Back to Journals » Drug Design, Development and Therapy » Volume 18

Advances in Nanotechnology for Enhancing the Solubility and Bioavailability of Poorly Soluble Drugs

Authors Liu Y, Liang Y, Yuhong J , Xin P, Han JL , Du Y, Yu X, Zhu R, Zhang M, Chen W , Ma Y

Received 14 November 2023

Accepted for publication 3 April 2024

Published 1 May 2024 Volume 2024:18 Pages 1469—1495

DOI https://doi.org/10.2147/DDDT.S447496

Checked for plagiarism Yes

Review by Single anonymous peer review

Peer reviewer comments 2

Editor who approved publication: Professor Anastasios Lymperopoulos



Yifan Liu,1 Yushan Liang,2 Jing Yuhong,3 Peng Xin,1 Jia Li Han,4 Yongle Du,5 Xinru Yu,3 Runhe Zhu,6 Mingxun Zhang,6 Wen Chen,7 Yingjie Ma7

1School of Medicine, Shandong University of Traditional Chinese Medicine, Jinan, Shandong, People’s Republic of China; 2School of Rehabilitation Medicine, Shandong University of Traditional Chinese Medicine, Jinan, Shandong, People’s Republic of China; 3School of Pharmacy, Shandong University of Traditional Chinese Medicine, Jinan, Shandong, People’s Republic of China; 4School of Health Sciences, Shandong University of Traditional Chinese Medicine, Jinan, Shandong, People’s Republic of China; 5School of Ophthalmology and Optometry, Shandong University of Traditional Chinese Medicine, Jinan, Shandong, People’s Republic of China; 6School of Traditional Chinese Medicine, Shandong University of Traditional Chinese Medicine, Jinan, Shandong, People’s Republic of China; 7First Clinical Medical College, Shandong University of Traditional Chinese Medicine, Jinan, Shandong, People’s Republic of China

Correspondence: Yingjie Ma; Wen Chen, The First Clinical College, Shandong University of Traditional Chinese Medicine; Reproductive and Genetic Center of Integrative Medicine, Affiliated Hospital of Shandong University of Traditional Chinese Medicine, 42, Wenhua Road (W), Lixia District, Jinan, 250014, People’s Republic of China, Email [email protected]; [email protected]

Abstract: This manuscript offers a comprehensive overview of nanotechnology’s impact on the solubility and bioavailability of poorly soluble drugs, with a focus on BCS Class II and IV drugs. We explore various nanoscale drug delivery systems (NDDSs), including lipid-based, polymer-based, nanoemulsions, nanogels, and inorganic carriers. These systems offer improved drug efficacy, targeting, and reduced side effects. Emphasizing the crucial role of nanoparticle size and surface modifications, the review discusses the advancements in NDDSs for enhanced therapeutic outcomes. Challenges such as production cost and safety are acknowledged, yet the potential of NDDSs in transforming drug delivery methods is highlighted. This contribution underscores the importance of nanotechnology in pharmaceutical engineering, suggesting it as a significant advancement for medical applications and patient care.

Keywords: nanotechnology, solubility and bioavailability, poorly soluble drugs, drug delivery systems, pharmaceutical engineering

Introduction

Currently, one of the primary challenges faced by the pharmaceutical industry is the poor water solubility and insufficient bioavailability of drugs.1,2 Current data suggest that approximately 40% of commercially available pharmaceuticals, as well as a significant majority of investigational drugs, struggle with low solubility.3,4 This issue can lead to compromised bioavailability and diminished therapeutic effectiveness, often necessitating increased dosages to achieve the desired medicinal impact.5 The difficulties in dissolving and releasing poorly soluble drugs have restricted the bioavailability of oral solid dosage forms — the most widely used and patient-compliant method of drug delivery — hindering the development and application of many new compound drugs.6,7 Due to low bioavailability, patients need to consume higher drug dosages to achieve the desired therapeutic effects.8 However, increasing the dosage can result in more side effects, potentially harming the physical and mental health of patients and decreasing their medication compliance.9 The issue of low water solubility not only poses significant challenges to medical development but can also lead to various clinical problems, such as variability in patient responses, difficulty in maintaining a safe therapeutic index, increased costs, and the potential risk of toxicity or inefficacy.10 Therefore, effectively addressing the poor solubility and low bioavailability of drugs has always been a focal point and a significant challenge in pharmaceutical and medical research. To overcome these issues, the development of nanomedicine delivery systems, as an innovative drug delivery strategy, has emerged, breaking through the traditional bottlenecks associated with the solubility and bioavailability of drugs.

Nanomedicine delivery systems primarily encompass two fundamental aspects: Firstly, based on pathological changes, these systems can precisely transport drugs to the specified lesion sites, thereby maximizing therapeutic effectiveness and significantly reducing damage to healthy tissues. Secondly, they can control the rate of drug release, ensuring that drug concentration in the blood remains within a safe and effective range, thus mitigating or avoiding toxic and adverse reactions.11 Composed of nanoparticles smaller than 100 nanometers, these delivery systems feature high permeability and potent retention effects, enhanced drug solubility, multifunctionality, controlled drug release mechanisms, and specific targeting capabilities towards diseased cells.12 Through nanotechnology-based structural modifications of drugs, these systems not only achieve precise control over drug release and increase drug stability but also prevent premature degradation of the drug molecules before reaching the lesion, thereby enhancing bioavailability and prolonging circulation time of the drug.11

Definition and Classification of Poorly Soluble Drugs

Research indicates that the absorption and bioavailability of a drug are influenced not only by the properties of the drug itself but also by factors such as dissolution rate, pH value, route of administration, and first-pass effect.13 Particularly, poorly soluble drugs with significant pharmacological activities often fail to achieve their full therapeutic potential due to their slow dissolution rates and low solubility,14,15 leading to issues such as the need for increased drug dosages in clinical settings.5 It is estimated that about 40% of the compounds on the market and the majority of candidate drugs are poorly soluble,3,16 mainly categorized under Class II and IV of the Biopharmaceutical Classification System (BCS).17,18 The BCS guides the setting of dissolution standards for formulations, aiming to reduce the requirements for in vivo bioequivalence (BE).19,20 Moreover, the theoretical framework of BCS enables formulation scientists to develop drug formulations based on the physicochemical and biopharmaceutical properties of the drug, rather than solely relying on experience. Precise formulation design is required for Class II and IV drugs in the BCS, especially in the case of oral administration. Figure 1 by figdraw elaborately illustrates the classification of the BCS and the feasible formulation choices based on this system.

Figure 1 Biopharmaceutics Classification System (BCS) and Feasible Formulation Choices Based on BCS.

Figure 2 The Basic Principles of Nanodrug Delivery Systems. (A) Working Principles of Nanodrug Delivery Systems. (B) Two structures of liposomes:SLN and NLC. (C) The absorption process of nanodrug delivery systems using SLN as carriers to encapsulate TKIs in the gastrointestinal tract. (D) The advantages of NLC as a carrier for Northaritin. (E) Polymer micelles and vesicles. (F) FA-PGA-PTX micelles can selectively enter FR positive cancer cells through receptor mediated endocytosis. (G) Structural basis of hydrophobic and hydrophilic drugs encapsulated in polymer vesicles. (H) Multiphase nanoemulsion. (I) Enhanced permeability and retention effect. (J) Degradation of nano hydrogel under photochemical conditions. (K) Working Principle of Inorganic Nanocarriers. (L) The Structure and Advantages of Dendritic Polymers.

Figure 3 Mechanism of improving the solubility of insoluble drugs through nano drug delivery systems. (A) Transnasal administration using liposome nanodelivery system. (B) The mechanism of receptor mediated transport and adsorption mediated transport breaking through the blood-brain barrier. (C) Surface modification technology to improve drug targeting. (D) Direct and indirect modification methods. (E) Transferrin and Tamoxifen Modified Polymer Dendritic Polymer PAMAM for the Treatment of Brain Glioma. (F) Carrier mediated technology.

Figure 4 Mechanism of Enhanced Cellular Uptake by Nanomedicine Drug Carriers.

Figure 5 pH-Responsive Mechanism.

Figure 6 Altering Lipid Surface Charge in Nano Drug Delivery Systems.

Figure 7 Mechanism of Phagocytosis Avoidance by PEG-Modified Lipid Molecules.

Figure 8 Physiological Barriers During Drug Delivery to the Target.

Figure 9 Basic Mechanisms of Nanocarrier Action.

Figure 10 Types and Mechanisms of Specific Barriers.

Figure 11 Multifunctional Nanoparticles for Drug Delivery.

Table 1 Key Differences Between Emulsion and Nanoemulsion

Table 2 Selected Exemplary Cases in This Field

The Biopharmaceutics Classification System (BCS) categorizes active pharmaceutical ingredients (APIs) by three principal parameters: the dose number (Do), quantifying the API’s administered dose relative to its solubility; the dissolution number (Dn), indicating the rate of API dissolution in the gastrointestinal tract; and the absorption number (An), assessing the extent and rate of API absorption. Firstly, the absorption number is determined by the ratio of the effective permeability (Peff) to the radius of the intestine (R) multiplied by the residence time (Tsi) (Equation 1). Secondly, the dissolution number (Dn) is the ratio of residence time to dissolution time, which involves the solubility (Cs), diffusion rate (D), density, initial particle radius (r), and intestinal transit time (Tsi) (Equation 2). Finally, the dose number (Do) is the ratio of the dose concentration to the drug solubility (Equation 3). The fraction of the drug absorbed in the solution, F, follows an exponential function and can be calculated using Equation 4.

(1)

(2)

(3)

(4)

Note: Cs represents solubility, M is the dose, and V0 denotes the volume of water consumed with the dose, typically set at 250 mL.

Using digoxin and griseofulvin as examples, it is evident that for digoxin, a smaller particle size (indicating a higher Dn)) can substantially improve the likelihood of its complete absorption. In contrast, digoxin with a larger particle size (representing a lower Dn) is likely to face challenges in dissolution, affecting its absorption profile.21,22 Micronization’s effect on griseofulvin’s absorption fraction is relatively modest; thus, adjusting the dose number (Do) to reach the saturation point for complete absorption is crucial for this medication. The dose number (Do) fluctuates based on the dose to solubility concentration ratio.With the drug dose being fixed, the required dissolution volume for a single dose of griseofulvin is exceptionally high,23 thus constraining the dose to a very low level. In this scenario, it becomes imperative to manipulate the only variable – solubility. Should the enhancement in solubility not be significant, griseofulvin might be classified as a drug with solubility limitations.

Class 1BCSII Drugs

Drugs classified under BCS Class II possess molecular characteristics of low solubility and high permeability. Examples of such drugs include morphine, chlorpromazine, and procaine.11 For BCS Class II drugs, the dissolution rate is a primary limiting factor for bioavailability, where minor changes in the dissolution rate can lead to significant increases/decreases in bioavailability.24 Over time, the dissolution rate progressively alters the actual concentration of a drug in solution. Moreover, several factors influence the dissolution rate of a drug, as defined by the Noyes-Whitney equation, including effective surface area, diffusion coefficient, diffusion layer thickness, saturation solubility, the amount of drug dissolved, and the volume of dissolution medium.25

Current strategies for improving the dissolution rates of BCS Class II drugs include reducing particle size,26 self-emulsification,27 pH modification,28 and crystal modification.29 Taking pH modification and self-emulsification as examples, firstly, water solubility represents the lipophilic and hydrophilic properties of a molecule. Enhancing the medium’s lipophilicity or the molecule’s hydrophilicity to improve water solubility is considered a viable strategy. Data collected by varying pH within a range of 1 to 8 is used to establish In Vitro-In Vivo Correlation (IVIVC).30 The use of organic solvents is seen as a less favorable approach due to the inability to achieve correlation and the fact that controlled release components in controlled release formulations are often determined by the solvent. Also, there is a risk of drug precipitation when pH variations are not suitable Moreover, the complexity of clinical application of pH modification is significantly increased by the changes in pH over time, location, and surfactant concentration.

As for self-emulsification, its advantage over non-physiological pH alterations or organic solvents lies in its ability to mimic the in-vivo environment of the intestine. Bile salts and phospholipids in the gastrointestinal tract can improve the wetting31 and solubility32,33 of many lipophilic substances. Additionally, interactions between intestinal fluids, lipids, and food can enhance the formation of emulsions.34 In vitro, various synthetic surfactants can serve as dissolution media,35 like Tween 20 (TW20), Sodium Lauryl Sulfate (SLS), and Dodecyl Trimethyl Ammonium Bromide (DTAB). However, there are some differences between emulsions and nanoemulsions, as outlined in Table 1. Singh et al significantly enhanced the oral bioavailability of primaquine by preparing it as a nanoemulsion, demonstrating its potential for treating early-stage malaria with minimal toxicity.36 Yet, these methods still have limitations. For instance, the formation of salts for neutral compounds is not feasible and can lead to aggregation and other negative effects;37 strategies like reducing particle size are not applicable for extremely fine powders with poor wettability.

BCS Class IV Drugs

Drugs classified under BCS Class IV are characterized by their low solubility and low permeability, with examples including aluminum hydroxide and acetazolamide. The pharmacokinetic patterns of these drugs can be influenced by various gastrointestinal factors such as gastric emptying, the motility phase, bacteria, enzyme activity, and intraluminal viscosity. Intraluminal viscosity becomes particularly crucial when the drug dose fails to dissolve and/or be absorbed during transit.38,39 Drug permeability can also change due to physiological factors. Under the dual constraints of solubility and permeability, considering physiological factors like gastric emptying and gastrointestinal transit times to enhance absorption presents significant challenges in the development and formulation of BCS Class IV drugs. Methods to enhance the permeability of BCS Class IV drugs are still in the early stages of research, and their safety is yet to be established.

Applying formulation methods similar to those used for BCS Class II drugs to BCS Class IV drugs can lead to their dissolution and absorption in the gastrointestinal tract. However, this approach might still be limited due to issues with permeability. Despite their solubilization in the gastrointestinal environment, the inherently poor permeability of BCS Class IV drugs can restrict their effective absorption, presenting a unique set of challenges for their formulation and therapeutic efficacy.

Fundamental Principles of Nanomedicine Drug Delivery Systems

Working Principles of Nanomedicine Drug Delivery Systems

Nanoparticle drug delivery systems utilize nanomaterials as carriers, leveraging their ultra-small size, high-energy catalytic activity of surface atoms, and protective capabilities for the encapsulated drugs.40–42 These systems enable drugs to bypass physiological barriers,43 reduce or avoid immune clearance and the effects of bodily fluids on the drug,44,45 and target drugs for cellular or subcellular-level slow release.46,47 The overarching goals of these strategies are to enhance drug utilization48 and minimize drug toxicity and side effects49 (see Figure 2A). These systems offer a promising avenue in drug delivery by addressing some of the key limitations of conventional drug delivery methods, particularly for drugs that face challenges in solubility, stability, and targeted delivery.

Types and Characteristics of Nanomedicine Drug Delivery Systems

Liposomal Nanocarriers

Liposomes are closed vesicles with a lipid bilayer structure, containing an aqueous core, capable of carrying both lipophilic and hydrophilic substances. Their structure allows for various drug transport mechanisms, such as passive diffusion and membrane fusion. Recent advancements have led to the development of new systems like ethosomal nanogels, which have shown efficacy in the topical drug delivery for skin cancer treatment. These systems are widely utilized to carry hydrophilic drugs due to their non-toxicity, biodegradability, and excellent biocompatibility. These innovative formulations highlight the progressive role of liposomes in nanotechnology-based therapeutics.50,51

Liposomal nanocarriers are primarily categorized into Solid Lipid Nanoparticles (SLNs) and Nanostructured Lipid Carriers (NLCs), as illustrated in Figure 2B. SLNs, with particle sizes ranging from 50 to 1000 nm, are solid colloidal drug delivery systems. They offer benefits such as minimal carrier toxicity, controlled drug release, and targeted delivery. Tyrosine Kinase Inhibitors (TKIs), commonly used oral anticancer agents, are weakly basic and exhibit poor stability and low bioavailability in the gastrointestinal environment. Encapsulating TKIs in SLNs protects the drugs from acidic degradation, aids in traversing physiological barriers in the gastrointestinal tract, and facilitates sustained release. The substantial surface area and adhesive properties of SLNs enhance their absorption in the gastrointestinal tract significantly52,53 (refer to Figure 2C). Further, modifying SLNs, for instance, with a Polyethylene Glycol (PEG) coating, endows them with a hydrophilic layer, minimizes the clearance by intestinal mucosal secretions, and augments the drug’s ability to penetrate mucosal barriers.54,55

NLCs, evolved from SLNs, are an innovative class of lipid nanocarriers composed of a mixture of solid (long-chain) and liquid (short-chain) lipids. The presence of liquid lipids lowers the melting point, allowing the carrier to remain solid at body temperature. This feature prevents the recrystallization of solid lipids during the drug loading process, thereby enhancing drug loading capacity and encapsulation efficiency. Thermodynamically, NLCs exhibit a more robust structure compared to SLNs.56,57 Norcantharidin (NCTD), an anti-cancer drug first discovered in China, faces clinical limitations due to its strong side effects, poor absorption, and weak targeting.58,59 Studies by Yan Z et al60 indicate that employing NLCs as carriers for NCTD improves the drug’s targeting to tumor tissues, extends its release duration, and increases its concentration in tumor sites, thereby exerting a notable inhibitory effect on HepG2 cells, promoting apoptosis in liver cancer cells. Compared to the oral administration of NCTD alone, NLCs significantly enhance the drug’s tumor inhibition rate, improve its utilization, and reduce side effects on normal cells (as shown in Figure 2D). These findings underscore the extensive potential applications of liposomal carriers in cancer therapy.

Furthermore, due to their strong skin permeability, high drug loading capacity, controllable drug release, and biocompatibility, SLNs and NLCs are extensively applied in dermatological drug delivery. The presence of solid lipids considerably reduces the exchange of active substances with the aqueous phase, aiding in the prevention of degradation of unstable chemical compounds. These properties provide novel therapeutic approaches for treating skin disorders such as psoriasis and acne, demonstrating a broad spectrum of applications.61,62

Polymer Nanocarriers

Polymer nanocarriers are versatile platforms in the field of drug delivery. These carriers, constructed from either naturally occurring macromolecules or carefully designed synthetic polymers, are renowned for their straightforward synthetic routes. Beyond simple polymeric nanoparticles, the field has seen the evolution of sophisticated structures such as polymer micelles and polymer vesicles, which are depicted in Figure 2E. These advanced forms are derived from the self-assembly of amphiphilic block copolymers in aqueous solutions, leading to nanostructures with distinct hydrophobic and hydrophilic regions.

Polymer micelles are characterized by their unique architecture with a hydrophobic core that can sequester poorly soluble drugs, surrounded by a hydrophilic shell that imparts stability and stealth features in the biological milieu. This design translates to enhanced drug loading capacities, improved encapsulation efficiencies, and favorable biocompatibility profiles, making them ideal candidates for therapeutic delivery. Their core-shell structure is not only crucial for protecting the active pharmaceutical ingredients from premature degradation but also for facilitating controlled release profiles, which is paramount in achieving sustained therapeutic effects.

Furthermore, the hydrophilic shell of the micelles can effectively mask the drug from the body’s immune surveillance system, particularly the reticuloendothelial system (RES), thereby minimizing opsonization and subsequent phagocytic uptake. This stealth property allows for the micelles to circulate for extended periods, which, in conjunction with the enhanced permeability and retention (EPR) effect, enables them to accumulate preferentially in tumor tissues through passive targeting mechanisms. The EPR effect is especially pronounced in pathological conditions with compromised lymphatic drainage, such as cancer, where it can be exploited to deliver high drug concentrations directly to the site of the tumor, enhancing the efficacy of the treatment while reducing systemic toxicity.63,64 In summary, polymer nanocarriers, particularly polymer micelles, offer a multifunctional and tunable platform for drug delivery applications, combining efficient drug encapsulation with the potential for targeted and controlled release, which are critical attributes in the advancement of nanomedicine. Drug encapsulation within these micelles can be achieved through covalent bonding, direct dissolution, self-assembly with solvent evaporation, dialysis, or emulsion solvent evaporation techniques. Lee et al65 investigated the self-assembled conjugated micelles of folate (FA)-PGA-PTX, demonstrating that, compared to free PTX, FA-PGA-PTX microparticles significantly reduced toxicity to normal cells while notably decreasing cell viability in folate receptor (FR)-positive MCF-7 cancer cells, indicating their excellent targeting ability and controlled drug release property, as seen in Figure 2F.

Polymer vesicles, formed from amphiphilic block polymers, possess a bilayer closed hollow structure, spherical or near-spherical in shape. Their internal layer, formed from hydrophilic block polymers, can encapsulate hydrophilic drugs, while the vesicle wall, composed of hydrophobic block segments, offers permeability and the capability to carry hydrophobic drugs. The outermost layer, also formed by hydrophilic segments, provides high encapsulation efficiency for water-soluble drugs as seen in Figure 2G. Wang et al66 developed amphiphilic polymer vesicles to co-deliver hydrophilic Doxorubicin (DOX) and hydrophobic Taxol (TAX). Compared with systems carrying DOX or TAX alone, these co-delivery systems exhibited a more pronounced inhibitory effect on tumor growth, demonstrating synergistic effects and significant clinical treatment implications. The molecular weight of these polymer vesicles, generally above 1000, ensures their high stability and sustained release characteristics. Chen et al67 polymer vesicles loaded with Ciprofloxacin (CIP) for treating bacterial keratitis (BK) revealed that polymer vesicle carriers could enhance ocular surface adhesion, corneal permeability, and bacterial targeting, thereby improving bioavailability and showing significant potential in treating BK and other bacterial infections.

Nanoemulsions

Nanoemulsions, primarily composed of water, oil, surfactants, and co-surfactants, typically range in size from 10 to 100 nm. They represent a low-viscosity, thermodynamically unstable, isotropic, and either transparent or semi-transparent colloidal dispersion system.68,69 The formation of nanoemulsions, involving the combination of two immiscible liquids via an emulsifier, results in either biphasic (O/W or W/O) or multiphasic systems70 (refer to Figure 2H). Nanoemulsions as drug carriers possess several advantages:①Enhanced Drug Stability: Encapsulation of drugs within the oil phase of nanoemulsions can shield them from hydrolysis and oxidation, thereby augmenting the stability of the pharmaceuticals.71②Increased Solubility and Bioavailability: As drug carriers, nanoemulsions can enhance the solubility and absorption rate of drugs, thereby improving their overall bioavailability. Ding L et al72 discovered that perfluorocarbon nanoemulsions, compared to poly-cationic/small interfering RNA (siRNA) complexes, demonstrated a more robust inhibition of pancreatic tumor growth due to the enhanced permeability and retention (EPR) effect, suggesting a promising future in the controlled therapy of pancreatic cancer.③Improved Patient Compliance: The liquid form of nanoemulsions facilitates multiple administration routes, including oral, intravenous, and topical applications. Niu Z et al73 found that nanoemulsion-based delivery of Coenzyme Q10 showed 1.8 to 2.8 times higher bioavailability when administered orally, compared to other mediums, highlighting the potential of nanoemulsions as oral drug carriers (refer to Figure 2).④Taste Masking: Nanoemulsions can conceal the bitter taste of oils and drugs, making them more palatable when used as oral drug carriers.71⑤Stability: Nanoemulsions are less prone to issues like flocculation, creaming, and sedimentation.74⑥However, nanoemulsions also exhibit certain limitations. For instance, their stability can be significantly affected by external factors such as temperature and pH.75 They are prone to processes like Ostwald ripening, droplet coalescence, and coagulation, which can lead to system destabilization and reduced shelf life.

Nanohydrogels

Nanohydrogels are hydrophilic yet water-insoluble polymers with a three-dimensional cross-linked network structure at the nanoscale. They are characterized by high biocompatibility, high water content, small surface-to-volume ratio, and low cytotoxicity. These properties endow nanohydrogels with numerous advantages such as efficient localized sustained-release, avoidance of macrophage phagocytosis, and enhanced cell recognition, making them suitable for various drug delivery strategies.76,77 Granata G et al78 found that self-assembling injectable nanohydrogels could protect curcumin from chemical or photochemical degradation (as shown in Figure 2J), and maintain a sustained release of the drug. By combining the mechanical properties of hydrogels with the advantages of nanomicelles in drug delivery, these self-assembling injectable nanohydrogels emerge as a novel material with broad application prospects in drug delivery.

Moreover, nanohydrogels can be synergistically combined with nanoparticles through physical encapsulation, electrostatic interactions, and covalent cross-linking. This formation of nano-composite hydrogels not only integrates the inherent characteristics of both components but also overcomes traditional drawbacks of nanohydrogels such as fragility and poor mechanical properties.79 For instance, El-RefaiE et al80 developed a new type of self-assembling hyaluronic acid gel composite with a core of elastic nanovesicles using 1% hyaluronic acid. When used for non-invasive transdermal delivery of hyaluronic acid, in vivo studies demonstrated that this composite gel of elastic nanovesicles enhanced the transdermal penetration into the knee joint sixfold compared to conventional hyaluronic acid gels.

Nanohydrogels, due to their intricate cross-linked nanostructures, present exciting opportunities in the biomedical field. These versatile carriers have been particularly notable for their role in targeted drug delivery and tissue engineering, given their excellent water retention capacity and tunable degradation rates. For instance, their hydrophilic nature allows for responsive drug release that can be triggered by environmental changes, such as pH, temperature, or ionic strength. This responsiveness can be tailored to ensure that the therapeutic agents are released in a controlled manner at the site of action, thus maximizing efficacy while minimizing side effects.Furthering their functionality, recent studies have highlighted the potential of nanohydrogels for dual delivery systems. These systems can simultaneously carry hydrophobic and hydrophilic drugs, releasing them in a synchronized manner, which is particularly advantageous for combination therapies. The application of such dual delivery systems has been studied in cancer therapy, where the coordinated release of chemotherapeutic agents and bioactive molecules has shown to significantly impede tumor growth while promoting healthy cell proliferation.

Another promising aspect of nanohydrogels is their application in regenerative medicine. Their biocompatibility and similarity to the natural extracellular matrix make them ideal scaffolds for cell growth and tissue regeneration. Modified nanohydrogels have been engineered to support the repair and regeneration of various tissues, including cartilage, nerve, and vascular systems. By incorporating growth factors or specific cell-signaling molecules, nanohydrogels can be transformed into bioactive platforms that not only deliver drugs but also actively participate in the healing and regenerative processes.While the body of research on nanohydrogels is vast, continued exploration into their interaction with biological systems, long-term biocompatibility, and the development of scalable manufacturing processes will further cement their place in the future of drug delivery and tissue engineering. By addressing these areas, the next generation of nanohydrogels could offer even more robust and precise delivery of therapeutics, opening new avenues for treatment strategies.

Inorganic Nanocarriers

Inorganic nanocarriers, composed of materials such as metals, metal oxides, and magnetic substances, constitute an inorganic nanoscale drug delivery system. They are advantageous due to their straightforward synthesis, ease of surface modification, high drug loading capacity, small size, large specific surface area, and good biocompatibility. Mesoporous silica, characterized by its interconnected pore structure, can reduce the resistance to drug diffusion, thus facilitating the movement of drugs into the dissolution medium, making it a commonly used inorganic nanocarrier.81 Zhang et al82 demonstrated that mesoporous silica enhances the dissolution rate and bioavailability of the hydrophobic drug Telmisartan (TEL) when orally administered. Compared to the commercial product Micardis, the relative bioavailability of TEL loaded onto MSNs was 154.4%±28.4%. In contrast to crude TEL powder, MSN-loaded TEL significantly improved in dissolution rate. Furthermore, permeability studies on the human colon cancer (Caco-2) cell line indicated that MSNs enhanced the drug’s permeability, reduced drug loss, and improved the absorption of orally administered drugs, as depicted in Figure 2K. This offers a promising new approach for the effective treatment of cancer.

Dendritic Polymer Nanocarriers

Dendritic polymer materials are a new type of artificial nanomaterials characterized by their three-dimensional, highly-ordered structure. They can be molecularly tailored in terms of size, shape, structure, and functional groups. Generally composed of an initiating core, internal repeat units, and terminal functional groups, these materials have evolved through several synthetic approaches, including divergent, convergent, a hybrid of divergent-convergent, and solid-phase synthesis methods. By attaching different functional groups to the peripheries of dendritic polymers, these materials can fulfill specific applications. They are noted for their high drug loading capacity, ease of surface modification, controlled drug release, increased drug solubility, and reduced adverse drug reactions, as illustrated in Figure 2L.83,84 Zhuo et al85 synthesized poly(amidoamine) (PAMAM) dendrimers with a cyclic core. When conjugated with the anticancer drug 5-fluorouracil, the free drug was slowly released in phosphate-buffered saline simulating the human body environment, thereby mitigating adverse drug reactions. Another study86 used dendritic polymers as gene carriers, protecting DNA from degradation and mediating about six times higher gene expression in tumors compared to the PEI transfectant, indicating high transfection efficiency and stability. Thus, dendritic polymers offer a promising, safe, and effective vector for gene therapy, introducing new avenues for drug delivery in gene treatment strategies.

Smart Responsive Nanomedicine Drug Delivery Systems

With the advancement of modern medical technology, smart responsive nano drug delivery systems (NDDS) are being extensively used in the treatment of various diseases such as cancer. These systems can be classified into different types based on their response to external stimuli, including pH, light, temperature, and enzymes. They can control the physicochemical properties of the drug delivery system by manipulating external factors like pH, enabling precise drug targeting.87,88

Cerium dioxide is known for its property of scavenging reactive oxygen species (ROS) and is frequently used to inhibit mitochondrial oxidative stress in the treatment of sepsis-induced acute kidney failure. However, due to the tendency of cerium dioxide nanoparticles to agglomerate and their lack of mitochondria-targeting, Hui Yu et al89 developed an ROS-responsive NDDS, which successfully targeted the mitochondria, effectively reducing oxidative stress and inflammation.

Owing to the variability in pH levels among different human body cells and tissues, Wang et al90 developed a pH-responsive LDP nanopolymer system. This system consistently releases in an environment with pH=7.4, while rapidly releasing in a pH=5.0 environment. Compared to free DOX, it showed enhanced toxicity against CAL-72 cells, indicating its ability to control drug release and prolong the drug’s circulation time. Hence, smart responsive NDDS not only improve drug targeting but also precisely control drug release. They hold promising prospects in treating complex microenvironment diseases like tumors.

Mechanisms of Nanomedicine Drug Delivery Systems for Improving Solubility of Poorly Soluble Drugs

Nanotechnology

Nanotechnology in drug delivery is categorized into two approaches: direct nanonization of the active pharmaceutical ingredient or the drug itself, and encapsulation of the drug within nanocarriers. Utilizing the ultra-small scale of nanoparticles allows drugs to overcome physiological barriers like the blood-brain barrier (BBB) and nasal transport barriers, offering novel administration routes. Nanocarriers, with their large surface area and ease of surface modification, can be attached with numerous functional groups, facilitating enhanced cellular entry and therapeutic efficacy, thus improving drug delivery efficiency.91,92

For instance, nasal drug delivery systems utilizing materials such as liposomes and nanoemulsions, enhanced for mucosal penetration, can improve drug adherence and permeation through the nasal epithelial barrier, significantly enhancing mucosal diffusion and absorption,93 thereby making nasal delivery more efficient as demonstrated in Figure 3A. This innovation provides new pathways and therapeutic methods in clinical medicine.Direct nanonization of drugs typically involves passive diffusion and unsaturated transport primarily through the pores of capillary endothelial cells. In contrast, the application of nanocarriers primarily utilizes receptor-mediated transport and adsorption-mediated transport under electric charge, constituting saturated transport methods to breach the BBB.94 For example, Sun et al95 developed butyl cyanoacrylate nanoparticles coated with Polysorbate-80, leveraging the surfactant’s adsorption of apolipoproteins for tight binding with apolipoprotein receptors on the BBB. Similarly, Lu et al96 found that cationized bovine serum albumin nanoparticles exhibited an eightfold increase in BBB permeability compared to their non-cationized counterparts, as depicted in Figure 3B.

Surface Modification Techniques

Nanocarrier surfaces can be modified through adsorption or covalent attachment of ligands, altering their surface charge, aggregation potential, hydrophilicity, and fluidity. These modifications enhance the targeting of particles to specific cells, increasing cellular uptake and, consequently, drug efficacy, as illustrated in Figure 3C. There are primarily two methods for active targeting modifications: one involves the direct modification by covalently bonding ligands to drug carriers with active functional groups;97 the other uses an indirect modification method where positively charged ligands are inserted into the membrane,98 as shown in Figure 3D.

Li et al99 modified poly(amidoamine) (PAMAM) dendrimers with transferrin and tamoxifen for the treatment of gliomas. The transferrin ligand enables the drug delivery system to cross the blood-brain barrier, while the tamoxifen ligand, targeting modifications, allows the delivery system to further accumulate in the tumor region once inside the brain. This dual-targeting approach not only improves drug efficacy but also reduces toxicity to normal cells, as depicted in Figure 3E. Zhao et al100 designed dendritic poly-L-lysine (DGL) nanoparticles modified with placental-like chondroitin sulfate A-binding peptides (plCSA-BP) to deliver HDZK-BYSB107 and Lingzhi red pigment (DGL/CSA-PNPs). The plCSA-BP specifically binds to chondroitin sulfate A, highly expressed in the trophoblast of the placenta, enabling DGL/CSA-PNPs to accumulate in choriocarcinoma tissues, thereby achieving effective tumor suppression.

Carrier-Mediated Techniques

Carrier-mediated drug delivery refers to the encapsulation of poorly soluble drugs in carriers like liposomes and polymeric micelles, exploiting their high solubility. This methodology facilitates the uptake of these drugs by target cells through passive diffusion, membrane fusion, and endocytosis, thus addressing the issue of drug solubility as illustrated in Figure 3F. Erlotinib (Er), a frontline agent in non-small cell lung cancer (NSCLC) therapy, is a small hydrophobic molecule inhibiting the epidermal growth factor receptor (EGFR). Its clinical efficacy and usability are hindered by low water solubility and permeability. Wang et al101 covalently bonded Er with azido-modified DNA strands, creating a novel anti-cancer drug delivery system. Leveraging the high solubility of the nano-DNA structure, the enhanced permeation and retention (EPR) effect, and robust intracellular uptake, this system significantly improves the bioavailability of Er.

Traditional nano drug delivery systems face challenges such as susceptibility to immune cell uptake and potential toxicity. In contrast, biomimetic nano drug delivery systems, encapsulating drugs within cell membranes, inherit functional properties from the source cells, including immune evasion, extended circulation, and targeted recognition capabilities. These systems are increasingly applied in clinical settings, extensively used for delivering chemotherapeutics, phototherapeutic agents, and cancer vaccines.102 For instance, Wang et al103 noted that nano-formulations wrapped in macrophage-cancer cell hybrid membranes exhibit superior immune evasion and homotypic adhesion, enabling prolonged in vivo retention and targeted accumulation, significantly inhibiting the growth of colorectal cancer cells. Hence, biomimetic nano delivery systems demonstrate extensive application potential in cancer treatment and management, heralding a promising future.

Mechanisms of Enhanced Bioavailability in Nanomedicine Drug Delivery Systems

The formidable efficacy of many drugs contrasts starkly with their low bioavailability upon entering the body, a disparity influenced by cellular physiological absorption barriers and drug stability, thus constraining their application.104 Nano drug delivery systems can enhance drug bioavailability by capitalizing on the material’s favorable properties such as pH-responsiveness, bioadhesion, biocompatibility, biodegradability, modifiability, and processability.

Enhancing Cellular Uptake

Microemulsion drug delivery systems enhance the permeability of mucosal and intestinal epithelial cells through the abundant use of surfactants in Self-Microemulsifying Drug Delivery Systems (SMEDDS), thereby promoting drug absorption.105 In nano drug delivery systems, lipid components can be degraded under the action of pancreatic enzymes and bile, increasing transmembrane transport and absorption of the drug.106 Solid Lipid Nanoparticles (SLNs) leverage their biocompatibility and biodegradability to protect drugs from chemical degradation, thereby enhancing cellular uptake.107,108 Nano drug delivery systems can augment active transport through pathways such as folate, bile acids, and betaine transporters.109 Nano carriers demonstrate Enhanced Permeability and Retention (EPR) effect, passively targeting tumor tissues, enriching in tumor sites, and increasing the likelihood of uptake by tumor cells.110 Furthermore, nano carriers, upon surface ligand modification, can actively target specific cells or tissues through receptor-mediated endocytosis, thus increasing the chances of uptake by target cells,111 as illustrated in Figure 4.

Promoting Intracellular Release

Certain emulsifiers in Self-Microemulsifying Drug Delivery Systems (SMEDDS) can inhibit the efflux action of P-glycoprotein (P-gp) on drugs, thereby enhancing intracellular drug absorption.112 Many delivery systems employ pH-responsive mechanisms to stably encapsulate drugs under suitable pH conditions, avoiding interaction with unfavorable environments. At neutral pH, they can degrade or swell to release the drug,113,114 as depicted in Figure 5. Nano drug delivery systems can be engineered with enzyme-sensitive formulations, such as PEGylated liposomes modified with the cell-penetrating peptide PF, which under the action of MMP-2, shed PEG-2, thereby exposing PF. This transformation results in the reversal of the liposome’s surface charge from negative to positive, facilitating uptake by tumor tissues,115 as shown in Figure 6. Controlled drug release can be achieved by manipulating external energy sources to influence functional nanoparticles. For example, magnetothermal effects generated by alternating magnetic fields control the release of loaded doxorubicin and magnetic Fe3O4 liposomes, enabling magnetothermal-chemotherapy combined treatment for ovarian cancer.116,117

Avoiding Early Metabolism of Drugs

Nano drug delivery systems can exploit the size difference between drugs and digestive enzymes to fabricate porous inorganic nanoparticle carriers, safeguarding the drug from enzymatic degradation.118,119 Additionally, these systems can utilize the hydrophobic interactions between the nanocarriers and digestive enzymes to shield the drug from degradation.120,121 Thermosensitive and biodegradable targeted nanoparticle hydrogels remain in a liquid state at room temperature but transform into a gel at body temperature. This conversion creates a depot system capable of sustained drug release over extended periods, enhancing in vivo retention and significantly improving clinical therapeutic efficacy.122 As depicted in Figure 7, unmodified traditional nano-drugs, upon entering the circulatory system, are predominantly phagocytized by hepatic macrophages, leading to a short half-life and low bioavailability. In contrast, nanoparticles modified with polyethylene glycol (PEG) exhibit increased solubility and markedly enhanced ability to evade macrophage recognition. This significant reduction in phagocytosis by hepatic macrophages prolongs blood circulation time, thereby effectively increasing the drug’s bioavailability.123

Advantages and Challenges of Nanomedicine Drug Delivery Systems

Advantages

The development of nanometer drug delivery systems (NDDS) plays a crucial role in addressing the inherent issues of poorly soluble drugs, such as enhancing solubility and stability while achieving controlled release, prolonging circulation time, and facilitating targeted delivery.124 By intelligently designing drug delivery systems, it is possible to enhance drug efficacy without altering the drug’s chemical structure.125,126 This supports the emergence of various novel administration routes, including local intrabody administration, mucosal absorption delivery,127 and oral administration of peptide drugs.128 Taking mucosal absorption delivery as an example, the combination of PVCL-PVA-PEG with all four types of cellulose polymers (MC, HPMC, sodium CMC, and cationic HEC) can reversibly enhance viscosity at physiological nasal temperatures. Specifically, the thickening effect of PVCL-PVA-PEG combined with sodium CMC is synergistically enhanced. The cellulose derivatives facilitate the transport of drug molecules across the nasal epithelium without causing irreversible changes, simultaneously enhancing the performance characteristics of nasal drug formulations.

Enhancing Drug Stability

The hydrophobic nature of many pharmaceutical compounds significantly impedes their bioavailability. Encapsulation of these drugs within nanoparticles is a strategy to augment their solubility.129,130 For instance, nanocrystals of varying sizes exhibit different affinities and functionalities, with 660 nm crystals favoring follicular accumulation and 250 nm crystals optimizing rapid dissolution. Further, exposure to light, oxygen, moisture, and the enzymatic degradation within the body, as well as adverse physiological conditions, can lead to premature drug decomposition or alteration before reaching the intended target site. This premature degradation severely hampers the therapeutic efficacy of drugs. Therefore, enhancing drug stability constitutes a critical approach to ensure the therapeutic performance of pharmaceuticals. Nanocarriers, characterized by their diminutive size, extensive surface area, and ease of modification, provide a more stable and secluded environment for the drugs. This sequestration significantly mitigates pre-target degradation or inactivation issues, thereby improving the stability of the drug during its delivery.126,131

Prolonging Circulation Time

The journey of drug molecules to their site of action entails traversing through several physiological barriers, including blood, tissue, cellular, and intracellular transport barriers, as illustrated in Figure 8.131 In the context of drug delivery, nanocarriers navigate these obstacles, particularly within the gastrointestinal tract. This process involves a complex interplay of mechanisms such as the endocytosis by epithelial cells, phagocytosis in the m-cell rich regions of Peyer’s Patches, absorption across intestinal interstitial spaces, and paracellular uptake under pathological conditions.132,133

Nanocarriers (NCs) primarily enter cells either passively or actively. Passive penetration occurs through the plasma membrane, while active penetration chiefly involves endocytic pathways, such as pinocytosis or phagocytosis.134 Furthermore, parameters like the shape, size, flexibility, stiffness, and surface functional groups of Nanodrug Delivery Systems (NDDS) significantly influence cellular uptake efficiency.135–137 Due to size constraints and other factors, biopharmaceutical drugs typically cannot permeate cell membranes through osmosis. In contrast, the dimensions of nano drug carriers typically range between 10–1000 nm,138 much smaller than cells and various cellular organelles. Therefore, employing NCs for encapsulation facilitates the traversal across the plasma membrane,139,140 reaching intracellular target sites.

Modifying the surface characteristics of nanoparticles, such as with Polyethylene Glycol (PEG) modifications, can create “stealth” nanoparticles,44,141 impeding the Reticuloendothelial System (RES) from phagocytosing drugs entering the bloodstream.142–144 This modification significantly extends the drug’s half-life in the blood. The primary intracellular barrier during delivery is the endosome/lysosome system. Functionally modified nanocarriers utilize mechanisms like the proton sponge effect to disrupt, destabilize, or fuse with the lysosomal membrane, thereby preventing drug degradation or inactivation in the presence of various enzymes and acidic conditions, as shown in Figure 9.

More fragile or functionally critical organs often develop specific barriers, such as the blood-brain barrier (BBB) safeguarding the central nervous system, the blood-testis barrier, and the placental barrier, as depicted in Figure 10. The BBB, in particular, poses a significant challenge due to its highly selective nature. The tight junctions of its endothelial cells, reinforced by the actions of astrocytes and pericytes, greatly hinder the delivery of pharmaceuticals to the central nervous system.145 However, effective strategies can circumvent the BBB, such as surface modifications designed based on the transference of transferrin (Tf) across the BBB.146,147 Studies indicate that carbon-based nanomaterials (NMs) can cross the BBB through the facilitative actions of the olfactory mucosa and nerves, but this comes with a concomitant increase in potential risks.148

Targeted Delivery and Combination Therapy with Multiple Drugs

Nanodrug delivery systems primarily encapsulate drugs either internally within the carrier or through modifications on the carrier surface. Nanomaterials, endowed with multiple unique characteristics such as surface effects, catalytic properties, and chemical reactivity, can be tailored based on their features and closely linked information about pathological sites. This precision enables the concentrated delivery of drugs at the disease locus, effectively preventing premature drug release, thus enhancing therapeutic efficacy while minimizing toxic side effects on healthy tissues.149 Generally, targeted delivery encompasses both passive and active targeting strategies. Passive targeting relies on the intrinsic properties of the nanodrug delivery system and the physiological or pathological characteristics of the target site to facilitate effective drug accumulation, such as pH-responsive polymer nanoparticles, protease-degradable nanogels, or nanoparticles incorporating disulfide/selenide linkages susceptible to GSH reduction.150 Active targeting, on the other hand, is achieved through the specific interaction and binding of the nanodrug delivery system’s surface with targeted molecules or proteins at the desired site (eg, ligand-receptor interactions, antibody-antigen interactions, lectin-sugar interactions), leading to the selective concentration of the drug at targeted tissues or cells.151 This active targeting is further categorized into biologically directed, in vivo stimulus-directed, and ex vivo stimulus-directed types.

Additionally, the unique physicochemical properties of nanomaterials, particularly their exceptional optical, electrical, thermal, and magnetic characteristics, enable their use as probes for disease diagnostics.152 From another perspective, the high surface energy and surface reactivity of nanomaterials also allow them to function as carriers for controlled drug release, integrating multiple functionalities, as illustrated in Figure 11. Furthermore, the same nanocarrier can be engineered to load multiple drugs, facilitating combination therapy to enhance treatment outcomes.

Challenges

Cost Issues

The cost considerations of nanodrug delivery systems encompass various aspects such as investment in equipment and technology, raw material costs, expenses related to quality control and testing, the impact of economies of scale, and the costs associated with safety evaluations. For instance, in terms of equipment and technology, the formulation of nanoparticles often necessitates the use of organic solvents, ultrasonication, high-speed homogenization, grinding, emulsification, crosslinking, filtration, centrifugation, and lyophilization techniques, all requiring significant financial and technical expertise. Regarding raw materials, the use of metallic nanoparticles, nanomicelles, and similar materials, whose preparation and purification can substantially increase the production costs of nanodrug delivery systems.

When it comes to quality control and testing, due to the small pore size of standard filtration membranes, larger particle size distributions and particles around 220 nm can pose significant filtration challenges. If the average particle size is not significantly lower than 220 nm, a substantial amount of active ingredients might be lost during the filtration process.153 To reduce the costs of nanodrug carriers and ensure their effective and intended use in vivo, innovative approaches and design principles are required.

Safety Concerns

Nanocarriers, despite their efficiency in encapsulating, adsorbing, or covalently linking drug molecules to enhance local drug concentration and reduce dosing frequency, present unique challenges due to their special properties. These include potential adverse reactions in the body, such as immune responses and toxicity. Drugs delivered via nanoentities may behave differently from those administered in normal or conventional forms. For instance, poly(d,l-lactide-co-glycolide) (PLGA) nanoparticles administered intravenously showed a significantly reduced uptake by liver Kupffer cells in rats, compared to free drugs. Moreover, nanoparticles are prone to accumulation in the body, potentially harming healthy tissues.

Utilizing magnetic nanoparticles (NPs) can have harmful consequences. For example, drugs and NPs crossing the blood-brain barrier (BBB) need careful consideration for effective treatment of brain disorders. The medicinal properties of iron oxide NP ions and the abundance of lysophosphatidic acid (LPA) can temporarily disrupt tight junctions, allowing iron oxide NP ions to enter brain cells. This opens up opportunities for treating neurological disorders but also poses a risk of unintended iron accumulation in the brain. Normally, iron oxide NP ions have a half-life of 6 minutes in plasma, primarily depositing in the liver and spleen. In experiments with mice treated with LPA-modified iron oxide NP anions in the brain and spleen, no signs of peripheral immune cell infiltration were found, and microglia and astrocytes showed no significant activation. These findings suggest that temporary disruption of the BBB might be a safe and effective way to enhance the delivery of iron oxide NP anions to the brain. Hence, the application of nanotechnology in drug delivery systems must meet biocompatibility and biosafety requirements to avoid adverse reactions in humans.154 There is an urgent need to develop nanomaterials with good biocompatibility and degradability to mitigate the limitations their use imposes on the industrialization and clinical application of nanodrug delivery systems.155,156

Sterile manufacturing of nanodrugs also presents challenges. For instance, when involving biomaterials, the risk of nanoparticle destruction increases with sterilization techniques like gamma irradiation or autoclaving.157 Environmental safety in nanoparticle manufacturing is another concern. During dry material processing, nanoparticles can disperse into the air, potentially causing lung toxicity.158 Extreme caution is thus required in the manufacturing processes of nanomedical drugs. Personal protective equipment is essential during manufacturing because some nanoparticles can penetrate skin barriers, posing a potential risk of skin contact.159 Nanoparticles manufactured entirely in a liquid environment, much like standard pharmaceutical liquid production, might significantly reduce environmental impacts.

Lack of Standardization and Scalability

The majority of nanomedicines exhibit complex properties, divergent from those produced by conventional pharmaceutical methodologies. Traditional drug manufacturing often fails to replicate the nano-scale three-dimensional, multi-component systems inherent to these medicines. A comprehensive understanding of key components and their interactions is imperative. This knowledge facilitates the early identification of critical characteristics during the development of nanomedicines, enabling the selection of appropriate large-scale production techniques. These techniques must establish crucial process steps and analytical standards to ensure the reproducibility of the final products. Currently, a unified standard for the preparation and evaluation of nanomaterials is lacking. Despite a broad array of nanomedicines being approved, the absence of specific regulatory guidelines for their development and characterization hinders their clinical potential.

Pharmaceuticals are generally regulated by the U.S.160 Food and Drug Administration (FDA) under two primary legislations: the Federal Food, Drug, and Cosmetic Act (FDCA), which encompasses all chemically synthesized drugs and devices, and the Public Health Service Act (PHSA), which covers biologically-derived therapeutic products.161,162 Nanomedicines are categorized by the FDA as combination products, assigned traditional regulatory pathways with additional specific requirements to ensure their safety and efficacy. For instance, nanoscale formulations of paclitaxel and Doxil® have received FDA approval as novel anticancer agents, classified as combination products. The current debate over the adequacy of regulatory frameworks and procedures has led to heightened scrutiny of the intrinsic risks associated with nanotechnology and products containing nanoparticles. These concerns include the toxicity of nanoparticles, the potential unintended effects due to their capability to penetrate the Blood-Brain Barrier (BBB), and the long-term implications of nanoparticle use.163,164

Furthermore, the production of nanomaterials demands more precise technologies to ensure the stability and consistency of the drug delivery systems. The complexity of the preparation process is a major barrier to the scalable manufacture and commercial application of novel drug delivery systems. Designing more refined industrial equipment and optimizing the production process are crucial to addressing these current challenges.

Latest Research Developments and Case Studies

Advancements in nanotechnology have significantly contributed to the research of drug delivery systems, offering valuable insights.165 This includes the development of liposomes, nanoparticles, polymeric micelles, and polymer-drug conjugates.166 To date, nanocarrier systems have been successfully applied in delivering poorly soluble drugs,167 with 51 nanotechnology-based products implemented in clinical practices.168–170

Compared to conventional drug delivery methods, nanomedicine primarily emphasizes targeted drug delivery, supporting effective treatments for various infectious diseases, cancers, diabetes, and neurodegenerative disorders such as Alzheimer’s disease.171,172 Several successful nanocarriers, including dendrimers, nanoparticles, liposomes, and carbon nanotubes, have been commercialized. These carriers achieve precise targeting and rapid accumulation at the site of action without reliance on any biological system intervention.173–175 However, challenges remain in optimizing delivery module efficiency and drug properties, such as biocompatibility and biodegradability.176 Numerous ongoing clinical trials are expected to propel the advancement of novel nanotherapeutics and highlight future directions for improvement in this field.174,175 This article selects typical examples from this area for analysis, as illustrated in Table 2.

Conclusion and Outlook

With the advancement of nanomedicine, nanoscale drug delivery systems (NDDS) have emerged as pivotal tools for enhancing the solubility and bioavailability of poorly soluble drugs, thereby advancing solutions to critical issues in pharmaceutical engineering related to drug solubility and bioavailability. This review comprehensively summarizes the research on NDDS for improving the solubility and bioavailability of insoluble drugs. Initially, we revisit the definition, classification, characteristics of poorly soluble drugs, and the fundamental principles, types, and features of NDDS. Poorly soluble drugs are categorized into BCS Class II and IV,47 both exhibiting low solubility but differing in permeability - BCS Class II drugs have higher permeability than BCS Class IV,177 leading to distinct improvement strategies for each class.

NDDS utilize the ultrafine size, catalytic activity, and protective nature of nanomaterials41,178,179 for drug delivery, thereby enhancing drug utilization and reducing adverse drug reactions.48 The systems can be classified into six types:49 lipid-based nanoparticles (LNPs), polymer-based nanoparticles, nanoemulsions, nanogels, inorganic nanoparticles, and dendrimer-based nanoparticles. LNPs have been shown to induce apoptosis in liver cancer cells, enhancing drug efficacy with promising research implications in cancer treatment.60 Polymer-based nanoparticles exhibit excellent bacterial targeting, enhancing ocular adhesion and corneal permeability,69 suggesting greater potential in treating bacterial keratitis and other bacterial infections. Nanoemulsions, as oral drug carriers, have demonstrated significantly higher bioavailability than other mediums,81 indicating their potential as superior oral drug carriers. Nanogels offer localized, sustained release of drugs.86 Inorganic nanoparticles, catalyzing the formation of highly cytotoxic substances like BCS Class IV, synergize with chemotherapy drugs like Doxorubicin (DOX) to kill cancer cells,180 offering new avenues for efficient cancer treatment. Dendrimer-based nanoparticles, post combination with anticancer drug 5-fluorouracil, gradually release free 5-fluorouracil,95 reducing adverse drug reactions and presenting broad application prospects as novel nanomaterials.

Addressing the characteristic challenges of poorly soluble drugs, our analysis of nanonization techniques, surface modification technologies, and carrier-mediated tactics concludes that NDDS can precisely enhance the solubility of these drugs. Nanonization techniques break through biological barriers like the blood-brain barrier and nasal transport barriers, thereby improving drug delivery efficiency101,102 and introducing new administration routes. Surface modification, altering the physicochemical properties of nanoparticle surfaces, enhances cellular targeting and consequently drug utilization.110 Carrier-mediated technology involves encapsulating poorly soluble substances within high-solubility carriers, leading to cellular uptake via passive diffusion, membrane fusion, and endocytosis.117 Additionally, NDDS enhance cellular uptake, promote intracellular release, and prevent premature drug metabolism, thus improving bioavailability.

Despite challenges in production costs and safety,119–121 nanomaterials possess unique advantageous properties not shared by the majority of other drug delivery materials,124 promising to overcome the limitations of traditional delivery methods and materials. Notably, NDDS have recently garnered increasing research attention in the medical field.

Traditional drug delivery methods often require frequent or high dosage administration to achieve therapeutic effects, leading to decreased overall efficacy and patient compliance. In contrast, the application of nanotechnology-based drug delivery systems is increasingly favored and researched. Studies suggest that controlling the particle size and modifying components within nanoscale drug delivery systems can increase drug solubility, control drug release, prolong circulation time, reduce drug clearance rates, selectively enhance cellular uptake, and minimize adverse reactions, thereby enhancing therapeutic outcomes.181

In terms of drug particle size, the efficiency of drug delivery can be enhanced by controlling the size of the nanoparticles. For instance, Mistry et al182 demonstrated the significance of reducing particle size in drug delivery systems by comparing the intranasal delivery efficiency of nanoparticles measuring 100 nm and 200 nm. Controlling the diameter of carbon nanoparticles (NPs) to 150 nm, for example, allows them to enter the lymphatic system without entering the bloodstream, directing them to regional lymph nodes.183 This indicates that controlling nanoparticle size can determine drug transport pathways, enhancing target specificity.

In terms of surface modification of drugs, firstly, encapsulating drugs with materials such as polysaccharides, solid lipid bodies, or magnetic substances can develop modified coating materials that prolong circulation time, increase absorption rate, reduce drug side effects, and enhance bioavailability.183 Secondly, adsorption or covalent attachment of ligands can be used for surface modification,110 altering physical and chemical properties like surface charge, aggregation potential, hydrophilicity, and fluidity. This effectively enhances the targeting of the nanoparticles to cells, improving drug absorption and efficient delivery of the loaded drug. Thirdly, modifiable polymers such as polyethylene glycol (PEG) can be used for drug modification, enhancing the mucosal penetration capability of the nanoparticles while ensuring high safety, thereby increasing drug absorption efficiency.

Therefore, leveraging the unique attributes of nanotechnology to devise novel inventions that enhance the utilization of poorly soluble drugs will be a reliable cornerstone in the field of nanomedicine. Overall, despite some challenges, nanoscale drug delivery systems still offer promising strategies and insights for addressing issues of drug insolubility and low bioavailability.

Abbreviations

BCS, The Biopharmaceutics Classification System; IVIVC, In Vivo Correlation; SLS, Sodium Lauryl Sulfate; DTAB, Dodecyl Trimethyl Ammonium Bromide; PEG, Polyethylene Glycol; NCTD, Norcantharidin; DOX, Doxorubicin; TAX, Taxol; CIP, Ciprofloxacin; TEL, Telmisartan; SMEDDS, Microemulsifying Drug Delivery Systems; EPR, Retention; NDDS, Nanodrug Delivery Systems; RES, Reticuloendothelial System; FDA, Drug Administration; FDCA, Cosmetic Act; PHSA, Public Health Service Act; BBB, Brain Barrier.

Author Contributions

All authors made a significant contribution to the work reported, whether that is in the conception, study design, execution, acquisition of data, analysis and interpretation, or in all these areas; took part in drafting, revising or critically reviewing the article; gave final approval of the version to be published; have agreed on the journal to which the article has been submitted; and agree to be accountable for all aspects of the work.

Disclosure

The authors declare no conflicts of interest in this work.

References

1. Prabhu P, Patravale V. Dissolution enhancement of atorvastatin calcium by co-grinding technique. Drug Delivery Transl Res. 2016;6(4):380–391. doi:10.1007/s13346-015-0271-x

2. Khalid Q, Ahmad M, Minhas MU, et al. Synthesis of β-cyclodextrin hydrogel nanoparticles for improving the solubility of dexibuprofen: characterization and toxicity evaluation. Drug Dev Ind Pharm 2017;43(11):1873–1884. doi:10.1080/03639045.2017.1350703

3. Jermain SV, Brough C, Williams RO, et al. Amorphous solid dispersions and nanocrystal technologies for poorly water-soluble drug delivery - An update. Int J Pharm. 2018;535(1–2):379–392. doi:10.1016/j.ijpharm.2017.10.051

4. Liu Y, Wang T, Ding W, et al. Dissolution and oral bioavailability enhancement of praziquantel by solid dispersions. Drug Delivery Transl Res. 2018;8(3):580–590. doi:10.1007/s13346-018-0487-7

5. Baghel S, Cathcart H, O’Reilly NJ, et al. Polymeric Amorphous Solid Dispersions: a Review of Amorphization, Crystallization, Stabilization, Solid-State Characterization, and Aqueous Solubilization of Biopharmaceutical Classification System Class II Drugs. J Pharmaceut Sci. 2016;105(9):2527–2544. doi:10.1016/j.xphs.2015.10.008

6. Hua S. Advances in Oral Drug Delivery for Regional Targeting in the Gastrointestinal Tract - Influence of Physiological, Pathophysiological and Pharmaceutical Factors. Front Pharmacol. 2020;11(524). doi:10.3389/fphar.2020.00524

7. Shreya AB, Raut SY, Managuli RS, et al. Active Targeting of Drugs and Bioactive Molecules via Oral Administration by Ligand-Conjugated Lipidic Nanocarriers: recent Advances. AAPS Pharm Sci Tech. 2018;20(1):15. doi:10.1208/s12249-018-1262-2

8. Siddiqui K, Waris A, Akber H, et al. Physicochemical Modifications and Nano Particulate Strategies for Improved Bioavailability of Poorly Water Soluble Drugs. Pharm nanotechnol. 2017;5(4):276–284. doi:10.2174/2211738506666171226120748

9. Rubin KM, Vona K, Madden K, et al. Side effects in melanoma patients receiving adjuvant interferon alfa-2b therapy: a nurse’s perspective. Supportive Care Cancer. 2012;20(8):1601–1611. doi:10.1007/s00520-012-1473-0

10. Da Silva FL, Marques MB, Kato KC, et al. Nanonization techniques to overcome poor water-solubility with drugs. Expert Opin Drug Discov. 2020;15(7):853–864. doi:10.1080/17460441.2020.1750591

11. Prakash S. Nano-based drug delivery system for therapeutics: a comprehensive review. Biomed Phys Eng Express. 2023;9(5):10.1088/2057–1976/acedb2. doi:10.1088/2057-1976/acedb2

12. Harder BG, et al. Developments in Blood-Brain Barrier Penetrance and Drug Repurposing for Improved Treatment of Glioblastoma. Front Oncol. 2018;8(462). doi:10.3389/fonc.2018.00462

13. Zhai X, Lademann J, Keck CM, et al. Nanocrystals of medium soluble actives--novel concept for improved dermal delivery and production strategy. Int J Pharm. 2014;470(1–2):141–150. doi:10.1016/j.ijpharm.2014.04.060

14. Kataoka M, Yonehara A, Minami K, et al. Control of Dissolution and Supersaturation/Precipitation of Poorly Water-Soluble Drugs from Cocrystals Based on Solubility Products: a Case Study with a Ketoconazole Cocrystal. Mol Pharmaceut. 2023;20(8):4100–4107. doi:10.1021/acs.molpharmaceut.3c00237

15. Haering B, Seyferth S, Schiffter HA, et al. The tangential flow absorption model (TFAM) - A novel dissolution method for evaluating the performance of amorphous solid dispersions of poorly water-soluble actives. Eur J Pharm Biopharm 2020;154:74–88. doi:10.1016/j.ejpb.2020.06.013

16. Yang M, Chen T, Wang L, et al. High dispersed phyto-phospholipid complex/TPGS 1000 with mesoporous silica to enhance oral bioavailability of tanshinol. Colloids Surf B. 2018;170:187–193. doi:10.1016/j.colsurfb.2018.06.013

17. Kawabata Y, Wada K, Nakatani M, et al. Formulation design for poorly water-soluble drugs based on biopharmaceutics classification system: basic approaches and practical applications. Int J Pharm. 2011;420(1):1–10. doi:10.1016/j.ijpharm.2011.08.032

18. Takagi T, Ramachandran C, Bermejo M, et al. A provisional biopharmaceutical classification of the top 200 oral drug products in the United States, Great Britain, Spain, and Japan. Mol Pharmaceut. 2006;3(6):631–643. doi:10.1021/mp0600182

19. Amidon GL, Lennernäs H, Shah VP, et al. A theoretical basis for a biopharmaceutic drug classification: the correlation of in vitro drug product dissolution and in vivo bioavailability. Pharm Res. 1995;12(3):413–420. doi:10.1023/a:1016212804288

20. Bhatt S, ROY D, KUMAR M, et al. Development and Validation of In Vitro Discriminatory Dissolution Testing Method for Fast Dispersible Tablets of BCS Class II Drug. Turkish j Pharm Sci. 2020;17(1):74–80. doi:10.4274/tjps.galenos.2018.90582

21. Fraser EJ, Leach RH, Poston JW, et al. Dissolution and bioavailability of digoxin tablets. J Pharm Pharmacol. 1973;25(12):968–973. doi:10.1111/j.2042-7158.1973.tb09988.x

22. Johnson BF, O’Grady J, Bye C, et al. The influence of digoxin particle size on absorption of digoxin and the effect of propantheline and metoclopramide. Br J Clin Pharmacol 1978;5(5):465–467. doi:10.1111/j.1365-2125.1978.tb01657.x

23. Chi-Yuan W, Benet LZ. Predicting drug disposition via application of BCS: transport/absorption/ elimination interplay and development of a biopharmaceutics drug disposition classification system. Pharm Res. 2005;22(1):11–23. doi:10.1007/s11095-004-9004-4

24. Löbenberg R, Amidon GL. Modern bioavailability, bioequivalence and biopharmaceutics classification system. New scientific approaches to international regulatory standards. Eur J Pharm Biopharm 2000;50(1):3–12. doi:10.1016/s0939-6411(00)00091-6

25. Hattori Y, Haruna Y, Otsuka M, et al. Dissolution process analysis using model-free Noyes-Whitney integral equation. Colloids Surf B. 2013;102:227–231. doi:10.1016/j.colsurfb.2012.08.017

26. Kesisoglou F, Mitra A. Crystalline nanosuspensions as potential toxicology and clinical oral formulations for BCS II/IV compounds. AAPS J. 2012;14(4):677–687. doi:10.1208/s12248-012-9383-0

27. Yang SG. Biowaiver extension potential and IVIVC for BCS ClassII drugs by formulation design:Case study for cyclosporine self-microemulsifying formulation. Arch Pharm Res. 2010;33(11):1835–1842. doi:10.1007/s12272-010-1116-2

28. Adachi M, Hinatsu Y, Kusamori K, et al. Improved dissolution and absorption of ketoconazole in the presence of organic acids as pH-modifiers. Eur j Pharm Sci. 2015;76:225–230. doi:10.1016/j.ejps.2015.05.015

29. Chi L, Wu D, Li Z, et al. Modified Release and Improved Stability of Unstable BCSII Drug by Using Cyclodextrin Complex as Carrier To Remotely Load Drug into Niosomes. Mol Pharm. 2016;13(1):113–124. doi:10.1021/acs.molpharmaceut.5b00566

30. Verma H, Garg R. Development of a bio-relevant pH gradient dissolution method for a high-dose, weakly acidic drug, its optimization and IVIVC in Wistar rats: a case study of magnesium orotate dihydrate. Magnesium Res. 2022;35(3):88–95. doi:10.1684/mrh.2022.0505

31. Efentakis M, Dressman JB. Gastric juice as a dissolution medium: surface tension and pH. Eur j Drug Metab Pharm. 1998;23(2):97–102. doi:10.1007/BF03189322

32. Mithani SD, Bakatselou V, TenHoor CN, et al. Estimation of the increase in solubility of drugs as a function of bile salt concentration. Pharm Res. 1996;13(1):163–167. doi:10.1023/a:1016062224568

33. Bakatselou V, Oppenheim RC, Dressman JB, et al. Solubilization and wetting effects of bile salts on the dissolution of steroids. Pharm Res. 1991;8(12):1461–1469. doi:10.1023/a:1015877929381

34. Charman WN, Porter CJH, Mithani S, et al. Physiochemical and physiological mechanisms for the effects of food on drug absorption: the role of lipids and pH. J Pharmaceut Sci. 1997;86(3):269–282. doi:10.1021/js960085v

35. La Sorella G, Sperni L, Canton P, et al. Selective Hydrogenations and Dechlorinations in Water Mediated by Anionic Surfactant-Stabilized Pd Nanoparticles. J Org Chem. 2018;83(14):7438–7446. doi:10.1021/acs.joc.8b00314

36. Singh KK, Vingkar SK. Formulation, antimalarial activity and biodistribution of oral lipid nanoemulsion of primaquine.Int. J Pharm. 2008;347(1–2):567.

37. Kalepu S, Nekkanti V. Insoluble drug delivery strategies: review of recent advances and business prospects. Acta Pharma Sin. 2015;5(5):442–453. doi:10.1016/j.apsb.2015.07.003

38. Hirose R, Sugano K. Effect of Food Viscosity on Drug Dissolution. Pharm Res. 2024;41(1):105–112. doi:10.1007/s11095-023-03620-y

39. Amidon GL, Leesman GD, Elliott RL, et al. Improving intestinal absorption of water-insoluble compounds: a membrane metabolism strategy. J Pharmaceut Sci. 1980;69(12):1363–1368. doi:10.1002/jps.2600691203

40. Mistry A, Stolnik S, Illum L, et al. Nanoparticles for direct nose-to-brain delivery of drugs. Int J Pharm. 2009;379(1):146–157. doi:10.1016/j.ijpharm.2009.06.019

41. Yuan S, Zhang Q. Application of One-Dimensional Nanomaterials in Catalysis at the Single-Molecule and Single-Particle Scale. Front Chem. 2021;9(812287). doi:10.3389/fchem.2021.812287

42. Hegewald AB, Breitwieser K, Ottinger SM, et al. Extracellular miR-574-5p Induces Osteoclast Differentiation via TLR 7/8 in Rheumatoid Arthritis. Front Immunol. 2020;11:585282. doi:10.3389/fimmu.2020.585282

43. Han L, Jiang C. Evolution of blood-brain barrier in brain diseases and related systemic nanoscale brain-targeting drug delivery strategies. Acta Pharma Sin. 2021;11(8):2306–2325. doi:10.1016/j.apsb.2020.11.023

44. Parodi A, Quattrocchi N, van de Ven AL, et al. Synthetic nanoparticles functionalized with biomimetic leukocyte membranes possess cell-like functions. Nat Nanotechnol. 2013;8(1):61–68. doi:10.1038/nnano.2012.212

45. Ugwoke MI, AGU R, VERBEKE N, et al. Nasal mucoadhesive drug delivery: background, applications, trends and future perspectives. Adv Drug Delivery Rev 2005;57(11):1640–1665. doi:10.1016/j.addr.2005.07.009

46. Kanazawa T, Taki H, Tanaka K, et al. Cell-penetrating peptide-modified block copolymer micelles promote direct brain delivery via intranasal administration. Pharm Res. 2011;28(9):2130–2139. doi:10.1007/s11095-011-0440-7

47. Rafiei P, Haddadi A. Pharmacokinetic Consequences of PLGA Nanoparticles in Docetaxel Drug Delivery. Pharm nanotechnol. 2017;5(1):3–23. doi:10.2174/2211738505666161230110108

48. Sridhar V, Gaud R, Bajaj A, et al. Pharmacokinetics and pharmacodynamics of intranasally administered selegiline nanoparticles with improved brain delivery in Parkinson’s disease. Nanomedicine. 2018;14(8):2609–2618. doi:10.1016/j.nano.2018.08.004

49. Guo Q, Chang Z, Khan NU, et al. Nanosizing Noncrystalline and Porous Silica Material-Naturally Occurring Opal Shale for Systemic Tumor Targeting Drug Delivery. ACS Appl Mater Interfaces. 2018;10(31):25994–26004. doi:10.1021/acsami.8b06275

50. Aslam M, Javed MN, Deeb HH, et al. Lipid Nanocarriers for Neurotherapeutics: introduction, Challenges, Blood-brain Barrier, and Promises of Delivery Approaches. CNS Neurol Disord Drug Targets. 2022;21(10):952–965. doi:10.2174/1871527320666210706104240

51. Large DE, Abdelmessih RG, Fink EA, et al. Liposome composition in drug delivery design, synthesis, characterization, and clinical application. Adv Drug Delivery Rev 2021;176:113851. doi:10.1016/j.addr.2021.113851

52. Harde H, Das M, Jain S, et al. Solid lipid nanoparticles: an oral bioavailability enhancer vehicle. Expert Opin Drug Delivery. 2011;8(11):1407–1424. doi:10.1517/17425247.2011.604311

53. Damgé C, Michel C, Aprahamian M, et al. New approach for oral administration of insulin with polyalkylcyanoacrylate nanocapsules as drug carrier. Diabetes. 1988;37(2):246–251. doi:10.2337/diab.37.2.246

54. Ponchel G, Montisci M-J, Dembri A, Durrer C, Duchêne D. Assia Dembri, Carlo Durrer, Dominique Duchêne,Mucoadhesion of colloidal particulate systems in the gastro-intestinal tract,European. J Pharm Biopharmaceutics. 1997;44(1):25–31. doi:10.1016/S0939-6411(97)00098-2

55. Suk JS, Xu Q, Kim N, et al. PEGylation as a strategy for improving nanoparticle-based drug and gene delivery. Adv Drug Delivery Rev 2016;99:28–51. doi:10.1016/j.addr.2015.09.012

56. Souto EB, Baldim I, Oliveira WP, et al. SLN and NLC for topical, dermal, and transdermal drug delivery. Expert Opin Drug Delivery. 2020;17(3):357–377. doi:10.1080/17425247.2020.1727883

57. Müller RH, Radtke M, Wissing SA, et al. Solid lipid nanoparticles (SLN) and nanostructured lipid carriers (NLC) in cosmetic and dermatological preparations. Adv Drug Delivery Rev 2002;54(1):S131–55. doi:10.1016/s0169-409x(02)00118-7

58. Xiu-Yan L, Guan Q-X, Shang Y-Z, et al. Metal-organic framework IRMOFs coated with a temperature-sensitive gel delivering norcantharidin to treat liver. World j Gastroenterol. 2021;27(26):4208–4220. doi:10.3748/wjg.v27.i26.4208

59. Pan M-S, Cao J, Fan Y-Z, et al. Insight into norcantharidin, a small-molecule synthetic compound with potential multi-target anticancer activities. ChinMed. 2020;15:55. doi:10.1186/s13020-020-00338-6

60. Yan Z, Yang K, Tang X, et al. Norcantharidin Nanostructured Lipid Carrier (NCTD-NLC) Suppresses the Viability of Human Hepatocellular Carcinoma HepG2 Cells and Accelerates the Apoptosis. J Immunol Res. 2022;2022:3851604. doi:10.1155/2022/3851604

61. Akhter MH, Ahmad A, Ali J, et al. Formulation and Development of CoQ10-Loaded s-SNEDDS for Enhancement of Oral Bioavailability. J Pharm Innov. 2014;9:121–131. doi:10.1007/s12247-014-9179-0

62. Soni K, Mujtaba A, Akhter MH, et al. Optimisation of ethosomal nanogel for topical nano-CUR and sulphoraphane delivery in effective skin cancer therapy. J Microencapsulation. 2020;37(2):91–108. doi:10.1080/02652048.2019.1701114

63. Katari O, Jain S. Solid lipid nanoparticles and nanostructured lipid carrier-based nanotherapeutics for the treatment of psoriasis. Expert Opin Drug Delivery. 2021;18(12):1857–1872. doi:10.1080/17425247.2021.2011857

64. Carmen Gómez-Guillén M, Montero MP. Pilar Montero,Enhancement of oral bioavailability of natural compounds and probiotics by mucoadhesive tailored biopolymer-based nanoparticles: a review. Food Hydrocoll. 2021;118:106772. doi:10.1016/j.foodhyd.2021.106772

65. Denora N, Trapani A, Laquintana V, et al. Recent advances in medicinal chemistry and pharmaceutical technology--strategies for drug delivery to the brain. Curr Top Med Chem 2009;9(2):182–196. doi:10.2174/156802609787521571

66. Mikhail AS, Allen C. Block copolymer micelles for delivery of cancer therapy: transport at the whole body, tissue and cellular levels. J Control Release. 2009;138(3):214–223. doi:10.1016/j.jconrel.2009.04.010

67. Lee Y. Preparation and characterization of folic acid linked poly(L-glutamate) nanoparticles for cancer targeting. Macromol Res. 2006;14:387–393. doi:10.1007/BF03219099

68. Wang H, Zhao Y, Wu Y, et al. Enhanced anti-tumor efficacy by co-delivery of doxorubicin and paclitaxel with amphiphilic methoxy PEG-PLGA copolymer nanoparticles. Biomaterials. 2011;32(32):8281–8290. doi:10.1016/j.biomaterials.2011.07.032

69. Chen Q, Han X, Liu L, et al. Multifunctional Polymer Vesicles for Synergistic Antibiotic-Antioxidant Treatment of Bacterial Keratitis. Biomacromolecules. 2023;24(11):5230–5244. doi:10.1021/acs.biomac.3c00754

70. Md S, Alhakamy NA, Neamatallah T, et al. Development, Characterization, and Evaluation of α-Mangostin-Loaded Polymeric Nanoparticle Gel for Topical Therapy in Skin Cancer. Gels. 2021;7(4):230. doi:10.3390/gels7040230

71. Shumin L, Xu Z, Alrobaian 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(5):2205–2221. doi:10.1002/bab.2279

72. Karim S, Akhter MH, Burzangi AS, 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. doi:10.3390/gels8040219

73. Mir MA, Akhter MH, Afzal O, et al. Design-of-Experiment-Assisted Fabrication of Biodegradable Polymeric Nanoparticles: in Vitro Characterization, Biological Activity, and In Vivo Assessment. ACS Omega. 2023;8(42):38806–38821. doi:10.1021/acsomega.3c01153

74. Vozza G, Khalid M, Byrne HJ, Ryan SM, Jesus M. Frias,Nutraceutical formulation, characterisation, and in-vitro evaluation of methylselenocysteine and selenocystine using food derived chitosan:zein nanoparticles. Food Res Int. 2019;120:295–304. doi:10.1016/j.foodres.2019.02.028

75. Pauluk D, Krause Padilha A, Maissar Khalil N, Mainardes RM. Rubiana Mara Mainardes,Chitosan-coated zein nanoparticles for oral delivery of resveratrol: formation, characterization, stability, mucoadhesive properties and antioxidant activity. Food Hydrocoll. 2019;94:411–417. doi:10.1016/j.foodhyd.2019.03.042

76. Xia L, Cong Z, Liu Z, et al. Improvement of the solubility, photostability, antioxidant activity and UVB photoprotection of trans-resveratrol by essential oil based microemulsions for topical application. J Drug Delivery Sci Technol. 2018;48:346–354. doi:10.1016/j.jddst.2018.10.017

77. Chatterjee B, Gorain B, Mohananaidu K, et al. Targeted drug delivery to the brain via intranasal nanoemulsion: available proof of concept and existing. Int j Pharma. 2019;565:258–268. doi:10.1016/j.ijpharm.2019.05.032

78. Pandey P, Gulati N, Makhija M, et al. Nanoemulsion: a Novel Drug Delivery Approach for Enhancement of Bioavailability. Recent Patents Nanotechnol. 2020;14(4):276–293. doi:10.2174/1872210514666200604145755

79. Jiang Y, Zhang Y, Liang R, et al. Beta-carotene chemical stability in Nanoemulsions was improved by stabilized with beta-lactoglobulin-catechin conjugates through free radical method. J Agr Food Chem. 2015;63(1):297–303. doi:10.1021/jf5056024

80. Ding L, Tang S, Yu A, et al. Nanoemulsion-Assisted siRNA Delivery to Modulate the Nervous Tumor Microenvironment in the Treatment of Pancreatic Cancer. ACS Appl Mater Interfaces. 2022;14(8):10015–10029. doi:10.1021/acsami.1c21997

81. Niu Z, Acevedo-Fani A, McDowell A, et al. Nanoemulsion structure and food matrix determine the gastrointestinal fate and in vivo bioavailability of coenzyme Q10. J Control Release. 2020;327:444–455. doi:10.1016/j.jconrel.2020.08.025

82. Bouchemal K, Briançon S, Perrier E, et al. Nano-emulsion formulation using spontaneous emulsification: solvent, oil and surfactant. Int j Pharm. 2004;280(1–2):241–251. doi:10.1016/j.ijpharm.2004.05.016

83. Weiss M, Steiner DF, Philipson LH, et al. Insulin Biosynthesis, Secretion, Structure, and Structure-Activity Relationships. In: Feingold KR, Anawalt B, Blackman MR, editors. Endotext. South Dartmouth (MA): MDText.com, Inc.; 2014.

84. Leber N, Nuhn L, Zentel R, et al. Cationic Nanohydrogel Particles for Therapeutic Oligonucleotide Delivery. Macromol biosci. 2017;17(10):10.1002/mabi.201700092. doi:10.1002/mabi.201700092

85. Deng S, Gigliobianco M, Mijit E, et al. Dually Cross-Linked Core-Shell Structure Nanohydrogel with Redox-Responsive Degradability for Intracellular Delivery. Pharmaceutics. 2021;13(12):2048. doi:10.3390/pharmaceutics13122048

86. Granata G, Petralia S, Forte G, et al. Injectable supramolecular nanohydrogel from a micellar self-assembling calix[4] arene derivative and curcumin for a sustained drug release. Mater Sci Eng. 2020;111:110842. doi:10.1016/j.msec.2020.110842

87. Raemdonck K, Demeester J, De Smedt S. Advanced nanogel engineering for drug delivery. Soft Matter. 2009;5(4):707–715. doi:10.1039/B811923F

88. El-Refaie WM, Elnaggar YSR, El-Massik MA, et al. Novel Self-assembled, Gel-core Hyaluosomes for Non-invasive Management of Osteoarthritis: in-vitro Optimization, Ex-vivo and In-vivo Permeation. Pharm Res. 2015;32(9):2901–2911. doi:10.1007/s11095-015-1672-8

89. Udeni Gunathilake TMS, Ching YC, Chuah CH. Enhancement of Curcumin Bioavailability Using Nanocellulose Reinforced Chitosan Hydrogel. Polymers. 2017;9(2):64. doi:10.3390/polym9020064

90. Ahmad I, Farheen M, Kukreti A, et al. Natural Oils Enhance the Topical Delivery of Ketoconazole by Nanoemulgel for Fungal Infections. ACS omega. 2023;8(31):28233–28248. doi:10.1021/acsomega.3c01571

91. Yanchen H, Zhi Z, Zhao Q, et al. 3D cubic mesoporous silica microsphere as a carrier for poorly soluble drug carvedilol. Microporous Mesoporous Mater. 2012;147(1):94–101. doi:10.1016/j.micromeso.2011.06.001

92. Zhang Y, Wang J, Bai X, et al. Mesoporous silica nanoparticles for increasing the oral bioavailability and permeation of poorly water soluble drugs. Mol Pharmaceut. 2012;9(3):505–513. doi:10.1021/mp200287c

93. Pawlaczyk M, Schroeder G. Dual-Polymeric Resin Based on Poly (methyl vinyl ether-alt-maleic anhydride) and PAMAM Dendrimer as a Versatile Supramolecular Adsorbent. ACS Appl Polymer Mater. 2021;3(2):956–967. doi:10.1021/acsapm.0c01254

94. Liu X, Peng Y, Qu X, et al. Multi-walled carbon nanotube-chitosan/poly(amidoamine)/DNA nanocomposite modified gold electrode for determination of dopamine and uric acid under coexistence of ascorbic acid. J Electroanal Chem. 2011;654(1–2):72–78. doi:10.1016/j.jelechem.2011.01.024

95. Zhuo RX, Du B, Lu ZR. In vitro release of 5-fluorouracil with cyclic core dendritic polymer. J Control Release. 1999;57(3):249–257. doi:10.1016/s0168-3659(98)00120-5

96. Luo K, Li C, Li L, et al. Arginine functionalized peptide dendrimers as potential gene delivery vehicles. Biomaterials. 2012;33(19):4917–4927. doi:10.1016/j.biomaterials.2012.03.030

97. Conghua Y, Qinghe X, Yang D, Wang M. Miao Wang,A novel pH-responsive charge reversal nanospheres based on acetylated histidine-modified lignin for drug delivery. Ind Crops Prod. 2022;186:115193. doi:10.1016/j.indcrop.2022.115193

98. Guo F, Jiao Y, Du Y, et al. Enzyme-responsive nano-drug delivery system for combined antitumor therapy. Int J Biol Macromol. 2022;220:1133–1145. doi:10.1016/j.ijbiomac.2022.08.123

99. Hui Y, Jin F, Liu D, et al. ROS-responsive nano-drug delivery system combining mitochondria-targeting ceria nanoparticles with atorvastatin for acute kidney injury. Theranostics. 2020;10(5):2342–2357. doi:10.7150/thno.40395

100. Wang G, Maciel D, Wu Y, et al. Amphiphilic polymer-mediated formation of laponite-based nanohybrids with robust stability and pH sensitivity for anticancer drug delivery. ACS Appl Mater Interfaces. 2014;6(19):16687–16695. doi:10.1021/am5032874

101. Ghadiri M, Young PM, Traini D. Strategies to Enhance Drug Absorption via Nasal and Pulmonary Routes. Pharmaceutics. 2019;11(3):113. doi:10.3390/pharmaceutics11030113

102. Pawar B, Vasdev N, Gupta T, et al. Current Update on Transcellular Brain Drug Delivery. Pharmaceutics. 2022;14(12):2719. doi:10.3390/pharmaceutics14122719

103. Sosnik A, Das Neves J. Bruno Sarmento.Mucoadhesive polymers in the design of nano-drug delivery systems for administration by non-parenteral routes: a review. Prog Polym Sci. 2014;39(12):2030–2075.

104. Costa CP, Moreira JN, Sousa Lobo JM, et al. Intranasal delivery of nanostructured lipid carriers, solid lipid nanoparticles and nanoemulsions: a current overview of in vivo studies. Acta Pharmaceutica Sinica B. 2021;11(4):925–940. doi:10.1016/j.apsb.2021.02.012

105. Bhaskar S, Tian F, Stoeger T, et al. Multifunctional Nanocarriers for diagnostics, drug delivery and targeted treatment across blood-brain barrier: perspectives on tracking and neuroimaging. Particle Fibre Toxicol. 2010;7:3. doi:10.1186/1743-8977-7-3

106. Sun W, Xie C, Wang H, et al. Specific role of polysorbate 80 coating on the targeting of nanoparticles to the brain. Biomaterials. 2004;25(15):3065–3071. doi:10.1016/j.biomaterials.2003.09.087

107. Wei L, Tan Y-Z, Hu K-L, et al. Cationic albumin conjugated pegylated nanoparticle with its transcytosis ability and little toxicity against blood-brain barrier. Int J Pharm. 2005;295(1–2):247–260. doi:10.1016/j.ijpharm.2005.01.043

108. Chai Z, Hu X, Wei X, et al. A facile approach to functionalizing cell membrane-coated nanoparticles with neurotoxin-derived peptide for brain-targeted drug delivery. J Control Release. 2017;264:102–111. doi:10.1016/j.jconrel.2017.08.027

109. Fang RH, Hu C-MJ, Chen KNH, et al. Lipid-insertion enables targeting functionalization of erythrocyte membrane-cloaked nanoparticles. Nanoscale. 2013;5(19):8884–8888. doi:10.1039/c3nr03064d

110. Yan L, He H, Jia X, et al. A dual-targeting nanocarrier based on poly(amidoamine) dendrimers conjugated with transferrin and tamoxifen for treating brain gliomas. Biomaterials. 2012;33(15):3899–3908. doi:10.1016/j.biomaterials.2012.02.004

111. Habban Akhter M, Beg S, Tarique M, Malik A, Afaq S, Choudhry H. Salman Hosawi,Receptor-based targeting of engineered nanocarrier against solid tumors: recent progress and challenges ahead. Biochimica et Biophysica Acta. 2021;1865(2):129777. doi:10.1016/j.bbagen.2020.129777

112. Spanjers JM, Städler B. Cell Membrane Coated Particles. Adv Biosyst 2020;4(11):e2000174. doi:10.1002/adbi.202000174

113. Ricordel C, Labalette-Tiercin M, Lespagnol A, et al. EFGR-mutant lung adenocarcinoma and Li-Fraumeni syndrome: report of two cases and review of the literature. Lung Cancer. 2015;87(1):80–84. doi:10.1016/j.lungcan.2014.11.005

114. Farheen M, Akhter MH, Chitme H, et al. Surface-Modified Biobased Polymeric Nanoparticles for Dual Delivery of Doxorubicin and Gefitinib in Glioma Cell Lines. ACS omega. 2023;8(31):28165–28184. doi:10.1021/acsomega.3c01375

115. Wang Z-H, Liu J-M, Zhao N, et al. Cancer Cell Macrophage Membrane-Camouflaged Persistent-Luminescent Nanoparticles for Imaging-Guided Photothermal Therapy of Colorectal Cancer. ACS Appl Nano Mater 2020;XXXX. doi:10.1021/acsanm.0c01433

116. Zhao K, Li D, Cheng G, et al. Targeted Delivery Prodigiosin to Choriocarcinoma by Peptide-Guided Dendrigraft Poly-l-lysines Nanoparticles. Int j Mol Sci. 2019;20(21):5458. doi:10.3390/ijms20215458

117. Wang Y, Cheng J, Zhao D, et al. Designed DNA nanostructure grafted with erlotinib for non-small-cell lung cancer therapy. Nanoscale. 2020;12(47):23953–23958. doi:10.1039/d0nr06945k

118. Gong Y, Mohd S, Wu S, et al. pH-responsive cellulose-based microspheres designed as an effective oral delivery system for insulin. ACSO Mega. 2021;6:2734–2741.

119. Cikrikci S, Mert H, Oztop MH. Development of pH sensitive alginate/gum tragacanth based hydrogels for oral insulin delivery. J Agric Food Chem. 2018;66:11784–11796. doi:10.1021/acs.jafc.8b02525

120. Qi X, Yuan Y, Zhang J, et al. Oral administration of salecan based hydrogels for controlled insulin delivery. J Agric Food Chem. 2018;66:10479–10489. doi:10.1021/acs.jafc.8b02879

121. Zhang L, Qin H, Li J, et al. Preparation and characterization of layer-by-layer hypoglycemic nanoparticles with pH-sensitivity for oral insulin delivery. J Mater Chem. 2018;6:7451–7461.

122. Shabana AM, Kambhampati SP, Hsia R-C, et al. Thermosensitive and Biodegradable Hydrogel Encapsulating Targeted Nanoparticles for the Sustained Co-Delivery of Gemcitabine and Paclitaxel to Pancreatic Cancer Cells. Int J Pharm. 2020;593(6):120139. doi:10.1016/j.ijpharm.2020.120139

123. Deodhar S, Dash AK, North EJ, et al. Development and In Vitro Evaluation of Long Circulating Liposomes for Targeted Delivery of Gemcitabine and Irinotecan in Pancreatic Ductal Adenocarcinoma. AAPS Pharm Sci Tech. 2020;21(6). doi:10.1208/s12249-020-01745-6

124. Patra JK, Das G, Fraceto LF, et al. Nano based drug delivery systems: recent developments and future prospects. J Nanobiotechnol. 2018;16(1):71. doi:10.1186/s12951-018-0392-8

125. Ruan S, Li J, Ruan H, et al. Microneedle-mediated nose-to-brain drug delivery for improved Alzheimer’s disease treatment. J Controlled Release. 2024;366:712–731. doi:10.1016/j.jconrel.2024.01.013

126. Jahangirian H, Ghasemian Lemraski E, Webster TJ, et al. A review of drug delivery systems based on nanotechnology and green chemistry: green nanomedicine. Int j Nanomed. 2017;12:2957–2978. doi:10.2147/IJN.S127683

127. Hansen K, Kim G, Desai K-GH, et al. Feasibility Investigation of Cellulose Polymers for Mucoadhesive Nasal Drug Delivery Applications. Mol Pharmaceut. 2015;12(8):2732–2741. doi:10.1021/acs.molpharmaceut.5b00264

128. Mignani S, El Kazzouli S, Bousmina M, et al. Expand classical drug administration ways by emerging routes using dendrimer drug delivery systems: a concise overview. Adv Drug Delivery Rev 2013;65(10):1316–1330. doi:10.1016/j.addr.2013.01.001

129. Miyata K, Christie RJ, Kataoka K. Polymeric micelles for nano-scale drug delivery. React Funct Polym. 2011;3(71). doi:10.1016/j.reactfunctpolym.2010.10.009

130. Wei X, Ling P, Zhang T, et al. Polymeric micelles, a promising drug delivery system to enhance bioavailability of poorly water-soluble drugs. J Drug Delivery. 2013;2013:340315. doi:10.1155/2013/340315

131. Lam P-L, Wong W-Y, Bian Z, Chui C-H, Gambari R. Gambari R.Recent advances in green nanoparticulate systems for drug delivery: efficient delivery and safety concern. Nanomedicine. 2017;12:357–385. doi:10.2217/nnm-2016-0305

132. Powell JJ, Faria N, Thomas-McKay E, et al. Origin and fate of dietary nanoparticles and microparticles in the gastrointestinal tract. J Autoimmun. 2010;34(3):J226–33. doi:10.1016/j.jaut.2009.11.006

133. Turner JR. Intestinal mucosal barrier function in health and disease. Nat Rev Immunol. 2009;9(11):799–809. doi:10.1038/nri2653

134. Sahay G, Alakhova DY, Kabanov AV, et al. Endocytosis of nanomedicines. J Controlled Release. 2010;145(3):182–195. doi:10.1016/j.jconrel.2010.01.036

135. Liang Q, Bie N, Yong T, et al. The softness of tumour-cell-derived microparticles regulates their drug-delivery efficiency. Nat Biomed Eng. 2019;3(9):729–740. doi:10.1038/s41551-019-0405-4

136. Della Camera G, Madej M, Ferretti AM, et al. Personalised Profiling of Innate Immune Memory Induced by Nano-Imaging Particles in Human Monocytes. Front Immunol. 2021;12:692165. doi:10.3389/fimmu.2021.692165

137. Wang J, Chen H-J, Hang T, et al. Physical activation of innate immunity by spiky particles. Nature Nanotechnol. 2018;13(11):1078–1086. doi:10.1038/s41565-018-0274-0

138. Ejigah V, Owoseni O, Bataille-Backer P, et al. Approaches to Improve Macromolecule and Nanoparticle Accumulation in the Tumor Microenvironment by the Enhanced Permeability and Retention Effect. Polymers. 2022;14(13):2601. doi:10.3390/polym14132601

139. Xiao M, Hu M, Dong M, et al. Folic Acid Decorated Zeolitic Imidazolate Framework (ZIF-8) Loaded with Baicalin as a Nano-Drug Delivery System for Breast Cancer Therapy. Int j Nanomed. 2021;16:8337–8352. doi:10.2147/IJN.S340764

140. Kabanov AV, Lemieux P, Vinogradov S, et al. Pluronic block copolymers: novel functional molecules for gene therapy. Adv Drug Delivery Rev 2002;54(2):223–233. doi:10.1016/s0169-409x(02)00018-2

141. Jin P, Sha R, Zhang Y, et al. Blood Circulation-Prolonging Peptides for Engineered Nanoparticles Identified via Phage Display. Nano Lett. 2019;19(3):1467–1478. doi:10.1021/acs.nanolett.8b04007

142. Aggarwal P, Hall JB, McLeland CB, et al. Nanoparticle interaction with plasma proteins as it relates to particle biodistribution, biocompatibility and therapeutic efficacy. Adv Drug Delivery Rev 2009;61(6):428–437. doi:10.1016/j.addr.2009.03.009

143. Hadidi N, Kobarfard F, Nafissi-Varcheh N, et al. PEGylated single-walled carbon nanotubes as nanocarriers for cyclosporin A delivery. AAPS Pharm Sci Tech. 2013;14(2):593–600. doi:10.1208/s12249-013-9944-2

144. Alimohammadi E, Nikzad A, Khedri M, et al. Molecular Tuning of the Nano-Bio Interface: alpha-Synuclein’s Surface Targeting with Doped Carbon Nanostructures. ACS Appl Bio Mater 2021;4(8):6073–6083. doi:10.1021/acsabm.1c00421

145. Wohlfart S, Gelperina S, Kreuter J, et al. Transport of drugs across the blood-brain barrier by nanoparticles. J Controlled Release. 2012;161(2):264–273. doi:10.1016/j.jconrel.2011.08.017

146. Shen J, Zhao Z, Shang W, et al. Ginsenoside Rg1 nanoparticle penetrating the blood–brain barrier to improve the cerebral function of diabetic rats complicated with cerebral infarction. Int j Nanomed. 2017;12:6477–6486. doi:10.2147/IJN.S139602

147. Wiley DT, Webster P, Gale A, et al. Transcytosis and brain uptake of transferrin-containing nanoparticles by tuning avidity to transferrin receptor. Proc Natl Acad Sci USA. 2013;110(21):8662–8667. doi:10.1073/pnas.1307152110

148. Liu Y, Zhang T, Li G, et al. Radiosensitivity enhancement by Co-NMS-mediated mitochondrial impairment in glioblastoma. J Cell Physiol. 2020;235(12):9623–9634. doi:10.1002/jcp.29774

149. Slepicka P, Kasalkova NS, Siegel J, et al. Nano-structured and functionalized surfaces for cytocompatibility improvement and bactericidal action. Biotechnol Adv 2015;33(6):1120–1129. doi:10.1016/j.biotechadv.2015.01.001

150. Luo X, Wu S, Xiao M, et al. Advances and Prospects of Prolamine Corn Protein Zein as Promising Multifunctional Drug Delivery System for Cancer Treatment. Int j Nanomed. 2023;18:2589–2621. doi:10.2147/IJN.S402891

151. Canal F, Vicent MJ, Pasut G, et al. Relevance of folic acid/polymer ratio in targeted PEG-epirubicin conjugates. J Controlled Release. 2010;146(3):388–399. doi:10.1016/j.jconrel.2010.05.027

152. Yang Y, He Y, Deng Z, et al. Intelligent Nanoprobe: acid-Responsive Drug Release and In Situ Evaluation of Its Own Therapeutic Effect. Anal Chem 2020;92(18):12371–12378. doi:10.1021/acs.analchem.0c02099

153. Desai N. Challenges in development of nanoparticle-based therapeutics. AAPS J. 2012;14(2):282–295. doi:10.1208/s12248-012-9339-4

154. Higaki M, Ishihara T, Izumo N, et al. Treatment of experimental arthritis with poly(D, L-lactic/glycolic acid) nanoparticles encapsulating betamethasone sodium phosphate. Ann Rheumatic Dis. 2005;64(8):1132–1136. doi:10.1136/ard.2004.030759

155. McCully M, Sanchez-Navarro M, Teixido M, et al. Peptide Mediated Brain Delivery of Nano- and Submicroparticles: a Synergistic Approach. Curr Pharm Des 2018;24(13):1366–1376. doi:10.2174/1381612824666171201115126

156. Tinkle S, McNeil SE, Mühlebach S, et al. Nanomedicines: addressing the scientific and regulatory gap. Ann N Y Acad Sci. 2014;1313:35–56. doi:10.1111/nyas.12403

157. Zcan I, et al. Effects of sterilization techniques on the PEGylated poly (γ-benzyl-L-glutamate) (PBLG) nanoparticles. Acta Pharma Sci. 2009;51(3):211–218.

158. Song Y, Li X, Du X, et al. Exposure to nanoparticles is related to pleural effusion, pulmonary fibrosis and granuloma. Europ resp J. 2009;34(3):559–567. doi:10.1183/09031936.00178308

159. Nel A, Xia T, Madler L, et al. Toxic potential of materials at the nanolevel. Science. 2006;311(5761):622–627. doi:10.1126/science.1114397

160. Wacker MG, Proykova A, Santos GML, et al. Dealing with nanosafety around the globe-Regulation vs. innovation. Int J Pharm. 2016;509(1–2):95–106. doi:10.1016/j.ijpharm.2016.05.015

161. Mühlebach S. Regulatory challenges of nanomedicines and their follow-on versions: a generic or similar approach? Adv. Drug Delivery Rev. 2018;131:122–131. doi:10.1016/j.addr.2018.06.024

162. Paradise J. Regulating Nanomedicine at the Food and Drug Administration. AMA j Ethics. 2019;21(4):E347–355. doi:10.1001/amajethics.2019.347

163. Resnik DB, Tinkle SS. Ethical issues in clinical trials involving nanomedicine. Contemporary Clinical Trials. 2007;28(4):433–441. doi:10.1016/j.cct.2006.11.001

164. Bawa R, Johnson S. The ethical dimensions of nanomedicine. Med Clin North Am. 2007;91(5):881–887. doi:10.1016/j.mcna.2007.05.007

165. Zhang J, Hu K, Di L, et al. Traditional herbal medicine and nanomedicine: converging disciplines to improve therapeutic efficacy and human health. Adv. Drug Delivery Rev. 2021;178:113964. doi:10.1016/j.addr.2021.113964

166. Majidinia M, Mirza‐Aghazadeh‐Attari M, Rahimi M, et al. Overcoming multidrug resistance in cancer: recent progress in nanotechnology and new horizons. IUBMB Life. 2020;72(5):855–871. doi:10.1002/iub.2215

167. Shen C-Y, Xu P-H, Shen B-D, et al. Nanogel for dermal application of the triterpenoids isolated from Ganoderma lucidum (GLT) for frostbite treatment. Drug Delivery. 2016;23(2):610–618. doi:10.3109/10717544.2014.929756

168. Ventola CL. Progress in Nanomedicine: approved and Investigational Nanodrugs. PT. 2017;42(12):742–755.

169. Sainz V, Conniot J, Matos AI, et al. Regulatory aspects on nanomedicines. Biochem. Biophys. Res. Commun. 2015;468(3):504–510. doi:10.1016/j.bbrc.2015.08.023

170. Agrahari V, Agrahari V. Facilitating the translation of nanomedicines to a clinical product: challenges and opportunities. Drug Discovery Today. 2018;23(5):974–991. doi:10.1016/j.drudis.2018.01.047

171. Zhang N, Feng N, Xin X, et al. Nano-drug delivery system with enhanced tumour penetration and layered anti-tumour efficacy. Nanomedicine. 2022;45:102592. doi:10.1016/j.nano.2022.102592

172. Ouyang Q, Meng Y, Zhou W, et al. New advances in brain-targeting nano-drug delivery systems for Alzheimer’s disease. J Drug Targeting. 2022;30(1):61–81. doi:10.1080/1061186X.2021.1927055

173. Filipczak N, Yalamarty SSK, Li X, et al. Developments in Treatment Methodologies Using Dendrimers for Infectious Diseases. Molecules. 2021;26(11):3304. doi:10.3390/molecules26113304

174. Cardoso RV, Pereira PR, Freitas CS, et al. Trends in Drug Delivery Systems for Natural Bioactive Molecules to Treat Health Disorders: the Importance of Nano-Liposomes. Pharmaceutics. 2022;14(12):2808. doi:10.3390/pharmaceutics14122808

175. Manzari MT, Shamay Y, Kiguchi H, et al. Targeted drug delivery strategies for precision medicines. Nature Rev Mater. 2021;6(4):351–370. doi:10.1038/s41578-020-00269-6

176. Taylor J, Sharp A, Rannard SP, et al. Nanomedicine strategies to improve therapeutic agents for the prevention and treatment of preterm birth and future directions. Nanoscale Adv. 2023;5(7):1870–1889. doi:10.1039/d2na00834c

177. Savjani KT, Gajjar AK, Savjani JK, et al. Drug solubility: importance and enhancement techniques. ISRN Pharmaceutics. 2012;2012:195727. doi:10.5402/2012/195727

178. Mistry A, Stolnik S, Illum L, et al. Nose-to-Brain Delivery: investigation of the Transport of Nanoparticles with Different Surface Characteristics and Sizes in Excised Porcine Olfactory Epithelium. Mol Pharmaceut. 2015;12(8):2755–2766. doi:10.1021/acs.molpharmaceut.5b00088

179. Hegewald AB, et al. Extracellular miR-574-5p Induces Osteoclast Differentiation via TLR 7/8 in Rheumatoid Arthritis. Front Immunol. 2020;11(585282). doi:10.3389/fimmu.2020.585282

180. Motoyama K, Onodera R, Okamatsu A, et al. Potential use of the complex of doxorubicin with folate-conjugated methyl- β -cyclodextrin for tumor-selective cancer chemotherapy. J Drug Targeting. 2014;22(3):211–219. doi:10.3109/1061186X.2013.856012

181. Date AA, Hanes J, Ensign LM, et al. Nanoparticles for oral delivery: design, evaluation and state-of-The-art. J Controlled Release. 2016;240:504–526. doi:10.1016/j.jconrel.2016.06.016

182. Xiao S, Tang Y, Lv Z, et al. Nanomedicine - advantages for their use in rheumatoid arthritis theranostics. J Controlled Release. 2019;316:302–316. doi:10.1016/j.jconrel.2019.11.008

183. Wang S, Lv J, Meng S, et al. Recent Advances in Nanotheranostics for Treat-to-Target of Rheumatoid Arthritis. Adv. Healthcare Mater. 2020;9(6):e1901541. doi:10.1002/adhm.201901541

Creative Commons License © 2024 The Author(s). This work is published and licensed by Dove Medical Press Limited. The full terms of this license are available at https://www.dovepress.com/terms.php and incorporate the Creative Commons Attribution - Non Commercial (unported, v3.0) License. By accessing the work you hereby accept the Terms. Non-commercial uses of the work are permitted without any further permission from Dove Medical Press Limited, provided the work is properly attributed. For permission for commercial use of this work, please see paragraphs 4.2 and 5 of our Terms.