumor tissue differ physically and biochemically from those in normal tissues. To improve the effectiveness of the drug delivery to the TME, the nanocarrier must penetrate or should have improved cellular uptake or enhanced drug release at cancer sites [92,170]. Newly designed stimuli-responsive drug delivery systems can overcome the barrier of TME and deliver the required amount of drug to the targeted sites. One such method is a supramolecular architecture based on peptides. In response to the TME, supramolecular architectures based on peptides can convert structurally and allow therapeutics for controlled release. This dissertation emphatically introduces peptide assemblies with a stimulus-responsive structural conversion to acids, high temperatures, and high oxidative potentials in tumor tissues. Functional moieties that respond to cellular stimuli such aspH, glutathione, adenosine triphosphate, reactive oxygen species, enzymes, and inflammatory factors can act as targeting strategies [171,172,173]. Further, the dense ECM prevents larger NPs (>100 nm) from penetrating the tumor after they have extravagated from vessels. After blood circulation, particles with a relatively large size can shrink in size due to internal stimuli such as enzymes, acidic pH, and hypoxia. Using peptides or favorable ligands, NPs entrap and form corona at the TME. This strategy helps to internalize the entrapped drug by interacting with the moiety and reducing its size by engulfing it with TME [170,174,175,176,177,178,179].
The strategy to enhance cellular uptake can be the conversion of charge or the detachment of the NP shell. It eliminates long circulation times and cellular uptake by modifying its surface charge. TME cues such as redox potential, acidic pH, and overexpressed enzymes can stimulate the responsive bonds on the surface of NPs to achieve charge reversal (Table 1). A few examples are pH-sensitive PEG coatings and enzyme-sensitive nanovectors [180]. As nanovectors accumulate at tumor sites via the EPR effect, overexpressed MMP-9 can detach the PEG corona to expose RGD and facilitate cellular internalization. The PEG coating cleaves with the TME and removes the PEG shell to enhance cellular uptake [92].
Another strategy is that on-demand drug release may result from a hydrophilic–hydrophobic switch triggered by TME signals (Table 1). As a result of the protonation and de-protonation polymers present in the NPs, they can switch from hydrophilic to hydrophobic and trigger drug release at the targeted sites. Cleavage with a sensitive linker to TME can trigger the rapid release of an entrapped drug from a nanocarrier [181,182].
Here is a brief overview of strategies to overcome drug delivery challenges in tumor microenvironments (Table 1).
Table 1. Strategy, process, and mechanism with a few examples to overcome the barrier of TME.
Table 1. Strategy, process, and mechanism with a few examples to overcome the barrier of TME.
StrategiesProcessMechanismExampleRef.
Modulation of TMEVascular network remodulationVascular network disruption and decompressionCo-administration of combretastatin A4 (CA4) NPs with doxorubicin, CA4 NPs with Imiquimod, nanocomposite hydrogel antitumor therapy, and near-infrared radiation[112,113,114,115,116]
Normalizing the vascular networkAnti-VEGF-receptor-2 antibody DC101 modulates NPs andnitric oxide delivery with nanocarriers[105,117,118]
Regulation of stromaExtracellular matrix (ECM) targeting[92,94]
ECM synthesisMetelimumab (transforming growth factor-β ligand-blocking antibody) conjugate NPs can enhance loaded drug effectivity.
Prolyl-4-hydroxylase inhibitors (which inhibit collagen synthesis in vascular smooth muscle cells) conjugate NPs can enhance drug effectivity
[123,124,125,126,127,128]
ECM degradationInhibition of hyaluronidases, collagenase enzymes, and putrescine regulates ECM degradation. Conjugating these particles into the loaded NPs can enhance the drug’s effectivity[122,130,131]
ECM signalingVolociximab inhibits angiogenesis by interfering with integrin α binding with fibronectin in tumor vasculature. Co-administration of volociximab with other tumor-mimicking drugs can be a more effective therapeutic target[123,129]
ECM mimickingPreparing artificial extracellular matrix (AECM) based on transformable laminin (LN)-mimic peptides and hydrogel-fabricated NP-loaded drugs can enhance effectivity[92,132,133]
Reducing cancer-associated fibroblast (CAF) activity[92]
CAFs disruptionVimentin expression increased the migration and invasion of cancer cells. Preparing artemisinin (which inhibits vimentin expression) as a capping agent for the NPs loaded with drugs can be useful.
Fibroblast activation protein is overexpressed in the stroma. N-(4-quinolinoyl)-Gly-(2-cyanopyrrolidine)-capping NPs inhibit FAPs overexpression with the active drug, which may be useful to regulate stroma
[134,135,136,137,138]
Reprogramming CAFsCAFs act as either immune suppressive or supportive agents. Angiotensin receptor blockers (ARB) reduce latent CAF activity. ARB nanoconjugates can enhance immune-supportive activity[139,140]
Hypoxia manipulationOxygen supply elevationUsing theranostic conversion nanoprobe MnO2 NPs. In the tumor cell, excessive amounts of H2O2 and lactic acid are produced. Theranostic MnO2 reacts with acidic H2O2 and produces Mn2+ and enhanced O2 production[141,142]
Decreases oxygen consumptionEncapsulating photothermal therapy with electron transport chain hindering agents through NPs[92,140,145]
Using hypoxia-activated prodrugs (HAP)HAPs are activated through spontaneous electron oxidoreductases. HAP agents combined with targeted therapy with checkpoint blockers increase the influx into the hypoxic zone[143,144,145]
pH manipulationAcidity neutralizing agentsSodium potassium citrate increased blood HCO3- levels in oral doses and neutralized the TME pH[146,147,148,149]
Controlling pH regulatory enzymesAs in the tumor microenvironment, acidic pH affects the chemotherapeutic drug efficacy. By regulating pH, the efficacy can be enhanced. Few drugs are carbonic anhydrase IX/XII and proton pump inhibitors[146,147]
Immunity modulationTyro3, Axl, and Mertk receptor (TAM) regulationTAM overexpression increases cell survival and decreases apoptotic signaling. Again, TAM down-streaming promotes metastasis via migration and invasion. The immunosuppressive nature of the TAM arises from the polarization of macrophages M1 to M2. M2 releases immune-suppressive cytokines. M2 blocking agents and M1 reprogramming agents can regulate this immunity suppression. Using small-molecule tyrosine inhibitors and TAM receptor targeted ligands is useful to regulate it[152,153,154,155]
Regulatory T-cell (Treg) inhibitionTreg cells regulate T-cell immune responses to maintain cell homeostasis. But in the TME, Treg cells decrease the entry of T cells. Transforming growth factor-β (TGF-β) inhibitors and anti-PD-L1 antibodies reduce the TGF-β signal to promote T-cell infiltration into the TME. Again, cytotoxic T-lymphocyte-associated protein-4 (CTLA-4) antibodies remove the Treg cells and can enhance T cell functions[152,156]
Myeloid-derived suppressor cell(MDSC) inhibitionMDSC induces immune suppression by inhibiting T-cell, NK-cell, and macrophage functions. Targeting and inhibiting phosphatidylinositol 3-kinase (PI3K)δ, PD-L1 or CTLA-4, and multi-kinase MDSC can be controlled[157,158]
Enhancement of active targetingSurface ligand modificationFolate discs enhanced the permeability and photothermal efficacy[153,166]
Biomimetically modified NPsCancer cell membrane-coated NPs can carry antigens and drugs to the target[164,165]
Tumor microenvironment responsive
drug delivery system
Enhanced tumor penetration of carrier NPsFunctional moieties
sensitive to a variety of
Tumor cellular stimuli
In response to the TME, supramolecular architectures based on peptides can convert structurally and allow therapeutics for controlled release. This dissertation emphatically introduces peptide assemblies with a stimulus-responsive structural conversion to acids, high temperatures, and high oxidative potentials in tumor tissues[153,166,167]
Particle size modificationAfter blood circulation, particles with a large size can shrink in size due to internal stimuli, such as enzymes, acidic pH, and hypoxia. Using peptides or other favorable ligands, NP entraps to form corona at TME[170,174,175,176,177,178,179]
Enhancement of cellular uptakeConversion of surface chargesIt helps to eliminate long circulation times and cellular uptake by modifying its surface charge. An example is a pH-sensitive PEG coating[180]
Detachments of shell of the NPsAs nano-vectors accumulate at tumor sites via the EPR effect, overexpressed MMP-9 can detach the PEG corona to expose peptide RGD to facilitate cellular internalization[92]
Elevate the drug release at cancer sitePolymer switches between hydrophilic–hydrophobic triggered by TME signalsProtonation and de-protonation polymers present in the NPs can switch from hydrophilic to hydrophobic and trigger drug release at the targeted sites. Poly(2-(diisopropylamino)ethyl methacrylate) can trigger the drug release[181,182]
Cleavage with a sensitive linkerHypoxia-sensitive linker[144,181]

5. Novel Nanocarriers Based Treatment Approach

Cancerous cell proliferation and migration profiles are different from those of normal cells. A therapeutic dosage form should enter the TME to control cancerous cell proliferation and migration. The penetration of conventional dosage forms into the TME is less due to its heterogeneity and the above-mentioned other factors. In addition, traditional drug delivery systems are less specific for cancer cells. Due to the lack of specificity and less penetration into the TME, the required concentration of the drug does not reach the cancer cells. Non-eliminated cancer cells alter metabolic signaling pathways and drug metabolism, inactivate drugs, suppress apoptosis, alter epigenetics, change drug targets, enhance DNA repair, alter epithelial–mesenchymal transition, and enhance gene amplification. As a result, cancer cells cause MDR, survive, rocket, and migrate [60,183,184,185,186,187].
As the nanocarriers have a diverse range (from 01–1000nanometers) and can be tuned according to the requirements of the (<200 nm) targeted site, the study and use of nanocarrier-based targeted drug delivery have increased. Again, the nanoparticulate nanocarriers can incorporate multiple targeting agents to enhance bioavailability, drug delivery, absorption, targeting precision, and stimulus technique. Understanding and identifying cancer cells’ physiochemical behavior can help optimize nanocarriers. In addition, the release pattern of drugs from nanocarriers determines the effectiveness of nanocarrier-based drug delivery systems [60,183,187].
Nanocarriers can be classified as organic, inorganic, or hybrid based on the components used in their development [188].

5.1. Organic Nanocarriers

The biocompatible nano-structurally dispersed, versatile, and less toxic organic carriers can be synthesized through either non-covalent or covalent interaction between the drug and adjuvants. Almost all organic nanocarriers contain carbon as a primary component. Commonly used adjuvants for organic nanocarrier synthesis are lipids, polymers, surfactants, proteins, and polysaccharides [189]. Depending upon the compound used for the physical synthesis of organic nanocarriers for delivering the drugs to the lung, they are classified into lipid-based nanocarriers (solid lipid NPs, liposomes, micelles, and lipid nano-capsules) and non-lipid-based nanocarriers (mesoporous NPs, polymeric NPs, dendrimers, and metallic NPs) [189,190].

5.1.1. Lipid Based Nanocarriers

As a carrier system, lipid NPs are biocompatible and biodegradable, and their toxicity is lower than that of polymeric NPs. In addition, it improves solubility and absorption to enhance bioavailability and pharmacokinetic parameters. These lipidic nanocarriers are classified into solid lipid NPs (SLN), nanostructured lipid nanocarriers (NLC), liposomes, lipidic nanocapsules, and niosomes [191]. In a recent clinical update, ceranib-2, a ceramidase inhibitor-loaded lipid NP, increased penetration through the membrane and bioavailability [192], as shown in Table 2.

Solid Lipid Based NPs

Solid lipid nanoparticles (SLN) are a surrogate of the colloidal drug delivery system, which can carry lipophilic and hydrophilic drugs, nucleic acids, and proteins to the targeted site. The size range of SLNs is 40–1000 nm [193]. It is a versatile, biocompatible, stable nanocarrier system with less toxicity. It is suitable for both active and passive targeting. Solid lipid NPs are prepared by dispersing the melted solid lipid in water, followed by the addition of emulsifying agents through different homogenization techniques or micro-emulsification. Supercritical fluid, solvent emulsification/evaporation, double emulsion, and spray drying methods can be used to prepare SLNs [190]. Primary solid lipids used in the SLN preparation are fatty acids, mono-, di-, triglycerides, or waxes. These biodegradable lipids of SLN can offer sustained release of drugs deep into the lungs and are for the pulmonary drug delivery system. Solid lipid NPs have a larger surface area and can load higher doses of active medicament. As per the requirement, in SLN, the drug can be incorporated into the matrix, shell, or core as shown in the Figure 4. SLN can be used in the preparation of oral dosage forms. Recently, studies have shown the higher transfection efficacy of cationic SLNs for the p53 gene targeting lung cancer [194]. A high-melting-point triglyceride in the SLN formulation is more efficient in the tumor cell environment [195]. Clinical updates indicate that folic acid-modified silymarin SLN enhances internalization through folate receptors in TME [196], as shown in Table 3. The main disadvantages of SLNs are their lower drug-loading efficacy and drug expulsion during storage. It can be rectified by mixing lipids with oil in a 70:30 to 99.9:0.01 ratio. SLNs can be optimized further by using appropriate ligands to overcome the TME, other than passive targeting [190,197,198].
In a study, researchers found that inhalable epirubicin-loaded SLN caused more cytotoxicity than epirubicin solution in the A549 cell line [199]. SLN loaded with docetaxel also prevented tumor growth and lung metastasis in 4T1 murine mammary carcinoma cells [200]. In a study, researchers found that the dual drug curcumin and paclitaxel-loaded SLN showed the highest tumor inhibitory action (78.42%) in the A549 cell line compared to other cell lines rather than the drugs separately administered. As well as enhancing P-glycoprotein efflux, this formulation reverses the MDR pathway and down-regulates NF-kB [201], as shown in Table 3. Enhanced green fluorescence protein plasmids and doxorubicin-loaded transferrin-conjugated SLN show improved anticancer activity [202], as shown in Table 4.

Liposomes

Liposomes are spherical vesicles with an aqueous core surrounded by natural phospholipids or synthetic amphiphiles and sterols in one or more bilayers with particle sizes ranging from 25 to 2500 nm, as shown in the Figure 5 [203]. This lipid-based drug delivery carrier is suitable for hydrophilic and lipophilic drugs as it has aqueous and lipidic layers. It can deliver macromolecules such asDNA, proteins, imaging, and active chemotherapeutic agents. It is a non-toxic, stable, high-vascular-density, and adjustable surface nanocarrier with a higher retention time in the targeted site. The half-life of this bilayer formulation is short in the systemic circulation. The preparation of liposomes generally begins with drying lipids from organic solvents and dispersing them in aqueous media, followed by purification and analysis. The composition of a bilayer determines the rigidity, fluidity, and charge of the layer. Long-chain acyl-functional phospholipids form the rigid, impermeable bilayer structure of the liposome. Unsaturated phosphatidylcholine shapes a flexible, permeable liposome. The commonly used phospholipids in liposome preparation are phosphatidylethanolamine and phosphatidylcholine. Microfluidizers, membrane extrusion, sonication, and homogenization techniques can control liposome size and size distribution. This nanocarrier NP can be used for active, passive, pH, magnetic, stimuli-responsive, and thermo-responsive targeting. Liposomes can enhance the loaded drug’s efficacy at the targeted site, therapeutic index, and stability. It also reduces the loaded drug’s toxicity and exposure to sensitive tissue [197,204,205]. Biofunctionalization liposomes enhance loaded drug efficacy in resisting lung cancer therapy through active targeting, as shown in Table 2 [206]. Again, in another clinical update (Table 2), researchers found that irinotecan and veliparib-loaded nano-liposomal intravenous formulations show combinational synergy for PARP and topoisomerase-1 inhibition along with better efficacy [207]. The disadvantages of liposomes are lower solubility, a shorter half-life, leakage of encapsulated drugs, oxidation and hydrolysis, and a higher production cost. Limitations and benefits of liposome drug carriers depend on liposome interaction with cells and their fate in vivo after administration. The interactions of liposomes with the cell surfaces take place either through adsorption or endocytosis. A liposome can be classified according to its functional modification: Conventional, PEGylated, ligand targeting, and theranostic [197,204,205,208,209]. These differently modulated liposomes can overcome the biophysiochemical difficulties of the active medicaments to reach the targeted sites. As well as liposome-loaded drugs suppressing the TME, soluble mediators in liposomal drug delivery systems inhibit TME immunity [208,209].

Conventional Liposome

Conventional liposomes and first-generation liposomes consist of an aqueous core encased in lipidic bilayers of cationic, anionic, or neutral phospholipids and cholesterol as shown in the Figure 5. Commonly used lipids and phospholipids for the preparation of conventional liposomes are 1,2-di-stearoryl-sn-glycero-3-phosphatidyl choline (DSPC), sphingomyelin, egg phosphatidylcholine, and monosialoganglioside. The main disadvantage of conventional liposomes is that they are rapidly eliminated. It occurs due to plasma opsonization and sequestration by reticuloendothelial macrophages [195,200,201]. In a study reported in Table 3, the researchers found that Honokiol-loaded liposomes show antitumor activity and induce apoptosis through the degradation of HSP90 client proteins to inhibit Akt and Erk1/2, which are mutant or wild-type EGFR signaling cascade effectors [210]. In another study reported in Table 3, researchers found that nano-liposomes loaded with the antioxidant Chinese herbal drug baicalin after intravenous administration to rabbits showed the highest drug accumulation in the lung and a higher survival time than the baicalin solution [211].

PEGylated Liposome

The sterically stabilized PEGylated liposomes were developed to improve stability and systemic circulation time by blending a hydrophilic polymer-polyethylene glycol (PEG), into the liposome preparation as shown in the Figure 5. Encapsulated PEG in the liposome formulation improved the efficacy of the entrapped drugs by creating a steric barrier that enhanced permeability and retention in the TME. PEGylating liposomes can overcome the opsonization of serum components, rapid recognition, and uptake by the reticuloendothelial system. It accumulates the active medicaments in the circulation and enhances the circulation time. If the accumulation increases exponentially, it interacts with other sites and causes toxicity [204,209]. It acts through passive targeting. According to a study, PEGylated liposomes accumulate in tumors via the EPR effect, but the mechanism of action differs in drug-resistant tumors. In hyper-permeable drug-resistant tumors, PEGylated liposomes penetrate deep into tumor cells. In hypo-permeable tumors, it enhances the proximity of tumor vasculature, thereby inhibiting angiogenesis [212,213]. The PEGylated NPs containing doxorubicin circulate longer with a higher intratumoral drug concentration, resulting in better therapeutic results [214]. A new study reported in Table 4 found that paclitaxel-loaded PEGylated liposomes enhance the efficacy of paclitaxel and reduce neuropathic pain associated with paclitaxel [215].

Ligand Targeted Liposome

The ligand-targeted liposome is a unique type of liposome in which monoclonal antibodies, proteins, growth factors, glycoproteins, and carbohydrates are chosen to couple with the liposome to target according to the overexpression at the disease site in an active targeting manner. Antibodies or fragments of antibodies are highly selective for this nanocarrier. In addition to entrapping more active drugs, it can act as a sustained-release agent at the target site. To improve the efficacy of the ligand-targeted liposome selection of targeted receptors, internalization versus non-internalization behavior of antigen, ligand selection, therapeutic agent selection, and location of the targeted site play an important role. This nanocarrier system can be classified as antibody fragment-targeted, receptor-targeted (folate receptor, transferrin receptor, sigma receptor), peptide-targeted, or multidrug resistance reversal-targeted as shown in the Figure 5. This nanocarrier is accessible if the target is in the blood and lymph nodes [216,217,218]. As shown in Table 4, the administration of the lipid stearic acid peptidomimetic conjugate SA-5 with doxorubicin ligand-liposome results in an antiproliferative effect in human epidermal growth factor receptor-2-mutated NSCLC [219]. Recently, researchers have tried dual ligand (CPP33 peptide and monoclonal anti-CA IX antibody)-modified liposome encapsulation with organic hetero-heptacyclic triptolide. They found that the dual ligand-modified liposome loaded with triptolide increased the cytotoxicity of triptolide with tumor-specific targeting and penetration without causing systemic toxicity, as shown in Table 4 [220].

Theranostic Liposome

Theranostic liposomes are hybrid liposome dosage forms that combine diagnosis profiles with targeted therapy to create a tailored treatment profile as shown in the Figure 5. Besides imaging purposes, it protects from systemic clearances. Commonly used nanosize imaging agents to prepare this type of liposome are iron oxide, quantum dots, and gold NPs. Here, imaging agents are covalently bonded to liposome surfaces. The active drugs are encapsulated at the core or embedded in the lipophilic bilayer shell [221]. In a study, researchers found that a dual-layered liposomal–gold liposome induces photothermal effects in cancer cells, as shown in Table 3 [222]. Another study revealed that folate-targeted theranostic liposomes containing paclitaxel and vinorelbine prevented NSCLC metastasis and cancer proliferation (Table 4) [223].

Micelles

A micelle is a supramolecular assembly of surfactant phospholipids in water to form a colloidal dispersion. It has both hydrophilic and hydrophobic parts as shown in the Figure 6. The hydrophilic group exists in the center, whereas the hydrophobic group exists in the external solvent. In the inverse type of micelle, the hydrophobic group exists in the center, whereas the hydrophilic group exists in external solvents. In terms of structure and function, self-assembling micelle structures are similar to biological transport systems in that they can protect insoluble hydrophobic drugs. This type of nanocarrier can carry low-molecular-mass hydrophobic drugs, proteins, and genes. As the micelle has a size range of 50 nm, it helps deliver the drugs to the systemic circulation through tissue penetration and cellular uptake for active medicament accumulation, permeability, and retention. This can improve the encapsulated drug’s delivery to the targeted site. But the main disadvantage of the micelle drug delivery system is its shorter stability and premature drug release when it comes into contact with the systemic circulation as it dilutes. Using a covalent crosslinking strategy such as corona formation around the micelle, dimerization, and di-functional crosslinking can stabilize it. Another method to stabilize the micelle is the complexation of the micelle core [224,225]. One study has found that the micelle reprogrammed the CAFs to modulate the entry of APIs into the TME [226]. As reported in Table 2, pH-sensitive epirubicin conjugated micelles with anticancer drugs synergistically enhance the efficacy of epirubicin in resistant and metastasizing cancer [227]. In 2007, a few researchers reported that cremophor-free paclitaxel-loaded PLGA-b-methoxy PEG polymeric micelles with cis-platin showed better efficacy for the advanced stage of NSCLC [228]. Furthermore, researchers found that PLGA-PEG-maleimide micelles prepared by microfluidics exhibited a higher degree of cytotoxicity in NSCLC when loaded with docetaxel, as reported in Table 3 [229].

Lipidic Nanocapsule

A lipidic nanocapsule is a hybrid biomimetic nanocarrier. It comprises medium-chain triglycerides encased in an aqueous phase using a PEGylated surfactant as shown in the Figure 7. Sometimes lecithin and co-surfactant are also used to prepare lipidic nanocapsules. Both active and passive targeting are possible with this form of the nanocarrier. Lipidic nanocapsules can incorporate both lipophilic and hydrophilic drugs. It can show adjuvant effects such asp-glycoprotein inhibition to favor higher anticancer APIs in the TME. Modification with a ligand can enhance the lipidic nanocapsules’ efficacy [230,231,232]. This formulation has better physical stability and a smaller particle size distribution (20–100 nm), which can be used for drug delivery through different routes. The disadvantages of lipid nanocapsules are the low encapsulation capacity of lipophilic drugs, leaky vessels, and instability in biological fluids [231,232]. According to researchers, lipidic nanocarriers containing erlotinib and modified with PEG polypeptide are cytotoxic to lung cancer cell lines HCC-827 and NCI-H358 (Table 4) [233]. Another study revealed that tretinoin-encapsulated lipid nanocapsules can overcome tretinoin resistance in the A549 cell line (Table 3) [234].

Nanostructured Lipid Nanocarrier

Nanostructured lipid carriers are biocompatible, unstructured nanocarrier systems made from biocompatible lipids, surfactants, and co-surfactants as shown in the Figure 8. To prepare NLCs, solid and liquid lipids are mixed at 70:30 to 99.9:0.1 with surfactants ranging from 1.5% to 5% (w/v). Commonly used lipids for the formulation of NLCs are fatty acids, glycerids (mono, di, and tri), steroids, and waxes. To prepare NLCs, micro-emulsification, solvent displacement methods, and high-pressure homogenization techniques are used. The average size of NLCs is 40–1000 nm with a spherical morphology. The selection of lipids and surfactants can influence the physicochemical properties and quality of the materials, including particle size and drug loading. Due to its stability, high drug load, P-GP efflux inhibition, and diverse theranostic capabilities, it became a topic of interest for drug delivery studies [193,235,236,237]. A549 cells treated with transferrin ligand-conjugated NLCs entrapped in a plasmid containing enhanced green fluorescence protein demonstrated better efficacy for gene therapy in lung cancer treatment, as found by a few researchers (Table 4) [238]. In another study, an inhalable drug delivery system using NLCs loaded with doxorubicin, siRNA, and conjugated LHRH peptide showed better control of cancer cells than intravenous injections (Table 3) [239]. Further, researchers found that NLC fabricated with doxorubicin and sorafenib can stimulate PD-1 expression, down-regulate Treg cells, activate effector T cells, and regenerate the immune response (Table 4) while inhibiting esophagus tumors by bypassing the TME [240].

5.1.2. Non-Lipid-Based Organic Nanocarriers

Non-lipid-based nanocarriers are the newer classes of drug delivery systems. The need for non-lipid-based drug delivery arises as the lipid nanocarrier has limitations such as hydrophilic drug loading, accumulation at non-targeted sites (spleen and liver), and reticuloendothelial clearance. Further, this new class has advantages such as multiple therapeutic targeting, higher toxicity, etc. Depending upon the particle used for the preparation, these non-lipid organic nanocarriers are classified into polymeric, dendrimers, mesoporous, and metallic NPs [197].

Polymeric Nanocarriers

Polymeric NP carriers are small (1–1000 nm), adjustable, rapidly absorbable, and versatile colloidal carrier systems to control the release of the entrapped active drug within the polymeric shell. Polymeric NPs can be classified into polymeric nanocapsules (reservoir systems) and nanospheres (matrix systems). Preparation methods for polymeric NPs include solvent evaporation and diffusion, nanoprecipitation, and reverse salting. Generally, the nanoprecipitation method is used to prepare polymeric nanocapsules. The stability of this nanocarrier depends on the adsorption of the active medicament onto the NP surface and the presence of surfactants as shown in the Figure 9. Microbial contamination is one of the challenges of this type of formulation. This problem can be resolved by adding preservatives, spray drying, or lyophilization. The drug delivery system is suitable for cancerous cell treatment using drug–nucleic acid combinations. These NPs can induce antitumor immunity in CD8+ T cells by regulating the lymphatic system and activating dendritic cells in TME [241,242,243]. The advantages of polymeric NPs include multiple therapeutic targeting and independent control of drug release. The main disadvantages of polymeric NPs are the synchronization of the pharmacokinetics and biodistribution of loaded compounds [241,242]. Novoselova M.V. et al. (2020) have found that the internalization of polymeric multilayer capsules in lung cancer cells is 75% higher than in healthy lungs. Embedding gemcitabine and clodronate in polymeric multilayer capsules inhibited macrophage-induced tumor growth (Table 3) [244]. In another study, silibinin, a low-water-soluble drug encapsulated in polycaprolactone/Pluronic F68 NPs, showed sustained release in the systemic circulation for up to 48 h, inhibited tumor growth, and improved the drug’s efficacy [245]. In a clinical update, the researchers found that polymeric NPs loaded with docetaxel can overcome drug resistance to refractory cancer (Table 2) [246]. Another clinical update, as shown in Table 2, is that polymeric micelles loaded with anticancer drugs are capable of releasing drugs whose AUC, Cmax, and volume of distribution are unstable [247]. As reported in Table 2, polymeric NPs entrapped with hypoxia-responsive photosensitizers and chemotherapeutic drugs produce reactive oxygen species that enhance the efficacy and photodynamic response of cancer treatments [248].

Dendrimers

Dendrimers, arborols, or cascade molecules are a 1–100nm-size, three-dimensional, highly ordered, monodisperse, globular polymeric, symmetric macromolecular, hyperbranched macromolecular, tailored carrier system suitable for targeting drugs. The macromolecular dendrimers consist of a core, a repeating mantle, and a functionalized group corona [249,250,251]. Dendrimers’ tunable surfaces facilitate covalent modification to create stable micelle-type structures suitable for non-covalent encapsulation of APIs. Dendrimers are more stable under high shear stress, dilution, temperature, and pressure than liposomes and micelles [252,253]. It can act through both active and passive targeting. The PEGylation of dendrimers shields against attacking proteases and improves the water solubility of the dendrimers and their loaded drugs. The covalent conjugation of mAbs with the multivalent hydrophobic inner core (encapsulated with the APIs) of dendrimers can cause toxicity to the targeted cells as shown in the Figure 10. A few interesting biomedical dendrimers are polyamidoamine and polypropylene imine. The amine group of corona produces more toxicity and limits its use [249,250,251]. Poly(propylene amine), polyglycerol, and polyethylene imine dendrimers release APIs in a pH-dependent manner. Commonly, at pH 5–6, these dendrimers release the drug at a faster rate as compared to pH 7.4 [252,253]. Generally, dendrimers are prepared using either divergent or convergent methods. In both, the dendrimer grows outward from a multifunctional core molecule. The core molecule reacts with monomer molecules containing one reactive and two dormant groups. Then, the activated new periphery of the molecule reacts with more monomers. Encapsulation, electrostatic interaction, and covalent conjugation methods load drugs onto dendrimers [108,249,250,251].
Further, according to the need, a modified dendrimer can overcome the first-pass effect, immune clearance, cell penetration, and off-target interactions. These unique characters enhance the circulation time followed by the maximum amount of entrapped drugs to the targeted site [108]. Researchers found that siRNA and cis-di-amine platinum-loaded polyamidoamine dendrimers conjugated with folic acid nanocarriers for targeting H1299 performed better than individual therapies and showed negligible toxicity to normal MRC9 lung fibroblast cells (Table 4) [254]. Further, in a clinical update [Table 2], the researcher found that a dendrimer loaded with bromoenol lactone inhibitors improves the drug’s solubility, tolerability, and therapeutic index [255].

Polystyrene NP Carriers

Polystyrenes are biocompatible polymers prepared by the polymerization of styrene monomers. As a result of external stimuli, it undergoes rapid and reversible phase transitions. It results in a desired drug release pattern from the formulation. Again, as per the previous reports, polystyrene had better penetration through the skin, respiratory tract, and digestive tract by forming a protein corona around it in the biological system. The Environmental Protection Agency report states that more than 1000 mg/m3 of chronic exposure to styrene is toxic for humans [256,257]. If the dose exceeds the limit, the chance of tumor formation increases, as styrene’s epoxide metabolite is genotoxic and can adduct DNA in humans [258]. Polystyrene NPs (10 nm) as shown in the Figure 11, can pass the blood–brain barrier and deposit in the alveoli, and their toxicity varies according to their size [259]. Polystyrene NPs can enhance reactive oxygen species production in the TME to inhibit cell growth [256,260]. It also disturbs metabolic activity, and amino-modified polystyrene causes cytotoxicity [261,262]. Further, no research suggests the carcinogenic nature of modified polystyrene NPs in a limited dose. Due to the complexity of drug delivery to the lungs, modified polystyrene NPs may be used. Research has found that polystyrene NP surfaces functionalized with carboxy, amino, and pristine accumulate inside cells and cause cytotoxicity and genotoxicity in A549 NSCLC [263].

Carbon Nanotubes

Carbon nanotubes, forged and rolledup with graphene, are cylindrical-shaped single-walled or multi-walled drug carrier NPs with an inner diameter of 0.4 to a few nm and an outer surface diameter varying from 2–30 nm as shown in the Figure 12. Low-temperature chemical vapor deposition (CVD) accurately synthesizes CNTs. After synthesis, purification of the CNTs is necessary to reduce larger graphite particles. CNTs have similar photothermal effects as other metal NPs. They have strong optical absorption due to their specific optical properties. To enhance drug loading, CNTs have π-electro-conjugated surfaces that interact with hydrophobic APIs. CNTs’ limitations as drug carriers include their lower aqueous solubility, physiochemical characteristic-dependent pharmacokinetic profile, and toxicity [82,264,265]. Generally, smaller-sized CNTs are not better suited for treating lung cancer, as they may cause pleural lesions, inflammation, pleural fibrosis, and malignant mesothelioma [266]. In addition to cancer diagnosis, CNTs can be usedin imaging, augmenting tissue engineering, and delivering drugs. Recently, efforts have been made to reduce toxicity by using CNT-based hydrogel [267]. Researchers reported that an ELISA test comparing untreated NSCLC A543 cells with multi-walled carbon nanotubes conjugated with bromocriptine was lethal (Table 4) [268]. In another study, researchers found that the higher concentration of multi-walled carbon nanotubes causes nuclear condensation and DNA laddering. It generates reactive oxygen species that induce oxidative stress and apoptosis in A549 cells [269]. In a clinical update of Table 2, researchers found that carbon nanotube-based anticancer drug delivery can suppress drug resistance [270]. Another clinical update found that a protein–CNT complex can bind to tumor vasculature endothelial cancer cells and destroy them through the PTT effect [271].

5.2. Inorganic Nanocarriers

Multidrug resistance (MDR) is one of the factors contributing to therapeutic failures in lung cancer treatment. A solution to the MDR lies in the use of inorganic nanocarriers. Nanocarriers can carry compounds with low molecular weights as well as imaging agents. Combining these nanocarriers with photothermal and interference agents can further enhance their efficacy. Their unique physiochemical properties, stability, inertness, and biocompatibility further improve their applications [272,273,274].

5.2.1. Metallic NP Carrier

Metallic nanocarriers are colloidal 10–1000 nm systems where the active medicaments are either dispersed or encapsulated in a shell or structure of a metal-based cavity or covalently attached to the surface of the metal cavity. These metal nanocarriers act as hypothermic agents. It reacts through surface plasmon resonance. Metallic NPs efficiently absorb near-infrared (NIR) light for photothermal therapy (PTT). So, it is an attractive option for cancer-targeted drug delivery. Again, surface modification of metal nanocarriers can optimize drug delivery to the targeted site. The metallic NP has diverse applications, from diagnosis to treatment. It can act via active and passive targeting in the cellular and subcellular regions. A drug’s efficacy at a targeted site depends on the physicochemical properties of metal nanocarriers. The main disadvantage of metallic NPs is their toxicity due to their accumulation in different tissues, which leads to stimulation or suppression of the immune response. It also causes acute toxicity through oxidative stress. The persistence of oxidative stress can cause DNA damage or genotoxicity. Different forms of metal NPs include pure metal NPs (gold, silver, copper, titanium, platinum, zinc, magnesium, iron, and alginate); metal oxide NPs (titanium dioxide, silver oxide, and zinc oxide); doped metal/metal oxide/metal nonmaterial; metal sulfide; and metal–organic frameworks (MOFs) nonmaterial, as shown in the Figure 13 [275,276]. A recent clinical study found that sorafenib-entrapped metal cluster-doped protein NPs enhance drug efficacy and bioavailability through enhanced optical contrast, magnetic contrast, and modulation of zeta potential [277]. Another study in Table 2 found that phospholipids containing cis-platin prodrug entrapped in MnO2 NPs generate a glutathione oxidation–reduction reaction to cause hyperpyrexia and activate photothermal effects to treat lung cancer [278]. In a recent clinical update, researchers reported that anticancer drug-loaded metal–organic frameworks cause dual effects of photothermal nature along with the inhibitory properties that can treat cancer [279]. Further, a clinical update of Table 2 found that antitumor drugs loaded in a hybrid metal–organic framework modified with cholesterol oxidase can catalyze the overexpressed cholesterol and overcome MDR [280].

Gold NP

Gold NPs (GNPs) are 5 to 400 nm in size and vary in shape as shown in the Figure 14; they are optoelectric, mildly antibacterial, and targeted drug delivery carriers. Their antibacterial activity depends on the intensification of ROS generation in the microbial cells. Other biomedical applications of GNPs are photodynamic immunotherapy for cancer treatment, diagnostic agents, etc. The photothermal activity of GNPs is due to the excitement of electrons when irradiated with laser light. GNPs can be synthesized using the bottom-up reduction method of chloroauric acid (HAuCl4). Commonly used reducing agents are sodium citrate, borohydride, polyalcohol, amines, etc. [281,282,283]. The reported absorption of GNPs in oral administration is low. IV administration of GNPs accumulates in the spleen, liver, and lung, and elimination is less. GNPs increase glucose and catalytic enzymes (alanine aminotransferase and aspartate transaminase). They also affect liver function [284]. In a study, researchers found that methotrexate conjugated GNPs in a lower dose inhibit tumor growth compared to methotrexate (without loading or conjugated) in Lewis lung carcinoma (Table 3) [285]. In another study, researchers reported significant cytotoxicity and apoptosis in lung cancer stem cells when aluminum (III) phthalocyanine chloride tetra sulfonic acid and anti-CD133 antibody bioconjugate GNPs were administered (Table 4) [286]. On the A549 cell line, researchers found that silibinin-conjugated gold NPs released pH-responsively enhanced silibinin efficacy up to 4–5 times (Table 3) [287]. In a recent clinical update of Table 2, researchers have found that T-cell, microRNA, or peptide-conjugated or entrapped gold NPs enhance the EPR effect and its photothermal nature to inhibit cancer cell growth [288].

Silver NP

Silver NPs are 1–100 nm-sized, stable, catalytic, and high-conductance NPs. It regenerates reactive oxygen species or releases silver ions from its surfaces. This NP has antibacterial, antifungal, and anti-inflammatory properties. NPs of silver penetrate the skin less than other inorganic metals [289,290]. Oral administration of silver NPs accumulates in different organs as silver ions, especially in the liver and spleen. The retention time of silver NPs in the brain and testis is longer [291]. In H1299 lung cancer cells, these NPs can cause cell apoptosis and inhibition of nuclear factor transcriptional activity in a dose-dependent manner [292]. Again, as reported, Toxicodendron vernicifluum-modified silver NPs can cause 82.5% of cancer cell apoptosis in A549 lung cancer cells [293]. In another study, as reported in Table 3, embelin biofunctionalized silver NPs exhibit significantly lower necrotic cells than apoptotic cells in A549 cancer cells in a dose-dependent manner [294]. In Table 3, as compared with cis-platin, Juniperus Chinensis leaf extract fabricated into biofunctionalized silver NPs showed better anti-proliferation and apoptotic effects on the A549 lung cancer cell line [295].

Platinum NP

Platinum NPs are variedly shaped (spherical, rods, tetrahedral, and cubes) and have a 2–100 nm stable brownish-red or black colloidal or suspension dispersion system. It acts as an automotive catalytic converter for hydrogenation in the chemical industry. In addition, platinum NPs are used as drug delivery and imaging agents [296,297,298,299]. In vitro, analysis of platinum NPs inhibits the growth of A549 cancer cells in a dose-dependent manner [300]. Again, blood-triggered platinum NP sactas an anticancer agent by forming the protein corona [301]. Another study revealed that platinum NPs reverse the oxidative stress in lung adenocarcinomas in the A549 lung cell line [299,300]. Again, biofabricated platinum NPs are biocompatible and have catalytic and anticancer activity [167,300]. In another study, researchers found that poly(lactic-co-glycolic acid) and polyethylene glycol-modified platinum NPs loaded with anti-EGFR showed better efficacy in triple-negative breast cancer [302]. Furthermore, porous Au–Pt NPs loaded with doxorubicin modified with c-RGD exhibit better drug release patterns and enhanced anticancer properties via photo-induced and photothermal processes. The scavenging activity of NPs enhances drug-induced oxidative stress (Table 4) [303].

5.2.2. Metal Oxide NP

The redox-reactive metal oxides of a size range less than 200 nm are another choice of nanocarriers for antibiotic-resistant wound healing, along with growth factors. A metal oxide is stably tunable, has a high surface area, and is a porous NP with antimicrobial, antifungal, and antiviral properties [304,305]. Again, researchers have found that CuO, NiO, and Fe2O3 generate ROS in normoxia and hypoxia that cause toxicity to the tumor environment in lung cancer patients [306].

Zinc Oxide NP

Metal oxides such as ZnO are less than 100 nm rod, hexagonal, tripod, spherical, and different-shaped photographic catalytic metal oxide NPs that can absorb and reflect ultraviolet rays. This nature is helpful for bio-imaging purposes. NPs of zinc demonstrate biological activity such as apoptosis upon activation with light. The anticancer, antibacterial, and antimicrobial activity of zinc oxide NPs is due to excess ROS production. ZnO NPs cause cell apoptosis in colon carcinoma by altering mitochondrial IL8 release function [307,308,309,310,311]. The exact mechanism also works against NSCLC [154,310,311]. Zinc NPs loaded with cis-platin and gemcitabine enhanced the inhibition of tumor formation and the apoptotic nature of the drugs. The formulation also decreases the total cell viability in the A549 cell line, as reported in Table 3 [310]. In a recent clinical update in Table 2, researchers found that rapamycin-loaded zinc–organic frameworks work by inhibiting the mTOR pathway. It also enhances the sensitivity of chemotherapy [312].

Iron Oxide NP

Iron oxide NPs are highly reactive, rapidly oxidized (in the presence of oxygen and water), superparamagnetic, 1–100 nm NPs. Among its many uses, it is suitable for magnetic imaging, bimolecular separation, and targeted drug delivery. Three oxides of iron NPs are magnetite (Fe3O4), maghemite (γ-Fe2O3), and hematite (α-Fe2O3) [154]. Iron oxide NPs interact with the immune cells to modulate the immune response. Further, the PTT activity of iron oxide can act as an antitumor response [295]. Initially, iron NPs caused higher cytotoxicity due to the quicker release of Fe ions. A carbohydrate or polysaccharide (carboxymaltose) complexed with iron NPs can control the release of Fe ions to saturate transferrin. The action mechanism of this NP starts with the uptake by the RES system, followed by degradation of the polymeric or carbohydrate shell through macrophages to conserve it as ferritin or excrete it through ferroportin-1 [313]. Iron oxides can be used for cancer diagnosis purposes using their photothermal activity. With FDA-approved ferumoxytol co-incubated with macrophage treatment for metastatic lung cancer in the liver and lungs, early mammary cancers have shown caspase-3-mediated apoptosis. It also increases the pro-inflammatory response in M1 macrophages [313,314]. Researchers also found that the proliferation of lung cancer cells was slowed by SPIONs coated with silica monolayers, as shown in Table 3 [315]. In another study, researchers discovered that iron oxide nanoflowers loaded in thermo-sensitive fluorescent liposomes could enhance the efficacy of cytotoxic drugs against lung cancer by causing hypothermia [316].

Copper Oxide NP

Copper oxide NPs are 1-100-sized antimicrobial, antibacterial, catalytically reactive, high surface-to-volume ratio NP carrier systems prepared from copper salt in the presence of surfactant [275,317]. It can be prepared from plant extracts such as Euphorbia nivulia’s latex, Magnolia kobus leaf, Calotropis procera latex, etc. [318]. Copper oxide (CuO), the oxide NP of copper, is effective for cancer cells. Copper displayed a dose-dependent degradation of DNA molecules by generating oxygen [319]. Researchers have reported that copper oxide promotes anticancer activity in A549 cell lines via I, II, and IV HDAC mRNA expression [320]. Further, researchers have found that biofilm-producing bacteria and cancer cells do not tolerate actinomycetes mediating CuO NPs [321].

Titanium Dioxide NP

Titanium dioxide NPs are 1–100 nm-sized photocatalytic nanocarriers. Photocatalysis and self-cleaning mechanisms apply to optics, materials science, electronics, catalysts, pigments, and biology [322,323]. TiO2 NPs exhibit antimicrobial activity via photocatalytic free oxide and peroxide formation [322,324]. According to the researchers metal-doped TiO2 improves antimicrobial properties. It modifies light absorption to enhance photocatalytic properties [325]. The anticancer activity of TiO2 NPs is also due to the production of radical oxides to reduce oxidative stress. In a study, researchers reported that TiO2nanosquares, nanotubes, and fine particles have immunomodulatory effects. These TiO2 NPs inhibit tumor angiogenesis via proinflammatory responses [325]. According to Behnam MA et al.(2018), PEGylated TiO2 NPs destroy solid tumors through photothermal effects [326]. According to another study, TiO2 inhibits lung cancer proliferation, DNA damage, and apoptosis through the intrinsic mitochondrial pathway [327]. YSA peptide-conjugated mesoporous titanium peroxide NPs loaded with cantharidin produced reactive oxygen species and increased photodynamic lung cancer apoptosis [328]. In a recent clinical update reported in Table 2, researchers have found that antitumor drugs entrapped in bionic titanium dioxide synergistically generate reactive oxygen species to enhance the loaded drug’s efficacy [329].

Magnesium Oxide NP

A magnesium NP is typically a black, spherical NP with a size range of less than 100 nm and a specific surface area of 30 to 70 m2/g [330]. Among all magnesium NPs, magnesium oxide NPs (MgONP) have antibacterial activity in a concentration-dependent manner. It causes physical injury to the cell wall along with ROS production to damage the DNA. Magnesium oxide NPs also reduce the tobacco bacterial wilt index [331,332,333]. In addition to being antimicrobial and photocatalytic NPs, green MgONPs exhibit a high level of cytotoxicity against the MCF-7 breast cancer cell line [334]. MgONPs also form complexes with human serum albumin and induce cytotoxicity against K562 cell lines [335]. Using pH-sensitive polymer-coated Mg nanoflowers for photoacoustic and bubble-enhanced ultrasound imaging can break the polymer shell for hydrogen generation in acidic TME [336].

5.2.3. Metal Sulfide NP

Photovoltaic, electrical, and optoelectrical metal sulfide NPs are varying size (01–100 nm) and shape (spheres, hexagons, cubic, etc.). Colloidal solvothermal synthesis can produce them [337,338,339,340]. These biocompatible NPs convert light, enhance radiation, and activate the immune system. Metal sulfides also enhance the Fenton catalysis process. It also enhances the EPR effect [341,342,343,344,345,346,347]. Recently, metal sulfides have been tried in cancer treatments using PTT, immunotherapy, radiotherapy, and targeted drug delivery mechanisms with limited success [339,344,345]. Recently, researchers found that PEG-surfaced copper sulfide NPs enhance immunotherapy efficacy via PTT and antigen capture [346]. Metal sulfide can be used for diagnostic purposes such as photoacoustic and multimodal imaging.

5.2.4. Metal–Organic NPs

Metal–organic NPs are high-internal-surface-area-coordinated porous polymeric clusters of inorganic compounds with organic ligands. These NPs are suitable for chemical sensing, separation, drug delivery, catalysis, and storage. The surface modification of metal–organic frameworks using biofunctionalization according to their targeted sites can enhance their efficacy at the target sites. The prime challenge for synthesizing these NPs is the non-uniformity of the prepared NPs and the slow nucleation rate. A commonly used process to obtain uniformly smaller NPs in the range of 10–100 nm burst nucleation is adopted, followed by termination of precursors using depletion [347,348,349]. A newer treatment regimen for cancer is photodynamic therapy (PDT). Its effectiveness at the targeted sites depends on the presence of light, a photosensitizer (PS), and oxygen molecules. The porphyrinic metal–organic framework can naturally harvest light, transport oxygen, catalyze reactions, and transfer electrons to instrument PDT cancer therapy. PS-based metal–organic framework NPs with precise spatial arrangements can improve PDT efficacy. Bioconjugations with ligands to the metal–organic framework can enhance the selectivity of the NPs to the targeted site. Metal–organic frameworks can deliver chemotherapeutics and nucleic acids [347,348,349]. Recently, researchers found that RGD peptide biofunctionalized metal–organic framework loaded with doxorubicin enhances the loaded drug efficacy 4–5 times (Table 4) [85].

5.2.5. Quantum Dots

Quantum dots are colloidal electronic and fluorescent nanocrystals with an average diameter of 2–100 nm and unique electronic, physical, and photo-physical properties as shown in the Figure 15. It is known as an artificial atom due to its disjunctive electronic energy level. Unlike isolated atoms, artificial atoms have disjunctive electronic energy levels. As a semiconductor heterostructure, a quantum dot traps charge carriers (electrons and holes) in a volume approximately equal to the quantum mechanical wavelength of its components. Due to these properties of quantum dots, interest has increased in biomedical, bio-sensing, intracellular protein tracking, tissue engineering, drug delivery, and bioterrorism purposes. Again, in cancer treatment, drug delivery to the tumor microenvironment is challenging. Smaller quantum dots increase permeability, and a higher surface area enhances targeting efficacy [350,351]. Recently, researchers found that modified graphene QDs have higher accumulation in the hypoxia-induced oral squamous tumor cell microenvironment with low systemic toxicity [352]. Conjugating QDs with an active drug improved internalization in tumors resistant to drugs [353]. The QD nanocrystals modified with folic acid and 11-mercaptoundecanoic acid show cytotoxicity, genotoxicity, and migration-inhibitory activity against A549 lung cancer cells [354]. Hyaluronic acid-conjugated ZnO-based pH-responsive doxorubicin QD shows a synergistic effect on tumor growth, as shown in Table 4 [355].

5.2.6. Magnetic NP

Magnetic NPs (1–100 nm) are magnetic-field-manipulating NPs that connect to form clusters with magnetic nano-chains composed of magnetic materials (iron, nickel, and cobalt) and chemical functional components, as shown in the Figure 16. A magnetic NP consists of a magnetic core, a protective coating, and a surface functionality linker. Commonly used surface functionality linkers are synthetic organic polymers, silica, gold, and organic polymers [356]. Maghemite Fe2O3 or magnetite Fe3O4 magnetic NPs exhibit excellent MRI contrast properties. They also provide the required systemic toxicity and can be used as a catalytic nonmaterial, magnetic colloidal photonic crystals, resonance imaging, targeting biomedicine, etc. [357,358]. In cancer theranostics, the use of magnetic NPs has increased due to their larger surface area, smaller size, magnetic resonance imaging capability, ease of synthesis, ease of decoration, lesser toxicity, and better delivery vehicles. Different preparation methods may vary the shapes and sizes of magnetic NPs. A magnetic field can target and release entrapped drugs within a magnetic NP by altering its surface charges. As reported in Table 3, on an A549 lung cancer cell line, superparamagnetic iron oxide polymer (SPION) conjugated NPs loaded with doxorubicin are biocompatible without causing systemic toxicity [358]. Again, silica-coated SPION delays the proliferation of cancer cells [359].

5.2.7. Ceramic NP

Ceramic NPs (CNs) are 50 nm diameter, temperature-resistive, inorganic nanocarrier systems prepared from metal (iron, calcium, titanium, silicon, etc.) oxides, carbides, phosphate, carbonates, albumin, and silica by successive heating and cooling. The preparation methods include sol-gels, low-temperature combustion, aerogels, hypothermal, Pechini-citrate gels, and microemulsions [360]. CNs can carry drugs that are enzyme, pH, or temperature sensitive. Besides photocatalysis, they can be used for dye degradation, imaging, and photodegradation. These CNs can be amorphous, hollow, porous, or polycrystalline [361]. Ceramic NPs can accumulate in the smaller capillaries, especially in the lung, and can cause risks in the circulation process. It can also affect the opsonization process. It can produce more reactive oxygen species and worsen the cancer environment. Mesoporous silica, calcium phosphate, carbon allotropes, and iron oxide are some of the ceramic NP carriers [362].

5.2.8. Mesoporous Silica Nanocarrier

Mesoporous silica NPs are solid, tunable, and porous nanocarriers with high encapsulation capacity through endocytosis. These NPs have uniform pore size ranges of 2–6 nm. There are three types of MSNs—ordered MSNs, hollow MSNs, and core/shell MSNs. A hollow MSN can load more drugs than the others. Surface functionalization can enhance NPs’ physicochemical properties. A few techniques for preparing MSNs are growth quench, confinement techniques, separation of confinement, and growth techniques. Functionalization can be done with co-condensation, multifunctionalization, and grafting methods. The surface modification allows this NP to target both actively and passively [363,364,365,366]. Human cells are more likely to internalize 50 nm MSNs, although smaller particles exhibit longer circulation times. As particle size influences cytotoxicity, micrometric particles of 1 mm are less toxic than nanometric particles of 200 nm. Again, cationic NPs are more immunogenic and cytotoxic than neutral or anionic ones. In melanoma treatment, the FDA has approved multimodal silica NPs [363]. Conjugating ligands such as folic acid, DNA aptamers, transferrin, and antibodies with MSNs can enhance the efficacy of photodynamic targeted therapy for cancer. Researchers have found that MSN injection before anti-PD-1 resensitizes to overcome tumor resistance improves anti-PD-1 activity and protects immunity [366,367]. Researchers also found that siRNA co-delivered with chemotherapeutic drugs loaded in MSNs synergistically enhanced their efficacy and survivin protein inhibition [367,368]. In another study, folic acid-modified MSNs loaded with multidrug-resistant protein-1 siRNA and myricetin reduced cell viability, suppressed tumors, and up-regulated the expression levels of cleaved caspase-3 and PARP in the cancer cell lines A549 and NCI-H1299 [368]. In a clinical update of Table 2, researchers have found that an antitumor drug loaded in pH-responsive mesoporous silica-coated gold NPs can cause a photothermal effect in addition to the loaded drug mechanism to produce anticancer activity specifically in the tumor cells [369].

5.3. Hybrid Nanocarrier

The advantages and disadvantages of a variety of drug nanocarriers are discussed above. Recently, adding a combinational approach can mimic the disadvantages of nanocarriers and increase their efficacy. So, the concept of hybrid nanocarriers has arrived. These hybrid systems combine the benefits of different structural components to synergize the outcome of the therapy. Erosion and degradation are the processes by which the hybrid NP releases the entrapped active medicaments from the core. Multiple layers of lipids, polymers, and organic–inorganic compounds may protect the core materials, along with the solubility and permeability modifications of the entrapped active ingredients [370]. Recently, curcumin and survivin shRNA loaded in polymeric hybrid NPs with PLGA-conjugated triblock polymers (W5R4K-PEG2K-PHIS) showed better penetration into the TME and synergistic tumor suppression action [371].
Table 2. Patents of nanocarriers containing payloads and their clinical status in cancer.
Table 2. Patents of nanocarriers containing payloads and their clinical status in cancer.
DrugsNanocarriersDosage FormKey TargetApprove StatusApproved ByRemarksPatent NoRefs.
Ceranib-2Lipid NPNanoemulsionCeramidase inhibitorsApprovedWorld
Patent
Enhances penetration through the cell membrane and increases bioavailabilityWO2020018049A2[192]
SilymarinSolid lipid NPIntravenous injectionFolic acidPendingChinese patentFolic acid modified silymarin SLN enhances internalization in TMECN111195239A[196]
Anticancer drugLiposomeSubcutaneousActive targetingApprovedChinese patent Biofunctionalization further enhances the loaded drug efficacyCN105726483B[206]
Irinotecan, veliparibNanoliposomeIntravenousPARP and topoisomerase-1 inhibitionGrantedJapanese patent Nano-liposomal formulation shows combinational synergy along with better efficacyJP2018528184A[207]
Anticancer drugEpirubicin conjugated polymeric micelleIntravenous and oralEpirubicin resistant cancerGrantedUnited States patent pH-sensitive epirubicin-conjugated micelle with anticancer drug synergistically enhances the efficacy of epirubicin in resistant and metastasizing cancerUS10220026B2[227]
DocetaxelPolymeric
NP
IntravenousDrug resistant cancerGrantedWorld patentRefractory cancerWO2014210485A1[246]
Bromoenol lactone inhibitorDendrimersIntravenous infusionInhibit bromoenol lactoneGrantedWorld patentBromoenol lactone inhibitor covalently attached dendrimers enhance the solubility, improve tolerability, and increase therapeutic indexWO2018154004A9[255]
Anticancer DrugPolymeric micelleIntravenous Endogenous proteinGrantedWorld patentFacilitates drug release, especially in unstable, low AUC, low Cmax, high volume of distribution, critical micelle concentration above theoretical Cmax of the drugWO2014165829A2[247]
Anticancer DrugCarbon nanotubesParenteral administrationDrug resistance decreasesGrantedUnited States patentDecreases drug resistanceUS20150196650A1[270]
ProteinSingle-walled carbon nanotubesParenteral administrationImmune stimulantGrantedUnited States patent Bind to tumor vasculature and endothelial cancer cellsUS20100184669A1[271]
T cellGold NPSystemic
administration
T-cell receptor proteinAbandonedUnited States patent Conjugation or entrapment of the gold NP enhances the EPR effect, and then the photothermal effect inhibits the growth of cancer cellsUS20140086828A1[289]
SorafenibMetal-cluster-doped protein NPIntravenousEGFRGrantedWorldwideMetal cluster-doped protein NP enhances the drug efficacy and bioavailability by enhancing optical contrast, and magnetic contrast, modulation of zeta potentialWO2014087413A1[277]
Phospholipids containing cis-platin prodrug MnO2 NPIntravenousMultidrug resistant cancerPendingChinese patent Tumor cells carry platinum through endocytosis. In the uptake of drugs, MnO2 can generate a glutathione oxidation–reduction reaction to cause hyperpyrexia and activate photothermal effects to treat lung cancerCN111214488A[278]
Anticancer drugMetal–organic frameworkIntravenousDouble effects:
Metal–organic framework photothermal effect
Anticancer drug inhibits cancer through a specific mechanism
GrantedChinese patentPhotothermal effects, in addition to the loaded drugs inhibitory action, can treat cancerCN110652497A[279]
Antitumor drugHybrid metal–organic framework modified with cholesterol oxidase IntravenousCatalyze the oxidation reaction of cholesterol GrantedChinese patentHybrid metal–organic framework can catalyze the overexpression of cholesterol and overcome multidrug resistanceCN112274648A[280]
Photosensitizer and chemotherapeutic drugPolymeric NPIntravenousHypoxia responsive GrantedChinese patentHypoxic response polymer NP helps generate reactive oxygen species that enhance the chemotherapeutic drug’s efficacy along with the photodynamic responseCN108653288B[247]
Antitumor drugMesoporous silica-coated gold NPIntravenouspH-responsive antitumor drug carrierGrantedChinese patentPhotothermal effectCN107412195B[370]
RapamycinZinc–organic frameworkIntravenousmTOR pathwayPendingChinese patent Inhibits mTOR pathway and enhances the sensitivity of chemotherapyCN110693883A[312]
Antitumor drugBionic titanium dioxideIntravenousGenerate reactive oxygen speciesGranted Drugive oxygen speciesChinese patent Reactive oxygen species can enhance the antitumor drug’s efficacyCN109646675B[329]
Table 3. Non-functional nanocarriers in lung cancer.
Table 3. Non-functional nanocarriers in lung cancer.
DrugNanocarrierCompositionCell LineInVitro Character ResultsRemarksRefs.
Particle Size (nm)Zeta Potential (mV)Drug Release
Paclitaxel + curcuminSolid lipid NPHydrogenated soybean phospholipids; 1,2-distearoyl-sn-glycero-3phosphoethanolamine-N[methoxy(polyethylene glycol)-2000]; polyvinyl pyrrolidone k15A549121.8 ± 1.6930.4 ± 1.25 Improved tumor inhibition. Reduces P-glycoprotein efflux, reverses MDR, and down-regulates the NF-κB pathway[201]
HonokiolLiposomeSodium per-carbonate, cholesterol, PEG2000-DSPEH1975, HCC827130 ± 20−20.0 to −30.0Sustained mannerShows time-dependent inhibition of degradation of HSP90 client proteins to inhibit Akt and Erk1/2, which are mutant or wild-type EGFR signaling cascade effectors[209]
BaicalinNanoliposomePhospholipon90H, Tween-80, citric acid, NaHCO3A549131.7 ± 11.7 Sustained release for 24 h up to 89.6 + 2.1%, stable for 12 monthsBaicalin, the antioxidant, has antitumor activity[210]
GoldTheranostic liposomeDistearoyl phosphatidylcholine, 1,2-distearoyl-sn-glycero-3-phosphoethanolamine (methoxy(polyethylene glycol)-5000), 1,2-distearoyl-sn-glycero-3-phosphoethanolamine-N-[amino(polyethylene glycol)-2000], cholesterol 72.84 ± 22.49 −20 to −40Sustained releaseLiposomal gold liposomes act via photothermal effect, and their stability is also enhanced[221]
DocetaxelMicellePLGA-PEG-MalA54972 + 1NeutralSustained releaseHigher cytotoxicity in NSCLC[229]
TretinoinLipidic nanocapsulePoly(e-caprolactone), sorbitan monostearate, f polysorbate 80A54925012.7 ± 0.9Sustained releaseHigher cytotoxicity through cell cycle arrest at the G1phase[233]
Gemcitabine and clodronatePolymeric multilayer nanocapsulesPoly-L-arginine hydrochloride, dextran sulfate sodium salt, tetramethylrhodamine isothiocyanate mixed isomers, rhodamine B, boric acid, glycerol, ethylene diamine tetraacetic acid disodium salt (EDTA), clodronate disodium tetrahydrateA549~250–500NeutralSustained releasePMC inhibited macrophage-induced tumor growth[244]
siRNA and different chemotherapeutic agentsMesoporous silica NP A549172-21Sustained releaseCombination of siRNA with chemotherapeutic agents shows synergistic effect with restraint of survivin effect[368]
SilibininPolymeric NPSilibinin (SB), polyvinyl alcohol (Mw 30,000–70,000 kDa), polycaprolactone (PCL), inhalable grade lactoseA549108 ± 3.21–397 ± 3.19NeutralSustained releasePCL/Pluronic F68 NPs loade silibinin significantly inhibited tumor growth in lung cancer-induced rats after inhalable administration[243]
MethotrexateGold NPMethotrexate, HAuCl4, sodium citrate, phosphate buffer 7.4A549, QU-DB14.3−7.3 ± 2.5 Gold NP, through PTT effect, enhances the drug’s efficacy[285]
SilibininGold NPHAuCl4, trisodium citrate dehydrate, silibinin, DMSOA549163 ± 5−22.2 ± 0.458 Silibinin-conjugated gold NPs released pH-responsively enhanced silibinin efficacy up to 4-5 times[287]
EmbelinSilver NPEmbelin, silver
nitrate
A54925−5.42 Embelin-biofunctionalized silver NPs exhibit significantly lower necrotic cells than apoptotic cells in A549 cancer cells in a dose-dependent manner[294]
Juniperus chinensis leaf extractsSilver NPJuniperus chinensis leaf extracts, silver nitriteA549, HEK29398.21 ± 1.54−26.5 Juniperus Chinensis leaf extract fabricated biofunctionalized silver NPs showed better antiproliferation and apoptotic effects[295]
Cis-platin, gemcitabine Zinc NPZinc oxide NP, methanol, tri-ethylamine, cis-platin, gemcitabineA549 21 ± 0.4 NASustained releaseNP loaded with cis-platin, gemcitabine inhibits tumor formation and enhances the apoptotic nature of the drugs[310]
Iron NPIron NP modified with silica layer NP Superparamagnetic iron (II,III) oxide NPs (SPIONs), tetraethyl orthosilicate, hexadecyltrimethylammonium bromide A549BEAS-2B 101.3 ± 2.8 −26.1 ± 0.1 Sustained releaseDelays the proliferation of cancer cells[315]
Table 4. Functional nanocarriers in lung cancer.
Table 4. Functional nanocarriers in lung cancer.
DrugNanocarriersReceptorsLigandCompositionCell LineIn Vitro Character ResultRemarksRefs.
Particle Size (nm)Zeta Potential
(MV)
Drug Release
Enhanced green fluorescence protein plasmid (pEGFP)+ doxorubicinTransferrin-conjugated SLNTransferrinTransferrinEnhanced green fluorescence protein plasmid (pEGFP)-N1
Soya lecithin
Human transferrin
A54926742SustainedImproves anticancer activity[202]
PaclitaxelPEGylated large liposomeBlocks cell cycle in the G2/M phasePEGLipo-Cat-PEG phosphatidylcholine, cholesterol, stearylamine, and DSPE-PEG2000A549,
LL2
180 SustainedAntitumor activity with painful neuropathy reduction[199]
DoxorubicinPeptidomimetic conjugate (SA-5) liposomeBlocks human epidermal growth factor receptor-2 (HER2) lipid stearic acid peptidomimetic conjugate SA-5Lipid dipalmitoylphosphatidylcholine
Poly(ethylene glycol) distearoylphosphatidylethanolamine
Cholesterol
BT474
A 549
CALC3
107.19−13.38 mVSustainedAntiproliferativeactivity[215]
TriptolideCPP33 peptide and monoclonal anti-CA IX antibody)-modified liposome3D tumor spheroidsCPP33 peptide, monoclonal anti-CA IX antibodyAnti-CA IX antibody, CPP33 peptide with a terminal cysteine, soybean lecithin, NBD-DPPE, DSPE-PEG-MALA549137.6 ± 0.8  SustainedTumor-specific targeting and increasing tumor cell penetrationwithout causing systemic toxicity[220]
ErlotinibPEGylated lipidic nanocapsuleEGFR PEGylated polypeptidePoly(ethylene glycol)-b-poly(L-aspartic acid), lecithin, sunflower oil, castor oil, Tween-20, and Span 20HCC-827 and NCI-H358-20∼200−20Sustained releaseHigher cytotoxicity than erlotinib without loading in any nanocarrier[233]
Plasmid-containing enhanced green fluorescence proteinTransferrin-nanostructured lipid carriersGene deliveryTransferrinSoya lecithin, Maleimide-PEG2000-COOH, human transferrin (iron-free), stearic acid, L-a-phosphatidylethanolamineA549157+15.9 ± 1.9Sustained releaseGene targeting drug delivery[238]
Doxorubicin, sorafenibFolic acid Nanostructured lipid carrierImmunotherapy Folic acidFolic acid, soya lecithin, maleimide-PEG2000-COOH, stearic acid 100 Sustained releaseHelps overcome the TME, immune response enhancement, cytotoxicity[240]
siRNA, cis-diamine platinumFolic acid -conjugated polyamidoamine dendrimersFolate receptor-α inhibitionFolic acidFolic acid, H1299, A 543280+14.5–17.2Sustained releaseSuitable for co-deliveryof si-RNA along with cytotoxicity[254]
siRNA, myricetinFolic acid conjugated mesoporous silica NPMultidrug resistance protein-1, folate receptorFolic acidFolic acid, tetraethylorthosilicate, cetyltrimethylammonium bromide, myricetin A549, NCI-H1299109.9Neutral Sustained releaseIt accumulates in TME and prevents colony formation by enhancing the cancer cells’ radiosensitivity[368]
BromocriptineCarboxyl or Hydroxyl conjugated multiwalled carbon nanotubesDopamine receptorCarboxyl or hydroxyl groupCarbon nanotubes, thionyl chloride, tetrahydrofuran A549, QU-DB26.3–32.6 Sustained ReleaseBromocriptine act via dopamine receptor and cause cancer cell apoptosis[268]
Gold NPAluminum (III) phthalocyanine chloride tetra sulfonic acid and anti-CD133 bioconjugated goldNPPhotodynamic effectAluminum (III) phthalocyanine chloride tetra sulfonic acid and anti-CD133 antibodyAluminum (III) phthalocyanine chloride tetra sulfonic acid, anti-CD133 antibodyA54963.91 nm−14.7 The bioconjugate enhance the gold NPs’ photothermal activity[286]
DoxorubicinAu–Pt NPPhotothermal/photodynamic cRGD, AuGold(III) chloride trihydrate (HAuCl4·3H2O), Pluronic® F-127 (F-127), silver nitrate (AgNO3), ascorbic acid, potassium tetrachloroplatinate(II) (K2PtCl4), methyl thiazolyl tetrazolium (MTT), calcein AM, and PI, thiol poly-(ethylene glycol) succinimidylglutaramide, doxorubicinMDA-MB23178.4–85.3−14.8Sustained releasePorous Au–Pt NPs loaded with doxorubicin modified with cRGD exhibit better drug release patterns, as well as enhanced anticancer properties[303]
CantharidinMesoporous titanium peroxide NPsPhotodynamic, increased reactive oxygen speciesYSATetrabutyl titanate, (3-aminopropyl) trimethoxysilanetitanium butoxide, hydrogen peroxide, heptanoic acid, ethanol, N-(3-dimethylaminopropyl)-N-ethylcarbodiimide hydrochloride, doxorubicin and N-hydroxysuccinimide. Cantharidin (CTD)A549150−21.77 YSA-modified mesoporous titanium peroxide NPs loaded with cantharidin produced reactive oxygen species and increased photodynamic lung cancer apoptosis[329]
DoxorubicinMetal–organic frameworkEnhanced the loading drug efficacy up to 5 times without affecting the normal cellsRGDDiphenyl carbomate, KOH, gamma cyclodextrin, RGD peptide, NHS, EDC, low-molecular-weight heparinA549150−25.6Sustained releaseThe loaded drug efficacy enhanced the targeted sites[85]
DoxorubicinQuantum dotsFolate receptorFolic acid, 11-mercaptoundecanoic acidSodium dihydrogen phosphate, disodium hydrogen phosphate, sodium chloride, potassium chloride, silver nitrate, indium(III) chloride, zinc stearate, 1-dodecanethiol, sulfur, 1-octadecene, oleylamine, MUA, dimethyl sulfoxide, cysteine, lipoic acid, NHS, EDC, doxorubicin hydrochloride, folic acidA54911–19−15.5 ± 3.5Sustained releaseQD nanocrystals modified with folic acid and 11-mercaptoundecanoic acid showed improved cytotoxicity, genotoxicity, and migration inhibitory activity against A549 lung cancer cells[355]
DoxorubicinQuantum dotsOverexpressed glycoprotein CD44Hyaluronic acidDicarboxyyl-terminated poly(ethylene glycol), hyaluronic acid, zinc acetate, magnesium acetate, sodium hydroxide, dimethyl sulfoxide, anhydrous N,N-dimethylformamide, doxorubicinA549 −0.0521 −1.90Sustained releaseShows synergistic effect of Zn2+and doxorubicin for antitumor activity[354]

6. Conclusions

Lung cancer has a lower survival rate due to the complexity of delivering the active drugs to the targeted sites. Biological barriers, behavioral nature, and tumor heterogeneity impact the delivery of drugs to lung cancer. There have been many attempts to overcome the barriers through different therapeutic approaches such as chemotherapy, immunity modulation therapy, radiation therapy, chemotherapy, stereotactic body radiotherapy, etc. Recently, other than the above therapeutic options, interest in targeted drug delivery systems has increased as adjuvant therapy in both the early and late stages of disease progression. The reason is that most of the above-mentioned conventional therapies became resistant after a certain period, and the therapeutic accumulation in the intracellular region was enough to cause toxicity in the tumor microenvironment. In addition to that, conventional therapies are unlikely to enter the tumor microenvironment.
In our study of nanocarrier-based targeting drug delivery to overcome the TME barriers, we have found that particle size and active targeting using receptor-based bioconjugating agents play roles in bypassing the TME to enhance the targeting efficacy of the loaded drug.
Different studies on lung cancer found that the inflammatory mediators were overexpressed, especially IL-6. A high-affinity protein can block it. The nanocarrier biofunctionalized with proteins such as RGD can be useful for targeting cancer cells. Further, folic acid deficiency promotes IL-6/JAK-1/pSTAT3 interactions in astrocytes after ischemia-reperfusion. So, folic acid biofunctionalized nanocarriers may be another approach to improving the targeting precision. Further, the PEGylation of NPs reduces the interaction with serum proteins and enhances the stability of the nanocarrier in the reticuloendothelial system.
The selection of the nanocarrier for lung cancer treatment primarily depends upon factors such as the physiochemical nature of the drug, solubility, permeability, molecular weight, and aqueous stability. Secondly, the nanocarrier’s specificity, viz., particle size and surface geometry, biocompatibility, and nano-toxicity are important considerations in designing nanocarriers. Biodegradable nanocarriers are often preferred as they can be metabolized and eliminated from the body, reducing the risk of long-term toxicity.
In this study, we have found multiple nanocarriers with different possibilities. Depending on the requirements and targeting strategy, nanocarriers can be modified to optimize the required outcome.

7. Future Prospective

In light of advances in nanotechnology, various research studies are underway to find more convenient cancer treatments. NSCLC remains a substantial clinical challenge, though chemotherapy and surgery are the only standards of care. Drug delivery to the targeted site remains challenging despite newer drugs for different histological subtypes and driver mutations. So, the emphasis on the nanocarrier drug delivery system as an add-on therapy to the current regime will lead to greater effectiveness. According to different studies, we found that biofunctionalized inorganic metal compounds with organic compound complex-loaded drugs may be a carrier system for NSCLC targeted therapy. Especially, with active targeting through surface modifications of receptors overexpressed in lung cancer cells (folic acid, peptide, somatostatin). The biofunctionalization of the nanocarrier enhances biosystem interaction, cellular uptake, immune system escape, and vascular alteration to penetrate the tumor microenvironment. The inorganic metal compounds have a photothermal effect that scavenges reactive oxygen species. Further, the loaded pathway-blocking agents can inhibit rapid cancer cell growth.
A redox-sensitive prodrug and pH-responsive carbon nanotubes can ensure active form in conditions that suit the tumor microenvironment. In addition, modified organic andinorganic NP carrier systems may also be useful in lung cancer-targeted therapies.
In this study, we have discussed the different nanocarrier systems and their TME bypassing strategies. This study will help to develop new targeted therapeutics using a modified bioconjugate hybrid nanocarrier that can act through active targeting by bypassing TME. Further, this study will give an idea of different nanocarrier efficacies in a concise form, along with their mechanisms. It will help to compare nanocarriers in diverse conditions for developing personalized therapy.

Author Contributions

Conceptualization, M.H.A., K.S. and I.A.; methodology, K.S., O.A. and M.H.A.; software, P.G.; validation, O.A., A.S.A.A. and M.A.A.; formal analysis, K.S. and M.J.; investigation, A.-H.E. and M.H.A.; resources, K.S. and I.A.; data curation, K.S. and M.H.A.; writing—original draft preparation, M.H.A. and I.A.; writing—review and editing, K.S., O.A., M.J., A.-H.E. and P.G.; visualization, I.A., M.J. and M.A.A.; supervision, M.H.A.; project administration, M.H.A. and A.S.A.A.; funding acquisition, A.-H.E., M.J. and O.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors are thankful to the Deanship of Scientific Research, King Khalid University, Abha, Saudi Arabia, for financially supporting this work through the Large Research Group Project under Grant no.R.G.P.2/317/44.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. WHO. WHO Global Report on Trends in Prevalence of Tobacco Smoking 2000–2025, 2nd ed. Available online: https://apps.who.int/iris/handle/10665/272694 (accessed on 2 December 2020).
  2. Allemani, C.; Matsuda, T.; Di Carlo, V.; Harewood, R.; Matz, M.; Nikšić, M.; Bonaventure, A.; Valkov, M.; Johnson, C.J.; Estève, J.; et al. Global surveillance of trends in cancer survival 2000–14 (CONCORD-3): Analysis of individual records for 37 513 025 patients diagnosed with one of 18 cancers from 322 population-based registries in 71 countries. Lancet 2018, 391, 1023–1075. [Google Scholar] [CrossRef]
  3. Goss, P.E.; Strasser-Weippl, K.; Lee-Bychkovsky, B.L.; Fan, L.L.J.; Chavarri-Guerra, Y.; Liedke, P.E.R.; Pramesh, C.S.; Badovinac-Crnjevic, T.; Sheikine, Y.; Chen, Z.; et al. Challenges to effective cancer control in China, India, and Russia. Lancet Oncol. 2014, 15, 489–538. [Google Scholar] [CrossRef]
  4. Van der Heyden, J.H.A.; Schaap, M.M.; Kunst, A.E.; Esnaola, S.; Borrell, C.; Cox, B.; Leinsalu, M.; Stirbu, I.; Kalediene, R.; Deboosere, P.; et al. Socioeconomic inequalities in lung cancer mortality in 16 European populations. Lung Cancer 2009, 63, 322–330. [Google Scholar] [CrossRef]
  5. Walters, S.; Maringe, C.; Coleman, M.P.; Peake, M.D.; Butler, J.; Young, N.; Bergström, S.; Hanna, L.; Jakobsen, E.; Kölbeck, K.; et al. Lung cancer survival and stage at diagnosis in Australia, Canada, Denmark, Norway, Sweden and the UK: A population-based study 2004–2007. Thorax 2013, 68, 551–564. [Google Scholar] [CrossRef]
  6. Zhang, J.; Fujimoto, J.; Zhang, J.; Wedge, D.C.; Song, X.; Zhang, J.; Seth, S.; Chow, C.W.; Cao, Y.; Gumbs, C.; et al. Intratumor heterogeneity in localized lung adenocarcinomas delineated by multiregion sequencing. Science 2014, 346, 256–259. [Google Scholar] [CrossRef] [PubMed]
  7. Alberg, A.J.; Brock, M.V.; Ford, J.G.; Samet, J.M.; Spivack, S.D. Epidemiology of lung cancer: Diagnosis and management of lung cancer, 3rd ed: American College of Chest Physicians evidence-based clinical practice guidelines. Chest 2013, 143 (Suppl. S5), e1S–e29S. [Google Scholar] [CrossRef]
  8. Devesa, S.S.; Bray, F.; Vizcaino, A.P.; Parkin, D.M. International lung cancer trends by histologic type: Male: Female differences diminishing and adenocarcinoma rates rising. Int. J. Cancer 2005, 117, 294–299. [Google Scholar] [CrossRef]
  9. Hasin, D.S. US Epidemiology of Cannabis Use and Associated Problems. Neuropsychopharmacol. Off. Publ. Am. Coll. Neuropsychopharmacol. 2018, 43, 195–212. [Google Scholar] [CrossRef] [PubMed]
  10. Krewski, D.; Lubin, J.H.; Zielinski, J.M.; Alavanja, M.; Catalan, V.S.; Field, R.W.; Klotz, J.B.; Létourneau, E.G.; Lynch, C.F.; Lyon, J.L.; et al. A combined analysis of North American case-control studies of residential radon and lung cancer. J. Toxicol. Environ. Health Part A 2006, 69, 533–597. [Google Scholar] [CrossRef] [PubMed]
  11. Bade, B.C.; Dela Cruz, C.S. Lung Cancer 2020: Epidemiology, Etiology, and Prevention. Clin. Chest Med. 2020, 41, 1–24. [Google Scholar] [CrossRef]
  12. Liu, P.; He, K.; Li, Y.; Wu, Q.; Yang, P.; Wang, D. Exposure to mercury causes formation of male-specific structural deficits by inducing oxidative damage in nematodes. Ecotoxicol. Environ. Saf. 2012, 79, 90–100. [Google Scholar] [CrossRef] [PubMed]
  13. Brenner, D.R.; McLaughlin, J.R.; Hung, R.J. Previous lung diseases and lung cancer risk: A systematic review and meta-analysis. PLoS ONE 2011, 6, e17479. [Google Scholar] [CrossRef] [PubMed]
  14. Caramori, G.; Adcock, I.M.; Casolari, P.; Ito, K.; Jazrawi, E.; Tsaprouni, L.; Villetti, G.; Civelli, M.; Carnini, C.; Chung, K.F.; et al. Unbalanced oxidant-induced DNA damage and repair in COPD: A link towards lung cancer. Thorax 2011, 66, 521–527. [Google Scholar] [CrossRef]
  15. Matakidou, A.; Eisen, T.; Houlston, R.S. Systematic review of the relationship between family history and lung cancer risk. Br. J. Cancer 2005, 93, 825–833. [Google Scholar] [CrossRef] [PubMed]
  16. Caliri, A.W.; Tommasi, S.; Besaratinia, A. Relationships among smoking, oxidative stress, inflammation, macromolecular damage, and cancer. Mutat. Research. Rev. Mutat. Res. 2021, 787, 108365. [Google Scholar] [CrossRef]
  17. Yamaoka, K.; Kataoka, T. Confirmation of efficacy, elucidation of mechanism, and new search for indications of radon therapy. J. Clin. Biochem. Nutr. 2022, 70, 87–92. [Google Scholar] [CrossRef]
  18. Valavanidis, A.; Vlachogianni, T.; Fiotakis, K. Tobacco smoke: Involvement of reactive oxygen species and stable free radicals in mechanisms of oxidative damage, carcinogenesis and synergistic effects with other respirable particles. Int. J. Environ. Res. Public Health 2009, 6, 445–462. [Google Scholar] [CrossRef]
  19. Coluzzi, E.; Leone, S.; Sgura, A. Oxidative Stress Induces Telomere Dysfunction and Senescence by Replication Fork Arrest. Cells 2019, 8, 19. [Google Scholar] [CrossRef]
  20. Liu, G.; Cheresh, P.; Kamp, D.W. Molecular basis of asbestos-induced lung disease. Annu. Rev. Pathol. 2013, 8, 161–187. [Google Scholar] [CrossRef]
  21. Gout, T. Role of ATP binding and hydrolysis in the gating of the cystic fibrosis transmembrane conductance regulator. Ann. Thorac. Med. 2012, 7, 115–121. [Google Scholar] [CrossRef]
  22. Yang, H.; Rivera, Z.; Jube, S.; Nasu, M.; Bertino, P.; Goparaju, C.; Franzoso, G.; Lotze, M.T.; Krausz, T.; Pass, H.I.; et al. Programmed necrosis induced by asbestos in human mesothelial cells causes high-mobility group box 1 protein release and resultant inflammation. Proc. Natl. Acad. Sci. USA 2010, 107, 12611–12616. [Google Scholar] [CrossRef]
  23. Klebe, S.; Leigh, J.; Henderson, D.W.; Nurminen, M. Asbestos, Smoking and Lung Cancer: An Update. Int. J. Environ. Res. Public Health 2019, 17, 258. [Google Scholar] [CrossRef] [PubMed]
  24. Kou, F.; Wu, L.; Ren, X.; Yang, L. Chromosome Abnormalities: New Insights into Their Clinical Significance in Cancer. Mol. Ther. Oncolytics 2020, 17, 562–570. [Google Scholar] [CrossRef] [PubMed]
  25. Weir, B.A.; Woo, M.S.; Getz, G.; Perner, S.; Ding, L.; Beroukhim, R.; Lin, W.M.; Province, M.A.; Kraja, A.; Johnson, L.A.; et al. Characterizing the cancer genome in lung adenocarcinoma. Nature 2007, 450, 893–898. [Google Scholar] [CrossRef]
  26. Davies, H.; Bignell, G.R.; Cox, C.; Stephens, P.; Edkins, S.; Clegg, S.; Teague, J.; Woffendin, H.; Garnett, M.J.; Bottomley, W.; et al. Mutations of the BRAF gene in human cancer. Nature 2002, 417, 949–954. [Google Scholar] [CrossRef]
  27. Ni, X.; Zhuo, M.; Su, Z.; Duan, J.; Gao, Y.; Wang, Z.; Zong, C.; Bai, H.; Chapman, A.R.; Zhao, J.; et al. Reproducible copy number variation patterns among single circulating tumor cells of lung cancer patients. Proc. Natl. Acad. Sci. USA 2013, 110, 21083–21088. [Google Scholar] [CrossRef]
  28. Cristofanilli, M.; Budd, G.T.; Ellis, M.J.; Stopeck, A.; Matera, J.; Miller, M.C.; Reuben, J.M.; Doyle, G.V.; Allard, W.J.; Terstappen, L.W.; et al. Circulating tumor cells, disease progression, and survival in metastatic breast cancer. N. Engl. J. Med. 2004, 351, 781–791. [Google Scholar] [CrossRef]
  29. Krebs, M.G.; Sloane, R.; Priest, L.; Lancashire, L.; Hou, J.M.; Greystoke, A.; Ward, T.H.; Ferraldeschi, R.; Hughes, A.; Clack, G.; et al. Evaluation and prognostic significance of circulating tumor cells in patients with non-small-cell lung cancer. J. Clin. Oncol. Off. J. Am. Soc. Clin. Oncol. 2011, 29, 1556–1563. [Google Scholar] [CrossRef] [PubMed]
  30. Magbanua, M.J.; Sosa, E.V.; Scott, J.H.; Simko, J.; Collins, C.; Pinkel, D.; Ryan, C.J.; Park, J.W. Isolation and genomic analysis of circulating tumor cells from castration resistant metastatic prostate cancer. BMC Cancer 2012, 12, 78. [Google Scholar] [CrossRef]
  31. Seto, M.; Honma, K.; Nakagawa, M. Diversity of genome profiles in malignant lymphoma. Cancer Sci. 2010, 101, 573–578. [Google Scholar] [CrossRef]
  32. Dalmay, T.; Edwards, D.R. MicroRNAs and the hallmarks of cancer. Oncogene 2006, 25, 6170–6175. [Google Scholar] [CrossRef]
  33. Hassanpour, S.H.; Dehghani, M.A. Review of cancer from perspective of molecular. J. Cancer Res. Pract. 2017, 4, 127–129. [Google Scholar] [CrossRef]
  34. Lazebnik, Y. What are the hallmarks of cancer? Nat. Rev. Cancer 2010, 10, 232–233. [Google Scholar] [CrossRef]
  35. Macé, A.; Kutalik, Z.; Valsesia, A. Copy Number Variation. Methods Mol. Biol. 2018, 1793, 231–258. [Google Scholar] [CrossRef]
  36. Zappa, C.; Mousa, S.A. Non-small cell lung cancer: Current treatment and future advances. Transl. Lung Cancer Res. 2016, 5, 288–300. [Google Scholar] [CrossRef] [PubMed]
  37. Landesman-Milo, D.; Ramishetti, S.; Peer, D. Nanomedicine as an emerging platform for metastatic lung cancer therapy. Cancer Metastasis Rev. 2015, 34, 291–301. [Google Scholar] [CrossRef]
  38. Travis, W.D.; Brambilla, E.; Nicholson, A.G.; Yatabe, Y.; Austin, J.H.M.; Beasley, M.B.; Chirieac, L.R.; Dacic, S.; Duhig, E.; Flieder, D.B.; et al. The 2015 World Health Organization Classification of Lung Tumors: Impact of Genetic, Clinical and Radiologic Advances Since the 2004 Classification. J. Thorac. Oncol. Off. Publ. Int. Assoc. Study Lung Cancer 2015, 10, 1243–1260. [Google Scholar] [CrossRef]
  39. Bychkov, A.; WHO Classification. Website. Available online: https://www.pathologyoutlines.com/topic/lungtumorWHO.html (accessed on 3 April 2023).
  40. Kim, E.S. Chemotherapy Resistance in Lung Cancer. Adv. Exp. Med. Biol. 2016, 893, 189–209. [Google Scholar] [CrossRef]
  41. Manser, R.; Wright, G.; Hart, D.; Byrnes, G.; Campbell, D.A. Surgery for early stage non-small cell lung cancer. Cochrane Database Syst. Rev. 2005, 2005, CD004699. [Google Scholar] [CrossRef] [PubMed]
  42. Barnes, H.; See, K.; Barnett, S.; Manser, R. Surgery for limited-stage small-cell lung cancer. Cochrane Database Syst. Rev. 2017, 4. [Google Scholar] [CrossRef]
  43. Lang-Lazdunski, L. Surgery for nonsmall cell lung cancer. Eur. Respir. Rev. 2013, 22, 382–404. [Google Scholar] [CrossRef] [PubMed]
  44. Demaria, S.; Golden, E.B.; Formenti, S.C. Role of Local Radiation Therapy in Cancer Immunotherapy. JAMA Oncol. 2015, 1, 1325–1332. [Google Scholar] [CrossRef] [PubMed]
  45. Baskar, R.; Lee, K.A.; Yeo, R.; Yeoh, K.W. Cancer and radiation therapy: Current advances and future directions. Int. J. Med. Sci. 2012, 9, 193–199. [Google Scholar] [CrossRef]
  46. Falls, K.C.; Sharma, R.A.; Lawrence, Y.R.; Amos, R.A.; Advani, S.J.; Ahmed, M.M.; Vikram, B.; Coleman, C.N.; Prasanna, P.G. Radiation-Drug Combinations to Improve Clinical Outcomes and Reduce Normal Tissue Toxicities: Current Challenges and New Approaches: Report of the Symposium Held at the 63rd Annual Meeting of the Radiation Research Society, 15–18 October 2017; Cancun, Mexico. Radiat. Res. 2018, 190, 350–360. [Google Scholar] [CrossRef]
  47. Majeed, H.; Gupta, V. Adverse Effects Of Radiation Therapy. In StatPearls; StatPearls Publishing: Treasure Island, FL, USA, 2023. [Google Scholar]
  48. Huang, C.Y.; Ju, D.T.; Chang, C.F.; Reddy, M.P.; 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. [Google Scholar] [CrossRef]
  49. Schiller, J.H.; Harrington, D.; Belani, C.P.; Langer, C.; Sandler, A.; Krook, J.; Zhu, J.; Johnson, D.H.; Eastern Cooperative Oncology Group. Comparison of four chemotherapy regimens for advanced non-small-cell lung cancer. N. Engl. J. Med. 2002, 346, 92–98. [Google Scholar] [CrossRef]
  50. Sebastian, N.T.; Xu-Welliver, M.; Williams, T.M. Stereotactic body radiation therapy (SBRT) for early stage non-small cell lung cancer (NSCLC): Contemporary insights and advances. J. Thorac. Dis. 2018, 10 (Suppl. S21), S2451–S2464. [Google Scholar] [CrossRef] [PubMed]
  51. VlaskouBadra, E.; Baumgartl, M.; Fabiano, S.; Jongen, A.; Guckenberger, M. Stereotactic radiotherapy for early stage non-small cell lung cancer: Current standards and ongoing research. Transl. Lung Cancer Res. 2021, 10, 1930–1949. [Google Scholar] [CrossRef]
  52. Feng, L.R.; Suy, S.; Collins, S.P.; Lischalk, J.W.; Yuan, B.; Saligan, L.N. Comparison of Late Urinary Symptoms Following SBRT and SBRT with IMRT Supplementation for Prostate Cancer. Curr. Urol. 2018, 11, 218–224. [Google Scholar] [CrossRef]
  53. Arvizo, R.R.; Miranda, O.R.; Moyano, D.F.; Walden, C.A.; Giri, K.; Bhattacharya, R.; Robertson, J.D.; Rotello, V.M.; Reid, J.M.; Mukherjee, P. Modulating pharmacokinetics, tumor uptake and biodistribution by engineered nanoparticles. PLoS ONE 2011, 6, e24374. [Google Scholar] [CrossRef]
  54. Decuzzi, P.; Lee, S.; Bhushan, B.; Ferrari, M. A theoretical model for the margination of particles within blood vessels. Ann. Biomed. Eng. 2005, 33, 179–190. [Google Scholar] [CrossRef]
  55. Tan, J.; Shah, S.; Thomas, A.; Ou-Yang, H.D.; Liu, Y. The influence of size, shape and vessel geometry on nanoparticle distribution. Microfluid. Nanofluidics 2013, 14, 77–87. [Google Scholar] [CrossRef] [PubMed]
  56. Singh, A.P.; Biswas, A.; Shukla, A.; Maiti, P. Targeted therapy in chronic diseases using nanomaterial-based drug delivery vehicles. Sig. Transduct. Target. Ther. 2019, 4, 33. [Google Scholar] [CrossRef] [PubMed]
  57. Steven, A.; Fisher, S.A.; Robinson, B.W. Immunotherapy for lung cancer. Respirology 2016, 21, 821–833. [Google Scholar] [CrossRef] [PubMed]
  58. Bradbury, P.A.; Shepherd, F.A. Immunotherapy for Lung Cancer. J. Thorac. Oncol. 2008, 3, S164–S170. [Google Scholar] [CrossRef]
  59. Ferrell, B.; Koczywas, M.; Grannis, F.; Harrington, A. Palliative care in lung cancer. Surg. Clin. N. Am. 2011, 91, 403–417. [Google Scholar] [CrossRef]
  60. Mansoori, B.; Mohammadi, A.; Davudian, S.; Shirjang, S.; Baradaran, B. The Different Mechanisms of Cancer Drug Resistance: A Brief Review. Adv. Pharm. Bull. 2017, 7, 339–348. [Google Scholar] [CrossRef]
  61. Neubert, R.H.H. Potentials of new nanocarriers for dermal and transdermal drug delivery. Eur. J. Pharm. Biopharm. 2011, 77, 1–2. [Google Scholar] [CrossRef]
  62. Din, F.U.; 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, 7291–7309. [Google Scholar] [CrossRef]
  63. Madni, A.; Tahir, N.; Rehman, M.; Raza, A.; Mahmood, M.A.; Khan, M.I.; Kashif, P.M. Hybrid Nano-carriers for potential drug delivery. Adv. Technol. Deliv. Ther. 2017, 54–87. [Google Scholar] [CrossRef]
  64. Dhaliwal, A.; Zheng, G. Improving accessibility of EPR-insensitive tumor phenotypes using EPR-adaptive strategies: Designing a new perspective in nanomedicine delivery. Theranostics 2019, 9, 8091–8108. [Google Scholar] [CrossRef]
  65. Blanco, E.; Shen, H.; Ferrari, M. Principles of design for overcoming biological barriers to drug delivery. Nat. Biotechnol. 2015, 33, 941–951. [Google Scholar] [CrossRef] [PubMed]
  66. Kay, M.A.; Glorioso, J.C.; Naldini, L. Viral vectors for gene therapy: The art of turning infectious agents into vehicles of therapeutics. Nat. Med. 2001, 7, 33–40. [Google Scholar] [CrossRef] [PubMed]
  67. Peer, D.; Karp, J.M.; Hong, S.; Farokhzad, O.C.; Margalit, R.; Langer, R. Nanocarriers as an emerging platform for cancer therapy. Nat. Nanotechnol. 2007, 2, 751–760. [Google Scholar] [CrossRef] [PubMed]
  68. Rawal, S.; Patel, M. Bio-Nanocarriers for Lung Cancer Management: Befriending the Barriers. Nano-Micro Lett. 2021, 13, 142. [Google Scholar] [CrossRef]
  69. Newman, S.P. Drug delivery to the lungs: Challenges and opportunities. Ther. Deliv. 2017, 8, 647–661. [Google Scholar] [CrossRef]
  70. Plaunt, A.J.; Nguyen, T.L.; Corboz, M.R.; Malinin, V.S.; Cipolla, D.C. Strategies to Overcome Biological Barriers Associated with Pulmonary Drug Delivery. Pharmaceutics 2022, 14, 302. [Google Scholar] [CrossRef]
  71. Kim, S.M.; Faix, P.H.; Schnitzer, J.E. Overcoming key biological barriers to cancer drug delivery and efficacy. J. Control. Release Off. J. Control. Release Soc. 2017, 267, 15–30. [Google Scholar] [CrossRef]
  72. Tang, L.; Yang, X.; Yin, Q.; Cai, K.; Wang, H.; Chaudhury, I.; Yao, C.; Zhou, Q.; Kwon, M.; Hartman, J.A.; et al. Investigating the optimal size of anticancer nanomedicine. Proc. Natl. Acad. Sci. USA 2014, 111, 15344–15349. [Google Scholar] [CrossRef]
  73. Ruffin, R.E.; Dolovich, M.B.; Wolff, R.K.; Newhouse, M.T. The effects of preferential deposition of histamine in the human airway. Am. Rev. Respir. Dis. 1978, 117, 485–492. [Google Scholar] [CrossRef]
  74. Lin, Y.-S.; Tsay, R.-Y. Drug Release from a Spherical Matrix: Theoretical Analysis for a Finite Dissolution Rate Affected by Geometric Shape of Dispersed Drugs. Pharmaceutics 2020, 12, 582. [Google Scholar] [CrossRef] [PubMed]
  75. He, Y.; Xu, J.; Sun, X.; Ren, X.; Maharjan, A.; York, P.; Su, Y.; Li, H.; Zhang, J. Cuboidal tethered cyclodextrin frameworks tailored for hemostasis and injured vessel targeting. Theranostics 2019, 9, 2489–2504. [Google Scholar] [CrossRef] [PubMed]
  76. Yokoi, K.; Tanei, T.; Godin, B.; van de Ven, A.L.; Hanibuchi, M.; Matsunoki, A.; Alexander, J.; Ferrari, M. Serum biomarkers for personalization of nanotherapeutics-based therapy in different tumor and organ microenvironments. Cancer Lett. 2014, 345, 48–55. [Google Scholar] [CrossRef]
  77. Yokoi, K.; Kojic, M.; Milosevic, M.; Tanei, T.; Ferrari, M.; Ziemys, A. Capillary-wall collagen as a biophysical marker of nanotherapeutic permeability into the tumor microenvironment. Cancer Res. 2014, 74, 4239–4246. [Google Scholar] [CrossRef] [PubMed]
  78. Beg, S.; Almalki, W.H.; Khatoon, F.; Alharbi, K.S.; Alghamdi, S.; Akhter, M.H.; Khalilullah, H.; Baothman, A.A.; Hafeez, A.; Rahman, M.; et al. Lipid/polymer-based nanocomplexes in nucleic acid delivery as cancer vaccines. Drug Discov. Today 2021, 26, 1891–1903. [Google Scholar] [CrossRef] [PubMed]
  79. Tenzer, S.; Docter, D.; Kuharev, J.; Musyanovych, A.; Fetz, V.; Hecht, R.; Schlenk, F.; Fischer, D.; Kiouptsi, K.; Reinhardt, C.; et al. Rapid formation of plasma protein corona critically affects nanoparticle pathophysiology. Nat. Nanotechnol. 2013, 8, 772–781. [Google Scholar] [CrossRef]
  80. Nel, A.E.; Mädler, L.; Velegol, D.; Xia, T.; Hoek, E.M.; Somasundaran, P.; Klaessig, F.; Castranova, V.; Thompson, M. Understanding biophysicochemical interactions at the nano-bio interface. Nat. Mater. 2009, 8, 543–557. [Google Scholar] [CrossRef]
  81. Sahay, G.; Alakhova, D.Y.; Kabanov, A.V. Endocytosis of nanomedicines. J. Control. Release Off. J. Control. Release Soc. 2010, 145, 182–195. [Google Scholar] [CrossRef]
  82. Salvati, A.; Pitek, A.S.; Monopoli, M.P.; Prapainop, K.; Bombelli, F.B.; Hristov, D.R.; Kelly, P.M.; Åberg, C.; Mahon, E.; Dawson, K.A. Transferrin-functionalized nanoparticles lose their targeting capabilities when a biomolecule corona adsorbs on the surface. Nat. Nanotechnol. 2013, 8, 137–143. [Google Scholar] [CrossRef]
  83. Docter, D.; Distler, U.; Storck, W.; Kuharev, J.; Wünsch, D.; Hahlbrock, A.; Knauer, S.K.; Tenzer, S.; Stauber, R.H. Quantitative profiling of the protein coronas that form around nanoparticle. Nat. Protoc. 2014, 9, 2030–2044. [Google Scholar] [CrossRef]
  84. Lee, J.J.; Kim, H.J.; Yang, C.S.; Kyeong, H.H.; Choi, J.M.; Hwang, D.E.; Yuk, J.M.; Park, K.; Kim, Y.J.; Lee, S.G.; et al. A high-affinity protein binder that blocks the IL-6/STAT3 signaling pathway effectively suppresses non-small cell lung cancer. Mol. Ther. J. Am. Soc. Gene Ther. 2014, 22, 1254–1265. [Google Scholar] [CrossRef]
  85. He, Y.; Xiong, T.; He, S.; Sun, H.; Huang, C.; Ren, X.; Zhang, J. Pulmonary targeting crosslinked cyclodextrin metal–organic frameworks for lung cancer therapy. Adv. Funct. Mater. 2021, 31, 2004550. [Google Scholar] [CrossRef]
  86. Cheng, M.; Liang, X.; Shi, L.; Zhang, Q.; Zhang, L.; Gong, Z.; Luo, S.; Wang, X.; Zhang, X. Folic acid deficiency exacerbates the inflammatory response of astrocytes after ischemia-reperfusion by enhancing the interaction between IL-6 and JAK-1/pSTAT3. CNS Neurosci. Ther. 2023, 29, 1537–1546. [Google Scholar] [CrossRef] [PubMed]
  87. Hoffmann, M.; Gerlach, S.; Hoffmann, C.; Richter, N.; Hersch, N.; Csiszár, A.; Merkel, R.; Hoffmann, B. PEGylation and folic-acid functionalization of cationic lipoplexes-Improved nucleic acid transfer into cancer cells. Front. Bioeng. Biotechnol. 2022, 10, 1066887. [Google Scholar] [CrossRef]
  88. Rodriguez, P.L.; Harada, T.; Christian, D.A.; Pantano, D.A.; Tsai, R.K.; Discher, D.E. Minimal “Self” peptides that inhibit phagocytic clearance and enhance delivery of nanoparticles. Science 2013, 339, 971–975. [Google Scholar] [CrossRef] [PubMed]
  89. Parodi, A.; Quattrocchi, N.; van de Ven, A.L.; Chiappini, C.; Evangelopoulos, M.; Martinez, J.O.; Brown, B.S.; Khaled, S.Z.; Yazdi, I.K.; Enzo, M.V.; et al. Synthetic nanoparticle NPs functionalized with biomimetic leukocyte membranes possess cell-like functions. Nat. Nanotechnol. 2013, 8, 61–68. [Google Scholar] [CrossRef]
  90. Prapainop, K.; Witter, D.P.; Wentworth, P., Jr. A chemical approach for cell-specific targeting of nanomaterials: Small-molecule-initiated misfolding of nanoparticle corona proteins. J. Am. Chem. Soc. 2012, 134, 4100–4103. [Google Scholar] [CrossRef]
  91. Overchuk, M.; Zheng, G. Overcoming obstacles in the tumor microenvironment: Recent advancements in nanoparticle delivery for cancer theranostics. Biomaterials 2018, 156, 217–237. [Google Scholar] [CrossRef]
  92. Zhou, Y.; Chen, X.; Cao, J.; Gao, H. Overcoming the biological barriers in the tumor microenvironment for improving drug delivery and efficacy. J. Mater. Chem. B 2020, 8, 6765–6781. [Google Scholar] [CrossRef]
  93. Junttila, M.R.; de Sauvage, F.J. Influence of tumour micro-environment heterogeneity on therapeutic response. Nature 2013, 501, 346–354. [Google Scholar] [CrossRef]
  94. Azzi, S.; Hebda, J.K.; Gavard, J. Vascular permeability and drug delivery in cancers. Front. Oncol. 2013, 3, 211. [Google Scholar] [CrossRef] [PubMed]
  95. Fang, J.; Nakamura, H.; Maeda, H. The EPR effect: Unique features of tumor blood vessels for drug delivery, factors involved, and limitations and augmentation of the effect. Adv. Drug Deliv. Rev. 2011, 63, 136–151. [Google Scholar] [CrossRef] [PubMed]
  96. Roma-Rodrigues, C.; Mendes, R.; Baptista, P.V.; Fernandes, A.R. Targeting Tumor Microenvironment for Cancer Therapy. Int. J. Mol. Sci. 2019, 20, 840. [Google Scholar] [CrossRef]
  97. Hinshaw, D.C.; Shevde, L.A. The Tumor Microenvironment Innately Modulates Cancer Progression. Cancer Res. 2019, 79, 4557–4566. [Google Scholar] [CrossRef] [PubMed]
  98. Li, S.; Xu, Z.; Alrobaian, M.; Afzal, O.; Kazmi, I.; Almalki, W.H.; Altamimi, A.S.A.; Al-Abbasi, F.A.; Alharbi, K.S.; Altowayan, W.M.; et al. EGF-functionalized lipid-polymer hybrid nanoparticles of 5-fluorouracil and sulforaphane with enhanced bioavailability and anticancer activity against colon carcinoma. Biotechnol. Appl. Biochem. 2022, 69, 2205–2221. [Google Scholar] [CrossRef]
  99. Ahmad, J.; Ameeduzzafar; Ahmad, M.Z.; Akhter, H.M. Surface-Engineered Cancer Nanomedicine: Rational Design and Recent Progress. Curr. Pharm. Des. 2020, 26, 1181–1190. [Google Scholar] [CrossRef]
  100. Akhter, M.H.; Rizwanullah, M.; Ahmad, J.; Ahsan, M.J.; Mujtaba, M.A.; Amin, S. Nanocarriers in advanced drug targeting: Setting novel paradigm in cancer therapeutics. Artif. Cells Nanomed. Biotechnol. 2018, 46, 873–884. [Google Scholar] [CrossRef]
  101. Schito, L.; Semenza, G. Hypoxia and Breast Cancer Metastasis. In Hypoxia and Cancer. Cancer Drug Discovery and Development; Melillo, G., Ed.; Springer: New York, NY, USA, 2014. [Google Scholar] [CrossRef]
  102. Tashkin, D.P.; Ferguson, G.T. Combination bronchodilator therapy in the management of chronic obstructive pulmonary disease. Respir. Res. 2013, 14, 49. [Google Scholar] [CrossRef]
  103. Tammela, T.; Zarkada, G.; Wallgard, E.; Murtomäki, A.; Suchting, S.; Wirzenius, M.; Waltari, M.; Hellström, M.; Schomber, T.; Peltonen, R.; et al. Blocking VEGFR-3 suppresses angiogenic sprouting and vascular network formation. Nature 2008, 454, 656–660. [Google Scholar] [CrossRef]
  104. Goel, S.; Wong, A.H.; Jain, R.K. Vascular normalization as a therapeutic strategy for malignant and nonmalignant disease. Cold Spring Harb. Perspect. Med. 2012, 2, a006486. [Google Scholar] [CrossRef]
  105. Rodrigo, G.J.; Plaza, V.; Castro-Rodríguez, J.A. Comparison of three combined pharmacological approaches with tiotropium monotherapy in stable moderate to severe COPD: A systematic review. Pulm. Pharmacol. Ther. 2012, 25, 40–47. [Google Scholar] [CrossRef]
  106. Hanjani, N.A.; Esmaelizad, N.; Zanganeh, S.; Gharavi, A.T.; Heidarizadeh, P.; Radfar, M.; Omidi, F.; MacLoughlin, R.; Doroudian, M. Emerging role of exosomes as biomarkers in cancer treatment and diagnosis. Crit. Rev.Oncol. Hematol. 2022, 169, 103565. [Google Scholar] [CrossRef] [PubMed]
  107. Su, S.M.; Kang, P. Recent Advances in Nanocarrier-Assisted Therapeutics Delivery Systems. Pharmaceutics 2020, 12, 837. [Google Scholar] [CrossRef]
  108. Zhang, Z.; Conniot, J.; Amorim, J.; Jin, Y.; Prasad, R.; Yan, X.; Fan, K.; Conde, J. Nucleic acid-based therapy for brain cancer: Challenges and strategies. J. Control. Release Off. J. Control. Release Soc. 2022, 350, 80–92. [Google Scholar] [CrossRef]
  109. Vaughan, H.J.; Green, J.J.; Tzeng, S.Y. Cancer-Targeting Nanoparticles for Combinatorial Nucleic Acid Delivery. Adv. Mater. 2020, 32, e1901081. [Google Scholar] [CrossRef] [PubMed]
  110. Guo, Y.; Cao, X.; Zheng, X.; Abbas, S.K.J.; Li, J.; Tan, W. Construction of nanocarriers based on nucleic acids and their applications in nanobiology delivery systems. Natl. Sci. Rev. 2022, 9, nwac006. [Google Scholar] [CrossRef] [PubMed]
  111. Mendes, B.B.; Conniot, J.; Avital, A.; Yao, D.; Jiang, X.; Zhou, X.; Sharf-Pauker, N.; Xiao, Y.; Adir, O.; Liang, H.; et al. Nanodelivery of nucleic acids. Nat. Rev. Methods Prim. 2022, 2, 24. [Google Scholar] [CrossRef]
  112. Song, W.; Tang, Z.; Zhang, D.; Yu, H.; Chen, X. Coadministration of Vascular Disrupting Agents and Nanomedicines to Eradicate Tumors from Peripheral and Central Regions. Small 2015, 11, 3755–3761. [Google Scholar] [CrossRef]
  113. Sengupta, S.; Eavarone, D.; Capila, I.; Zhao, G.; Watson, N.; Kiziltepe, T.; Sasisekharan, R. Temporal targeting of tumour cells and neovasculature with a nanoscale delivery system. Nature 2005, 436, 568–572. [Google Scholar] [CrossRef]
  114. Shen, N.; Wu, J.; Yang, C.; Yu, H.; Yang, S.; Li, T.; Chen, J.; Tang, Z.; Chen, X. Combretastatin A4 Nanoparticles Combined with Hypoxia-Sensitive Imiquimod: A New Paradigm for the Modulation of Host Immunological Responses during Cancer Treatment. Nano Lett. 2019, 19, 8021–8031. [Google Scholar] [CrossRef]
  115. Liang, Y.; Hao, Y.; Wu, Y.; Zhou, Z.; Li, J.; Sun, X.; Liu, Y.N. Integrated Hydrogel Platform for Programmed Antitumor Therapy Based on Near Infrared-Triggered Hyperthermia and Vascular Disruption. ACS Appl. Mater. Interfaces 2019, 11, 21381–21390. [Google Scholar] [CrossRef]
  116. Siemann, D.W. The unique characteristics of tumor vasculature and preclinical evidence for its selective disruption by Tumor-Vascular Disrupting Agents. Cancer Treat. Rev. 2011, 37, 63–74. [Google Scholar] [CrossRef]
  117. Liu, L.; Schuetze, R.; Gerberich, J.L.; Lopez, R.; Odutola, S.O.; Tanpure, R.P.; Charlton-Sevcik, A.K.; Tidmore, J.K.; Taylor, E.A.-S.; Kapur, P.; et al. Demonstrating Tumor Vascular Disrupting Activity of the Small-Molecule Dihydronaphthalene Tubulin-Binding Agent OXi6196 as a Potential Therapeutic for Cancer Treatment. Cancers 2022, 14, 4208. [Google Scholar] [CrossRef] [PubMed]
  118. Yang, T.; Xiao, H.; Liu, X.; Wang, Z.; Zhang, Q.; Wei, N.; Guo, X. Vascular Normalization: A New Window Opened for Cancer Therapies. Front. Oncol. 2021, 11, 719836. [Google Scholar] [CrossRef] [PubMed]
  119. Magnussen, A.L.; Mills, I.G. Vascular normalisation as the stepping stone into tumour microenvironment transformation. Br. J. Cancer 2021, 125, 324–336. [Google Scholar] [CrossRef] [PubMed]
  120. Chauhan, V.P.; Stylianopoulos, T.; Martin, J.D.; Popović, Z.; Chen, O.; Kamoun, W.S.; Bawendi, M.G.; Fukumura, D.; Jain, R.K. Normalization of tumour blood vessels improves the delivery of nanomedicines in a size-dependent manner. Nat. Nanotechnol. 2012, 7, 383–388. [Google Scholar] [CrossRef]
  121. Sung, Y.C.; Jin, P.R.; Chu, L.A.; Hsu, F.F.; Wang, M.R.; Chang, C.C.; Chiou, S.J.; Qiu, J.T.; Gao, D.Y.; Lin, C.C.; et al. Delivery of nitric oxide with a nanocarrier promotes tumour vessel normalization and potentiates anti-cancer therapies. Nat. Nanotechnol. 2019, 14, 1160–1169. [Google Scholar] [CrossRef]
  122. Bremnes, R.M.; Dønnem, T.; Al-Saad, S.; Al-Shibli, K.; Andersen, S.; Sirera, R.; Camps, C.; Marinez, I.; Busund, L.T. The Role of Tumor Stroma in Cancer Progression and Prognosis: Emphasis on Carcinoma-Associated Fibroblasts and Non-small Cell Lung Cancer. J. Thorac. Oncol. 2011, 6, 209–217. [Google Scholar] [CrossRef] [PubMed]
  123. Järveläinen, H.; Sainio, A.; Koulu, M.; Wight, T.N.; Penttinen, R. Extracellular matrix molecules: Potential targets in pharmacotherapy. Pharmacol. Rev. 2009, 61, 198–223. [Google Scholar] [CrossRef]
  124. Denton, C.P.; Merkel, P.A.; Furst, D.E.; Khanna, D.; Emery, P.; Hsu, V.M.; Silliman, N.; Streisand, J.; Powell, J.; Akesson, A.; et al. Clinical Trials Consortium. Recombinant human anti-transforming growth factor beta1 antibody therapy in systemic sclerosis: A multicenter, randomized, placebo-controlled phase I/II trial of CAT-192. Arthritis Rheum. 2007, 56, 323–333. [Google Scholar] [CrossRef]
  125. Schuster, L.; Seifert, O.; Vollmer, S.; Kontermann, R.E.; Schlosshauer, B.; Hartmann, H. Immunoliposomes for Targeted Delivery of an Antifibrotic Drug. Mol. Pharm. 2015, 12, 3146–3157. [Google Scholar] [CrossRef]
  126. Dykman, L.A.; Khlebtsov, N.G. Gold nanoparticles in chemo-, immuno-, and combined therapy: Review. Biomed. Opt. Express 2019, 10, 3152–3182. [Google Scholar] [CrossRef]
  127. Nwogu, J.I.; Geenen, D.; Bean, M.; Brenner, M.C.; Huang, X.; Buttrick, P.M. Inhibition of collagen synthesis with prolyl 4-hydroxylase inhibitor improves left ventricular function and alters the pattern of left ventricular dilatation after myocardial infarction. Circulation 2001, 104, 2216–2221. [Google Scholar] [CrossRef]
  128. Li, H.; Li, J.; Jiang, X.; Liu, S.; Liu, Y.; Chen, W.; Yang, J.; Zhang, C.; Zhang, W. Melatonin enhances atherosclerotic plaque stability by inducing prolyl-4-hydroxylase α1 expression. J. Hypertens. 2019, 37, 964–971. [Google Scholar] [CrossRef] [PubMed]
  129. Almokadem, S.; Belani, C.P. Volociximab in cancer. Expert Opin. Biol. Ther. 2012, 12, 251–257. [Google Scholar] [CrossRef] [PubMed]
  130. Ingber, D.E. Can cancer be reversed by engineering the tumor microenvironment? Semin. Cancer Biol. 2008, 18, 356–364. [Google Scholar] [CrossRef] [PubMed]
  131. Li, W.; Wang, L.; Sun, T.; Tang, H.; Bui, B.; Cao, D.; Wang, R.; Chen, W. Characterization of nanoparticles combining polyamine detection with photodynamic therapy. Commun. Biol. 2021, 4, 803. [Google Scholar] [CrossRef]
  132. Hu, X.X.; He, P.P.; Qi, G.B.; Gao, Y.J.; Lin, Y.X.; Yang, C.; Yang, P.P.; Hao, H.; Wang, L.; Wang, H. Transformable Nanomaterials as an Artificial Extracellular Matrix for Inhibiting Tumor Invasion and Metastasis. ACS Nano 2017, 11, 4086–4096. [Google Scholar] [CrossRef]
  133. Geckil, H.; Xu, F.; Zhang, X.; Moon, S.; Demirci, U. Engineering hydrogels as extracellular matrix mimics. Nanomedicine 2010, 5, 469–484. [Google Scholar] [CrossRef]
  134. von Ahrens, D.; Bhagat, T.D.; Nagrath, D.; Maitra, A.; Verma, A. The role of stromal cancer-associated fibroblasts in pancreatic cancer. J. Hematol. Oncol. 2017, 10, 76. [Google Scholar] [CrossRef]
  135. Ngan, C.Y.; Yamamoto, H.; Seshimo, I.; Tsujino, T.; Man-i, M.; Ikeda, J.I.; Konishi, K.; Takemasa, I.; Ikeda, M.; Sekimoto, M.; et al. Quantitative evaluation of vimentin expression in tumour stroma of colorectal cancer. Br. J. Cancer 2007, 96, 986–992. [Google Scholar] [CrossRef]
  136. Jiu, Y.; Peränen, J.; Schaible, N.; Cheng, F.; Eriksson, J.E.; Krishnan, R.; Lappalainen, P. Vimentin intermediate filaments control actin stress fiber assembly through GEF-H1 and RhoA. J. Cell Sci. 2017, 130, 892–902. [Google Scholar] [CrossRef]
  137. Javed, R.; Zia, M.; Naz, S.; Aisida, S.O.; Ain, N.U.; Ao, Q. Role of capping agents in the application of nanoparticles in biomedicine and environmental remediation: Recent trends and future prospects. J. Nanobiotechnology 2020, 18, 172. [Google Scholar] [CrossRef]
  138. Jansen, K.; Heirbaut, L.; Cheng, J.D.; Joossens, J.; Ryabtsova, O.; Cos, P.; Maes, L.; Lambeir, A.M.; De Meester, I.; Augustyns, K.; et al. Selective Inhibitors of Fibroblast Activation Protein (FAP) with a (4-Quinolinoyl)-glycyl-2-cyanopyrrolidine Scaffold. ACS Med. Chem. Lett. 2013, 4, 491–496. [Google Scholar] [CrossRef] [PubMed]
  139. Chauhan, V.P.; Chen, I.X.; Tong, R.; Ng, M.R.; Martin, J.D.; Naxerova, K.; Wu, M.W.; Huang, P.; Boucher, Y.; Kohane, D.S.; et al. Reprogramming the microenvironment with tumor-selective angiotensin blockers enhances cancer immunotherapy. Proc. Natl. Acad. Sci. USA 2019, 116, 10674–10680. [Google Scholar] [CrossRef]
  140. Ren, Y.; Jia, H.H.; Xu, Y.Q.; Zhou, X.; Zhao, X.H.; Wang, Y.F.; Song, X.; Zhu, Z.Y.; Sun, T.; Dou, Y.; et al. Paracrine and epigenetic control of CAF-induced metastasis: The role of HOTAIR stimulated by TGF-ß1 secretion. Mol. Cancer 2018, 17, 5. [Google Scholar] [CrossRef]
  141. Fan, W.; Bu, W.; Shen, B.; He, Q.; Cui, Z.; Liu, Y.; Zheng, X.; Zhao, K.; Shi, J. Intelligent MnO2 Nanosheets Anchored with Upconversion Nanoprobes for Concurrent pH-/H2O2-Responsive UCL Imaging and Oxygen-Elevated Synergetic Therapy. Adv. Mater. 2015, 27, 4155–4161. [Google Scholar] [CrossRef]
  142. Liu, Y.; Jiang, Y.; Zhang, M.; Tang, Z.; He, M.; Bu, W. Modulating Hypoxia via Nanomaterials Chemistry for Efficient Treatment of Solid Tumors. Acc. Chem. Res. 2018, 51, 2502–2511. [Google Scholar] [CrossRef]
  143. Baran, N.; Konopleva, M. Molecular Pathways: Hypoxia-Activated Prodrugs in Cancer Therapy. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2017, 23, 2382–2390. [Google Scholar] [CrossRef]
  144. Jayaprakash, P.; Ai, M.; Liu, A.; Budhani, P.; Bartkowiak, T.; Sheng, J.; Ager, C.; Nicholas, C.; Jaiswal, A.R.; Sun, Y.; et al. Targeted hypoxia reduction restores T cell infiltration and sensitizes prostate cancer to immunotherapy. J. Clin. Investig. 2018, 128, 5137–5149. [Google Scholar] [CrossRef] [PubMed]
  145. Fan, Y.T.; Zhou, T.J.; Cui, P.F.; He, Y.J.; Chang, X.; Xing, L.; Jiang, H.L. Modulation of intracellular oxygen pressure by dual-drug nanoparticles to enhance photodynamic therapy. Adv. Funct. Mater. 2019, 29, 1806708. [Google Scholar] [CrossRef]
  146. Hao, G.; Xu, Z.P.; Li, L. Manipulating extracellular tumour pH: An effective target for cancer therapy. RSC Adv. 2018, 8, 22182–22192. [Google Scholar] [CrossRef]
  147. Ando, H.; Eshima, K.; Ishida, T. Neutralization of Acidic Tumor Microenvironment (TME) with Daily Oral Dosing of Sodium Potassium Citrate (K/Na Citrate) Increases Therapeutic Effect of Anti-cancer Agent in Pancreatic Cancer Xenograft Mice Model. Biol. Pharm. Bull. 2021, 44, 266–270. [Google Scholar] [CrossRef] [PubMed]
  148. Pilon-Thomas, S.; Kodumudi, K.N.; El-Kenawi, A.E.; Russell, S.; Weber, A.M.; Luddy, K.; Damaghi, M.; Wojtkowiak, J.W.; Mulé, J.J.; Ibrahim-Hashim, A.; et al. Neutralization of Tumor Acidity Improves Antitumor Responses to Immunotherapy. Cancer Res. 2016, 76, 1381–1390. [Google Scholar] [CrossRef]
  149. Iessi, E.; Logozzi, M.; Mizzoni, D.; Di Raimo, R.; Supuran, C.T.; Fais, S. Rethinking the Combination of Proton Exchanger Inhibitors in Cancer Therapy. Metabolites 2017, 8, 2. [Google Scholar] [CrossRef]
  150. De Milito, A.; Fais, S. Tumor acidity, chemoresistance and proton pump inhibitors. Future Oncol. 2005, 1, 779–786. [Google Scholar] [CrossRef]
  151. Ray, P.; Ferraro, M.; Haag, R.; Quadir, M. Dendritic Polyglycerol-Derived Nano-Architectures as Delivery Platforms of Gemcitabine for Pancreatic Cancer. Macromol. Biosci. 2019, 19, e1900073. [Google Scholar] [CrossRef] [PubMed]
  152. Gadiyar, V.; Patel, G.; Davra, V. Immunological role of TAM receptors in the cancer microenvironment. Int. Rev. Cell Mol. Biol. 2020, 357, 57–79. [Google Scholar] [CrossRef]
  153. Phuengkham, H.; Ren, L.; Shin, I.W.; Lim, Y.T. Nanoengineered Immune Niches for Reprogramming the Immunosuppressive Tumor Microenvironment and Enhancing Cancer Immunotherapy. Adv. Mater. 2019, 31, e1803322. [Google Scholar] [CrossRef] [PubMed]
  154. Zanganeh, S.; Hutter, G.; Spitler, R.; Lenkov, O.; Mahmoudi, M.; Shaw, A.; Pajarinen, J.S.; Nejadnik, H.; Goodman, S.; Zanganeh, S.; et al. Iron oxide nanoparticles inhibit tumour growth by inducing pro-inflammatory macrophage polarization in tumour tissues. Nat. Nanotechnol. 2016, 11, 986–994. [Google Scholar] [CrossRef]
  155. Mulens-Arias, V.; Rojas, J.M.; Pérez-Yagüe, S.; Morales, M.P.; Barber, D.F. Polyethylenimine-coated SPIONs trigger macrophage activation through TLR-4 signaling and ROS production and modulate podosome dynamics. Biomaterials 2015, 52, 494–506. [Google Scholar] [CrossRef]
  156. Li, H.; Zhou, L.; Zhou, J.; Li, Q.; Ji, Q. Underlying mechanisms and drug intervention strategies for the tumour microenvironment. J. Exp. Clin. Cancer Res. 2021, 40, 97. [Google Scholar] [CrossRef]
  157. Groth, C.; Hu, X.; Weber, R.; Fleming, V.; Altevogt, P.; Utikal, J.; Umansky, V. Immunosuppression mediated by myeloid-derived suppressor cells (MDSCs) during tumour progression. Br. J. Cancer 2019, 120, 16–25. [Google Scholar] [CrossRef]
  158. Mi, Y.; Guo, N.; Luan, J.; Cheng, J.; Hu, Z.; Jiang, P.; Jin, W.; Gao, X. The Emerging Role of Myeloid-Derived Suppressor Cells in the Glioma Immune Suppressive Microenvironment. Front. Immunol. 2020, 11, 737. [Google Scholar] [CrossRef]
  159. Dijk, M.V.; Mundt, C.A.; Ritter, G.; Schaer, D.; Wolchok, J.D.; Merghoub, T.; Savitsky, D.A.; Wilson, N.S. Anti-Ctla-4 Antibodies and Methods of Use Thereof. US20190135919A1, 9 May 2019. [Google Scholar]
  160. Amano, T.; Yu, H.; Amano, M.; Leyder, E.; Badiola, M.; Ray, P.; Kim, J.; Ko, A.C.; Achour, A.; Weng, N.P.; et al. Controllable self-replicating RNA vaccine delivered intradermally elicits predominantly cellular immunity. IScience 2023, 26, 106335. [Google Scholar] [CrossRef]
  161. Arbeit, R.D.; Ragan, P.M. Methods for Treating Cancer. JP2019502741A, 31 January 2019. [Google Scholar]
  162. Bertrand, N.; Wu, J.; Xu, X.; Kamaly, N.; Farokhzad, O.C. Cancer nanotechnology: The impact of passive and active targeting in the era of modern cancer biology. Adv. Drug Deliv. Rev. 2014, 66, 2–25. [Google Scholar] [CrossRef] [PubMed]
  163. Tahmasbi Rad, A.; Chen, C.W.; Aresh, W.; Xia, Y.; Lai, P.S.; Nieh, M.P. Combinational Effects of Active Targeting, Shape, and Enhanced Permeability and Retention for Cancer Theranostic Nanocarriers. ACS Appl. Mater. Interfaces 2019, 11, 10505–10519. [Google Scholar] [CrossRef] [PubMed]
  164. Jin, J.; Bhujwalla, Z.M. Biomimetic Nanoparticles Camouflaged in Cancer Cell Membranes and Their Applications in Cancer Theranostics. Front. Oncol. 2019, 9, 1560. [Google Scholar] [CrossRef] [PubMed]
  165. Sushnitha, M.; Evangelopoulos, M.; Tasciotti, E.; Taraballi, F. Cell Membrane-Based Biomimetic Nanoparticles and the Immune System: Immunomodulatory Interactions to Therapeutic Applications. Front. Bioeng. Biotechnol. 2020, 8, 627. [Google Scholar] [CrossRef]
  166. Ullah, S.; Ahmad, A.; Wang, A.; Raza, M.; Jan, A.U.; Tahir, K.; Rahman, A.U.; Qipeng, Y. Bio-fabrication of catalytic platinum nanoparticles and their in vitro efficacy against lungs cancer cells line (A549). J. Photochem. Photobiol. B Biol. 2017, 173, 368–375. [Google Scholar] [CrossRef]
  167. Ullah, S.; Azad, A.K.; Nawaz, A.; Shah, K.U.; Iqbal, M.; Albadrani, G.M.; Al-Joufi, F.A.; Sayed, A.A.; Abdel-Daim, M.M. 5-Fluorouracil-Loaded Folic-Acid-Fabricated Chitosan Nanoparticles for Site-Targeted Drug Delivery Cargo. Polymers 2022, 14, 2010. [Google Scholar] [CrossRef]
  168. Mao, K.; Zhang, W.; Yu, L.; Yu, Y.; Liu, H.; Zhang, X. Transferrin-Decorated Protein-Lipid Hybrid Nanoparticle Efficiently Delivers Cisplatin and Docetaxel for Targeted Lung Cancer Treatment. Drug Des. Dev. Ther. 2021, 15, 3475–3486. [Google Scholar] [CrossRef]
  169. Yadav, B.; Chauhan, M.; Shekhar, S.; Kumar, A.; Mehata, A.K.; Nayak, A.K.; Dutt, R.; Garg, V.; Kailashiya, V.; Muthu, M.S.; et al. RGD-decorated PLGA nanoparticles improved effectiveness and safety of cisplatin for lung cancer therapy. Int. J. Pharm. 2023, 633, 122587. [Google Scholar] [CrossRef]
  170. Akhter, M.H.; Beg, S.; Tarique, M.; Malik, A.; Afaq, S.; Choudhry, H.; Hosawi, S. Receptor-based targeting of engineered nanocarrier against solid tumors: Recent progress and challenges ahead. Biochim. Et Biophys. Acta. Gen. Subj. 2021, 1865, 129777. [Google Scholar] [CrossRef]
  171. Zhang, W.; Yu, L.; Ji, T.; Wang, C. Tumor Microenvironment-Responsive Peptide-Based Supramolecular Drug Delivery System. Front. Chem. 2020, 8, 549. [Google Scholar] [CrossRef] [PubMed]
  172. He, Q.; Chen, J.; Yan, J.; Cai, S.; Xiong, H.; Liu, Y.; Peng, D.; Mo, M.; Liu, Z. Tumor microenvironment responsive drug delivery systems. Asian J. Pharm. Sci. 2020, 15, 416–448. [Google Scholar] [CrossRef] [PubMed]
  173. Morales-Cruz, M.; Delgado, Y.; Castillo, B.; Figueroa, C.M.; Molina, A.M.; Torres, A.; Milián, M.; Griebenow, K. Smart Targeting To Improve Cancer Therapeutics. Drug Des. Dev. Ther. 2019, 13, 3753–3772. [Google Scholar] [CrossRef]
  174. Zhang, Z.; Wang, H.; Tan, T.; Li, J.; Wang, Z.; Li, Y. Rational Design of Nanoparticles with Deep Tumor Penetration for Effective Treatment of Tumor Metastasis. Adv. Funct. Mater. 2018, 28, 1801840. [Google Scholar] [CrossRef]
  175. Hu, C.; Cun, X.; Ruan, S.; Liu, R.; Xiao, W.; Yang, X.; Yang, Y.; Yang, C.; Gao, H. Enzyme-triggered size shrink and laser-enhanced NO release nanoparticles for deep tumor penetration and combination therapy. Biomaterials 2018, 168, 64–75. [Google Scholar] [CrossRef]
  176. Chen, J.; Ding, J.; Wang, Y.; Cheng, J.; Ji, S.; Zhuang, X.; Chen, X. Sequentially Responsive Shell-Stacked Nanoparticles for Deep Penetration into Solid Tumors. Adv. Mater. 2017, 29, 1701170. [Google Scholar] [CrossRef]
  177. Kumari, R.; Sunil, D.; Ningthoujam, R.S. Hypoxia-responsive nanoparticle based drug delivery systems in cancer therapy: An up-to-date review. J. Control. Release Off. J. Control. Release Soc. 2020, 319, 135–156. [Google Scholar] [CrossRef]
  178. Niu, Y.; Zhu, J.; Li, Y.; Shi, H.; Gong, Y.; Li, R.; Huo, Q.; Ma, T.; Liu, Y. Size shrinkable drug delivery nanosystems and priming the tumor microenvironment for deep intratumoral penetration of nanoparticles. J. Control. Release Off. J. Control. Release Soc. 2018, 277, 35–47. [Google Scholar] [CrossRef]
  179. Zhang, X.; An, L.; Tian, Q.; Lin, J.; Yang, S. Tumor-microenvironment activated second near-infrared agents for tumor imaging and therapy. J. Mater. Chem. B 2020, 8, 4738–4747. [Google Scholar] [CrossRef]
  180. Han, H.; Valdepérez, D.; Jin, Q.; Yang, B.; Li, Z.; Wu, Y.; Pelaz, B.; Parak, W.J.; Ji, J. Dual Enzymatic Reaction-Assisted Gemcitabine Delivery Systems for Programmed Pancreatic Cancer Therapy. ACS Nano 2017, 11, 1281–1291. [Google Scholar] [CrossRef]
  181. Li, S.; Wu, W.; Xiu, K.; Xu, F.; Li, Z.; Li, J. Doxorubicin loaded pH-responsive micelles capable of rapid intracellular drug release for potential tumor therapy. J. Biomed. Nanotechnol. 2014, 10, 1480–1489. [Google Scholar] [CrossRef]
  182. Thambi, T.; Deepagan, V.G.; Yoon, H.Y.; Han, H.S.; Kim, S.H.; Son, S.; Jo, D.G.; Ahn, C.H.; Suh, Y.D.; Kim, K.; et al. Hypoxia-responsive polymeric nanoparticles for tumor-targeted drug delivery. Biomaterials 2014, 35, 1735–1743. [Google Scholar] [CrossRef] [PubMed]
  183. Hossen, S.; Hossain, M.K.; Basher, M.K.; Mia, M.N.H.; Rahman, M.T.; Uddin, M.J. Smart nanocarrier-based drug delivery systems for cancer therapy and toxicity studies: A review. J. Adv. Res. 2018, 15, 1–18. [Google Scholar] [CrossRef]
  184. Wang, J.; Seebacher, N.; Shi, H.; Kan, Q.; Duan, Z. Novel strategies to prevent the development of multidrug resistance (MDR) in cancer. Oncotarget 2017, 8, 84559–84571. [Google Scholar] [CrossRef] [PubMed]
  185. Housman, G.; Byler, S.; Heerboth, S.; Lapinska, K.; Longacre, M.; Snyder, N.; Sarkar, S. Drug resistance in cancer: An overview. Cancers 2014, 6, 1769–1792. [Google Scholar] [CrossRef]
  186. Sever, R.; Brugge, J.S. Signal transduction in cancer. Cold Spring Harb. Perspect. Med. 2015, 5, a006098. [Google Scholar] [CrossRef] [PubMed]
  187. Ding, C.; Tong, L.; Feng, J.; Fu, J. Recent Advances in Stimuli-Responsive Release Function Drug Delivery Systems for Tumor Treatment. Molecules 2016, 21, 1715. [Google Scholar] [CrossRef] [PubMed]
  188. Alves, A.C.S.; Bruinsmann, F.A.; Guterres, S.S.; Pohlmann, A.R. Organic Nanocarriers for Bevacizumab Delivery: An Overview of Development, Characterization and Applications. Molecules 2021, 26, 4127. [Google Scholar] [CrossRef]
  189. Peng, Y.; Bariwal, J.; Kumar, V.; Tan, C.; Mahato, R.I. Organic nanocarriers for delivery and targeting of therapeutic agents for cancer treatment. Adv. Ther. 2020, 3, 1900136. [Google Scholar] [CrossRef]
  190. Abdelaziz, H.M.; Freag, M.S.; Elzoghby, A.O. Solid lipid nanoparticle-based drug delivery for lung cancer. In Nanotechnology-Based Targeted Drug Delivery Systems for Lung Cancer; Academic Press: Cambridge, MA, USA, 2019; pp. 95–121. [Google Scholar] [CrossRef]
  191. Dhiman, N.; Awasthi, R.; Sharma, B.; Kharkwal, H.; Kulkarni, G.T. Lipid Nanoparticles as Carriers for Bioactive Delivery. Front. Chem. 2021, 9, 580118. [Google Scholar] [CrossRef] [PubMed]
  192. Kutlu, M.; Kus, G.; Ulukaya, E. Lipid Nanoparticles Loaded with Ceranib-2 as Anticancer Agents. WO Patent 2020018049A3, 27 February 2020. [Google Scholar]
  193. Naseri, N.; Valizadeh, H.; Zakeri-Milani, P. Solid Lipid Nanoparticles and Nanostructured Lipid Carriers: Structure, Preparation and Application. Adv. Pharm. Bull. 2015, 5, 305–313. [Google Scholar] [CrossRef]
  194. 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. Off. J. Arb. Fur Pharm. V 2008, 68, 545–554. [Google Scholar] [CrossRef] [PubMed]
  195. Naguib, Y.W.; Rodriguez, B.L.; Li, X.; Hursting, S.D.; Williams, R.O.; Cui, Z., 3rd. Solid lipid nanoparticle formulations of docetaxel prepared with high melting point triglycerides: In vitro and in vivo evaluation. Mol. Pharm. 2014, 11, 1239–1249. [Google Scholar] [CrossRef] [PubMed]
  196. Yan, Z. Preparation Method of Folic Acid Targeted Silymarin Solid Lipid Nanoparticles. China Patent 111195239A, 26 May 2020. [Google Scholar]
  197. Chamundeeswari, M.; Jeslin, J.; Verma, M.L. Nanocarriers for drug delivery applications. Environ. Chem. Lett. 2019, 17, 849–865. [Google Scholar] [CrossRef]
  198. Weber, S.; Zimmer, A.; Pardeike, J. Solid Lipid Nanoparticles (SLN) and Nanostructured Lipid Carriers (NLC) for pulmonary application: A review of the state of the art. Eur. J. Pharm. Biopharm. Off. J. Arb. Fur Pharm. V 2014, 86, 7–22. [Google Scholar] [CrossRef]
  199. Hu, L.; Jia, Y. Preparation and characterization of solid lipid nanoparticles loaded with epirubicin for pulmonary delivery. Die Pharm. Int. J. Pharm. Sci. 2010, 65, 585–587. [Google Scholar]
  200. da Rocha, M.C.O.; da Silva, P.B.; Radicchi, M.A.; Andrade, B.Y.G.; de Oliveira, J.V.; Venus, T.; Merker, C.; Estrela-Lopis, I.; Longo, J.P.F.; Báo, S.N. Docetaxel-loaded solid lipid nanoparticles prevent tumor growth and lung metastasis of 4T1 murine mammary carcinoma cells. J. Nanobiotechnol. 2020, 18, 43. [Google Scholar] [CrossRef]
  201. Pi, C.; Zhao, W.; Zeng, M.; Yuan, J.; Shen, H.; Li, K.; Su, Z.; Liu, Z.; Wen, J.; Song, X.; et al. Anti-lung cancer effect of paclitaxel solid lipid nanoparticles delivery system with curcumin as co-loading partner in vitro and in vivo. Drug Deliv. 2022, 29, 1878–1891. [Google Scholar] [CrossRef]
  202. Han, Y.; Zhang, P.; Chen, Y.; Sun, J.; Kong, F. Co-delivery of plasmid DNA and doxorubicin by solid lipid nanoparticles for lung cancer therapy. Int. J. Mol. Med. 2014, 34, 191–196. [Google Scholar] [CrossRef] [PubMed]
  203. Pilch, E.; Musiał, W. Liposomes with an Ethanol Fraction as an Application for Drug Delivery. Int. J. Mol. Sci. 2018, 19, 3806. [Google Scholar] [CrossRef]
  204. Sercombe, L.; Veerati, T.; Moheimani, F.; Wu, S.Y.; Sood, A.K.; Hua, S. Advances and Challenges of Liposome Assisted Drug Delivery. Front. Pharmacol. 2015, 6, 286. [Google Scholar] [CrossRef] [PubMed]
  205. Skupin-Mrugalska, P. Liposome-based drug delivery for lung cancer. In Nanotechnology-Based Targeted Drug Delivery Systems for Lung Cancer; Academic Press: Cambridge, MA, USA, 2019; pp. 123–160. [Google Scholar] [CrossRef]
  206. Kaihua, L.; Liang, H.; Meiling, Z. Active Targeting Liposome for Resisting Lung Cancer and Preparing Method and Application of Active Targeting Liposome. China Patent 105726483B, 3 February 2016. [Google Scholar]
  207. Blanchett, S.F.; Drummond, D.C.; Fitzgerald, J.B.; Moyo, V. Combination Therapy Using Liposomal Irinotecan and PARP Inhibitors for Cancer Treatment. Japanese Patent 2018528184A, 27 September 2018. [Google Scholar]
  208. Srivastava, M.K.; Zhu, L.; Harris-White, M.; Kar, U.K.; Huang, M.; Johnson, M.F.; Lee, J.M.; Elashoff, D.; Strieter, R.; Dubinett, S.; et al. Myeloid suppressor cell depletion augments antitumor activity in lung cancer. PLoS ONE 2012, 7, e40677. [Google Scholar] [CrossRef]
  209. Akbarzadeh, A.; Rezaei-Sadabady, R.; Davaran, S.; Joo, S.W.; Zarghami, N.; Hanifehpour, Y.; Samiei, M.; Kouhi, M.; Nejati-Koshki, K. Liposome: Classification, preparation, and applications. Nanoscale Res. Lett. 2013, 8, 102. [Google Scholar] [CrossRef] [PubMed]
  210. Yang, J.; Wu, W.; Wen, J.; Ye, H.; Luo, H.; Bai, P.; Tang, M.; Wang, F.; Zheng, L.; Yang, S.; et al. Liposomal honokiol induced lysosomal degradation of Hsp90 client proteins and protective autophagy in both gefitinib-sensitive and gefitinib-resistant NSCLC cells. Biomaterials 2017, 141, 188–198. [Google Scholar] [CrossRef]
  211. 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. 2016, 12, 251–261. [Google Scholar] [CrossRef]
  212. Grabarnick Portnoy, E.; Andriyanov, A.V.; Han, H.; Eyal, S.; Barenholz, Y. PEGylated Liposomes Remotely Loaded with the Combination of Doxorubicin, Quinine, and Indocyanine Green Enable Successful Treatment of Multidrug-Resistant Tumors. Pharmaceutics 2021, 13, 2181. [Google Scholar] [CrossRef]
  213. Kibria, G.; Hatakeyama, H.; Sato, Y.; Harashima, H. Anti-Tumor Effect via Passive Anti-angiogenesis of PEGylated Liposomes Encapsulating Doxorubicin in Drug Resistant Tumors. Int. J. Pharm. 2016, 509, 178–187. [Google Scholar] [CrossRef]
  214. Gabizon, A.; Martin, F. Polyethylene glycol-coated (pegylated) liposomal doxorubicin. Rationale for use in solid tumours. Drugs 1997, 54 (Suppl. S4), 15–21. [Google Scholar] [CrossRef] [PubMed]
  215. Jiménez-López, J.; Bravo-Caparrós, I.; Cabeza, L.; Nieto, F.R.; Ortiz, R.; Perazzoli, G.; Fernández-Segura, E.; Cañizares, F.J.; Baeyens, J.M.; Melguizo, C.; et al. Paclitaxel antitumor effect improvement in lung cancer and prevention of the painful neuropathy using large pegylated cationic liposomes. Biomed. Pharmacother. Biomed. Pharmacother. 2021, 133, 111059. [Google Scholar] [CrossRef]
  216. Sapra, P.; Tyagi, P.; Allen, T.M. Ligand-targeted liposomes for cancer treatment. Curr. Drug Deliv. 2005, 2, 369–381. [Google Scholar] [CrossRef] [PubMed]
  217. Yu, B.; Tai, H.C.; Xue, W.; Lee, L.J.; Lee, R.J. Receptor-targeted nanocarriers for therapeutic delivery to cancer. Mol. Membr. Biol. 2010, 27, 286–298. [Google Scholar] [CrossRef] [PubMed]
  218. Noble, G.T.; Stefanick, J.F.; Ashley, J.D.; Kiziltepe, T.; Bilgicer, B. Ligand-targeted liposome design: Challenges and fundamental considerations. Trends Biotechnol. 2014, 32, 32–45. [Google Scholar] [CrossRef]
  219. Naik, H.; Sonju, J.J.; Singh, S.; Chatzistamou, I.; Shrestha, L.; Gauthier, T.; Jois, S. Lipidated Peptidomimetic Ligand-Functionalized HER2 Targeted Liposome as Nano-Carrier Designed for Doxorubicin Delivery in Cancer Therapy. Pharmaceuticals 2021, 14, 221. [Google Scholar] [CrossRef]
  220. Lin, C.; Zhang, X.; Chen, H.; Bian, Z.; Zhang, G.; Riaz, M.K.; Tyagi, D.; Lin, G.; Zhang, Y.; Wang, J.; et al. Dual-ligand modified liposomes provide effective local targeted delivery of lung-cancer drug by antibody and tumor lineage-homing cell-penetrating peptide. Drug Deliv. 2018, 25, 256–266. [Google Scholar] [CrossRef]
  221. Muthu, M.S.; Feng, S.S. Theranostic liposomes for cancer diagnosis and treatment: Current development and pre-clinical success. Expert Opin. Drug Deliv. 2013, 10, 151–155. [Google Scholar] [CrossRef]
  222. Jeon, M.; Kim, G.; Lee, W.; Baek, S.; Jung, H.N.; Im, H.-J. Development of theranostic dual-layered Au-liposome for effective tumor targeting and photothermal therapy. J. Nanobiotechnol. 2021, 19, 262. [Google Scholar] [CrossRef]
  223. Karpuz, M.; Silindir-Gunay, M.; Ozer, A.Y.; Ozturk, S.C.; Yanik, H.; Tuncel, M.; Aydin, C.; Esendagli, G. Diagnostic and therapeutic evaluation of folate-targeted paclitaxel and vinorelbine encapsulating theranostic liposomes for non-small cell lung cancer. Eur. J. Pharm. Sci. Off. J. Eur. Fed. Pharm. Sci. 2021, 156, 105576. [Google Scholar] [CrossRef]
  224. Lu, Y.; Zhang, E.; Yang, J.; Cao, Z. Strategies to improve micelle stability for drug delivery. Nano Res. 2018, 11, 4985–4998. [Google Scholar] [CrossRef]
  225. Hanafy, N.A.N.; El-Kemary, M.; Leporatti, S. Micelles Structure Development as a Strategy to Improve Smart Cancer Therapy. Cancers 2018, 10, 238. [Google Scholar] [CrossRef] [PubMed]
  226. Panagi, M.; Mpekris, F.; Chen, P.; Voutouri, C.; Nakagawa, Y.; Martin, D.J.; Hiroko, H.; Hashimoto, H.; Demetriou, P.; Pierides, C.; et al. Polymeric micelles effectively reprogram the tumor microenvironment to potentiate nano-immunotherapy in mouse breast cancer models. Nat. Commun. 2022, 13, 7165. [Google Scholar] [CrossRef] [PubMed]
  227. Kinoh, H.; Kataoka, K.; Cabral, H.; Miura, Y.; Fukushima, S.; Nishiyama, N.; Chida, T. Micelle Containing Epirubicin-Complexed Block Copolymer and Anti-Cancer Agent, and Pharmaceutical Composition Containing Said Micelle Applicable to Treatment of Cancer, Resistant Cancer or Metastatic Cancer. US Patent 10220026B2, 5 March 2019. [Google Scholar]
  228. 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. Off. J. Eur. Soc. Med. Oncol. 2007, 18, 2009–2014. [Google Scholar] [CrossRef] [PubMed]
  229. Bao, Y.; Deng, Q.; Li, Y.; Zhou, S. Engineering docetaxel-loaded micelles for non-small cell lung cancer: A comparative study of microfluidic and bulk nanoparticle preparation. RSC Adv. 2018, 8, 31950–31966. [Google Scholar] [CrossRef]
  230. Lamprecht, A.; Benoit, J.P. Etoposide nanocarriers suppress glioma cell growth by intracellular drug delivery and simultaneous P-glycoprotein inhibition. J. Control. Release Off. J. Control. Release Soc. 2006, 112, 208–213. [Google Scholar] [CrossRef] [PubMed]
  231. Huynh, N.T.; Passirani, C.; Saulnier, P.; Benoit, J.P. Lipid nanocapsules: A new platform for nanomedicine. Int. J. Pharm. 2009, 379, 201–209. [Google Scholar] [CrossRef]
  232. Mouzouvi, C.R.A.; Umerska, A.; Bigot, A.K.; Saulnier, P. Surface active properties of lipid nanocapsules. PLoS ONE 2017, 12, e0179211. [Google Scholar] [CrossRef]
  233. Kim, J.; Ramasamy, T.; Choi, J.Y.; Kim, S.T.; Youn, Y.S.; Choi, H.G.; Yong, C.S.; Kim, J.O. PEGylated polypeptide lipid nanocapsules to enhance the anticancer efficacy of erlotinib in non-small cell lung cancer. Colloids Surf. B Biointerfaces 2017, 150, 393–401. [Google Scholar] [CrossRef] [PubMed]
  234. Schultze, E.; Ourique, A.; Yurgel, V.C.; Begnini, K.R.; Thurow, H.; de Leon, P.M.M.; Campos, V.F.; Dellagostin, O.A.; Guterres, S.R. Encapsulation in lipid-core nanocapsules overcomes lung cancer cell resistance to tretinoin. Eur. J. Pharm. Biopharm. 2014, 87, 55–63. [Google Scholar] [CrossRef] [PubMed]
  235. AkhoondZardini, A.; Mohebbi, M.; Farhoosh, R.; Bolurian, S. Production and characterization of nanostructured lipid carriers and solid lipid nanoparticles containing lycopene for food fortification. J. Food Sci. Technol. 2018, 55, 287–298. [Google Scholar] [CrossRef] [PubMed]
  236. Beloqui, A.; Solinís, M.Á.; Rodríguez-Gascón, A.; Almeida, A.J.; Préat, V. Nanostructured lipid carriers: Promising drug delivery systems for future clinics. Nanomed. Nanotechnol. Biol. Med. 2016, 12, 143–161. [Google Scholar] [CrossRef]
  237. Chauhan, I.; Yasir, M.; Verma, M.; Singh, A.P. Nanostructured Lipid Carriers: A Groundbreaking Approach for Transdermal Drug Delivery. Adv. Pharm. Bull. 2020, 10, 150–165. [Google Scholar] [CrossRef]
  238. Han, Y.; Li, Y.; Zhang, P.; Sun, J.; Li, X.; Sun, X.; Kong, F. Nanostructured lipid carriers as novel drug delivery system for lung cancer gene therapy. Pharm. Dev. Technol. 2016, 21, 277–281. [Google Scholar] [CrossRef] [PubMed]
  239. 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. Control. Release Off. J. Control. Release Soc. 2013, 171, 349–357. [Google Scholar] [CrossRef]
  240. Wang, J.Y.; Song, Y.Q.; Peng, J.; Luo, H.L. Nanostructured Lipid Carriers Delivering Sorafenib to Enhance Immunotherapy Induced by Doxorubicin for Effective Esophagus Cancer Therapy. ACS Omega 2020, 5, 22840–22846. [Google Scholar] [CrossRef]
  241. Peltonen, L.; Singhal, M.; Hirvonen, J. Principles of nanosized drug delivery systems. In Nanoengineered Biomaterials for Advanced Drug Delivery; Elsevier: Amsterdam, The Netherlands, 2020; pp. 3–25. [Google Scholar] [CrossRef]
  242. Madkour, L.H. Polymer nanoparticle drug-nucleic acid combinations. In Nucleic Acids as Gene Anticancer Drug Delivery Therapy; Academic Press: London, UK, 2019; pp. 241–255. [Google Scholar] [CrossRef]
  243. Thangam, R.; Patel, K.D.; Kang, H.; Paulmurugan, R. Advances in Engineered Polymer Nanoparticle Tracking Platforms towards Cancer Immunotherapy-Current Status and Future Perspectives. Vaccines 2021, 9, 935. [Google Scholar] [CrossRef] [PubMed]
  244. Novoselova, M.V.; Loh, H.M.; Trushina, D.B.; Ketkar, A.; Abakumova, T.O.; Zatsepin, T.S.; Kakran, M.; Brzozowska, A.M.; Lau, H.H.; Gorin, D.A.; et al. Biodegradable Polymeric Multilayer Capsules for Therapy of Lung Cancer. ACS Appl. Mater. Interfaces 2020, 12, 5610–5623. [Google Scholar] [CrossRef]
  245. Patel, P.; Raval, M.; Manvar, A.; Airao, V.; Bhatt, V.; Shah, P. Lung cancer targeting efficiency of Silibinin loaded Poly Caprolactone /Pluronic F68 Inhalable nanoparticles: In vitro and In vivo study. PLoS ONE 2022, 17, e0267257. [Google Scholar] [CrossRef] [PubMed]
  246. Wright, J. Docetaxel Polymeric Nanoparticles for Cancer Treatment. WO Patent 2014210485A1, 31 December 2014. [Google Scholar]
  247. Trieu, V.; Motamed, K. Nanoparticle Formulations. WO 2014165829A2, 31 December 2014. [Google Scholar]
  248. Xiaolong, Z.; Jingfeng, L.; Xiaolong, L.; Yongyi, Z.; Ming, W. Hypoxic Response Polymer Nanoparticle and Application Thereof. Chinese Patent 108653288B, 5 May 2020. [Google Scholar]
  249. Gorain, B.; Choudhury, H.; Pandy, M.; Nair, A.; Amin, M.C.I.M.; Molugulu, N.; Deb, P.K.; Tripathi, P.K.; Khurana, S.; Shukla, R.; et al. Dendrimer-Based Nanocarriers in Lung Cancer Therapy. Nanotechnol. Based Target. Drug Deliv. Syst. Lung Cancer. [CrossRef]
  250. Abbasi, E.; Aval, S.F.; Akbarzadeh, A.; Milani, M.; Nasrabadi, H.T.; Joo, S.W.; Hanifehpour, Y.; Nejati-Koshki, K.; Pashaei-Asl, R. Dendrimers: Synthesis, applications, and properties. Nanoscale Res. Lett. 2014, 9, 247. [Google Scholar] [CrossRef]
  251. Yang, H. Targeted nanosystems: Advances in targeted dendrimers for cancer therapy. Nanomed.Nanotechnol. Biol. Med. 2016, 12, 309–316. [Google Scholar] [CrossRef] [PubMed]
  252. Haag, R.; Kratz, F. Polymer therapeutics: Concepts and applications. Angew. Chem. 2006, 45, 1198–1215. [Google Scholar] [CrossRef] [PubMed]
  253. Haag, R. Supramolecular drug-delivery systems based on polymeric core-shell architectures. Angew. Chem. 2004, 43, 278–282. [Google Scholar] [CrossRef] [PubMed]
  254. Amreddy, N.; Babu, A.; Panneerselvam, J.; Srivastava, A.; Muralidharan, R.; Chen, A.; Zhao, Y.D.; Munshi, A.; Ramesh, R. Chemo-biologic combinatorial drug delivery using folate receptor-targeted dendrimer nanoparticles for lung cancer treatment. Nanomed. Nanotechnol. Biol. Med. 2018, 14, 373–384. [Google Scholar] [CrossRef]
  255. Mccoull, W.; Ashford, M.A.; Grant, I.; Hennessy, E.J.; Secrist, J.P.; Owen, D.; Kelly, B.; Giannis, M. Therapeutic Dendrimers. WO Patent 2018154004A9, 27 December 2018. [Google Scholar]
  256. Schirinzi, G.F.; Pérez-Pomeda, I.; Sanchís, J.; Rossini, C.; Farré, M.; Barceló, D. Cytotoxic effects of commonly used nanomaterials and microplastics on cerebral and epithelial human cells. Environ. Res. 2017, 159, 579–587. [Google Scholar] [CrossRef]
  257. Cohen, J.T.; Carlson, G.; Charnley, G.; Coggon, D.; Delzell, E.; Graham, J.D.; Greim, H.; Krewski, D.; Medinsky, M.; Monson, R.; et al. A comprehensive evaluation of the potential health risks associated with occupational and environmental exposure to styrene. J. Toxicol. Environ. Health. Part B Crit. Rev. 2002, 5, 1–265. [Google Scholar] [CrossRef]
  258. Huff, J.; Infante, P.F. Styrene exposure and risk of cancer. Mutagenesis 2011, 26, 583–584. [Google Scholar] [CrossRef]
  259. Kik, K.; Bukowska, B.; Sicińska, P. Polystyrene nanoparticles: Sources, occurrence in the environment, distribution in tissues, accumulation and toxicity to various organisms. Environ. Pollut. 2020, 262, 114297. [Google Scholar] [CrossRef]
  260. Weinberg, F.; Ramnath, N.; Nagrath, D. Reactive Oxygen Species in the Tumor Microenvironment: An Overview. Cancers 2019, 11, 1191. [Google Scholar] [CrossRef]
  261. Murali, K.; Kenesei, K.; Li, Y.; Demeter, K.; Környei, Z.; Madarász, E. Uptake and bio-reactivity of polystyrene nanoparticles is affected by surface modifications, ageing and LPS adsorption: In vitro studies on neural tissue cells. Nanoscale 2015, 7, 4199–4210. [Google Scholar] [CrossRef] [PubMed]
  262. Anguissola, S.; Garry, D.; Salvati, A.; O’Brien, P.J.; Dawson, K.A. High content analysis provides mechanistic insights on the pathways of toxicity induced by amine-modified polystyrene nanoparticles. PLoS ONE 2014, 9, e108025. [Google Scholar] [CrossRef] [PubMed]
  263. Shi, X.; Wang, X.; Huang, R.; Tang, C.; Hu, C.; Ning, P.; Wang, F. Cytotoxicity and Genotoxicity of Polystyrene Micro- and Nanoplastics with Different Size and Surface Modification in A549 Cells. Int. J. Nanomed. 2022, 17, 4509–4523. [Google Scholar] [CrossRef] [PubMed]
  264. Eatemadi, A.; Daraee, H.; Karimkhanloo, H.; Kouhi, M.; Zarghami, N.; Akbarzadeh, A.; Abasi, M.; Hanifehpour, Y.; Joo, S.W. Carbon nanotubes: Properties, synthesis, purification, and medical applications. Nanoscale Res. Lett. 2014, 9, 393. [Google Scholar] [CrossRef]
  265. Saleem, J.; Wang, L.; Chen, C. Carbon-Based Nanomaterials for Cancer Therapy via Targeting Tumor Microenvironment. Adv. Healthc. Mater. 2018, 7, e1800525. [Google Scholar] [CrossRef]
  266. Sheikhpour, M.; Naghinejad, M.; Kasaeian, A.; Lohrasbi, A.; Shahraeini, S.S.; Zomorodbakhsh, S. The Applications of Carbon Nanotubes in the Diagnosis and Treatment of Lung Cancer: A Critical Review. Int. J. Nanomed. 2020, 15, 7063–7078. [Google Scholar] [CrossRef]
  267. Simon, J.; Flahaut, E.; Golzio, M. Overview of Carbon Nanotubes for Biomedical Applications. Materials 2019, 12, 624. [Google Scholar] [CrossRef]
  268. Kamazani, F.M.; Sotoodehnejadnematalahi, F.; Siadat, S.D.; Pornour, M.; Sheikhpour, M.A. Success targeted nano delivery to lung cancer cells with multi-walled carbon nanotubes conjugated to bromocriptine. Sci. Rep. 2021, 11, 24419. [Google Scholar] [CrossRef]
  269. Srivastava, R.K.; Pant, A.B.; Kashyap, M.P.; Kumar, V.; Lohani, M.; Jonas, L.; Rahman, Q. Multi-walled carbon nanotubes induce oxidative stress and apoptosis in human lung cancer cell line-A549. Nanotoxicology 2011, 5, 195–207. [Google Scholar] [CrossRef]
  270. Khang, D.W.; Kang, S.S.; Choi, J.; Nam, T.H. Carbon Nanotube-Based Anti-Cancer Agent Capable of Suppressing Drug Resistance. US Patent 20150196650A1, 16 July 2015. [Google Scholar]
  271. Roger, G., Jr.; Harrison, R.G.; Resasco, D.E. Compositions and Methods for Cancer Treatment Using Targeted Carbon Nanotubes. US Patent 20100184669A1, 22 July 2010. [Google Scholar]
  272. Lin, G.; Mi, P.; Chu, C.; Zhang, J.; Liu, G. Inorganic Nanocarriers Overcoming Multidrug Resistance for Cancer Theranostics. Adv. Sci. 2016, 3, 1600134. [Google Scholar] [CrossRef]
  273. Ghosn, Y.; Kamareddine, M.H.; Tawk, A.; Elia, C.; El Mahmoud, A.; Terro, K.; El Harake, N.; El-Baba, B.; Makdessi, J.; Farhat, S. Inorganic Nanoparticles as Drug Delivery Systems and Their Potential Role in the Treatment of Chronic Myelogenous Leukaemia. Technol. Cancer Res. Treat. 2019, 18, 1533033819853241. [Google Scholar] [CrossRef]
  274. Chen, S.; Hao, X.; Liang, X.; Zhang, Q.; Zhang, C.; Zhou, G.; Shen, S.; Jia, G.; Zhang, J. Inorganic Nanomaterials as Carriers for Drug Delivery. J. Biomed. Nanotechnol. 2016, 12, 1–27. [Google Scholar] [CrossRef]
  275. Desai, N.; Momin, M.; Khan, T.; Gharat, S.; Ningthoujam, R.S.; Omri, A. Metallic nanoparticles as drug delivery system for the treatment of cancer. Expert Opin. Drug Deliv. 2021, 18, 1261–1290. [Google Scholar] [CrossRef] [PubMed]
  276. Chandrakala, V.; Aruna, V.; Angajala, G. Review on metal nanoparticles as nanocarriers: Current challenges and perspectives in drug delivery systems. Emergent Mater. 2022, 5, 1593–1615. [Google Scholar] [CrossRef] [PubMed]
  277. Amrita, V.V.P.U. Nanoparticles Comprising Sorafenib. World Patent 2014087413A1, 12 October 2014. [Google Scholar]
  278. Minji, S.; Chenggen, Q.; Xue, Y. Cisplatin Prodrug-Manganese Dioxide Nano Drug-Loading System and Preparation Method and Application Thereof. Chinese Patent 111214488A, 2 June 2020. [Google Scholar]
  279. Jun, G.; Yuan, L.; Licheng, W.; Han, H. Double-Effect Treatment Targeted Drug Delivery System and preparation Method and Application Thereof. Chinese Patent 110652497A, 7 January 2020. [Google Scholar]
  280. Bin, D.; Jingshu, H.; Mei, Z.; Yuehua, W.; Huizhen, M.; Yimeng, B.; Qingqing, J.; Mengmeng, Z.; Mengxue, P.; Lili, Y. Preparation Method of Cholesterol Oxidase Modified Hybrid Metal Organic Framework Tumor Targeting Nano Preparation. Chinese Patent 112274648A, 29 January 2021. [Google Scholar]
  281. Iqbal, M.; Usanase, G.; Oulmi, K.; Aberkane, F.; Bendaikha, T.; Fessi, H.; Zine, N.; Agusti, G.; Errachid, E.-S.; Elaissari, A. Preparation of gold nanoparticles and determination of their particles size via different methods. Mater. Res. Bull. 2016, 79, 97–104. [Google Scholar] [CrossRef]
  282. Yeh, Y.C.; Creran, B.; Rotello, V.M. Gold nanoparticles: Preparation, properties, and applications in bionanotechnology. Nanoscale 2012, 4, 1871–1880. [Google Scholar] [CrossRef]
  283. Mokammel, M.A.; Islum, M.J.; Muhammad, H. Nanoscale materials for self-cleaning and antibacterial applications. Ref. Modul. Mater. Sci. Mater. Eng. 2019, 315–324. [Google Scholar] [CrossRef]
  284. Bednarski, M.; Dudek, M.; Knutelska, J.; Nowiński, L.; Sapa, J.; Zygmunt, M.; Nowak, G.; Luty-Błocho, M.; Wojnicki, M.; Fitzner, K.; et al. The influence of the route of administration of gold nanoparticles on their tissue distribution and basic biochemical parameters: In vivo studies. Pharmacol. Rep. 2015, 67, 405–409. [Google Scholar] [CrossRef]
  285. 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. Pharm. 2007, 4, 713–722. [Google Scholar] [CrossRef]
  286. Crous, A.; Abrahamse, H. Effective Gold Nanoparticle-Antibody-Mediated Drug Delivery for Photodynamic Therapy of Lung Cancer Stem Cells. Int. J. Mol. Sci. 2020, 21, 3742. [Google Scholar] [CrossRef] [PubMed]
  287. Ravi, R.; Zeyaullah, M.; Ghosh, S.; Khan Warsi, M.; Baweja, R.; AlShahrani, A.M.; Mishra, A.; Ahmad, R. Use of gold nanoparticle-silibinin conjugates: A novel approach against lung cancer cells. Front. Chem. 2022, 10, 1018759. [Google Scholar] [CrossRef]
  288. Foster, A.E.; Carpin, L.B.; Bear, A.S.; Drezek, R.; Lin, Y.A. Modified Gold Nanoparticles for Therapy. US Patent 20140086828A1, 27 March 2014. [Google Scholar]
  289. Ciucă, A.G.; Grecu, C.I.; Rotărescu, P.; Gheorghe, I.; Bolocan, A.; Grumezescu, A.M.; Andronescu, E. Nanostructures for drug delivery: Pharmacokinetic and toxicological aspects. In Nanostructures for Drug Delivery; Elsevier: Amsterdam, The Netherlands, 2017; pp. 941–957. [Google Scholar] [CrossRef]
  290. Wang, M.; Marepally, S.K.; Vemula, P.K.; Xu, C. Inorganic Nanoparticles for Transdermal Drug Delivery and Topical Application. In Nanoscience in Dermatology; Academic Press: Cambridge, MA, USA, 2016; pp. 57–72. [Google Scholar] [CrossRef]
  291. van der Zande, M.; Vandebriel, R.J.; Van Doren, E.; Kramer, E.; Herrera Rivera, Z.; Serrano-Rojero, C.S.; Gremmer, E.R.; Mast, J.; Peters, R.J.; Hollman, P.C.; et al. Distribution, elimination, and toxicity of silver nanoparticles and silver ions in rats after 28-day oral exposure. ACS Nano 2012, 6, 7427–7442. [Google Scholar] [CrossRef] [PubMed]
  292. He, Y.; Du, Z.; Ma, S.; Liu, Y.; Li, D.; Huang, H.; Jiang, S.; Cheng, S.; Wu, W.; Zhang, K.; et al. Effects of green-synthesized silver nanoparticles on lung cancer cells in vitro and grown as xenograft tumors in vivo. Int. J. Nanomed. 2016, 11, 1879–1887. [Google Scholar] [CrossRef] [PubMed]
  293. Saravanakumar, K.; Chelliah, R.; MubarakAli, D.; Oh, D.H.; Kathiresan, K.; Wang, M.H. Unveiling the potentials of biocompatible silver nanoparticles on human lung carcinoma A549 cells and Helicobacter pylori. Sci. Rep. 2019, 9, 5787. [Google Scholar] [CrossRef] [PubMed]
  294. Jagtap, R.R.; Garud, A.; Puranik, S.S.; Rudrapal, M.; Ansari, M.A.; Alomary, M.N.; Alshamrani, M.; Salawi, A.; Almoshari, Y.; Khan, J.; et al. Biofabrication of Silver Nanoparticles (AgNPs) Using Embelin for Effective Therapeutic Management of Lung Cancer. Front. Nutr. 2022, 9, 960674. [Google Scholar] [CrossRef]
  295. Noorbazargan, H.; Amintehrani, S.; Dolatabadi, A.; Mashayekhi, A.; Khayam, N.; Moulavi, P.; Naghizadeh, M.; Mirzaie, A.; Mirzaei Rad, F.; Kavousi, M. Anti-cancer & anti-metastasis properties of bioorganic-capped silver nanoparticles fabricated from Juniperus chinensis extract against lung cancer cells. AMB Express 2021, 11, 61. [Google Scholar] [CrossRef]
  296. Bigall, N.C.; Härtling, T.; Klose, M.; Simon, P.; Eng, L.M.; Eychmüller, A. Monodisperse platinum nanospheres with adjustable diameters from 10 to 100 nm: Synthesis and distinct optical properties. Nano Lett. 2008, 8, 4588–4592. [Google Scholar] [CrossRef] [PubMed]
  297. Ramirez, E.; Eradès, L.; Philippot, K.; Lecante, P.; Chaudret, B. Shape Control of Platinum Nanoparticles. Adv. Funct. Mater. 2007, 17, 2219–2228. [Google Scholar] [CrossRef]
  298. Ensafi, A.A.; Saberi, Z.; Kazemifard, N. Functionalized nanomaterial-based medical sensors for point-of-care applications: An overview. Funct. Nanomater. Based Electrochem. Sens. 2022, 277–308. [Google Scholar] [CrossRef]
  299. Bendale, Y.; Bendale, V.; Natu, R.; Paul, S. Biosynthesized Platinum Nanoparticles Inhibit the Proliferation of Human Lung-Cancer Cells in vitro and Delay the Growth of a Human Lung-Tumor Xenograft in vivo: -In vitro and in vivo Anticancer Activity of bio-Pt NPs. J. Pharmacopunct. 2016, 19, 114–121. [Google Scholar] [CrossRef]
  300. Ismail, N.A.S.; Lee, J.X.; Yusof, F. Platinum Nanoparticles: The Potential Antioxidant in the Human Lung Cancer Cells. Antioxidants 2022, 11, 986. [Google Scholar] [CrossRef] [PubMed]
  301. Zeng, X.; Sun, J.; Li, S.; Shi, J.; Gao, H.; Sun Leong, W.; Wu, Y.; Li, M.; Liu, C.; Li, P.; et al. Blood-triggered generation of platinum nanoparticle functions as an anti-cancer agent. Nat. Commun. 2020, 11, 567. [Google Scholar] [CrossRef] [PubMed]
  302. Ruiz, A.L.; Arribas, E.V.; McEnnis, K. Poly (lactic-co-glycolic acid) encapsulated platinum nanoparticles for cancer treatment. Mater. Adv. 2022, 3, 2858–2870. [Google Scholar] [CrossRef]
  303. Yang, Q.; Peng, J.; Xiao, Y.; Li, W.; Tan, L.; Xu, X.; Qian, Z. Porous Au@Pt Nanoparticles: Therapeutic Platform for Tumor Chemo-Photothermal Co-Therapy and Alleviating Doxorubicin-Induced Oxidative Damage. ACS Appl. Mater. Interfaces 2018, 10, 150–164. [Google Scholar] [CrossRef] [PubMed]
  304. Alavi, M.; Rai, M. Topical delivery of growth factors and metal/metal oxide nanoparticles to infected wounds by polymeric nanoparticles: An overview. Expert Rev. Anti-Infect. Ther. 2020, 18, 1021–1032. [Google Scholar] [CrossRef]
  305. Nikolova, M.P.; Chavali, M.S. Metal Oxide Nanoparticles as Biomedical Materials. Biomimetics 2020, 5, 27. [Google Scholar] [CrossRef]
  306. Pandey, N.; Dhiman, S.; Srivastava, T.; Majumder, S. Transition metal oxide nanoparticles are effective in inhibiting lung cancer cell survival in the hypoxic tumor microenvironment. Chem. Biol. Interact. 2016, 254, 221–230. [Google Scholar] [CrossRef]
  307. Vimercati, L.; Cavone, D.; Caputi, A.; De Maria, L.; Tria, M.; Prato, E.; Ferri, G.M. Nanoparticles: An Experimental Study of Zinc Nanoparticles Toxicity on Marine Crustaceans. General Overview on the Health Implications in Humans. Front. Public Health 2020, 8, 192. [Google Scholar] [CrossRef]
  308. Naseer, M.; Aslam, U.; Khalid, B.; Chen, B. Green route to synthesize Zinc Oxide Nanoparticles using leaf extracts of Cassia fistula and Melia azadarach and their antibacterial potential. Sci. Rep. 2020, 10, 9055. [Google Scholar] [CrossRef]
  309. Siddiqi, K.S.; Ur Rahman, A.; Tajuddin, N.; Husen, A. Properties of Zinc Oxide Nanoparticles and Their Activity Against Microbes. Nanoscale Res. Lett. 2018, 13, 141. [Google Scholar] [CrossRef]
  310. Hu, C.; Du, W. Zinc oxide nanoparticles (ZnO NPs) combined with cisplatin and gemcitabine inhibits tumor activity of NSCLC cells. Aging 2020, 12, 25767–25777. [Google Scholar] [CrossRef] [PubMed]
  311. Tanino, R.; Amano, Y.; Tong, X.; Sun, R.; Tsubata, Y.; Harada, M.; Fujita, Y.; Isobe, T. Anticancer Activity of ZnO Nanoparticles against Human Small-Cell Lung Cancer in an Orthotopic Mouse Model. Mol. Cancer Ther. 2020, 19, 502–512. [Google Scholar] [CrossRef] [PubMed]
  312. Ping, L.; Mengran, X.; Yi, H. Preparation Method and Application of Rapamycin-Entrapped Zinc-Organic Framework Drug Delivery System. Chinese Patent 110693883A, 17 January 2020. [Google Scholar]
  313. Soetaert, F.; Korangath, P.; Serantes, D.; Fiering, S.; Ivkov, R. Cancer therapy with iron oxide nanoparticles: Agents of thermal and immune therapies. Adv. Drug Deliv. Rev. 2020, 163–164, 65–83. [Google Scholar] [CrossRef] [PubMed]
  314. Iacovita, C.; Fizeșan, I.; Pop, A.; Scorus, L.; Dudric, R.; Stiufiuc, G.; Vedeanu, N.; Tetean, R.; Loghin, F.; Stiufiuc, R.; et al. In Vitro Intracellular Hyperthermia of Iron Oxide Magnetic Nanoparticles, Synthesized at High Temperature by a Polyol Process. Pharmaceutics 2020, 12, 424. [Google Scholar] [CrossRef] [PubMed]
  315. Reczyńska, K.; Marszałek, M.; Zarzycki, A.; Reczyński, W.; Kornaus, K.; Pamuła, E.; Chrzanowski, W. Superparamagnetic Iron Oxide Nanoparticles Modified with Silica Layers as Potential Agents for Lung Cancer Treatment. Nanomaterials 2020, 10, 1076. [Google Scholar] [CrossRef]
  316. Theodosiou, M.; Sakellis, E.; Boukos, N.; Kusigerski, V.; Kalska-Szostko, B.; Efthimiadou, E. Iron oxide nanoflowers encapsulated in thermosensitive fluorescent liposomes for hyperthermia treatment of lung adenocarcinoma. Sci. Rep. 2022, 12, 8697. [Google Scholar] [CrossRef]
  317. Xu, V.W.; Nizami, M.Z.I.; Yin, I.X.; Yu, O.Y.; Lung, C.Y.K.; Chu, C.H. Application of Copper Nanoparticles in Dentistry. Nanomaterials 2022, 12, 805. [Google Scholar] [CrossRef]
  318. Al-Hakkani, M.F. Biogenic copper nanoparticles and their applications: A review. SN Appl. Sci. 2020, 2, 505. [Google Scholar] [CrossRef]
  319. Mabrouk, M.; Kenawy, S.H.; El-Bassyouni, G.E.; Ibrahim Soliman, A.A.E.; Aly Hamzawy, E.M. Cancer Cells Treated by Clusters of Copper Oxide Doped Calcium Silicate. Adv. Pharm. Bull. 2019, 9, 102–109. [Google Scholar] [CrossRef]
  320. Kalaiarasi, A.; Sankar, R.; Anusha, C.; Saravanan, K.; Aarthy, K.; Karthic, S.; Mathuram, T.L.; Ravikumar, V. Copper oxide nanoparticles induce anticancer activity in A549 lung cancer cells by inhibition of histone deacetylase. Biotechnol. Lett. 2018, 40, 249–256. [Google Scholar] [CrossRef]
  321. Zhao, X.; Maruthupandy, M.; Al-mekhlafi, F.A.; Chackaravarthi, G.; Ramachandran, G.; KanishaChelliah, C.K. Biological synthesis of copper oxide nanoparticles using marine endophytic actinomycetes and evaluation of biofilm producing bacteria and A549 lung cancer cells. J. King Saud Univ. Sci. 2022, 34, 3. [Google Scholar] [CrossRef]
  322. Gohari, G.; Mohammadi, A.; Akbari, A.; Panahirad, S.; Dadpour, M.R.; Fotopoulos, V.; Kimura, S. Titanium dioxide nanoparticles (TiO2 NPs) promote growth and ameliorate salinity stress effects on essential oil profile and biochemical attributes of Dracocephalummoldavica. Sci. Rep. 2020, 10, 912. [Google Scholar] [CrossRef] [PubMed]
  323. Grande, F.; Tucci, P. Titanium Dioxide Nanoparticles: A Risk for Human Health? Mini Rev. Med. Chem. 2016, 16, 762–769. [Google Scholar] [CrossRef] [PubMed]
  324. Wong, M.S.; Sun, D.S.; Chang, H.H. Bactericidal performance of visible-light responsive titania photocatalyst with silver nanostructures. PLoS ONE 2010, 5, e10394. [Google Scholar] [CrossRef] [PubMed]
  325. Latha, T.S.; Reddy, M.C.R.; Durbaka, P.V.; Muthukonda, S.V.; Lomada, D. Immunomodulatory properties of titanium dioxide nanostructural materials. Indian J. Pharmacol. 2017, 49, 458–464. [Google Scholar] [CrossRef] [PubMed]
  326. Behnam, M.A.; Emami, F.; Sobhani, Z.; Dehghanian, A.R. The application of titanium dioxide (TiO2) nanoparticles in the photo-thermal therapy of melanoma cancer model. Iran. J. Basic Med. Sci. 2018, 21, 1133–1139. [Google Scholar] [CrossRef] [PubMed]
  327. Wang, Y.; Cui, H.; Zhou, J.; Li, F.; Wang, J.; Chen, M.; Liu, Q. Cytotoxicity, DNA damage, and apoptosis induced by titanium dioxide nanoparticles in human non-small cell lung cancer A549 cells. Environ. Sci. Pollut. Res. Int. 2015, 22, 5519–5530. [Google Scholar] [CrossRef]
  328. Zheng, K.; Chen, R.; Sun, Y.; Tan, Z.; Liu, Y.; Cheng, X.; Leng, J.; Guo, Z.; Xu, P. Cantharidin-loaded functional mesoporous titanium peroxide nanoparticles for non-small cell lung cancer targeted chemotherapy combined with high effective photodynamic therapy. Thorac. Cancer 2020, 11, 1476–1486. [Google Scholar] [CrossRef]
  329. Qianhua, F.; Zhenzhong, Z.; Yuzhen, L.; Yutong, H.; Ning, W.; Hongling, Z. Pharmaceutical Composition of Cell Membrane Bionic Titanium Dioxide Nanoparticles. Chinese Patent 109646675B, 9 July 2021. [Google Scholar]
  330. Hopper, E.R.; Wayman, T.M.R.; Asselin, J.; Pinho, B.; Boukouvala, C.; Torrente-Murciano, L.; Ringe, E. Size Control in the Colloidal Synthesis of Plasmonic Magnesium Nanoparticles. J. Phys. Chem. C Nanomater. Interfaces 2022, 126, 563–577. [Google Scholar] [CrossRef]
  331. Krishnamoorthy, K.; Moon, J.Y.; Hyun, H.B.; Cho, S.K.; Kim, S.-J. Mechanistic investigation on the toxicity of MgO nanoparticles toward cancer cells. J. Mater. Chem. 2012, 22, 24610. [Google Scholar] [CrossRef]
  332. Cai, L.; Chen, J.; Liu, Z.; Wang, H.; Yang, H.; Ding, W. Magnesium Oxide Nanoparticles: Effective Agricultural Antibacterial Agent Against Ralstonia Solanacearum. Front. Microbiol. 2018, 9, 790. [Google Scholar] [CrossRef]
  333. Amina, M.; Al Musayeib, N.M.; Alarfaj, N.A.; El-Tohamy, M.F.; Oraby, H.F.; Al Hamoud, G.A.; Bukhari, S.I.; Moubayed, N.M.S. Biogenic green synthesis of MgO nanoparticles using Saussureacostus biomasses for a comprehensive detection of their antimicrobial, cytotoxicity against MCF-7 breast cancer cells and photocatalysis potentials. PLoS ONE 2020, 15, e0237567. [Google Scholar] [CrossRef] [PubMed]
  334. Behzadi, E.; Sarsharzadeh, R.; Nouri, M.; Attar, F.; Akhtari, K.; Shahpasand, K.; Falahati, M. Albumin binding and anticancer effect of magnesium oxide nanoparticles. Int. J. Nanomed. 2018, 14, 257–270. [Google Scholar] [CrossRef] [PubMed]
  335. Liu, L.; Wu, Y.; Ye, J.; Fu, Q.; Su, L.; Wu, Z.; Feng, J.; Chen, Z.; Song, J. Synthesis of magnesium nanoparticle for NIR-II-photoacoustic-imaging-guided synergistic burst-like and H2 cancer therapy. Chem 2022, 8, 2990–3007. [Google Scholar] [CrossRef]
  336. Mohammad, R.S.; Sarvi, M.N. Recent achievements in the microbial synthesis of semiconductor metal sulfide nanoparticles. Mater. Sci. Semicond. Process. 2015, 40, 293–301. [Google Scholar] [CrossRef]
  337. Rao, B.S.; Kumar, B.R.; Reddy, V.R.; Rao, T.S. Preparation and Characterization of CdS Nanoparticles by Chemical Coprecipitation Technique. Chalcogenide Lett. 2011, 8, 177–185. [Google Scholar] [CrossRef]
  338. Ajibade, P.A.; Mbese, J.Z. Synthesis and characterization of metal sulfides nanoparticles/poly (methyl methacrylate) nanocomposites. Int. J. Polym. Sci. 2014, 2014, 752394. [Google Scholar] [CrossRef]
  339. Zhou, M.; Zhang, R.; Huang, M.; Lu, W.; Song, S.; Melancon, M.P.; Tian, M.; Liang, D.; Li, C. A chelator-free multifunctional [64Cu]CuS nanoparticle platform for simultaneous micro-PET/CT imaging and photothermal ablation therapy. J. Am. Chem. Soc. 2010, 132, 15351–15358. [Google Scholar] [CrossRef]
  340. Tian, Q.; Jiang, F.; Zou, R.; Liu, Q.; Chen, Z.; Zhu, M.; Yang, S.; Wang, J.; Wang, J.; Hu, J. Hydrophilic Cu9S5 nanocrystals: A photothermal agent with a 25.7% heat conversion efficiency for photothermal ablation of cancer cells in vivo. ACS Nano 2011, 5, 9761–9771. [Google Scholar] [CrossRef]
  341. Hessel, C.M.; Pattani, V.P.; Rasch, M.; Panthani, M.G.; Koo, B.; Tunnell, J.W.; Korgel, B.A. Copper selenide nanocrystals for photothermal therapy. Nano Lett. 2011, 11, 2560–2566. [Google Scholar] [CrossRef]
  342. Zhou, M.; Tian, M.; Li, C. Copper-Based Nanomaterials for Cancer Imaging and Therapy. Bioconjugate Chem. 2016, 27, 1188–1199. [Google Scholar] [CrossRef]
  343. Fei, W.; Zhang, M.; Fan, X.; Ye, Y.; Zhao, M.; Zheng, C.; Li, Y.; Zheng, X. Engineering of bioactive metal sulfide nanomaterials for cancer therapy. J. Nanobiotechnol. 2021, 19, 93. [Google Scholar] [CrossRef]
  344. Diaz-Diestra, D.; Gholipour, H.M.; Bazian, M.; Thapa, B.; Beltran-Huarac, J. Photodynamic Therapeutic Effect of Nanostructured Metal Sulfide Photosensitizers on Cancer Treatment. Nanoscale Res. Lett. 2022, 17, 33. [Google Scholar] [CrossRef]
  345. Wang, R.; He, Z.; Cai, P.; Zhao, Y.; Gao, L.; Yang, W.; Zhao, Y.; Gao, X.; Gao, F. Surface-Functionalized Modified Copper Sulfide Nanoparticles Enhance Checkpoint Blockade Tumor Immunotherapy by Photothermal Therapy and Antigen Capturing. ACS Appl. Mater. Interfaces 2019, 11, 13964–13972. [Google Scholar] [CrossRef]
  346. Wang, S.; McGuirk, C.M.; d’Aquino, A.; Mason, J.A.; Mirkin, C.A. Metal-Organic Framework Nanoparticles. Adv. Mater. 2018, 30, e1800202. [Google Scholar] [CrossRef]
  347. Lismont, M.; Dreesen, L.; Wuttke, S. Metal-organic framework nanoparticles in photodynamic therapy: Current status and perspectives. Adv. Funct. Mater. 2017, 27, 1606314. [Google Scholar] [CrossRef]
  348. Wang, S.; Park, S.S.; Buru, C.T.; Lin, H.; Chen, P.-C.; Roth, E.W.; Farha, O.K.; Chad, A.M.; Chen, P.-C.; Roth, E.W.; et al. Colloidal crystal engineering with metal–organic framework nanoparticles and DNA. Nat. Commun. 2020, 11, 2495. [Google Scholar] [CrossRef] [PubMed]
  349. Cotta, A.M. Quantum Dots and Their Applications: What Lies Ahead? ACS Appl. Nano Mater. 2020, 3, 4920–4924. [Google Scholar] [CrossRef]
  350. Badıllı, U.; Mollarasouli, F.; Bakirhan, N.K.; Ozkan, Y.; Ozkan, S.A. Role of quantum dots in pharmaceutical and biomedical analysis, and its application in drug delivery. TrAC Trends Anal. Chem. 2020, 131, 116013. [Google Scholar] [CrossRef]
  351. Wei, Z.; Yin, X.; Cai, Y.; Xu, W.; Song, C.; Wang, Y.; Zhang, J.; Kang, A.; Wang, Z.; Han, W. Antitumor effect of a Pt-loaded nanocomposite based on graphene quantum dots combats hypoxia-induced chemoresistance of oral squamous cell carcinoma. Int. J. Nanomed. 2018, 13, 1505–1524. [Google Scholar] [CrossRef]
  352. Kulkarni, N.S.; Parvathaneni, V.; Shukla, S.K.; Barasa, L.; Perron, J.C.; Yoganathan, S.; Muth, A.; Gupta, V. Tyrosine kinase inhibitor conjugated quantum dots for non-small cell lung cancer (NSCLC) treatment. Eur. J. Pharm. Sci. Off. J. Eur. Fed. Pharm. Sci. 2019, 133, 145–159. [Google Scholar] [CrossRef]
  353. Ruzycka-Ayoush, M.; Kowalik, P.; Kowalczyk, A.; Bujak, P.; Nowicka, A.M.; Wojewodzka, M.; Kruszewski, M.; Grudzinski, I.P. Quantum dots as targeted doxorubicin drug delivery nanosystems in human lung cancer cells. Cancer Nano 2021, 12, 9. [Google Scholar] [CrossRef]
  354. Cai, X.; Luo, Y.; Zhang, W.; Du, D.; Lin, Y. pH-Sensitive ZnO Quantum Dots-Doxorubicin Nanoparticles for Lung Cancer Targeted Drug Delivery. ACS Appl. Mater. Interfaces 2016, 8, 22442–22450. [Google Scholar] [CrossRef]
  355. McBain, S.C.; Yiu, H.H.; Dobson, J. Magnetic nanoparticles for gene and drug delivery. Int. J. Nanomed. 2008, 3, 169–180. [Google Scholar] [CrossRef]
  356. Tadic, M.; Kralj, S.; Jagodic, M.; Hanzel, D.; Makovec, D. Magnetic properties of novel superparamagnetic iron oxide nanoclusters and their peculiarity under annealing treatment. Appl. Surf. Sci. 2014, 322, 255–264. [Google Scholar] [CrossRef]
  357. Kralj, S.; Makovec, D. Magnetic Assembly of Superparamagnetic Iron Oxide Nanoparticle Clusters into Nanochains and Nanobundles. ACS Nano 2015, 9, 9700–9707. [Google Scholar] [CrossRef]
  358. Akbarzadeh, A.; Samiei, M.; Joo, S.W.; Anzaby, M.; Hanifehpour, Y.; Nasrabadi, H.T.; Davaran, S. Synthesis, characterization and in vitro studies of doxorubicin-loaded magnetic nanoparticles grafted to smart copolymers on A549 lung cancer cell line. J. Nanobiotechnology 2012, 10, 46. [Google Scholar] [CrossRef]
  359. Carvalho, A.; Fernandes, A.R.; Baptista, P.V. Nanoparticles as delivery systems in cancer therapy. Appl. Target. Nano Drugs Deliv. Syst. 2019, 257–295. [Google Scholar] [CrossRef]
  360. Baeza, A. Ceramic Nanoparticles for Cancer Treatment. Bio-Ceram. Clin. Appl. 2014, 421–455. [Google Scholar] [CrossRef]
  361. Tiwari, A.; Rohiwal, S. Synthesis and Bioconjugation of Hybrid Nanostructures for Biomedical Applications. In Hybrid Nanostructures for Cancer Theranostics; Elsevier: Amsterdam, The Netherlands, 2019. [Google Scholar] [CrossRef]
  362. Singh, D.; Singh, S.; Sahu, J.; Srivastava, S.; Singh, M.R. Ceramic nanoparticles: Recompense, cellular uptake and toxicity concerns. Artif. Cells Nanomed. Biotechnol. 2016, 44, 401–409. [Google Scholar] [CrossRef]
  363. Pednekar, P.P.; Godiyal, S.C.; Jadhav, K.R.; Kadam, V.J. Mesoporous silica nanoparticles: A promising multifunctional drug delivery system. In Nanostructures for Cancer Therapy; Elsevier: Amsterdam, The Netherlands, 2017; pp. 593–621. [Google Scholar] [CrossRef]
  364. Rajani, C.; Borisa, P.; Karanwad, T.; Borade, Y.; Patel, V.; Rajpoot, K.; Tekade, R.K. Cancer-targeted chemotherapy: Emerging role of the folate anchored dendrimer as drug delivery nanocarrier. In Pharmaceutical Applications of Dendrimers; Elsevier: Amsterdam, The Netherlands, 2020; pp. 151–198. [Google Scholar] [CrossRef]
  365. Bharti, C.; Nagaich, U.; Pal, A.K.; Gulati, N. Mesoporous silica nanoparticles in target drug delivery system: A review. Int. J. Pharm. Investig. 2015, 5, 124–133. [Google Scholar] [CrossRef]
  366. Sun, M.; Gu, P.; Yang, Y.; Yu, L.; Jiang, Z.; Li, J.; Le, Y.; Chen, Y.; Ba, Q.; Wang, H. Mesoporous silica nanoparticles inflame tumors to overcome anti-PD-1 resistance through TLR4-NFκB axis. J. Immunother. Cancer 2021, 9, e002508. [Google Scholar] [CrossRef]
  367. Karim, S.; Akhter, M.H.; Burzangi, A.S.; Alkreathy, H.; Alharthy, B.; Kotta, S.; Md, S.; Rashid, M.A.; Afzal, O.; Altamimi, A.S.A.; et al. Phytosterol-Loaded Surface-Tailored Bioactive-Polymer Nanoparticles for Cancer Treatment: Optimization, In Vitro Cell Viability, Antioxidant Activity, and Stability Studies. Gels 2022, 8, 219. [Google Scholar] [CrossRef] [PubMed]
  368. Dilnawaz, F.; Sahoo, S.K. Augmented anticancer efficacy by si-RNA complexed drug loaded mesoporous silica nanoparticles in lung cancer therapy. ACS Appl. Nano Mater. 2018, 1, 730–740. [Google Scholar] [CrossRef]
  369. Song, Y.; Zhou, B.; Du, X.; Wang, Y.; Zhang, J.; Ai, Y.; Xia, Z.; Zhao, G. Folic acid (FA)-conjugated mesoporous silica nanoparticles combined with MRP-1 siRNA improves the suppressive effects of myricetin on non-small cell lung cancer (NSCLC). Biomed. Pharmacother. Biomed. Pharmacother. 2020, 125, 109561. [Google Scholar] [CrossRef]
  370. Jintao, Z.; Liping, L.; Juan, T.; Nam, X.; Kui, W.; Yuce, L.; Chen, G. PH-Responsive Antitumor Drug Carrier Material and Preparation and Application Thereof. Chinese Patent 107412195B, 18 September 2020. [Google Scholar]
  371. Xu, B.; Zhou, W.; Cheng, L.; Zhou, Y.; Fang, A.; Jin, C.; Zeng, J.; Song, X.; Guo, X. Novel Polymeric Hybrid Nanocarrier for Curcumin and Survivin shRNA Co-delivery Augments Tumor Penetration and Promotes Synergistic Tumor Suppression. Front. Chem. 2020, 8, 762. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Modulation of tumor blood vessels to enhance cancer therapy.
Figure 1. Modulation of tumor blood vessels to enhance cancer therapy.
Molecules 29 01076 g001
Figure 2. Vascular remodulation.
Figure 2. Vascular remodulation.
Molecules 29 01076 g002
Figure 3. Tumor progression through stromal regulation.
Figure 3. Tumor progression through stromal regulation.
Molecules 29 01076 g003
Figure 4. Solid lipid NP.
Figure 4. Solid lipid NP.
Molecules 29 01076 g004
Figure 5. Liposome.
Figure 5. Liposome.
Molecules 29 01076 g005
Figure 6. Micelle.
Figure 6. Micelle.
Molecules 29 01076 g006
Figure 7. Lipid nanocapsule.
Figure 7. Lipid nanocapsule.
Molecules 29 01076 g007
Figure 8. PEGylated nanostractured lipid nanocarrier.
Figure 8. PEGylated nanostractured lipid nanocarrier.
Molecules 29 01076 g008
Figure 9. Polymeric NP.
Figure 9. Polymeric NP.
Molecules 29 01076 g009
Figure 10. Dendrimer.
Figure 10. Dendrimer.
Molecules 29 01076 g010
Figure 11. Polystyrene NP.
Figure 11. Polystyrene NP.
Molecules 29 01076 g011
Figure 12. Carbon nanotube.
Figure 12. Carbon nanotube.
Molecules 29 01076 g012
Figure 13. Metal NP.
Figure 13. Metal NP.
Molecules 29 01076 g013
Figure 14. Gold NP.
Figure 14. Gold NP.
Molecules 29 01076 g014
Figure 15. Quantum dots.
Figure 15. Quantum dots.
Molecules 29 01076 g015
Figure 16. Magnetic NP.
Figure 16. Magnetic NP.
Molecules 29 01076 g016
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Sarma, K.; Akther, M.H.; Ahmad, I.; Afzal, O.; Altamimi, A.S.A.; Alossaimi, M.A.; Jaremko, M.; Emwas, A.-H.; Gautam, P. Adjuvant Novel Nanocarrier-Based Targeted Therapy for Lung Cancer. Molecules 2024, 29, 1076. https://doi.org/10.3390/molecules29051076

AMA Style

Sarma K, Akther MH, Ahmad I, Afzal O, Altamimi ASA, Alossaimi MA, Jaremko M, Emwas A-H, Gautam P. Adjuvant Novel Nanocarrier-Based Targeted Therapy for Lung Cancer. Molecules. 2024; 29(5):1076. https://doi.org/10.3390/molecules29051076

Chicago/Turabian Style

Sarma, Kangkan, Md Habban Akther, Irfan Ahmad, Obaid Afzal, Abdulmalik S. A. Altamimi, Manal A. Alossaimi, Mariusz Jaremko, Abdul-Hamid Emwas, and Preety Gautam. 2024. "Adjuvant Novel Nanocarrier-Based Targeted Therapy for Lung Cancer" Molecules 29, no. 5: 1076. https://doi.org/10.3390/molecules29051076

Article Metrics

Back to TopTop