Next Article in Journal
Charge Regulation of Poly(acrylic acid) in Solutions of Non-Charged Polymer and Colloids
Previous Article in Journal
Study on Flocculation Behavior of Cr(VI) Using a Novel Chitosan Functionalized with Thiol Groups
Previous Article in Special Issue
Enhancing the Antifungal Activity and Ophthalmic Transport of Fluconazole from PEGylated Polycaprolactone Loaded Nanoparticles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Polymeric Nanoparticles for Delivery of Natural Bioactive Agents: Recent Advances and Challenges

by
Mohammed Elmowafy
1,*,
Khaled Shalaby
1,
Mohammed H. Elkomy
1,
Omar Awad Alsaidan
1,
Hesham A. M. Gomaa
2,
Mohamed A. Abdelgawad
3 and
Ehab M. Mostafa
4
1
Department of Pharmaceutics, College of Pharmacy, Jouf University, Sakaka P.O. Box 2014, Saudi Arabia
2
Department of Pharmacology, College of Pharmacy, Jouf University, Sakaka P.O. Box 2014, Saudi Arabia
3
Department of Pharmaceutical Chemistry, College of Pharmacy, Jouf University, Sakaka P.O. Box 2014, Saudi Arabia
4
Department of Pharmacognosy, College of Pharmacy, Jouf University, Sakaka P.O. Box 2014, Saudi Arabia
*
Author to whom correspondence should be addressed.
Polymers 2023, 15(5), 1123; https://doi.org/10.3390/polym15051123
Submission received: 26 January 2023 / Revised: 16 February 2023 / Accepted: 20 February 2023 / Published: 23 February 2023
(This article belongs to the Special Issue Research Progress of Polymer Nanomaterials for Drug Delivery)

Abstract

:
In the last few decades, several natural bioactive agents have been widely utilized in the treatment and prevention of many diseases owing to their unique and versatile therapeutic effects, including antioxidant, anti-inflammatory, anticancer, and neuroprotective action. However, their poor aqueous solubility, poor bioavailability, low GIT stability, extensive metabolism as well as short duration of action are the most shortfalls hampering their biomedical/pharmaceutical applications. Different drug delivery platforms have developed in this regard, and a captivating tool of this has been the fabrication of nanocarriers. In particular, polymeric nanoparticles were reported to offer proficient delivery of various natural bioactive agents with good entrapment potential and stability, an efficiently controlled release, improved bioavailability, and fascinating therapeutic efficacy. In addition, surface decoration and polymer functionalization have opened the door to improving the characteristics of polymeric nanoparticles and alleviating the reported toxicity. Herein, a review of the state of knowledge on polymeric nanoparticles loaded with natural bioactive agents is presented. The review focuses on frequently used polymeric materials and their corresponding methods of fabrication, the needs of such systems for natural bioactive agents, polymeric nanoparticles loaded with natural bioactive agents in the literature, and the potential role of polymer functionalization, hybrid systems, and stimuli-responsive systems in overcoming most of the system drawbacks. This exploration may offer a thorough idea of viewing the polymeric nanoparticles as a potential candidate for the delivery of natural bioactive agents as well as the challenges and the combating tools used to overcome any hurdles.

1. Introduction

Natural products, such as plants, have been considered one of the major treatments for current diseases. Phytochemicals are substances obtained from herbal plants and are commonly used for non-nutritive benefits to the plants, such as protection from external dangers such as pollution, pressure, deficiency, UV, and microbial threats [1]. Out-of-date familiarity and updated findings verify comparable protective effects in humans. In particular, different phytochemicals have been well established for their antioxidants, anti-inflammatories (such as polyphenols, flavonoids, carotenoids, and allyl sulfide-containing compounds), hormones (such as isoflavones of soy), enzyme-regulating materials (protease inhibitors and indole-containing compounds), anti-infective agents (proanthocyanidins), and DNA replication modulators (saponins and capsaicin). Hence, attention has been paid to phytochemicals, particularly in pharmaceutical technology, to fabricate such compounds in easily administered and effective formulations.
However, phytochemicals suffer from several drawbacks that limit their applications in crude forms or at least lead to unambitious effects. These drawbacks include hydrophobicity, poor stability, reduced absorption and bioavailability, extensive metabolism, and fast excretion, leading to issued therapeutic effects, which can probably be overwhelmed by using nanotechnology [2].
With the alarming rise of illnesses, appropriate drug delivery has received much attention for treating and controlling diseases. Nanocarriers have been considered promising platforms for drug delivery that can enhance their therapeutic effect and overcome hurdles. They have offered an improved solubility and bioavailability of sparingly soluble drugs, enhanced preventive and pharmacological effects, and reduced adverse effects. They include nanospheres, nanocapsules, nanoemulsions, nanoliposomes, and nanoniosomes [3]. Polymeric nanocarriers (Figure 1) are manifested by their distinctive physicochemical structures, including solid polymeric nanoparticles, polymeric micelle, polymer conjugates, dendrimers, polymersomes, polyplexes, and lipomers [4]. In the last few decades, polymeric nanoparticles have received much attention and held much importance in the horizon of drug delivery. They were recognized as colloidal systems with particle sizes ranging from 1–1000 nm [5].
They are fabricated by natural, synthetic, or semisynthetic polymers. Polymeric nanocarriers synthesized from biodegradable and biocompatible polymers offer motivating preferences for controlled and targeted drug delivery [5,6,7]. Natural polymers are a concern due to their lower toxicity when compared to synthetic polymers [8], as natural polymers are biocompatible, biodegradable, safe, greatly stable, economical, and easily handled. In general, polymeric nanoparticles possess several merits (Figure 2), such as encapsulation of differently characterized actives, protection of the encapsulated drugs, and the possibility of targeting.
In this work, we have attempted to present a comprehensive review of utilizing polymeric nanoparticulate systems as promising carriers for natural bioactive agents with modern advances and persistent challenges for their effective delivery. So, this paper reviewed the formulation of polymeric nanoparticles loaded with natural bioactive agents used to treat and control different diseases. This review does not include the details of the fabrication techniques. However, it offers the principal approach to delivering natural bioactive agents via polymeric nanoparticles. To the best of our knowledge, this review addresses this essential topic, and therefore, it will assist the researchers in appreciating and formulating promising polymeric-based nanoparticles for loading natural bioactive agents by realizing the merits and challenges of the system, which could eventually be utilized in industrial development. Hence, the following aspects were considered: (a) an overview of various components of polymeric nanoparticles and methods for fabrication, (b) polymeric nanoparticles loaded with natural bioactive agents in the literature, and (c) the potential role of surface decoration and polymer functionalization in overcoming most of the drawbacks of these systems.

2. Components and Formulation Attributes

Principally, polymeric nanoparticles are simply composed of polymer(s), surface active agent(s), and aqueous phase. Polymers are the backbone of polymeric nanoparticles. Generally, they are made up of large repeating units designated as monomers. Currently, utilizing nanosystems for drug delivery has been fortified by the improvement/modification of actives’ properties, such as solubility, targetability, release pattern t0.5, pharmacokinetics, and pharmacodynamics. Accordingly, polymers are considered the cornerstone when formulating polymeric nanoparticles. Hence, it is critical for the researcher to be aware of the polymer properties, such as biocompatibility, biodegradability, stability, drug–polymer interaction, transition temperature, permeability, and safety, before starting formulation. In order to modulate the characteristics (such as release behavior, targetability, and biocompatibility) of the generated system, the functionalization of the polymer is the point of view. Changes in the polymer can be achieved by chemical modification, surface addition of targeting moiety, and incorporation of lipids to achieve the objectives of the researchers [9].
However, polymers are basically categorized into two classes according to their origin source; natural polymers and synthetic polymers. According to the process of development and components, the structural organization varies (Figure 3). In the case of nanocapsules, active agents might be encapsulated in the core or adsorbed at the polymeric surface, while in the case of nanospheres, active agents might be encapsulated in the matrix of the nanosphere or the surface may be adsorbed at the polymeric surface [10].

2.1. Natural Polymers

In the literature, there are many natural polymers frequently utilized for the development of polymeric nanoparticles, such as chitosan, alginate, gelatin, and albumin (Table 1).

2.1.1. Chitosan

Chitosan (Figure 4) is a natural polymer obtained by the deacetylation of chitin. It is obtained from different natural sources such as marines, insects, fungi, or animals. In the last few decades, it has gained attention as a drug delivery platform due to its excellent biocompatibility and biodegradability [39]. Chitosan is not soluble in neutral aqueous solutions, but it can be inherently dissolved in diluted acidic solutions due to the protonation of free amine groups. Consequently, it is soluble in diluted acetic, formic, citric, and other acids [40]. On the other hand, amine groups also affect chitosan applications in the pharmaceutical field as they particularly influence mucoadhesion, permeation enhancement, transfection, and in situ gelation of chitosan [39,41]. In general, chitosan has different promising biomedical properties, for instance, as a penetration enhancer in both paracellular (by opening the tight junctions) and transcellular transport and as a mucoadhesive, with additional wound-healing activities. Several methods are utilized to prepare chitosan nanoparticles, such as ionic gelation, microemulsion, and emulsification solvent diffusion. The main advantage of the reported methods is that they use much smaller amounts of organic solvent. The most popular techniques are mentioned below:

Ionic Gelation

Ionic gelation is a simple and widespread technique used for nanoparticle preparation. In this method, chitosan solution (polycationic) is prepared by dissolving in a diluted acidic solution, such as acetic acid, and thoroughly mixing (under mechanical stirring) with a polyanionic agent, such as tripolyphosphate, as a cross-linking agent leading to complexation between positive and negative charged ions and the formation of rounded shape particles. This technique is favored due to the lack of organic solvent and minimal opportunity to change the chemistry of the entrapped drug. However, this method is characterized by its poor stability in acidic environments and uneasiness in encapsulating high molecular weight actives [42,43].

Microemulsion

Herein, chitosan solution and glutaraldehyde are first mixed. Then, the mixture is added to a surfactant in an organic solvent. The resultant mixture is incessantly stirred until cross-linking is accomplished. After that, the removal of organic solvent can be carried out by evaporation. If desirable, the removal of extra surfactant can be carried out by centrifugation after adding calcium chloride to precipitate the excess surfactant. The preparation of nanoparticles by this method leads to the production of small particle sizes [44]. On the other hand, the incorporation of glutaraldehyde and organic solvent, as well as a longer period of processing, are the main drawbacks of this method [42].

Emulsification Solvent Diffusion

In this method, the preparation of oil in water emulsion is accomplished by mixing the drug (dissolved in organic solvent) solution stabilizer containing chitosan solution with continuous stirring. The resultant dispersion is further homogenized by high-pressure homogenization [45,46]. This method is favored for hydrophobic drugs as it results in high encapsulation efficiency. On the other hand, the incorporation of organic solvents and high shear forces are the main drawbacks of this method.

Coprecipitation

In this method, acidic chitosan solution is added to a high pH 8.5–9.0 solution (NH4OH), leading to the coprecipitation of nanoparticles. This method results in the production of highly homogenous small-sized nanoparticles with high entrapment efficiency [47,48].

2.1.2. Alginate

Alginate is an anionic water-soluble natural polymer. Chemically (Figure 4), it is 1–4 linked α-L-guluronic acid (G) and β-D-mannuronic acid (M). Owing to its biodegradability, biocompatibility, and mucoadhesion properties, it has gained attention as a drug delivery platform in the pharmaceutical field [49,50]. Mucoadhesive properties are principally attributed to the formation of strong hydrogen bonds with mucin glycoproteins via carboxyl hydroxyl interactions [51]. Generally, alginate nanoparticles are prepared by several techniques, such as ionic gelation, emulsion, covalent cross-linking, emulsification solvent displacement, and emulsion-solvent evaporation. The ionic gelation method is a broadly used technique to prepare alginate nanoparticles [52]. Ionic gelation of alginate was carried out due to its affinity towards multivalent cations such as Ca2+. This affinity leads to the formation of ionically or physically cross-linked alginate [50,53].

2.1.3. Gelatin

Gelatin is a protein obtained from collagen by conversion into an unoriented protein by the method of partial hydrolysis [54]. According to the pH used during collage hydrolysis, gelation is classified into two types: gelatin A (isoelectric point of 9), where collagen is hydrolyzed at acidic pH, and gelatin B (isoelectric point of 5), where collagen is hydrolyzed at alkaline pH. Because of its water solubility, it may be necessary to cross-link gelatin during the development of nanoparticles [55,56]. They are characterized by controlling the release of loaded drugs [57].

2.1.4. Albumin

Albumin is the most popular plasma protein. Its molecular weight is about 66,000 Da [51]. It is a biodegradable, nontoxic, and non-immunogenic protein. Therefore, it presents a potential carrier for drug delivery [49]. More precisely, it is considered a promising platform for cancer drug delivery as it can be preferentially entered into cancer cells (by endocytosis) to provide them with nitrogen and energy, allowing the loaded active to be delivered into cancers [58,59]. Nanoparticles prepared from albumin are generally stimuli-sensitive and highly bioactive substances [60].

2.2. Synthetic Polymers

In the literature, there are many synthetic polymers frequently utilized for the development of polymeric nanoparticles, such as polylactic acid, poly (lactide co-glycolides), polycaprolactone, and poly(amido amine) (Table 2). Different methods were used to prepare such types of nanoparticles (Figure 5).

2.2.1. Polylactic Acid (PLA)

PLA is an aliphatic polyester characterized by its good biodegradability and biocompatibility because it can degrade in situ through hydrolysis. During the hydrolysis process, water molecules break the ester bonds in the PLA backbone, leading to the formation of lactic acid and its short oligomers (Figure 4). These degradation byproducts are easily metabolized by the body, offering an inherent biocompatibility that diminishes the accomplishment of immune reactions. Additionally, it can simply be processed with typical development methodologies, such as injection and extrusion [89]. Furthermore, it displays a high mechanical strength that can be effectively used to develop a controlled-release drug delivery system [90]. Nanoparticle-based PLA can be prepared by different methods such as emulsion-based, nanoprecipitation, spray drying, and supercritical fluid methods.

2.2.2. Poly(lactic-co-glycolic Acid) (PLGA)

PLGA is considered one of the most effectively used biodegradable nanocarriers for the production of nanoparticles as it degrades inside the body by hydrolysis to yield lactic acid and glycolic acid, which are biodegradable metabolic products (Figure 4). As these two byproducts are endogenous and easily undergo biotransformation through the Krebs cycle, the toxicity accompanied by using PLGA for drug delivery is negligible [9,10]. It has recently been tested for use in cancer imaging and therapy [10]. However, the reassembly of PLGA into polymeric devices is still under consideration [91]. PLGA nanoparticles have generally been developed by solvent emulsion–evaporation [92], interfacial deposition [93], emulsification–diffusion [94], and the nanoprecipitation method [95].

2.2.3. Poly-ε-caprolactone (PCL)

PCL is a semi-crystalline aliphatic polyester (Figure 4) that is widely used as a biocompatible and biocompatible platform for the development of nanoparticles. Owing to its hydrophobic nature, the encapsulation of hydrophobic actives is quite easy. Compared with other polymeric materials, the biodegradation of PCL is slower, giving a greater chance for developing controlled release delivery systems [96,97,98]. It can be used alone or as a blend with other biocompatible polymers to obtain new desired properties, such as degradation kinetics, hydrophilicity, and mucoadhesion [99,100]. For example, copolymerization with PLA and PLGA enhanced the biodegradation property [101]. PCL nanoparticles have been generally developed by several techniques, such as the emulsion solvent evaporation technique and the nanoprecipitation and dialysis methods.

2.2.4. Poly(amidoamine) (PAMAM)

This type can develop dendritic polymers possessing good biocompatibility, water solubility, and non-immunogenicity [102]. It can be loaded with different types of drug molecules owing to its distinctive architecture (amide and amide functionality containing polymer). Unfortunately, PAMAM dendrimers were reported to have certain toxicity at cellular levels [103], especially uncoated ones [104]. Thus, surface modification becomes very important to improve their safety profile. The PEGylation of PAMAM has been considered one of the most efficient methods of drug delivery for the treatment of cancer [105]. For example, the PEGylation of PAMAM dendrimers enhanced efficacy and mitigated the toxicity of anticancer and gene delivery [103]. Unlike other polymers, PAMAM dendrimers are synthesized by stepwise synthesis procedures (such as using divergent or convergent methods or a combination of both methods [106,107,108]), which donate well-arranged structures and narrow polydispersity.

3. Why Do We Need Polymeric Nanoparticles for Delivery of Natural Bioactive Agents?

Natural bioactive agents generally face problems such as poor aqueous solubility, poor stability in alimentary canal conditions, extensive metabolism, short half-life, and low bioavailability [109]. The chemical structure of the natural bioactive agents regulates their absorption after oral administration and plasma concentrations. However, the determination of plasma concentration is a key indicator for establishing the bioavailability of natural bioactive agents and their metabolites. For example, nearly 10% of orally administered polyphenols, or their metabolites, are detected in urine and plasma [110]. Other problems associated with oral delivery are insufficient gastric transit period, poor intestinal permeability, and sensitivity to alimentary canal conditions (pH, enzymes, etc.), all of which obstruct the efficacy and therapeutic effectiveness of these natural bioactive agents [111]. If the natural bioactive agent is suffering from extensive metabolism, oral administration will result in a faint therapeutic effect due to the poor availability of active form in the plasma or the availability of pharmacologically inactive fractions.
Establishing a suitable drug delivery system is a major challenge for loading natural bioactive agents. Generally, compounding in nanocarriers is a promising strategy for the delivery of such compounds because the purpose of the nanosized delivery systems of natural bioactive agents is to improve their efficacy and overwhelm the drawbacks associated with their administration. In particular, polymeric nanoparticles are characterized by several properties that enable them to defeat most of the challenges facing the delivery of natural bioactive agents by different routes. For instance, for oral delivery, they can protect actives from degradation and convey them to the optimal regions within the GIT [112]. In addition, their small particle size results in possessing a large effective surface area, leading to an improvement in the dissolution rate of loaded drugs and the effective interaction with epithelial surfaces [9,113]. Furthermore, adjusting self-assembly conditions during preparing polymeric nanoparticles and selecting particular polymer categories can lead to an appropriate design by controlling their characteristics (such as particle diameter, zeta potential, and hydrophobicity) and drug release parameters (fast, controlled, sustained) [114]. In the case of parenteral use, surface modification of polymeric nanoparticles is feasible and not complicated, which leads to overcoming some obstacles associated with parenteral administration. In particular, surface decoration with targeting moieties and PEGylation can effectively lead to targeting behavior and longer residence time in general circulation, respectively. This can be achieved by using diverse polymers tailored with desired groups or coupling specific polymers to the designed polymeric nanoparticles [115]. Specifically, polymeric nanoparticles were reported to effectively deliver actives to the brain through susceptibility to adsorptive-facilitated transcytosis and receptor-mediated transcytosis across the blood–brain barrier, which can be achieved by targeting peptides and/or cell-penetrating moieties anchored to surfaces of polymeric nanoparticles [116]. Different mechanisms were also reported by Kreuter [117,118], such as endothelial cell transcytosis, endothelial cell endocytosis, high-concentration gradients created by polymeric nanoparticles, and membrane fluidization by included surfactants. In the case of topical delivery, polymeric nanoparticles improve skin permeation as their assembly on the surface of the skin or hair follicle can offer a reservoir to release the drug incessantly, creating a concentration gradient [119,120].

4. Natural Bioactive Agents Loaded Polymeric Nanoparticles

Various natural bioactive agents have been loaded into polymeric nanoparticles intended for different uses, as mentioned below.
Bioflavonoids are one of the most popular classes of natural bioactive agents that gained attention due to their promising therapeutic activities and health-promoting effects [121]. The health-improving effect enhances well-being and decreases the opportunity of emerging risk, which includes the immune-stimulant effect, cytotoxic effect, valuable cardiovascular influence, improved eyesight [122], and reduction in amyloidogenic-related cell death [123]. In addition, they can provide efficient anticancer activity by different mechanisms (Figure 6). Their actions were utilized to treat different diseases such as oxidative stress, inflammation, skin disorders, cardiovascular, respiratory, and diabetes.

4.1. Quercetin

Quercetin (3,3′,4′,5,7-penthydroxyflavone) is semi-lipophilic flavonol plentifully present in leafy vegetables, citrus fruits, tomatoes, berries, etc. It possesses several therapeutic effects such as antioxidant, anti-inflammatory, wound healing, antithrombotic, antitumor, antiprotozoal [124,125,126,127,128], and cytoprotective effects [129]. Among the different nanocarriers, polymeric nanoparticles may provide an effective way to encapsulate quercetin as they can improve bioavailability, therapeutic effectiveness, and in vivo stability. However, encapsulation of quercetin in polymeric nanoparticles was reported not to influence their efficacy. Kumari and coworkers developed quercetin-loaded PLA nanoparticles and reported a similar antioxidant activity to free quercetin [67]. Sunoqrot and Abujamous developed the pH-sensitive polymer Eudragit S100 loaded with quercetin by a nanoprecipitation technique to achieve potential colon-specific drug delivery [130]. The prepared formula was 66.8 nm in particle size, −5.2 mV zeta potential, and 41.8% encapsulation efficiency. It showed minimal in vitro quercetin release at an acidic pH and complete release at pH 7.2. In vitro cytotoxicity assay showed promising activity against CT26 murine colon carcinoma cells and significantly lower IC50 when compared to free quercetin. Zhou and coworkers quercetin loaded PLGA-TPGS nanoparticles to improve anticancer activity [131]. The prepared formula was 198.4 ± 7.8 nm particle size, −22.5 ± 2.5 mV zeta potential, and 82.3 ± 5.7% encapsulation efficiency. In vitro cytotoxicity assay exhibited a potential activity against triple-negative breast cancer cells. Interestingly, after oral administration, significant anticancer activity was found in 4T1-bearing mice with few metastatic lung colonies. Additionally, the inhibitory activity on the migration of urokinase-type plasminogen activator (breast cancer tumor marker) knockdown MDA-MB231 cells was significantly decreased. Elmowafy and coworkers prepared and optimized quercetin-loaded mesoporous silica nanoparticles in order to improve the physiochemical characteristics, particularly the dissolution rate and hence the bioavailability [132]. The optimized formulation (using 32 full factorial design) exhibited good physiochemical properties and enhanced the saturation solubility and dissolution rate. Pandey and coworkers developed quercetin-loaded PLA nanoparticles to improve anticancer activity against breast cancer [66]. The particle sizes of the prepared formulations varied from 32 ± 8 to 152 ± 9 nm with 62 ± 3% (w/w) encapsulation efficiency. In vitro cytotoxicity experiment showed that 5 days were sufficient to kill about 40% of breast cancer cells, verifying a sustained release manner. Zhang and coworkers developed quercetin-loaded chitosan nanoparticles to improve antioxidant activity and bioavailability [133]. They found that quercetin existed in nanoparticles in an amorphous state, which showed good antioxidant activity evaluated by scavenging DPPH radical and reducing the power of some plant extracts.

4.2. Curcumin

Curcumin is a natural hydrophobic yellow polyphenol derived from the turmeric rhizome. Accumulating studies have revealed that curcumin possesses a wide range of therapeutic activities. In the previous century, studies proved different pharmacological actions, including antihypercholesterolemia [134], antidiabetic [135], anti-inflammatory [136], and anti-oxidant [137] actions. Recently, different beneficial actions have been reported, such as metabolic disorders [138], inflammatory bowel diseases [139], anti-hepatotoxic [140], anti-proliferative [141,142], anti-carcinogenic [142], antimicrobial [143], and neuroprotective properties [144]. While it looks to have numerous therapeutic activities, its practical applications have been limited owing to limited aqueous solubility, poor bioavailability, rapid biotransformation, low penetrative and targeting power, and instability in alkaline conditions and in the presence of metal ions, heat, and light [145]. In order to improve the oral bioavailability of curcumin, Umerska and coworkers compared three polymeric nanoparticles, Eudragit® RLPO, PLGA, and PCL, loaded with curcumin [146]. The authors found that Eudragit® RLPO nanoparticles had smaller particle sizes, lower encapsulation efficiency, and faster release within 1 h. All developed formulations exhibited good compatibility with intestinal Caco-2 cells. They also concluded that PLGA and PCL-based nanoparticles are more suitable for parenteral administration. Jahagirdar and coworkers developed PLGA nanoparticles loaded with curcumin by facile in situ nanoprecipitation method to target macrophages and treat associated diseases such as intracellular infections [147]. The optimized formulation was 208.25 ± 7.55 nm in particle size, −26  ±  2.55 mV zeta potential, and 90.16 ± 1.17% encapsulation efficiency. Structural analyses (by DSC and FTIR) showed that curcumin was embedded in polymeric matrix in amorphous state. Flow cytometry and HPLC analyses verified fast and efficient uptake (three-fold enhancement in uptake when compared to free curcumin) by RAW 264.7 macrophages. Confocal microscopy confirmed localization of nanprticles in phago-lysosomal compartment. Udompornmongkol and Chiang developed curcumin loaded chitosan/gum arabic polymeric nanoparticles by an emulsification solvent diffusion technique to improve anti-colorectal cancer activity [148]. The developed formulation was 136.3 ± 3.9 nm in particle size, +48 mV zeta potential, 3.57 ± 0.69 loading efficiency, and 950.2 ± 0.03% encapsulation efficiency. Significant cell viability reduction against two models of human colorectal carcinoma cells (HCT116 (Duke’s type A)) and HT29 (Duke’s type B)) and cellular uptake (by fluorescence microscopy and flow cytometry). In addition, curcumin polymeric nanoparticles were more proficient in inducing apoptosis when compared to free curcumin. Nair and coworkers developed curcumin-loaded chitosan polymeric nanoparticles by ionic gelation (using tripolyphosphate as a cross-linker) technique to improve transdermal delivery [149]. The authors studied different factors (such as the chitosan/tripolyphosphate ratio and stirring rate) on the physicochemical characteristics of the developed batches. They found that a high chitosan/tripolyphosphate ratio produces larger particles as tripolyphosphate aided in the attraction of chitosan molecules by sturdier intramolecular interactions, leading to the formation of reduced particle sizes. A high stirring rate (1000 rpm) resulted in larger particle sizes due to the decrease in the repulsive forces between the particles and the improved aggregation. Additionally, curcumin polymeric nanoparticles showed enhanced permeation through Strat-M® membrane and percentage cell viability towards human keratinocyte (HaCat) when compared to free curcumin solution [149]. Recently, Walbi and coworkers developed and optimized (using a 3-factorial Box–Behnken statistical design) curcumin-loaded lecithin/chitosan polymeric nanoparticles to improve drug loading and controlled release behavior [150]. The optimized formulation was of 138.4 ± 2.10 nm particle size, +18.98 ± 0.72 mV zeta potential, and 77.40 ± 1.70% encapsulation efficiency. It also showed diffusion-controlled release behavior and good stability.
On the other hand, researchers paid attention to the functionalization of polymeric nanoparticles’ surfaces in order to improve efficacy and biocompatibility. Grabovac and Bernkop [151] synthesized thiolated chitosan tailoring curcumin PLGA nanoparticles. Thamake and coworkers [152] synthesized homobifunctional spacer, bis(sulfosuccinimidyl) suberate (BS3) tailoring curcumin PLGA nanoparticles and reported an enhancement linkage of Annexin A2 leading to proficient specific delivery to Annexin-A2-positive MDA-MB-231 cancer cells. Joseph and coworkers developed PEG-modified PLGA nanoparticles loaded with curcumin by nanoprecipitation technique to target the brain and control hypoxic–ischemic brain injury in neonatal rats using the Vannucci model [153]. Results showed that nanoparticles could overcome the impaired blood–brain barrier and improve the extravasation of nanoparticles into the brain parenchyma, leading to an efficient decrease in global injury by the neuroprotective effect, which was more marked in the penumbral region. Anitha and coworkers developed curcumin-loaded water-soluble chitosan derivatives (O-carboxymethyl chitosan and N,O-carboxymethyl chitosan) and showed good biocompatibility and induced apoptotic cancer cell death [154,155]. The development of multiple polymeric materials nanocomposites was also a point of interest. Das and coworkers prepared curcumin-loaded alginate/chitosan/Pluronic nanocomposite by ionotropic pre-gelation followed by polycationic crosslinking, showing cellular internalization inside the Hela cells (verified by fluorescent microscopy) [156]. A combination of natural and synthetic polymers in a single polymeric matrix was also studied by Liu and coworkers. They developed chitosan/PCL polymeric nanoparticles by the nanoprecipitation method, showed positively charged nanoparticles, and enhanced cellular uptake by Hela cells and human choroidal melanoma. The authors proposed that the combination of PCL with chitosan (bioadhesive polymer) led to efficient curcumin loading and release profiles, cellular interaction, and remarkable in vitro anticancer activity [157].

4.3. Kaempferol

Kaempferol (3,5,7-trihydroxy-2-(4-hydroxyphenyl)-4H-1-benzopyran-4-one) is frequently present in different plats such as onions, green tea, and grapes. Kaempferol has promising pharmacological actions such as antitumor, antioxidant, anti-inflammatory, and anti-allergic effects [158,159,160]. Like other flavonoids, its therapeutic applications are limited because of its limited aqueous solubility, poor oral bioavailability, extensive first-pass metabolism, and instability in water alkaline medium [161].
Luo and coworkers developed Kaempferol-loaded nonionic poly(ethylene oxide)-poly(propylene oxide)-poly(ethylene oxide) (PEO-PPO-PEO) and PLGA nanoparticles as chemopreventive of ovarian cancer [162]. The authors concluded the selectivity of PLGA nanoparticles to cancer cells was not offered by PEO-PPO-PEO nanoparticles (both cancerous and normal cells were affected). Kazmi and coworkers developed Kaempferol-loaded hydroxypropyl methylcellulose acetate succinate and Kollicoat MAE 30 DP nanoparticles to control induced hepatocellular carcinoma [163]. The authors prepared several batches and obtained a nanometric range of particle size, monodisperse particles, and negatively charged nanoparticles. They also found a significant decrease in serum levels of different liver biomarkers and an increase in antioxidant enzymes and proteins. Additionally, significant decreases in an mRNA expression of interleukin-1β, interleukin-6, and tumor necrosis factor-alpha verify a potential application against oxidative stress induced-liver injury. Ilk and coworkers developed Kaempferol loaded chitosan nanoparticles to modulate quorum sensing mediated by auto-inducers in model bioassay test systems to be used as new emerging strategy in antimicrobial chemotherapy [164]. Nanoparticles were developed by ionic gelation and were 192.27 ± 13.6 nm and +35 mV in value for particle size and zeta potential, respectively. Nanoparticles significantly prohibited the production of violacein pigment in Chromobacterium violaceum CV026 within the 30 storage days. In another work, Ilk and coworkers developed Kaempferol-loaded lecithin/chitosan nanoparticles by electrostatic self-assembly technique to attain sustained antifungal activity [165]. Nanoparticles significantly prohibited 67% of the growth of Fusarium oxysporium within the 60 storage days.

4.4. Resveratrol

Resveratrol (3,5,4′-trihydroxystilbene) is a natural non-flavonoid polyphenolic substance found in different plants, such as grapes, peanuts, and legumes [166]. It exists in two isomers: cis-(Z) and trans-(E) isomers, where the trans-isomer is considered a more stable form and biologically active than the cis-isomer because of its non-planar conformation [167,168]. However, the trans-isomer can be converted to the cis-isomer form upon exposure to UV-light [169,170]. Resveratrol exhibits several biological actions, such as cardioprotection, cancer prevention, platelet de-aggregation, antioxidant, anti-inflammatory, and vasorelaxant properties [171]. It can also improve the antiviral activity of zidovudine, zalcitabine, and didanosine [172].
Hung and coworkers developed PEG-modified PLA nanoparticles loaded with resveratrol to enhance drug stability, controlled delivery, and anticancer activity [173]. Cellular apoptosis and uptake were enhanced in nanoparticles when compared to free resveratrol (verified by flow cytometry and western blots). In vivo, retardation of tumor growth was observed during a long treatment period. Yee and coworkers developed methoxyPEG-modified PLA nanoparticles loaded with resveratrol to improve its plasmatic stability and reduce the metabolic rate [61].
The developed nanoparticles displayed good stability in plasma and decreased hepatic metabolic rate. The authors reported a correlation between enhanced plasma stability, reduced liver metabolic rate, and suppression of tumor growth in mice. Chen and coworkers developed PLGA nanoparticles loaded with both resveratrol and puerarin by the O/W emulsion method to improve their chemotherapeutic efficacy in spinal cord injury [174]. Nanoparticles were homogeneously dispersed with mean diameters varying from 238 nm to 274 nm and a zeta potential of −12.6  ±  2.1 mV. Optimized formulation showed 72−79% of drug release through 36 h and decreased free radicals production and oxidative stress. Zu and coworkers developed carboxymethyl chitosan nanoparticles loaded with resveratrol by emulsion cross-linking to improve its physicochemical properties, antioxidant activity, and pharmacokinetic profile [175]. Optimized formulation was of 155.3 ± 15.2 nm, −10.28 ± 6.4, and 44.5 ± 2.2% values for particle size, zeta potential, and encapsulation efficiency, respectively. It also showed improved solubility and antioxidant activity. In addition, it showed enhanced in vivo absorption, prolonged duration of action, and increased relative bioavailability by 3.5-fold when compared to free drugs. Sarma and coworkers developed chitosan–pectin core–shell nanoparticles loaded with resveratrol to improve its antioxidant activity and obtain prolonged release [176]. Liu and coworkers developed α-lactalbumin and chitosan core–shell nanoparticles loaded with resveratrol [177]. Nanoparticles were more stable against UV-light and heat and bioaccessible with prompt antioxidant activity.

4.5. Epigallocatechin-3-Gallate

Epigallocatechin-3-Gallate (a major constituent of green tea polyphenols) has been shown to be chemopreventive for various kinds of cancers, such as liver, prostate, stomach, and breast cancers [178,179,180,181].
Sanna and coworkers developed PLGA and PLGA–PEG nanoparticles loaded with (-)-epigallocatechin-3-Gallate for targeting prostatic cancer by nanoprecipitation method [182]. All prepared batches were homogeneously dispersed with mean diameters varying from 130 nm to 250 nm and negatively charged surface charges (ranging from −32.1  ±  8.80 to −21.3  ±  8.35 mV). In vitro, nanoparticles showed an enhanced anti-proliferative activity against PCa cell lines and apoptosis modulation when compared to the free drug. In vivo, functionalization of PLGA nanoparticles with low molecular weight molecules led to specific targeting of prostate-specific membrane antigen and improved anticancer activity in the mouse xenograft model of the prostatic tumor.
Alserihi and coworkers developed folic acid anchored PLGA–PEG nanoparticles loaded with (-)-epigallocatechin-3-Gallate by nanoprecipitation for prostatic cancer treatment [183]. The prepared nanoparticles were 128.42 ± 3.57 nm and 0.114 ± 0.03 and 49.82 ± 3.74% in particle size, polydispersity index, and encapsulation efficiency, respectively. Results also exhibited a significant enhancement in the antiproliferative activity against prostate cancer cell lines, particularly toward the prostate-specific membrane antigen, and an increase in the number of dead apoptotic cells in 22Rv1 spheroids. Shao and coworkers developed nanofiber meshes by electrospinning the PCL loaded with green tea polyphenol to increase the efficacy against different cancerous lines [184]. The prepared nanoparticles showed a controlled release pattern and longer stability of green tea polyphenol. It also showed significant inhibitory effects against HepG2 and A549.
Young-Jin and coworkers developed PCL nanofibers loaded with (-)-epigallocatechin-3-gallate and caffeic acid to increase the efficacy against gastric cancer cell lines (MKN-28). The prepared nanoparticles showed apoptosis via activated caspase-3 due to the H2O2 generation, which offered a suitable system for long-term cancer therapy [185]. Safer and coworkers fabricated chitosan nanoparticles loaded with (-)-epigallocatechin-3-Gallate by ionic gelation method in order to develop a nano-therapeutic delivery system against hepatic fibrosis [186]. The prepared nanoparticles were stable, positively charged, in the nanorange, and had isoelectric pH of 7.61. Moreno-Vásquez and coworkers fabricated (-)-epigallocatechin-3-gallate grafted-chitosan nanoparticles by nanoprecipitation method in order to maximize antibacterial and antioxidant potential [187]. FTIR finding verified the modification of the chitosan with (-)-epigallocatechin-3-gallate where nanoparticles fabricated offered lower minimal inhibitory concentration when compared to chitosan and (-)-epigallocatechin-3-gallate against Pseudomonas fluorescens. In addition, nanoparticles produced a significant enhancement of reactive oxygen species (p < 0.05) in two bacterial species and an increase in antioxidant activity. Hong and coworkers developed (-)-epigallocatechin-3-gallate-loaded self-assembled nanoparticles of chitosan and aspartic acid in order to improve its action against atherosclerosis [188], which was found to be comparable with simvastatin administration in the percentage of lipid deposition. Liang and coworkers developed (-)-epigallocatechin-3-gallate-loaded chitosan/β-Lactoglobulin nanoparticles of chitosan and aspartic acid in order to achieve a prolonged release of epigallocatechin-3-gallate in the gastrointestinal tract [189].

4.6. Silymarin

Silymarin is an active extract from milk thistle seeds that consist of 65–80% silymarin flavonolignans. It has been used as a natural medication to protect the liver and treat several liver diseases, such as hepatitis and cirrhosis. Pharmacologically, it has antioxidant, antilipid peroxidative, antifibrotic, and anti-inflammatory actions. Additionally, it has been widely studied as a chemopreventive agent against different cancers [190,191,192]. Snima and coworkers fabricated PLGA nanoparticles loaded with silymarin by single-step emulsion technique to improve the anticancer application in vitro. Nearly 60% of silymarin was entrapped into PLGA nanoparticles, which produced time- and dose-dependent cytotoxic effects against prostate cancer cells (PC-3) [193]. Azadpour and coworkers fabricated PLGA nanoparticles loaded with silymarin by the emulsification/solvent evaporation method to improve the anti-inflammatory effect against sepsis and septic shock [194]. Results showed that nanoparticles decreased pro-inflammatory cytokines and increased anti-inflammatory cytokine and M2-associated markers and proteins. Elsherbiny and coworkers fabricated pH-responsive alginate–PLGA nano/micro hydrogel nanoparticles loaded with silymarin to develop a biodegradable and controlled-release oral drug delivery system for silymarin [195]. Results showed that the developed alginate-based hydrogel microparticles encapsulating silymarin-loaded PLGA nanoparticles can improve the dissolution and oral bioavailability of silymarin.

4.7. Saponins

Saponins are a group of naturally derived plant glycosides that demonstrated considerable cytotoxic activity [196,197,198]. According to their origins, there are different types, such as Astragalus saponin, Ginseng saponin, and saponin extracts from the Quillaja saponaria Molina. Sanoj and coworkers developed saponin-loaded chitosan nanoparticles by ionic gelation as a biocompatible and biodegradable system to control and sustain delivery to prostate cancer cell lines. The cellular internalization studies showed a promising uptake by L929 and PC3 cell lines with potential selective cytotoxic activity against cancerous [26]. Van de Ven and coworkers developed saponin-loaded chitosan nanoparticles in order to improve its release and anticancer activity [26]. Results showed that nanoparticles have a particle size of 65 ± 7 nm with improved cytotoxicity on PC3 and KB cell lines and cellular uptake by L929 and PC3. Saharan and coworkers developed saponin β-aescin loaded PLGA nanoparticles in order to improve drug delivery in Leishmania-infected macrophages [199]. Nanoparticles were prepared by combined emulsification solvent evaporation/salting-out technique. Results showed that nanoparticles have a particle size of less than 300 nm with negative zeta potential. Additionally, they were up-taken by the macrophages and trafficked towards the lysosomes.

4.8. Oridonin

Oridonin is a natural diterpenoid that was reported to have several potential therapeutic actions against cancerous cells, including induction of apoptosis in lymphoid malignancies, inhibition of Nuclear factor-kB (NF-kB ), and down-regulation of the Bcl-2 family proteins [200]. Unfortunately, the clinical applications of oridonin are really limited owing to its poor aqueous solubility and narrow therapeutic index. Xing and coworkers developed oridonin-loaded PLA nanoparticles by an emulsion solvent diffusion method in order to improve its plasma residence period and drug delivery [201]. The prepared nanoparticles showed a good pharmacokinetic profile; particularly in the long period, the drug stayed in circulation. It also showed significant deposition of oridonin in spleen, liver and lung, while its renal and cardiac concentrations were noticeably dropped.

4.9. Paclitaxel

Paclitaxel is a natural product isolated from the bark of Pacific Yew (Taxus brevifolia). It is widely used as a chemotherapeutic agent in the treatment of different types of cancers, such as advanced ovarian, lung, and breast cancers [202,203]. The mechanism by which paclitaxel acts is the stabilization of microtubules, blocking the cancer cell cycle in the G2/M phase, and finally, the induction of apoptosis [204]. The major obstacle to paclitaxel delivery is its poor aqueous water solubility (~0.4 μg/mL). However, this drawback was overcome in the commercial product (Taxol®) by dissolving paclitaxel in Cremophor EL/ethanol (1/1, v/v). Nevertheless, Cremophor EL was reported to have various severe adverse reactions, including hypersensitivity, cardiotoxicity, and peripheral neuropathy [205].
Danhier and coworkers developed paclitaxel-loaded PEGylated PLGA nanoparticles by nanoprecipitation in order to develop Cremophor EL-free i.v. formulation to enhance the safety and therapeutic index of paclitaxel [206]. The prepared nanoparticles showed biphasic release pattern, significant in vitro anticancer efficacy against human cervix carcinoma cells (HeLa), and lower IC50 (5.5 µg/mL) when compared to commercial product (15.5 µg/mL), concentration-dependent cellular uptake and comparable apoptosis. In vivo, nanoparticles presented superior tumor growth inhibition outcomes on TLT tumors.
Fonseca and coworkers [207] developed paclitaxel-loaded PLGA nanoparticles by the interfacial deposition method in order to achieve the same target as the previous study. Results showed that the prepared nanoparticles were less than 200 nm in particle size, mono-dispersed, and carried a negative charge. The prepared nanoparticles showed a biphasic release pattern and significant in vitro anticancer efficacy against small cell lung cancer cell line (NCI-H69 SCLC). Su and coworkers [208] developed paclitaxel-loaded cholic acid-functionalized star-shaped PLGA-d-α-tocopheryl polyethylene glycol 1000 succinate nanoparticles by modified nanoprecipitation method in order to treat malignant melanoma. Authors reported that nanoparticle formulation overweighed commercial products in treating malignant melanoma in terms of in vitro cell toxicity and xenograft tumor model in vivo. Cui and coworkers [209] developed magnetic PLGA nanoparticles co-loaded with paclitaxel and curcumin by modified nanoprecipitation method in order to cross the blood–brain barrier effectively and treat glioma. A dual-targeting approach was utilized by a combination of magnetic guidance and active targeting transferrin receptor-binding peptide T7-mediated nanoparticles. The resultant nanoparticles resulted in a more than 10-time increase in cellular uptake studies, a more than 5-time enhancement in brain delivery compared to the non-targeting nanoparticles and high survival rate in mice bearing orthotopic glioma.
Abriata and coworkers [210] developed PCL nanoparticles loaded with paclitaxel by nanoprecipitation method for evaluation of the activity against ovarian cancer in vitro. The results showed that nanoparticles had a mono-modal particle distribution with a size of around 140 nm, negatively charged particles, and high entrapment efficiency. They also presented a high efficiency in reducing the cell viability of SKOV-3 cells. Bernabeu and coworkers [211] developed PCL-alpha tocopheryl polyethylene glycol 1000 succinate (TPGS) nanoparticles loaded with paclitaxel to fabricate long-circulating safe nanoparticles and evaluate the activity against breast cancer in vitro. The authors prepared nanoparticles by three different methods (nanoprecipitation, emulsion–solvent evaporation/homogenization, and emulsion–solvent evaporation/homogenization). The former method yielded smaller particles and higher encapsulation of paclitaxel when compared to other methods. The results presented a better efficiency in reducing cell viability of two human breast cancer cell lines (MCF-7 and MDA-MB-231) when compared to free paclitaxel solution and the commercial product Abraxane®). Gupta and coworkers [212] developed chitosan nanoparticles loaded with paclitaxel to achieve safe and effective delivery. The authors prepared nanoparticles by w/o nanoemulsion method, and the yielded formulation was of 226.7 ± 0.70 nm particle size, 0.345 ± 0.039 polydispersity index, 37.4 ± 0.77 mV surface charge, 79.24 ± 2.95% encapsulation efficiency, and 11.57 ± 0.81% loading capacity. The results showed promising efficiency in reducing cell viability of human breast cancer cell line MDA-MB-231, a reduced IC50, and extreme biocompatibility.

5. Polymer Functionalization and Surface Decoration

The functionalization of polymeric nanoparticles has garnered much attention in previous years through different methods of polymeric modification. The surface decoration, chemical synthesis, and the selection of the nanoparticles are completely dependent on the nature of the actives, the required duration of action, stability, permeability, the release of the actives, etc. [213]. However, the aims of polymeric functionalization and surface decoration depend on the objective of the work. It was reported that the functionalized polymeric nanoparticles are mostly classified into four generations; long circulating “stealth” nanoparticles, lectin-based polymeric nanoparticles, polysaccharide-based polymeric nanoparticles, and, lately, ligand-based polymeric nanoparticles [214]. Generally, they aim to prolong the release and plasma residence time, specifically deliver the actives by active targeting, enable nanoparticles to withstand harsh environments, and improve physicochemical properties and biocompatibility. On the other hand, researchers should take the following criteria into consideration during working on the functionalization of polymeric nanoparticles [215]:
  • Determination of the number of conjugate sites to calculate the bio-molecule ratio appropriately.
  • Avoidance of non-specific conjugation.
  • Adjustment of polymer/conjugate affinity.
  • Keeping the proper effectiveness of the physiochemical properties of the yielded polymer/conjugate.
  • High reproducibility of the method.
There are two main techniques used for the functionalization of the polymers: direct functionalization and post-functionalization. In the first technique, a reactive group is attached to the nanoparticle surface from one side and to the required active group from the other side. Frequently, polymers have been functionalized with different groups, such as thiols, disulfides, and amines [216,217,218].
Coester and coworkers fabricated biofunctionalized gelatin nanoparticles with sulfhydryl groups to improve the entrapment of biotinylated substances because the sulfhydryl group was covalently linked with proteins such as avidin [219]. Lectins are essential to and easily recognized by several cells. Therefore, they display selective non-covalent interaction with carbohydrates. However, they can be incorporated into polymeric nanoparticles by adsorption or covalent linkage. Rodrigues and coworkers fabricated biofunctionalized PCL-based nanoparticles as lectin-decorated or protein-loaded nanoparticles with PCL as a hydrophobic core and dextran as a hydrophilic corona [220].
Lectins were bovine serum albumin, and those obtained from Bauhinia monandra, Lens culinaris leaves, and the prepared nanoparticles were of nearly 200 nm particle size. They also conserved lectins’ hemagglutinating activity and enhanced Caco-2-cellular viability, proposing the likely use of this modified nanoparticle type for targeted oral delivery. In the second method, a similar strategy is followed, and the nature of the functionalizing moiety may not be compatible with good control over the size and dispersion in the vehicle used during the preparation [215]. As most biodegradable polymers are hydrophobic in nature, they are easily recognized by the mononuclear phagocytic system to remove them from the blood stream before reaching the target. However, the fabrication of nanoparticles of particle size less than 100 nm was reported to escape from mononuclear phagocytic system recognition [221]. This strategy is not usually guaranteed for most polymeric nanoparticles. Thus, surface coating with a hydrophilic polymer (such as polyethylene glycol) will offer the chance for nanoparticles to escape from mononuclear phagocytic system recognition by forming a hydrophilic shielding layer surrounding the nanoparticles, which enables deterring the attraction of opsonin proteins by steric repulsive forces, thus closing and suspending the first stage in the opsonization process [222]. In addition, and depending on the anchored polymer, surface modification can improve the targeting of tumors and resistance to GIT-harsh conditions. Hydrophilic polymers such as polyethylene glycol, Pluronics, Tween 80, TPGS, and Tween 20, and polysaccharides such as dextran are frequently used to modify the surface of conventional nanoparticles [223].
Surface coverage by hydrophilic polymers can be achieved by surface adsorption or by the use of block or branched copolymers [224,225]. Polyethylene glycol is a hydrophilic polymer and is commonly used for coating the surface of different polymeric nanoparticles [226] due to its good biocompatibility. It can also prolong the residence time of nanoparticles in plasma efficiently due to escaping from mononuclear phagocytic system recognition. Kim and coworkers fabricated methoxy polyethylene glycol surface decorated PCL nanoparticles [227]. Calvo and coworkers fabricated polyethylene glycol surface decorated polyhexadecyl cyanoacrylate nanoparticles to minimize the opsonization and recognition by macrophages and, thus, prolong the plasma half-life. PEGylated nanoparticles presented reasonably higher deposition in the spleen and brain than conventional non-PEGylated nanoparticles [228]. Tobio and coworkers fabricated a polyethylene glycol surface decorated PLA nanoparticles for oral administration to improve the stability and decrease the influence of GIT on the digestive enzymes [226]. However, the selection of the more relevant PEG chain is not an easy process as various brands of PEG chains possess dissimilar physicochemical properties (such as molecular weight, chain length, and several branch arms), which could influence the efficiency of the PEG chain. It was reported that PEG with a molecular weight of less than 2000 Da was not efficient enough to prevent protein adsorption and immuno-recognition [222,229,230]. On the other hand, no extra benefit was achieved upon using a PEG chain with a molecular weight higher than 5000 Da [231]. Pluronics (Poloxamers) were reported to improve plasma residence time and deposition in tumors of polyethylene oxide–PCL nanoparticles [232]. Tweens-coated nanoparticles were reported to adsorb certain compounds from the blood by endothelial cells facilitating the transport through the blood–brain barrier [233,234,235]. Dextran was also reported to prohibit protein adsorption of PCL nanoparticles [236]. TPGS was reported to provide milder environments to proteins when compared to uncoated PLA nanoparticles [237].

6. Clinical Trials

As polymeric nanoparticles have garnered much attention as an efficient drug delivery system, clinical trials using polymeric nanoparticles loaded with natural bioactive agents have been conducted. In addition, some formulations have been approved by Food and Drug administration (FDA). Paclitaxel albumin-stabilized nanoparticle has completed Phase II clinical trials for the treatment of ovarian cancer (NCT00499252 at https://clinicaltrials.gov/ and updated on 11 January 2018). On the other hand, two commercial products are now available in the market. Genexol PM® are paclitaxel-loaded PEG-poly(D,L-lactide) nanoparticles at a dose of 30 mg and have been approved in South Korea (2007) for the treatment of breast cancer, though it is still under phase II clinical trials in the USA [238]. Paclical® are paclitaxel-loaded poly (glutamic acid) nanoparticles for the treatment of ovarian cancer [239].

7. Challenges

Generally, using nanoparticles for drug delivery might be considered a risk as the particle size reduction is correlated to improved reactivity and toxicity [240]. However, other different factors can participate in such behavior, including morphology, chemical composition, hydrophobicity/hydrophilicity, and the surface charge of nanoparticles. Gatoo and coworkers studied the link between the physicochemical characteristics of nanoparticles and their induced toxicity. They found that the smaller the size of the particles, the higher the risk of potential toxicity, owing to their potential capability to traverse biological membranes and attain various tissues without being recognized by the reticuloendothelial system [241]. In particular, the safety/toxicity of polymeric nanoparticles is considered one of the major challenges. The toxicity of polymeric nanoparticles is attributable to the toxic monomer aggregation, homocompatibility, residual material attached to them, and toxic degradation products [242].
However, the toxicity of polymeric nanoparticles is particularly influenced by the quantum size, which is associated with oxidative stress, cytotoxicity, and genotoxicity [243]. In addition, the polymeric material using physicochemical properties of nanoparticle biodegradability and biocompatibility are also causal factors. Generally, the toxicological effects associated with natural polymer-based nanoparticles were established to be less than that associated with the administration of synthetic polymer-based nanoparticles [244]. In addition, using natural polymer blends or coating synthetic polymeric nanoparticles with natural polymer was reported to alleviate the toxicity [245,246,247,248,249]. The particle size of nanoparticles can also influence the manner of endocytosis, uptake, and efficacy of treatment in the endocytic pathway [250,251]. Commonly, as the size decreases, the opportunity for interaction with the biological membranes increases [252,253]. It was established that when particle size is bigger than 100 nm, particles can enter the cells via phagocytosis, whereas when the particle size is less than 100 nm, endocytosis is the main route of entry, leading to significant higher toxicity risk [254,255]. On the other hand, the incorporation of stabilizers (such as polyvinyl alcohol and Pluronic) was documented to exaggerate the toxicity of PLGA-based nanoparticles especially when used in high concentrations [256]. The surface charge might also have an impact in that regard. Regardless of their compositions, polymeric nanoparticles carrying positive charges are subjected to higher phagocytic uptake than negatively charged ones resulting in higher toxicological aspects [257]. Das and coworkers studied the differential pulmonary toxicity of positively and negatively charged PEG–PLA nanoparticles in vivo. Positively charged nanoparticles were reported to induce higher local and systemic toxicities [258]. Bancos and coworkers studied the effect of silica nanoparticles on macrophages and found that phagocytosis decreased and cytokine production increased [259]. Gonzalez and coworkers found that silica nanoparticles had cytotoxic effects on lung carcinoma cells [260]. Genotoxic effects on lung cells were associated with silica nanoparticle administration, which was reported by Maser and coworkers in a dose-dependent manner [261]. The production of reactive oxygen species was also one of the toxicological features of polymeric nanoparticles. Production of reactive oxygen species was reported in a dose-dependent manner upon administration of PLGA nanoparticles [256,262]. On the contrary, Patel and coworkers found no production of reactive oxygen species upon administration of PLGA nanoparticles [263]. Xifei and coworkers studied the effect of silica nanoparticles on oxidative stress and found the overproduction of reactive oxygen species with the exhaustion of glutathione at intracellular levels [264]. Immune reactions and the overproduction of inflammatory mediators are considered one of the most common toxicological aspects associated with polymeric nanoparticle administration. PLGA-polyethylene oxide nanoparticles were reported to increase GM–CSF levels (pro-inflammatory mediator) [265].
In addition, using PEG as a hydrophilic coating polymer has recently shown some demerits. It was reported that the PEG chain stimulated hypersensitivity reactions. It was also found that it was not biodegradable, which hurdles renal excretion and leads to the deposition of hepatic and lysosomes of healthy tissues [266]. Another drawback was an immune response by detecting anti-PEG antibodies (reaching 25% of the population) [267]. Additionally, repeated intravenous administration of PEGylated formulations was reported to induce splenic anti-PEG IgM, which deprived the formulation of its advantageous long circulation by “accelerated blood clearance” phenomenon [268].
Furthermore, scale-up production and continuous manufacturing are other challenges facing polymeric nanoparticles [269].

8. Recent Advances

Lipid–polymer hybrid nanocarriers have been recently used as alternatives to polymeric nanoparticles. They are hybrid systems containing both lipid and polymeric components. So, the system collects the benefits of the lipid-based nanoparticles, particularly biomimetic and biocompatibility natures, along with the merits of polymeric nanoparticles, particularly controlled release behavior and targeting [270]. The outer shell is represented by the lipid, while the inner core is represented by the polymer. The system offers a high payload [271], controlled release pattern, non-immunogenicity [272], and capability of surface decoration by desired ligands [273]. Quirós-Fallas and coworkers developed curcumin-loaded hybrid lipid polymeric nanoparticles and demonstrated efficient antioxidant activity, immune–stimulant response, and cytotoxic effect against both gastric and colon cancer cell lines [274]. Naziris and coworkers developed curcumin-loaded mixed lipid–polymer nanoparticles and found higher interaction of the system with bacterial membrane proteins leading to an efficient antibacterial effect [275]. Abosabaa and coworkers developed green tea extract-loaded hybrid chitosan–lipid nanoparticles. Results showed higher skin permeability and deposition of the system when compared to the solution. In addition, the formula showed promising anti-cellulite activity and reduction of adipocyte perimeter and fat layer thickness of the skin [276]. Patel and coworkers developed hybrid chitosan–lipid nanoparticles co-loaded with mycophenolic acid and quercetin. In vivo results showed higher half-life (2.17-fold when compared to free mycophenolic acid) and promising anti-breast cancer activity [277].
Stimuli-responsive polymeric nanosystem is another system that has garnered much attention [278]. The system has been efficiently utilized in controlled drug delivery, particularly in cancer therapy [279]. However, the stimuli may be internal (patho-physiological/patho-chemical condition), such as pH [280], reduced glutathione [281], and protease [282], or external (physical), such as light [283], temperature [284], and ultrasound [285]. Regarding internal stimuli, tumor tissues are manifested by a slightly low pH [286], a high amount of reduced glutathione [287], and the overexpression of protease enzyme [288]. On the other hand, near-infrared light [289], ultrasound [290], and temperature [291] were reported to control the release from some polymeric materials. Chen and coworkers developed dual hydrogen peroxide (as an extracellular trigger)/glutathione (as an intracellular trigger)-responsive thioketal polymeric nanoparticles for programmable paclitaxel release at the tumor site [292]. Among stimuli-responsive polymeric nanosystems, nanogel-based polymeric nanoparticles have established a high capability to uptake water in 3D nano-polymeric networks. Modern findings have revealed that the disruption of nanogels can be performed by using cross-linking agents sensible to temperature, light, and change in pH. Consequently, the loaded material can reach intracellular targets after the endocytosis of nanogel [293,294]. Howaili and coworkers developed curcumin-loaded dual thermo/pH-responsive plasmonic nanogel by grafting poly (N-isopropyl acrylamide) to chitosan with further incorporation into gold nanoparticles to provide simultaneous delivery and photothermal therapy of cancer. The results showed an enhanced chemotherapy efficacy when irradiated with near-infrared laser in vitro [295].

9. Perspectives and Conclusions

Throughout the whole world, attention has become increasingly directed toward the use of natural bioactive agents owing to their excellent therapeutic and preventive benefits. They possess several actions, such as antioxidant, anti-inflammatory, antimicrobial, and chemopreventive actions, which are not solely sufficient to provide to a patient. Thus, finding relevant formulation/technology to deliver them through different routes is the biggest challenge due to their drawbacks, such as poor solubility, poor bioavailability, and instability. At present, a wide range of polymeric nanoparticles (naturally or synthetically based) has been found to be relevant for the encapsulation or loading of different natural bioactive agents in order to enhance their bioavailability and therapeutic efficacy. Though polymeric nanoparticles were recognized by some researchers for their toxicity (particularly in high concentrations), surface functionalization is considered a key factor that can overcome such problems. In addition, using a lipid–polymer hybrid system and stimuli-responsive nanocarrier has opened a new horizon for more safe and site-specific drug release. So, the future of such new polymeric nanoparticles will enable scientists to provide safe and smart delivery of natural bioactive agents. Overall, the use of natural bioactive agent-loaded polymeric nanoparticles is a promising strategy, provided that the platform used fulfills the therapeutic and safety aspects.

Author Contributions

Conceptualisation, M.E., H.A.M.G., M.A.A. and E.M.M.; methodology, M.E.; writing—original draft preparation, M.E.; data curation, M.E., K.S. and O.A.A.; supervision, M.E.; review and editing, M.H.E.; visualisation, K.S. and O.A.A.; management, H.A.M.G., M.A.A. and E.M.M. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the Deanship of Scientific Research at Jouf University under Grant Number (DSR2022-RG-0142).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Authors acknowledge the financial support by the Deanship of Scientific Research at Jouf University. The authors also would like to express their gratitude to Alfred Fahr for his kind assistance and guidance.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Saxena, M.; Saxena, J.; Nema, R.; Singh, D.; Gupta, A. Phytochemistry of medicinal plants. J. Pharmacogn. Phytochem. 2013, 1, 6. [Google Scholar]
  2. Aqil, F.; Munagala, R.; Jeyabalan, J.; Vadhanam, M.V. Bioavailability of phytochemicals and its enhancement by drug delivery systems. Cancer Lett. 2013, 334, 133–141. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Jain, A.K.; Thareja, S. In vitro and in vivo characterization of pharmaceutical nanocarriers used for drug delivery. Artif. Cells Nanomed. Biotechnol. 2019, 47, 524–539. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Prabhu, H.R.; Patravale, V.B.; Joshi, M.D. Polymeric nanoparticles for targeted treatment in oncology: Current insights. Int. J. Nanomed. 2015, 10, 1001–1018. [Google Scholar]
  5. Gref, R.; Minamitake, Y.; Peracchia, M.T.; Trubetskoy, V.; Torchilin, V.; Langer, R. Biodegradable Long-Circulating Polymeric Nanospheres. Science 1994, 263, 1600–1603. [Google Scholar] [CrossRef] [Green Version]
  6. Borgheti-Cardoso, L.N.; Viegas, J.S.R.; Silvestrini, A.V.P.; Caron, A.L.; Praça, F.G.; Kravicz, M.; Bentley, M.V.L.B. Nanotechnology approaches in the current therapy of skin cancer. Adv. Drug Deliv. Rev. 2020, 153, 109–136. [Google Scholar] [CrossRef]
  7. Sinha, V.R.; Bansal, K.; Kaushik, R.; Kumria, R.; Trehan, A. Poly-ε-caprolactone microspheres and nanospheres: An overview. Int. J. Pharm. 2004, 278, 1–23. [Google Scholar] [CrossRef]
  8. Sabra, R.; Roberts, C.J.; Billa, N. Courier properties of modified citrus pectinate-chitosan nanoparticles in colon delivery of curcumin. Colloid Interface Sci. Commun. 2019, 32, 100192. [Google Scholar] [CrossRef]
  9. Kumari, A.; Yadav, S.K.; Yadav, S.C. Biodegradable polymeric nanoparticles based drug delivery systems. Colloids Surf. B Biointerfaces 2010, 75, 1–18. [Google Scholar] [CrossRef]
  10. Danhier, F.; Ansorena, E.; Silva, J.M.; Coco, R.; Le Breton, A.; Préat, V. PLGA-based nanoparticles: An overview of biomedical applications. J. Control. Release 2012, 161, 505–522. [Google Scholar] [CrossRef]
  11. Nan, W.; Ding, L.; Chen, H.; Khan, F.U.; Yu, L.; Sui, X.; Shi, X. Topical Use of Quercetin-Loaded Chitosan Nanoparticles against Ultraviolet B Radiation. Front. Pharmacol. 2018, 9, 826. [Google Scholar] [CrossRef]
  12. De Souza, G.M.; Gervasoni, L.F.; da Silva Rosa, R.; de Souza Iacia, M.V.M.; Nai, G.A.; Pereira, V.C.; Winkelströter, L.K. Quercetin-loaded chitosan nanoparticles as an alternative for controlling bacterial adhesion to urethral catheter. Int. J. Urol. 2022, 29, 1228–1234. [Google Scholar] [CrossRef] [PubMed]
  13. Jardim, K.V.; Siqueira, J.L.N.; Báo, S.N.; Parize, A.L. In vitro cytotoxic and antioxidant evaluation of quercetin loaded in ionic cross-linked chitosan nanoparticles. J. Drug Deliv. Sci. Technol. 2022, 74, 103561. [Google Scholar] [CrossRef]
  14. Choudhary, A.; Kant, V.; Jangir, B.L.; Joshi, V. Quercetin loaded chitosan tripolyphosphate nanoparticles accelerated cutaneous wound healing in Wistar rats. Eur. J. Pharmacol. 2020, 880, 173172. [Google Scholar] [CrossRef]
  15. Kabiriyel, J.; Jeyanthi, R.; Jayakumar, K.; Amalraj, A.; Arjun, P.; Shanmugarathinam, A.; Vignesh, G.; Mohan, C.R. Green synthesis of carboxy methyl chitosan based curcumin nanoparticles and its biological activity: Influence of size and conductivity. Carbohydr. Polym. Technol. Appl. 2023, 5, 100260. [Google Scholar] [CrossRef]
  16. Prasad, M.; Salar, A.; Salar, R.K. In vitro anticancer activity of curcumin loaded chitosan nanoparticles (CLCNPs) against Vero cells. Pharmacol. Res. Mod. Chin. Med. 2022, 3, 100116. [Google Scholar] [CrossRef]
  17. Ali, K.A.; El-Naa, M.M.; Bakr, A.F.; Mahmoud, M.Y.; Abdelgawad, E.M.; Matoock, M.Y. The dual gastro- and neuroprotective effects of curcumin loaded chitosan nanoparticles against cold restraint stress in rats. Biomed. Pharmacother. 2022, 148, 112778. [Google Scholar] [CrossRef]
  18. Sorasitthiyanukarn, F.N.; Muangnoi, C.; Thaweesest, W.; Rojsitthisak, P.; Rojsitthisak, P. Enhanced cytotoxic, antioxidant and anti-inflammatory activities of curcumin diethyl disuccinate using chitosan-tripolyphosphate nanoparticles. J. Drug Deliv. Sci. Technol. 2019, 53, 101118. [Google Scholar] [CrossRef]
  19. Duan, J.; Zhang, Y.; Han, S.; Chen, Y.; Li, B.; Liao, M.; Chen, W.; Deng, X.; Zhao, J.; Huang, B. Synthesis and in vitro/in vivo anti-cancer evaluation of curcumin-loaded chitosan/poly(butyl cyanoacrylate) nanoparticles. Int. J. Pharm. 2010, 400, 211–220. [Google Scholar] [CrossRef]
  20. Saha, M.; Saha, D.R.; Ulhosna, T.; Sharker, S.M.; Shohag, H.; Islam, M.S.; Ray, S.K.; Rahman, G.S.; Reza, H.M. QbD based development of resveratrol-loaded mucoadhesive lecithin/chitosan nanoparticles for prolonged ocular drug delivery. J. Drug Deliv. Sci. Technol. 2021, 63, 102480. [Google Scholar] [CrossRef]
  21. Chung, J.H.; Lee, J.-S.; Lee, H.G. Resveratrol-loaded chitosan–γ-poly (glutamic acid) nanoparticles: Optimization, solubility, UV stability, and cellular antioxidant activity. Colloids Surf. B Biointerfaces 2020, 186, 110702. [Google Scholar] [CrossRef] [PubMed]
  22. Ramalingam, P.; Ko, Y.T. Improved oral delivery of resveratrol from N-trimethyl chitosan-g-palmitic acid surface-modified solid lipid nanoparticles. Colloids Surf. B Biointerfaces 2016, 139, 52–61. [Google Scholar] [CrossRef] [PubMed]
  23. Bu, L.; Gan, L.-C.; Guo, X.-Q.; Chen, F.-Z.; Song, Q.; Zhao, Q.; Gou, X.-J.; Hou, S.-X.; Yao, Q. Trans-resveratrol loaded chitosan nanoparticles modified with biotin and avidin to target hepatic carcinoma. Int. J. Pharm. 2013, 452, 355–362. [Google Scholar] [CrossRef] [PubMed]
  24. Scalia, S.; Trotta, V.; Iannuccelli, V.; Bianchi, A. Enhancement of in vivo human skin penetration of resveratrol by chitosan-coated lipid microparticles. Colloids Surf. B Biointerfaces 2015, 135, 42–49. [Google Scholar] [CrossRef] [PubMed]
  25. Moghaddam, A.H.; Sangdehi, S.R.M.; Ranjbar, M.; Hasantabar, V. Preventive effect of silymarin-loaded chitosan nanoparticles against global cerebral ischemia/reperfusion injury in rats. Eur. J. Pharmacol. 2020, 877, 173066. [Google Scholar] [CrossRef]
  26. Rejinold, S.; Muthunarayanan, M.; Muthuchelian, K.; Chennazhi, K.P.; Nair, S.V.; Jayakumar, R. Saponin-loaded chitosan nanoparticles and their cytotoxicity to cancer cell lines in vitro. Carbohydr. Polym. 2011, 84, 407–416. [Google Scholar] [CrossRef]
  27. Xu, J.; Ma, L.; Liu, Y.; Xu, F.; Nie, J.; Ma, G. Design and characterization of antitumor drug paclitaxel-loaded chitosan nanoparticles by W/O emulsions. Int. J. Biol. Macromol. 2012, 50, 438–443. [Google Scholar] [CrossRef]
  28. Mondéjar-López, M.; Rubio-Moraga, A.; López-Jimenez, A.J.; Martínez, J.C.G.; Ahrazem, O.; Gómez-Gómez, L.; Niza, E. Chitosan nanoparticles loaded with garlic essential oil: A new alternative to tebuconazole as seed dressing agent. Carbohydr. Polym. 2021, 277, 118815. [Google Scholar] [CrossRef]
  29. Hadidi, M.; Pouramin, S.; Adinepour, F.; Haghani, S.; Jafari, S.M. Chitosan nanoparticles loaded with clove essential oil: Characterization, antioxidant and antibacterial activities. Carbohydr. Polym. 2020, 236, 116075. [Google Scholar] [CrossRef]
  30. Zhao, F.-Q.; Wang, G.-F.; Xu, D.; Zhang, H.-Y.; Cui, Y.-L.; Wang, Q.-S. Glycyrrhizin mediated liver-targeted alginate nanogels delivers quercetin to relieve acute liver failure. Int. J. Biol. Macromol. 2021, 168, 93–104. [Google Scholar] [CrossRef]
  31. Zakerikhoob, M.; Abbasi, S.; Yousefi, G.; Mokhtari, M.; Noorbakhsh, M.S. Curcumin-incorporated crosslinked sodium alginate-g-poly (N-isopropyl acrylamide) thermo-responsive hydrogel as an in-situ forming injectable dressing for wound healing: In vitro characterization and in vivo evaluation. Carbohydr. Polym. 2021, 271, 118434. [Google Scholar] [CrossRef] [PubMed]
  32. Wang, Y.; Li, Y.; He, L.; Mao, B.; Chen, S.; Martinez, V.; Guo, X.; Shen, X.; Liu, B.; Li, C. Commensal flora triggered target anti-inflammation of alginate-curcumin micelle for ulcerative colitis treatment. Colloids Surf. B Biointerfaces 2021, 203, 111756. [Google Scholar] [CrossRef] [PubMed]
  33. Karthikeyan, S.; Prasad, N.R.; Ganamani, A.; Balamurugan, E. Anticancer activity of resveratrol-loaded gelatin nanoparticles on NCI-H460 non-small cell lung cancer cells. Biomed. Prev. Nutr. 2013, 3, 64–73. [Google Scholar] [CrossRef]
  34. Karthikeyan, S.; Hoti, S.L.; Prasad, N.R. Resveratrol loaded gelatin nanoparticles synergistically inhibits cell cycle progression and constitutive NF-kappaB activation, and induces apoptosis in non-small cell lung cancer cells. Biomed. Pharmacother. 2015, 70, 274–282. [Google Scholar] [CrossRef]
  35. Ding, Y.; Zhang, S.; Sun, Z.; Tong, Z.; Ge, Y.; Zhou, L.; Xu, Q.; Zhou, H.; Wang, W. Preclinical validation of silibinin/albumin nanoparticles as an applicable system against acute liver injury. Acta Biomater. 2022, 146, 385–395. [Google Scholar] [CrossRef]
  36. Egil, A.C.; Kesim, H.; Ustunkaya, B.; Kutlu, Ö.; Ince, G.O. Self-assembled albumin nanoparticles for redox responsive release of curcumin. J. Drug Deliv. Sci. Technol. 2022, 76, 103831. [Google Scholar] [CrossRef]
  37. Fonseca, D.P.; Khalil, N.M.; Mainardes, R.M. Bovine serum albumin-based nanoparticles containing resveratrol: Characterization and antioxidant activity. J. Drug Deliv. Sci. Technol. 2017, 39, 147–155. [Google Scholar] [CrossRef]
  38. Zhao, Y.; Cai, C.; Liu, M.; Zhao, Y.; Wu, Y.; Fan, Z.; Ding, Z.; Zhang, H.; Wang, Z.; Han, J. Drug-binding albumins forming stabilized nanoparticles for co-delivery of paclitaxel and resveratrol: In vitro/in vivo evaluation and binding properties investigation. Int. J. Biol. Macromol. 2020, 153, 873–882. [Google Scholar] [CrossRef]
  39. Ali, A.; Ahmed, S. A review on chitosan and its nanocomposites in drug delivery. Int. J. Biol. Macromol. 2018, 109, 273–286. [Google Scholar] [CrossRef]
  40. Ravi Kumar, M.N.V. A review of chitin and chitosan applications. React. Funct. Polym. 2000, 46, 1–27. [Google Scholar] [CrossRef]
  41. Bravo-Osuna, I.; Vauthier, C.; Farabollini, A.; Palmieri, G.F.; Ponchel, G. Mucoadhesion mechanism of chitosan and thiolated chitosan-poly(isobutyl cyanoacrylate) core-shell nanoparticles. Biomaterials 2007, 28, 2233–2243. [Google Scholar] [CrossRef] [PubMed]
  42. Nagpal, K.; Singh, S.K.; Mishra, D.N. Chitosan Nanoparticles: A Promising System in Novel Drug Delivery. Chem. Pharm. Bull. 2010, 58, 1423–1430. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Gonçalves, I.; Henriques, A.P.; Seabra, C.; Martins, M.C. The potential utility of chitosan micro/nanoparticles in the treatment of gastric infection. Expert Rev. Anti-Infect. Ther. 2014, 12, 981–992. [Google Scholar] [CrossRef] [PubMed]
  44. Wang, Y.; Wang, X.; Luo, G.; Dai, Y. Adsorption of bovin serum albumin (BSA) onto the magnetic chitosan nanoparticles prepared by a microemulsion system. Bioresour. Technol. 2008, 99, 3881–3884. [Google Scholar] [CrossRef]
  45. El-Shabouri, M. Positively charged nanoparticles for improving the oral bioavailability of cyclosporin-A. Int. J. Pharm. 2002, 249, 101–108. [Google Scholar] [CrossRef]
  46. Niwa, T.; Takeuchi, H.; Hino, T.; Kunou, N.; Kawashima, Y. Preparations of biodegradable nanospheres of water-soluble and insoluble drugs with D,L-lactide/glycolide copolymer by a novel spontaneous emulsification solvent diffusion method, and the drug release behavior. J. Control. Release 1993, 25, 89–98. [Google Scholar] [CrossRef]
  47. Hembram, K.C.; Prabha, S.; Chandra, R.; Ahmed, B.; Nimesh, S. Advances in preparation and characterization of chitosan nanoparticles for therapeutics. Artif. Cells Nanomed. Biotechnol. 2014, 44, 305–314. [Google Scholar] [CrossRef]
  48. Tiyaboonchai, W. Chitosan nanoparticles: A promising system for drug delivery. Naresuan Univ. J. Sci. Technol. 2013, 11, 51–66. [Google Scholar]
  49. Stokke, B.T.; Drager, K.I.; Yuguchi, Y.; Urakawa, H.; Kajiwara, K. Small-angle X-ray scattering and rheological characterization of alginate gels. Macromol. Symp. 1997, 120, 91–101. [Google Scholar] [CrossRef]
  50. Smidsrood, O.; Draget, K.I. Alginate gelation technologies. Food Colloids Proteins Lipids Polysacch. 1997, 21, 279. [Google Scholar]
  51. Nair, L.S.; Laurencin, C.T. Biodegradable polymers as biomaterials. Prog. Polym. Sci. 2007, 32, 762–798. [Google Scholar] [CrossRef]
  52. Damelin, L.H.; Fernandes, M.A.; Tiemessen, C.T. Alginate microbead-encapsulated silver complexes for selective delivery of broad-spectrum silver-based microbicides. Int. J. Antimicrob. Agents 2015, 46, 394–400. [Google Scholar] [CrossRef] [PubMed]
  53. Draget, K.I.; Skjåk-Bræk, G.; Smidsrød, O. Alginate based new materials. Int. J. Biol. Macromol. 1997, 21, 47–55. [Google Scholar] [CrossRef] [PubMed]
  54. Keenan, T.R. Gelatin. In Kirk-Othmer Encyclopedia of Chemical Technology; John Wiley & Sons: Hoboken, NJ, USA, 2000. [Google Scholar] [CrossRef]
  55. Kuijpers, A.J.; Engbers, G.H.M.; Feijen, J.; De Smedt, S.C.; Meyvis, T.K.L.; Demeester, J.; Krijgsveld, J.; Zaat, S.A.J.; Dankert, J. Characterization of the Network Structure of Carbodiimide Cross-Linked Gelatin Gels. Macromolecules 1999, 32, 3325–3333. [Google Scholar] [CrossRef]
  56. Lin, M.; Meng, S.; Zhong, W.; Cai, R.; Du, Q.; Tomasik, P. Novel drug-loaded gelatin films and their sustained-release performance. J. Biomed. Mater. Res. Part B Appl. Biomater. 2009, 90B, 939–944. [Google Scholar] [CrossRef] [PubMed]
  57. Mannaa, S.; Lakshmia, U.S.; Racharlaa, M.; Sinhab, P.; Kanthala, L.K.; Kumara, S.P.N. Bioadhesive HPMC gel containing gelatin nanoparticles for intravaginal delivery of tenofovir. J. Appl. Pharm. Sci. 2016, 6, 22–29. [Google Scholar] [CrossRef] [Green Version]
  58. Chuang, V.T.G.; Kragh-Hansen, U.; Otagiri, M. Pharmaceutical Strategies Utilizing Recombinant Human Serum Albumin. Pharm. Res. 2002, 19, 569–577. [Google Scholar] [CrossRef]
  59. Wunder, A.; Muller-Ladner, U.; Stelzer, E.; Neumann, E.; Sinn, H.; Gay, S.; Fiehn, C. Albumin-based drug delivery as novel therapeutic approach for rheumatoid arthritis. Thromb. Haemost. 2003, 5, 9. [Google Scholar] [CrossRef] [Green Version]
  60. Shen, Y.; Li, W. HA/HSA co-modified erlotinib–albumin nanoparticles for lung cancer treatment. Drug Des. Dev. Ther. 2018, 12, 2285–2292. [Google Scholar] [CrossRef] [Green Version]
  61. Yee, Y.J.; Benson, H.A.E.; Dass, C.R.; Chen, Y. Evaluation of novel conjugated resveratrol polymeric nanoparticles in reduction of plasma degradation, hepatic metabolism and its augmentation of anticancer activity in vitro and in vivo. Int. J. Pharm. 2022, 615, 121499. [Google Scholar] [CrossRef]
  62. Kamel, R.; Abbas, H.; Shaffie, N.M. Development and evaluation of PLA-coated co-micellar nanosystem of Resveratrol for the intra-articular treatment of arthritis. Int. J. Pharm. 2019, 569, 118560. [Google Scholar] [CrossRef] [PubMed]
  63. Thu, H.P.; Nam, N.H.; Quang, B.T.; Son, H.A.; Toan, N.L.; Quang, D.T. In vitro and in vivo targeting effect of folate decorated paclitaxel loaded PLA–TPGS nanoparticles. Saudi Pharm. J. 2015, 23, 683–688. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Pulkkinen, M.; Pikkarainen, J.; Wirth, T.; Tarvainen, T.; Haapa-Aho, V.; Korhonen, H.; Seppälä, J.; Järvinen, K. Three-step tumor targeting of paclitaxel using biotinylated PLA-PEG nanoparticles and avidin–biotin technology: Formulation development and in vitro anticancer activity. Eur. J. Pharm. Biopharm. 2008, 70, 66–74. [Google Scholar] [CrossRef] [PubMed]
  65. Hu, Q.; Gao, X.; Gu, G.; Kang, T.; Tu, Y.; Liu, Z.; Song, Q.; Yao, L.; Pang, Z.; Jiang, X.; et al. Glioma therapy using tumor homing and penetrating peptide-functionalized PEG–PLA nanoparticles loaded with paclitaxel. Biomaterials 2013, 34, 5640–5650. [Google Scholar] [CrossRef]
  66. Pandey, S.K.; Patel, D.K.; Thakur, R.; Mishra, D.P.; Maiti, P.; Haldar, C. Anti-cancer evaluation of quercetin embedded PLA nanoparticles synthesized by emulsified nanoprecipitation. Int. J. Biol. Macromol. 2015, 75, 521–529. [Google Scholar] [CrossRef]
  67. Kumari, A.; Yadav, S.K.; Pakade, Y.B.; Singh, B.; Yadav, S.C. Development of biodegradable nanoparticles for delivery of quercetin. Colloids Surf. B Biointerfaces 2010, 80, 184–192. [Google Scholar] [CrossRef]
  68. Kost, B.; Svyntkivska, M.; Brzeziński, M.; Makowski, T.; Piorkowska, E.; Rajkowska, K.; Kunicka-Styczyńska, A.; Biela, T. PLA/β-CD-based fibres loaded with quercetin as potential antibacterial dressing materials. Colloids Surf. B Biointerfaces 2020, 190, 110949. [Google Scholar] [CrossRef]
  69. El-Naggar, M.E.; Al-Joufi, F.; Anwar, M.; Attia, M.F.; El-Bana, M.A. Curcumin-loaded PLA-PEG copolymer nanoparticles for treatment of liver inflammation in streptozotocin-induced diabetic rats. Colloids Surf. B Biointerfaces 2019, 177, 389–398. [Google Scholar] [CrossRef]
  70. Zamboni, F.; Ren, G.; Culebras, M.; O’Driscoll, J.; O’Dwyer, J.; Ryan, E.J.; Collins, M.N. Curcumin encapsulated polylactic acid nanoparticles embedded in alginate/gelatin bioinks for in situ immunoregulation: Characterization and biological assessment. Int. J. Biol. Macromol. 2022, 221, 1218–1227. [Google Scholar] [CrossRef]
  71. Niza, E.; Božik, M.; Bravo, I.; Clemente-Casares, P.; Sánchez, A.L.; Juan, A.; Klouček, P.; Alonso-Moreno, C. PEI-coated PLA nanoparticles to enhance the antimicrobial activity of carvacrol. Food Chem. 2020, 328, 127131. [Google Scholar] [CrossRef]
  72. Fernández, K.; Aburto, J.; von Plessing, C.; Rockel, M.; Aspé, E. Factorial design optimization and characterization of poly-lactic acid (PLA) nanoparticle formation for the delivery of grape extracts. Food Chem. 2016, 207, 75–85. [Google Scholar] [CrossRef] [PubMed]
  73. Seko, I.; Tonbul, H.; Tavukçuoğlu, E.; Şahin, A.; Akbas, S.; Yanık, H.; Öztürk, S.C.; Esendagli, G.; Khan, M.; Capan, Y. Development of curcumin and docetaxel co-loaded actively targeted PLGA nanoparticles to overcome blood brain barrier. J. Drug Deliv. Sci. Technol. 2021, 66, 102867. [Google Scholar] [CrossRef]
  74. Prabhuraj, R.; Bomb, K.; Srivastava, R.; Bandyopadhyaya, R. Selection of superior targeting ligands using PEGylated PLGA nanoparticles for delivery of curcumin in the treatment of triple-negative breast cancer cells. J. Drug Deliv. Sci. Technol. 2020, 57, 101722. [Google Scholar] [CrossRef]
  75. Khalil, N.M.; do Nascimentod, T.C.F.; Casa, D.M.; Dalmolin, L.F.; de Mattos, A.C.; Hoss, I.; Romano, M.A.; Mainardes, R.M. Pharmacokinetics of curcumin-loaded PLGA and PLGA–PEG blend nanoparticles after oral administration in rats. Colloids Surf. B Biointerfaces 2013, 101, 353–360. [Google Scholar] [CrossRef]
  76. Cetin, N.; Menevse, E.; Celik, Z.E.; Ceylan, C.; Rama, S.T.; Gultekin, Y.; Tekin, T.; Sahin, A. Evaluation of burn wound healing activity of thermosensitive gel and PLGA nanoparticle formulation of quercetin in Wistar albino rats. J. Drug Deliv. Sci. Technol. 2022, 75, 103620. [Google Scholar] [CrossRef]
  77. Anwer, K.; Al-Mansoor, M.A.; Jamil, S.; Al-Shdefat, R.; Ansari, M.N.; Shakeel, F. Development and evaluation of PLGA polymer based nanoparticles of quercetin. Int. J. Biol. Macromol. 2016, 92, 213–219. [Google Scholar] [CrossRef]
  78. Zhu, X.; Zeng, X.; Zhang, X.; Cao, W.; Wang, Y.; Chen, H.; Wang, T.; Tsai, H.-I.; Zhang, R.; Chang, D.; et al. The effects of quercetin-loaded PLGA-TPGS nanoparticles on ultraviolet B-induced skin damages in vivo. Nanomed. Nanotechnol. Biol. Med. 2015, 12, 623–632. [Google Scholar] [CrossRef]
  79. Sun, D.; Li, N.; Zhang, W.; Zhao, Z.; Mou, Z.; Huang, D.; Liu, J.; Wang, W. Design of PLGA-functionalized quercetin nanoparticles for potential use in Alzheimer’s disease. Colloids Surf. B Biointerfaces 2016, 148, 116–129. [Google Scholar] [CrossRef]
  80. Tsai, I.-L.; Tsai, C.-Y.; Kuo, L.-L.; Woung, L.-C.; Ku, R.-Y.; Cheng, Y.-H. PLGA nanoparticles containing Lingzhi extracts rescue corneal epithelial cells from oxidative damage. Exp. Eye Res. 2021, 206, 108539. [Google Scholar] [CrossRef]
  81. El-Naggar, M.E.; Abdelgawad, A.M.; Abdel-Sattar, R.; Gibriel, A.A.; Hemdan, B.A. Potential antimicrobial and antibiofilm efficacy of peppermint oil nanoemulsion loaded polycaprolactone nanofibrous scaffolds. Eur. Polym. J. 2022, 184, 111782. [Google Scholar] [CrossRef]
  82. Jummes, B.; Sganzerla, W.G.; da Rosa, C.G.; Noronha, C.M.; Nunes, M.R.; Bertoldi, F.C.; Barreto, P.L.M. Antioxidant and antimicrobial poly-ε-caprolactone nanoparticles loaded with Cymbopogon martinii essential oil. Biocatal. Agric. Biotechnol. 2020, 23, 101499. [Google Scholar] [CrossRef]
  83. Dezfouli, E.A.; Kiaie, S.H.; Danafar, H.; Nomani, A.; Sadeghizadeh, M. BTN-PEG-PCL nanoparticles for targeted delivery of curcumin: In vitro and in Ovo assessment. J. Drug Deliv. Sci. Technol. 2022, 74, 103382. [Google Scholar] [CrossRef]
  84. Rathinavel, S.; Korrapati, P.S.; Kalaiselvi, P.; Dharmalingam, S. Mesoporous silica incorporated PCL/Curcumin nanofiber for wound healing application. Eur. J. Pharm. Sci. 2021, 167, 106021. [Google Scholar] [CrossRef] [PubMed]
  85. Thuy, L.T.; Kang, N.; Choi, M.; Lee, M.; Choi, J.S. Dendrimeric micelles composed of polyamidoamine dendrimer-peptide-cholesterol conjugates as drug carriers for the treatment of melanoma and bacterial infection. J. Ind. Eng. Chem. 2022, 114, 361–376. [Google Scholar] [CrossRef]
  86. Acet, Ö.; Dikici, E.; Acet, B.; Odabaşı, M.; Mijakovic, I.; Pandit, S. Inhibition of bacterial adhesion by epigallocatechin gallate attached polymeric membranes. Colloids Surf. B Biointerfaces 2023, 221, 113024. [Google Scholar] [CrossRef]
  87. Kumar, S.S.D.; Surianarayanan, M.; Vijayaraghavan, R.; Mandal, A.B.; MacFarlane, D. Curcumin loaded poly(2-hydroxyethyl methacrylate) nanoparticles from gelled ionic liquid—In vitro cytotoxicity and anti-cancer activity in SKOV-3 cells. Eur. J. Pharm. Sci. 2013, 51, 34–44. [Google Scholar] [CrossRef]
  88. Younis, N.; Shaheen, M.A.; Abdallah, M.H. Silymarin-loaded Eudragit® RS100 nanoparticles improved the ability of silymarin to resolve hepatic fibrosis in bile duct ligated rats. Biomed. Pharmacother. 2016, 81, 93–103. [Google Scholar] [CrossRef]
  89. Casalini, T.; Rossi, F.; Castrovinci, A.; Perale, G. A Perspective on Polylactic Acid-Based Polymers Use for Nanoparticles Synthesis and Applications. Front. Bioeng. Biotechnol. 2019, 7, 259. [Google Scholar] [CrossRef]
  90. Dave, V.; Yadav, R.B.; Kushwaha, K.; Yadav, S.; Sharma, S.; Agrawal, U. Lipid-polymer hybrid nanoparticles: Development & statistical optimization of norfloxacin for topical drug delivery system. Bioact. Mater. 2017, 2, 269–280. [Google Scholar] [CrossRef]
  91. Song, X.; Zhao, Y.; Wu, W.; Bi, Y.; Cai, Z.; Chen, Q.; Li, Y.; Hou, S. PLGA nanoparticles simultaneously loaded with vincristine sulfate and verapamil hydrochloride: Systematic study of particle size and drug entrapment efficiency. Int. J. Pharm. 2008, 350, 320–329. [Google Scholar] [CrossRef]
  92. Zambaux, M.; Bonneaux, F.; Gref, R.; Dellacherie, E.; Vigneron, C. Preparation and characterization of protein C-loaded PLA nanoparticles. J. Control. Release 1999, 60, 179–188. [Google Scholar] [CrossRef] [PubMed]
  93. Reis, C.P.; Neufeld, R.J.; Ribeiro, A.J.; Veiga, F. Nanoencapsulation I. Methods for preparation of drug-loaded polymeric nanoparticles. Nanomed. Nanotechnol. Biol. Med. 2006, 2, 8–21. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Sahana, D.; Mittal, G.; Bhardwaj, V.; Kumar, M. PLGA Nanoparticles for Oral Delivery of Hydrophobic Drugs: Influence of Organic Solvent on Nanoparticle Formation and Release Behavior In Vitro and In Vivo Using Estradiol as a Model Drug. J. Pharm. Sci. 2008, 97, 1530–1542. [Google Scholar] [CrossRef] [PubMed]
  95. Barichello, J.M.; Morishita, M.; Takayama, K.; Nagai, T. Encapsulation of Hydrophilic and Lipophilic Drugs in PLGA Nanoparticles by the Nanoprecipitation Method. Drug Dev. Ind. Pharm. 1999, 25, 471–476. [Google Scholar] [CrossRef]
  96. Seremeta, K.P.; Chiappetta, D.A.; Sosnik, A. Poly (ε-caprolactone), Eudragit® RS 100 and poly (ε-caprolactone)/Eudragit® RS 100 blend submicron particles for the sustained release of the antiretroviral efavirenz. Colloids Surf. B Biointerfaces 2013, 102, 441–449. [Google Scholar] [CrossRef]
  97. Woodruff, M.A.; Hutmacher, D.W. The return of a forgotten polymer—Polycaprolactone in the 21st century. Prog. Polym. Sci. 2010, 35, 1217–1256. [Google Scholar] [CrossRef] [Green Version]
  98. Karuppuswamy, P.; Venugopal, J.R.; Navaneethan, B.; Laiva, A.L.; Ramakrishna, S. Polycaprolactone nanofibers for the controlled release of tetracycline hydrochloride. Mater. Lett. 2015, 141, 180–186. [Google Scholar] [CrossRef]
  99. Bilensoy, E.; Sarisozen, C.; Esendağlı, G.; Doğan, A.L.; Aktaş, Y.; Şen, M.; Mungan, N.A. Intravesical cationic nanoparticles of chitosan and polycaprolactone for the delivery of Mitomycin C to bladder tumors. Int. J. Pharm. 2009, 371, 170–176. [Google Scholar] [CrossRef]
  100. Payyappilly, S.S.; Panja, S.; Mandal, P.; Dhara, S.; Chattopadhyay, S. Organic solvent-free low temperature method of preparation for self assembled amphiphilic poly(ε-caprolactone)–poly(ethylene glycol) block copolymer based nanocarriers for protein delivery. Colloids Surf. B Biointerfaces 2015, 135, 510–517. [Google Scholar] [CrossRef]
  101. Koleske, J.V. Blends Containing Poly (ε-caprolactone) and Related Polymers. In Polymer Blends; Elsevier: Amsterdam, The Netherlands, 1978; pp. 369–389. [Google Scholar]
  102. Sur, S.; Rathore, A.; Dave, V.; Reddy, K.R.; Chouhan, R.S.; Sadhu, V. Recent developments in functionalized polymer nanoparticles for efficient drug delivery system. Nano-Struct. Nano-Objects 2019, 20, 100397. [Google Scholar] [CrossRef]
  103. Luong, D.; Kesharwani, P.; Deshmukh, R.; Amin, M.C.I.M.; Gupta, U.; Greish, K.; Iyer, A.K. PEGylated PAMAM dendrimers: Enhancing efficacy and mitigating toxicity for effective anticancer drug and gene delivery. Acta Biomater. 2016, 43, 14–29. [Google Scholar] [CrossRef] [PubMed]
  104. Albertazzi, L.; Gherardini, L.; Brondi, M.; Sato, S.S.; Bifone, A.; Pizzorusso, T.; Ratto, G.M.; Bardi, G. In Vivo Distribution and Toxicity of PAMAM Dendrimers in the Central Nervous System Depend on Their Surface Chemistry. Mol. Pharm. 2012, 10, 249–260. [Google Scholar] [CrossRef] [PubMed]
  105. Chanphai, P.; Tajmir-Riahi, H.A. DNA binding efficacy with functionalized folic acid-PAMAM nanoparticles. Chem. Biol. Interact. 2018, 290, 52–56. [Google Scholar] [CrossRef] [PubMed]
  106. Tomalia, D.A.; Frechet, J.M.J. Discovery of dendrimers and dendritic polymers: A brief historical perspective. J. Polym. Sci. Part A Polym. Chem. 2002, 40, 2719–2728. [Google Scholar] [CrossRef]
  107. Vögtle, F.; Richardt, G.; Werner, N. Dendrimer Chemistry: Concepts, Syntheses, Properties, Applications; John Wiley & Sons: Hoboken, NJ, USA, 2009. [Google Scholar]
  108. Walter, M.V.; Malkoch, M. Simplifying the synthesis of dendrimers: Accelerated approaches. Chem. Soc. Rev. 2012, 41, 4593–4609. [Google Scholar] [CrossRef] [PubMed]
  109. Anand, P.; Kunnumakkara, A.B.; Newman, R.A.; Aggarwal, B.B. Bioavailability of curcumin: Problems and promises. Mol. Pharm. 2007, 4, 807–818. [Google Scholar] [CrossRef]
  110. Scalbert, A.; Williamson, G. Dietary Intake and Bioavailability of Polyphenols. J. Nutr. 2000, 130, 2073S–2085S. [Google Scholar] [CrossRef] [Green Version]
  111. Wildman, R.E.; Wildman, R.; Wallace, T.C. (Eds.) Handbook of Nutraceuticals and Functional Foods; CRC Press: Boca Raton, FL, USA, 2002. [Google Scholar]
  112. Ensign, L.M.; Cone, R.; Hanes, J. Oral drug delivery with polymeric nanoparticles: The gastrointestinal mucus barriers. Adv. Drug Deliv. Rev. 2012, 64, 557–570. [Google Scholar] [CrossRef] [Green Version]
  113. He, C.; Yin, L.; Tang, C.; Yin, C. Size-dependent absorption mechanism of polymeric nanoparticles for oral delivery of protein drugs. Biomaterials 2012, 33, 8569–8578. [Google Scholar] [CrossRef]
  114. Kamaly, N.; Xiao, Z.; Valencia, P.M.; Radovic-Moreno, A.F.; Farokhzad, O.C. Targeted polymeric therapeutic nanoparticles: Design, development and clinical translation. Chem. Soc. Rev. 2012, 41, 2971–3010. [Google Scholar] [CrossRef]
  115. Valencia, P.M.; Pridgen, E.M.; Rhee, M.; Langer, R.; Farokhzad, O.C.; Karnik, R. Microfluidic Platform for Combinatorial Synthesis and Optimization of Targeted Nanoparticles for Cancer Therapy. ACS Nano 2013, 7, 10671–10680. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Patel, T.; Zhou, J.; Piepmeier, J.M.; Saltzman, W.M. Polymeric nanoparticles for drug delivery to the central nervous system. Adv. Drug Deliv. Rev. 2012, 64, 701–705. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Kreuter, J. Drug delivery to the central nervous system by polymeric nanoparticles: What do we know? Adv. Drug Deliv. Rev. 2014, 71, 2–14. [Google Scholar] [CrossRef] [PubMed]
  118. Kreuter, J. Nanoparticulate systems for brain delivery of drugs. Adv. Drug Deliv. Rev. 2001, 47, 65–81. [Google Scholar] [CrossRef] [PubMed]
  119. Rancan, F.; Papakostas, D.; Hadam, S.; Hackbarth, S.; Delair, T.; Primard, C.; Verrier, B.; Sterry, W.; Blume-Peytavi, U.; Vogt, A. Investigation of Polylactic Acid (PLA) Nanoparticles as Drug Delivery Systems for Local Dermatotherapy. Pharm. Res. 2009, 26, 2027–2036. [Google Scholar] [CrossRef]
  120. Watkinson, A.C.; Bunge, A.L.; Hadgraft, J.; Lane, M.E. Nanoparticles Do Not Penetrate Human Skin—A Theoretical Perspective. Pharm. Res. 2013, 30, 1943–1946. [Google Scholar] [CrossRef]
  121. Wang, T.; Li, Q.; Bi, K. Bioactive flavonoids in medicinal plants: Structure, activity and biological fate. Asian J. Pharm. Sci. 2018, 13, 12–23. [Google Scholar] [CrossRef]
  122. Rengasamy, K.R.; Khan, H.; Gowrishankar, S.; Lagoa, R.J.; Mahomoodally, F.M.; Khan, Z.; Suroowan, S.; Tewari, D.; Zengin, G.; Hassan, S.T.; et al. The role of flavonoids in autoimmune diseases: Therapeutic updates. Pharmacol. Ther. 2018, 194, 107–131. [Google Scholar] [CrossRef]
  123. Shaham-Niv, S.; Rehak, P.; Zaguri, D.; Levin, A.; Adler-Abramovich, L.; Vuković, L.; Král, P.; Gazit, E. Differential inhibition of metabolite amyloid formation by generic fibrillation-modifying polyphenols. Commun. Chem. 2018, 1, 25. [Google Scholar] [CrossRef] [Green Version]
  124. Amália, P.M.; Possa, M.N.; Augusto, M.C.; Francisca, L.S. Quercetin Prevents Oxidative Stress in Cirrhotic Rats. Dig. Dis. Sci. 2007, 52, 2616–2621. [Google Scholar] [CrossRef]
  125. Min, Y.-D.; Choi, C.-H.; Bark, H.; Son, H.-Y.; Park, H.-H.; Lee, S.; Park, J.-W.; Park, E.-K.; Shin, H.-I. Quercetin inhibits expression of inflammatory cytokines through attenuation of NF-κB and p38 MAPK in HMC-1 human mast cell line. Inflamm. Res. 2007, 56, 210–215. [Google Scholar] [CrossRef]
  126. Jeong, J.-H.; An, J.Y.; Kwon, Y.T.; Rhee, J.G.; Lee, Y.J. Effects of low dose quercetin: Cancer cell-specific inhibition of cell cycle progression. J. Cell Biochem. 2009, 106, 73–82. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Casagrande, R.; Georgetti, S.R.; Verri, W.A.; Borin, M.F.; Lopez, R.F.; Fonseca, M.J. In vitro evaluation of quercetin cutaneous absorption from topical formulations and its functional stability by antioxidant activity. Int. J. Pharm. 2007, 328, 183–190. [Google Scholar] [CrossRef] [PubMed]
  128. Sarkar, S.; Mandal, S.; Sinha, J.; Mukhopadhyay, S.; Das, N.; Basu, M. Quercetin: Critical Evaluation as an Antileishmanial Agent In Vivo in Hamsters Using Different Vesicular Delivery Modes. J. Drug Target. 2002, 10, 573–578. [Google Scholar] [CrossRef] [PubMed]
  129. Markham, R.J.F.; Erhardt, N.P.; Dininno, V.L.; Penman, D.; Bhatti, A.R. Flavonoids Protect Against T-2 Mycotoxins both in vitro and in vivo. Microbiology 1987, 133, 1589–1592. [Google Scholar] [CrossRef] [Green Version]
  130. Sunoqrot, S.; Abujamous, L. pH-sensitive polymeric nanoparticles of quercetin as a potential colon cancer-targeted nanomedicine. J. Drug Deliv. Sci. Technol. 2019, 52, 670–676. [Google Scholar] [CrossRef]
  131. Zhou, Y.; Chen, D.; Xue, G.; Yu, S.; Yuan, C.; Huang, M.; Jiang, L. Improved therapeutic efficacy of quercetin-loaded polymeric nanoparticles on triple-negative breast cancer by inhibiting uPA. RSC Adv. 2020, 10, 34517–34526. [Google Scholar] [CrossRef]
  132. Elmowafy, M.; Alruwaili, N.K.; Ahmad, N.; Kassem, A.M.; Ibrahim, M.F. Quercetin-Loaded Mesoporous Silica Nanoparticle–Based Lyophilized Tablets for Enhanced Physicochemical Features and Dissolution Rate: Formulation, Optimization, and In Vitro Evaluation. AAPS PharmSciTech 2020, 24, 6. [Google Scholar] [CrossRef]
  133. Zhang, Y.; Yang, Y.; Tang, K.; Hu, X.; Zou, G. Physicochemical characterization and antioxidant activity of quercetin-loaded chitosan nanoparticles. J. Appl. Polym. Sci. 2007, 107, 891–897. [Google Scholar] [CrossRef]
  134. Patil, T.N.; Srinivasan, M. Hypocholesteremic effect of curcumin in induced hypercholesteremic rats. Experiment 1971, 9, 167–169. [Google Scholar]
  135. Srinivasan, M. Effect of curcumin on blood sugar as seen in a diabetic subject. Indian J. Med. Sci. 1972, 26, 269–270. [Google Scholar] [PubMed]
  136. Srimal, R.C.; Dhawan, B.N. Pharmacology of diferuloyl methane (curcumin), a non-steroidal anti-inflammatory agent. J. Pharm. Pharmacol. 1973, 25, 447–452. [Google Scholar] [CrossRef] [PubMed]
  137. Sharma, O.P. Antioxidant activity of curcumin and related compounds. Biochem. Pharmacol. 1976, 25, 1811–1812. [Google Scholar] [CrossRef] [PubMed]
  138. Panahi, Y.; Alishiri, G.H.; Parvin, S.; Sahebkar, A. Mitigation of Systemic Oxidative Stress by Curcuminoids in Osteoarthritis: Results of a Randomized Controlled Trial. J. Diet. Suppl. 2016, 13, 209–220. [Google Scholar] [CrossRef]
  139. Sreedhar, R.; Arumugam, S.; Thandavarayan, R.A.; Karuppagounder, V.; Watanabe, K. Curcumin as a therapeutic agent in the chemoprevention of inflammatory bowel disease. Drug Discov. Today 2016, 21, 843–849. [Google Scholar] [CrossRef] [PubMed]
  140. García-Niño, W.R.; Pedraza-Chaverrí, J. Protective effect of curcumin against heavy metals-induced liver damage. Food Chem. Toxicol. 2014, 69, 182–201. [Google Scholar] [CrossRef]
  141. Joe, B.; Vijaykumar, M.; Lokesh, B.R. Biological Properties of Curcumin-Cellular and Molecular Mechanisms of Action. Crit. Rev. Food Sci. Nutr. 2004, 44, 97–111. [Google Scholar] [CrossRef]
  142. Wang, X.; Hang, Y.; Liu, J.; Hou, Y.; Wang, N.; Wang, M. Anticancer effect of curcumin inhibits cell growth through miR-21/PTEN/Akt pathway in breast cancer cell. Oncol. Lett. 2017, 13, 4825–4831. [Google Scholar] [CrossRef] [Green Version]
  143. Da Silva, A.C.; De Freitas Santos, P.D.; Do Prado Silva, J.T.; Leimann, F.V.; Bracht, L.; Gonçalves, O.H. Impact of curcumin nanoformulation on its antimicrobial activity. Trends Food Sci. Technol. 2018, 72, 74–82. [Google Scholar] [CrossRef]
  144. Esatbeyoglu, T.; Huebbe, P.; Ernst, I.M.A.; Chin, D.; Wagner, A.E.; Rimbach, G. Curcumin-From Molecule to Biological Function. Angew. Chem. Int. Ed. 2012, 51, 5308–5332. [Google Scholar] [CrossRef]
  145. Flora, G.; Gupta, D.; Tiwari, A. Nanocurcumin: A Promising Therapeutic Advancement over Native Curcumin. Crit. Rev. Ther. Drug Carrier Syst. 2013, 30, 331–368. [Google Scholar] [CrossRef] [PubMed]
  146. Umerska, A.; Gaucher, C.; Oyarzun-Ampuero, F.; Fries-Raeth, I.; Colin, F.; Villamizar-Sarmiento, M.G.; Maincent, P.; Sapin-Minet, A. Polymeric Nanoparticles for Increasing Oral Bioavailability of Curcumin. Antioxidants 2018, 7, 46. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Jahagirdar, P.; Gupta, P.K.; Kulkarni, S.P.; Devarajan, P.V. Polymeric curcumin nanoparticles by a facile in situ method for macrophage targeted delivery. Bioeng. Transl. Med. 2018, 4, 141–151. [Google Scholar] [CrossRef] [PubMed]
  148. Udompornmongkol, P.; Chiang, B.-H. Curcumin-loaded polymeric nanoparticles for enhanced anti-colorectal cancer applications. J. Biomater. Appl. 2015, 30, 537–546. [Google Scholar] [CrossRef]
  149. Nair, R.S.; Morris, A.; Billa, N.; Leong, C.-O. An Evaluation of Curcumin-Encapsulated Chitosan Nanoparticles for Transdermal Delivery. AAPS PharmSciTech 2019, 20, 69. [Google Scholar] [CrossRef]
  150. Walbi, I.A.; Ahmad, M.Z.; Ahmad, J.; Algahtani, M.S.; Alali, A.S.; Alsudir, S.A.; Aodah, A.H.; Albarqi, H.A. Development of a Curcumin-Loaded Lecithin/Chitosan Nanoparticle Utilizing a Box-Behnken Design of Experiment: Formulation Design and Influence of Process Parameters. Polymers 2022, 14, 3758. [Google Scholar] [CrossRef]
  151. Grabovac, V.; Bernkop-Schnürch, A. Development and In Vitro Evaluation of Surface Modified Poly(lactide-co-glycolide) Nanoparticles with Chitosan-4-Thiobutylamidine. Drug Dev. Ind. Pharm. 2007, 33, 767–774. [Google Scholar] [CrossRef]
  152. Thamake, S.I.; Raut, S.L.; Ranjan, A.P.; Gryczynski, Z.; Vishwanatha, J.K. Surface functionalization of PLGA nanoparticles by non-covalent insertion of a homo-bifunctional spacer for active targeting in cancer therapy. Nanotechnology 2010, 22, 35101. [Google Scholar] [CrossRef]
  153. Joseph, A.; Wood, T.; Chen, C.-C.; Corry, K.; Snyder, J.M.; Juul, S.E.; Parikh, P.; Nance, E. Curcumin-loaded polymeric nanoparticles for neuroprotection in neonatal rats with hypoxic-ischemic encephalopathy. Nano Res. 2018, 11, 5670–5688. [Google Scholar] [CrossRef]
  154. Anitha, A.; Maya, S.; Deepa, N.; Chennazhi, K.; Nair, S.; Tamura, H.; Jayakumar, R. Efficient water soluble O-carboxymethyl chitosan nanocarrier for the delivery of curcumin to cancer cells. Carbohydr. Polym. 2011, 83, 452–461. [Google Scholar] [CrossRef]
  155. Anitha, A.; Maya, S.; Deepa, N.; Chennazhi, K.P.; Nair, S.; Jayakumar, R. Curcumin-loaded N, O-carboxymethyl chitosan nanoparticles for cancer drug delivery. J. Biomater. Sci. Polym. Ed. 2012, 23, 1381–1400. [Google Scholar] [CrossRef] [PubMed]
  156. Das, R.K.; Kasoju, N.; Bora, U. Encapsulation of curcumin in alginate-chitosan-pluronic composite nanoparticles for delivery to cancer cells. Nanomed. Nanotechnol. Biol. Med. 2010, 6, 153–160. [Google Scholar] [CrossRef] [PubMed]
  157. Liu, J.; Xu, L.; Liu, C.; Zhang, D.; Wang, S.; Deng, Z.; Lou, W.; Xu, H.; Bai, Q.; Ma, J. Preparation and characterization of cationic curcumin nanoparticles for improvement of cellular uptake. Carbohydr. Polym. 2012, 90, 16–22. [Google Scholar] [CrossRef] [PubMed]
  158. Kim, S.-H.; Choi, K.-C. Anti-cancer effect and underlying mechanism(s) of kaempferol, a phytoestrogen, on the regulation of apoptosis in diverse cancer cell models. Toxicol. Res. 2013, 29, 229–234. [Google Scholar] [CrossRef] [Green Version]
  159. Li, H.; Yang, L.; Zhang, Y.; Gao, Z. Kaempferol inhibits fibroblast collagen synthesis, proliferation and activation in hypertrophic scar via targeting TGF-β receptor type I. Biomed. Pharmacother. 2016, 83, 967–974. [Google Scholar] [CrossRef]
  160. Wang, J.; Fang, X.; Ge, L.; Cao, F.; Zhao, L.; Wang, Z.; Xiao, W. Antitumor, antioxidant and anti-inflammatory activities of kaempferol and its corresponding glycosides and the enzymatic preparation of kaempferol. PLoS ONE 2018, 13, e0197563. [Google Scholar] [CrossRef]
  161. Deng, S.-P.; Yang, Y.-L.; Cheng, X.-X.; Li, W.-R.; Cai, J.Y. Synthesis, Spectroscopic Study and Radical Scavenging Activity of Kaempferol Derivatives: Enhanced Water Solubility and Antioxidant Activity. Int. J. Mol. Sci. 2019, 20, 975. [Google Scholar] [CrossRef] [Green Version]
  162. Luo, H.; Jiang, B.; Li, B.; Li, Z.; Jiang, B.H.; Chen, Y.C. Kaempferol nanoparticles achieve strong and selective inhibition of ovarian cancer cell viability. Int. J. Nanomed. 2012, 7, 3951–3959. [Google Scholar] [CrossRef] [Green Version]
  163. Kazmi, I.; Al-Abbasi, F.A.; Afzal, M.; Altayb, H.N.; Nadeem, M.S.; Gupta, G. Formulation and Evaluation of Kaempferol Loaded Nanoparticles against Experimentally Induced Hepatocellular Carcinoma: In Vitro and In Vivo Studies. Pharmaceutics 2021, 13, 2086. [Google Scholar] [CrossRef]
  164. Ilk, S.; Sağlam, N.; Özgen, M.; Korkusuz, F. Chitosan nanoparticles enhances the anti-quorum sensing activity of kaempferol. Int. J. Biol. Macromol. 2016, 94, 653–662. [Google Scholar] [CrossRef]
  165. Ilk, S.; Saglam, N.; Özgen, M. Kaempferol loaded lecithin/chitosan nanoparticles: Preparation, characterization, and their potential applications as a sustainable antifungal agent. Artif. Cells Nanomed. Biotechnol. 2017, 45, 907–916. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Amri, A.; Chaumeil, J.C.; Sfar, S.; Charrueau, C. Administration of resveratrol: What formulation solutions to bioavailability limitations? J. Control. Release 2012, 158, 182–193. [Google Scholar] [CrossRef] [PubMed]
  167. Fulda, S. Resveratrol and derivatives for the prevention and treatment of cancer. Drug Discov. Today 2010, 15, 757–765. [Google Scholar] [CrossRef] [PubMed]
  168. Rius, C.; Abu-Taha, M.; Hermenegildo, C.; Piqueras, L.; Cerda-Nicolas, J.-M.; Issekutz, A.; Estañ, L.; Cortijo, J.; Morcillo, E.; Orallo, F.; et al. Trans-but not Cis-resveratrol impairs angiotensin-II–mediated vascular inflammation through inhibition of NF-κB activation and peroxisome proliferator-activated receptor-γ upregulation. J. Immunol. 2010, 185, 3718–3727. [Google Scholar] [CrossRef] [Green Version]
  169. Soleas, G.J.; Diamandis, E.P.; Goldberg, D.M. Resveratrol: A molecule whose time has come? And gone? Clin. Biochem. 1997, 30, 91–113. [Google Scholar] [CrossRef] [PubMed]
  170. Trela, B.C.; Waterhouse, A.L. Resveratrol: Isomeric Molar Absorptivities and Stability. J. Agric. Food Chem. 1996, 44, 1253–1257. [Google Scholar] [CrossRef]
  171. Park, E.-J.; Pezzuto, J.M. The pharmacology of resveratrol in animals and humans. Biochim. Biophys. Acta (BBA)-Mol. Basis Dis. 2015, 1852, 1071–1113. [Google Scholar] [CrossRef] [Green Version]
  172. Heredia, A.; Davis, C.; Redfield, R. Synergistic inhibition of HIV-1 in activated and resting peripheral blood mononuclear cells, monocyte-derived macrophages, and selected drug-resistant isolates with nucleoside analogues combined with a natural product, resveratrol. JAIDS J. Acquir. Immune Defic. Syndr. 2000, 25, 246–255. [Google Scholar] [CrossRef]
  173. Jung, K.-H.; Lee, J.H.; Park, J.W.; Quach, C.H.T.; Moon, S.-H.; Cho, Y.S.; Lee, K.-H. Resveratrol-loaded polymeric nanoparticles suppress glucose metabolism and tumor growth in vitro and in vivo. Int. J. Pharm. 2015, 478, 251–257. [Google Scholar] [CrossRef]
  174. Chen, W.; Zhao, Z.; Zhao, S.; Zhang, L.; Song, Q. Resveratrol and Puerarin loaded polymeric nanoparticles to enhance the chemotherapeutic efficacy in spinal cord injury. Biomed. Microdevices 2020, 22, 1–10. [Google Scholar] [CrossRef]
  175. Zu, Y.; Zhang, Y.; Wang, W.; Zhao, X.; Han, X.; Wang, K.; Ge, Y. Preparation and in vitro/in vivo evaluation of resveratrol-loaded carboxymethyl chitosan nanoparticles. Drug Deliv. 2016, 23, 971–981. [Google Scholar] [CrossRef]
  176. Sarma, S.; Agarwal, S.; Bhuyan, P.; Hazarika, J.; Ganguly, M. Resveratrol-loaded chitosan–pectin core–shell nanoparticles as novel drug delivery vehicle for sustained release and improved antioxidant activities. R. Soc. Open Sci. 2022, 9, 210784. [Google Scholar] [CrossRef]
  177. Liu, Y.; Gao, L.; Yi, J.; Fan, Y.; Wu, X.; Zhang, Y. α-Lactalbumin and chitosan core–shell nanoparticles: Resveratrol loading, protection, and antioxidant activity. Food Funct. 2020, 11, 1525–1536. [Google Scholar] [CrossRef]
  178. Khan, N.; Adhami, V.M.; Mukhtar, H. Review: Green Tea Polyphenols in Chemoprevention of Prostate Cancer: Preclinical and Clinical Studies. Nutr. Cancer 2009, 61, 836–841. [Google Scholar] [CrossRef] [Green Version]
  179. Myung, S.K.; Bae, W.K.; Oh, S.M.; Kim, Y.; Ju, W.; Sung, J.; Lee, Y.J.; Ko, J.A.; Song, J.I.; Choi, H.J. Green tea consumption and risk of stomach cancer: A meta-analysis of epidemiologic studies. Int. J. Cancer 2009, 124, 670–677, Erratum in Int. J. Cancer 2009, 124, 1496. [Google Scholar] [CrossRef] [PubMed]
  180. Yang, C.S.; Lambert, J.D.; Ju, J.; Lu, G.; Sang, S. Tea and cancer prevention: Molecular mechanisms and human relevance. Toxicol. Appl. Pharmacol. 2007, 224, 265–273. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  181. Ogunleye, A.A.; Xue, F.; Michels, K.B. Green tea consumption and breast cancer risk or recurrence: A meta-analysis. Breast Cancer Res. Treat. 2009, 119, 477–484. [Google Scholar] [CrossRef] [PubMed]
  182. Sanna, V.; Singh, C.K.; Jashari, R.; Adhami, V.M.; Chamcheu, J.C.; Rady, I.; Sechi, M.; Mukhtar, H.; Siddiqui, I.A. Targeted nanoparticles encapsulating (−)-epigallocatechin-3-gallate for prostate cancer prevention and therapy. Sci. Rep. 2017, 7, srep41573. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Alserihi, R.F.; Mohammed, M.R.S.; Kaleem, M.; Khan, M.I.; Sechi, M.; Sanna, V.; Zughaibi, T.A.; Abuzenadah, A.M.; Tabrez, S. Development of (−)-epigallocatechin-3-gallate-loaded folate receptor-targeted nanoparticles for prostate cancer treatment. Nanotechnol. Rev. 2021, 11, 298–311. [Google Scholar] [CrossRef]
  184. Shao, S.; Li, L.; Yang, G.; Li, J.; Luo, C.; Gong, T.; Zhou, S. Controlled green tea polyphenols release from electrospun PCL/MWCNTs composite nanofibers. Int. J. Pharm. 2011, 421, 310–320. [Google Scholar] [CrossRef]
  185. Maya, S.; Sabitha, M.; Nair, S.V.; Jayakumar, R. Phytomedicine-Loaded Polymeric Nanomedicines: Potential Cancer Therapeutics. Multifaceted Dev. Appl. Biopolym. Biol. Biomed. Nanotechnol. 2012, 254, 203–239. [Google Scholar] [CrossRef]
  186. Safer, A.-M.; Leporatti, S.; Jose, J.; Soliman, M.S. Conjugation of EGCG and Chitosan NPs as A Novel Nano-Drug Delivery System. Int. J. Nanomed. 2019, 14, 8033–8046. [Google Scholar] [CrossRef] [Green Version]
  187. Moreno-Vásquez, M.J.; Plascencia-Jatomea, M.; Sánchez-Valdes, S.; Tanori-Córdova, J.C.; Castillo-Yañez, F.J.; Quintero-Reyes, I.E.; Graciano-Verdugo, A.Z. Characterization of Epigallocatechin-Gallate-Grafted Chitosan Nanoparticles and Evaluation of Their Antibacterial and Antioxidant Potential. Polymers 2021, 13, 1375. [Google Scholar] [CrossRef] [PubMed]
  188. Hong, Z.; Xu, Y.; Yin, J.-F.; Jin, J.; Jiang, Y.; Du, Q. Improving the Effectiveness of (−)-Epigallocatechin Gallate (EGCG) against Rabbit Atherosclerosis by EGCG-Loaded Nanoparticles Prepared from Chitosan and Polyaspartic Acid. J. Agric. Food Chem. 2014, 62, 12603–12609. [Google Scholar] [CrossRef] [PubMed]
  189. Liang, J.; Yan, H.; Yang, H.-J.; Kim, H.W.; Wan, X.; Lee, J.; Ko, S. Synthesis and controlled-release properties of chitosan/β-Lactoglobulin nanoparticles as carriers for oral administration of epigallocatechin gallate. Food Sci. Biotechnol. 2016, 25, 1583–1590. [Google Scholar] [CrossRef]
  190. Post-White, J.; Ladas, E.J.; Kelly, K.M. Advances in the Use of Milk Thistle (Silybum marianum). Integr. Cancer Ther. 2007, 6, 104–109. [Google Scholar] [CrossRef]
  191. Kroll, D.J.; Shaw, H.S.; Oberlies, N.H. Milk Thistle Nomenclature: Why It Matters in Cancer Research and Pharmacokinetic Studies. Integr. Cancer Ther. 2007, 6, 110–119. [Google Scholar] [CrossRef] [Green Version]
  192. Ramasamy, K.; Agarwal, R. Multitargeted therapy of cancer by silymarin. Cancer Lett. 2008, 269, 352–362. [Google Scholar] [CrossRef] [Green Version]
  193. Snima, K.S.; Arunkumar, P.; Jayakumar, R.; Lakshmanan, V.-K. Silymarin Encapsulated Poly(D, L-lactic-co-glycolic acid) Nanoparticles: A Prospective Candidate for Prostate Cancer Therapy. J. Biomed. Nanotechnol. 2014, 10, 559–570. [Google Scholar] [CrossRef]
  194. Azadpour, M.; Farajollahi, M.M.; Dariushnejad, H.; Varzi, A.M.; Varezardi, A.; Barati, M. Effects of synthetic silymarin-PLGA nanoparticles on M2 polarization and inflammatory cytokines in LPS-treated murine peritoneal macrophages. Iran. J. Basic Med. Sci. 2021, 24, 1446. [Google Scholar]
  195. El-Sherbiny, I.M.; Abdel-Mogib, M.; Dawidar, A.-A.M.; Elsayed, A.; Smyth, H.D.C. Biodegradable pH-responsive alginate-poly (lactic-co-glycolic acid) nano/micro hydrogel matrices for oral delivery of silymarin. Carbohydr. Polym. 2011, 83, 1345–1354. [Google Scholar] [CrossRef]
  196. Kerwin, S.M. Soy saponins and the anticancer effects of soybeans and soy-based foods. Curr. Med. Chem. Agents 2004, 4, 263–272. [Google Scholar] [CrossRef] [PubMed]
  197. Rao, A.V.; Sung, M.K. Saponins as anticarcinogens. J. Nutr. 1995, 125, 717S–724S. [Google Scholar] [CrossRef] [PubMed]
  198. Tian, F.; Zhang, X.; Tong, Y.; Yi, Y.; Zhang, S.; Liping, L.; Sun, P.; Lin, L.; Ding, J. PE, a new sulfated saponin from sea cucumber, Exhibits anti-angiogenic and anti-tumor activities in vitro and in vivo. Cancer Biol. Ther. 2005, 4, 874–882. [Google Scholar] [CrossRef] [Green Version]
  199. Van de Ven, H.; Vermeersch, M.; Vandenbroucke, R.; Matheeussen, A.; Apers, S.; Weyenberg, W.; De Smedt, S.; Cos, P.; Maes, L.; Ludwig, A. Intracellular drug delivery in Leishmania-infected macrophages: Evaluation of saponin-loaded PLGA nanoparticles. J. Drug Target. 2011, 20, 142–154. [Google Scholar] [CrossRef]
  200. Ikezoe, T.; Yang, Y.; Bandobashi, K.; Saito, T.; Takemoto, S.; Machida, H.; Togitani, K.; Koeffler, H.P.; Taguchi, H. Oridonin, a diterpenoid purified from Rabdosia rubescens, inhibits the proliferation of cells from lymphoid malignancies in association with blockade of the NF-κB signal pathways. Mol. Cancer Ther. 2005, 4, 578–586. [Google Scholar] [CrossRef] [Green Version]
  201. Xing, J.; Zhang, D.; Tan, T. Studies on the oridonin-loaded poly(d,l-lactic acid) nanoparticles in vitro and in vivo. Int. J. Biol. Macromol. 2007, 40, 153–158. [Google Scholar] [CrossRef]
  202. Cho, H.-Y.; Lee, C.K.; Lee, Y.-B. Preparation and Evaluation of PEGylated and Folate-PEGylated Liposomes Containing Paclitaxel for Lymphatic Delivery. J. Nanomater. 2015, 2015, 1–10. [Google Scholar] [CrossRef] [Green Version]
  203. Danhier, F.; Danhier, P.; De Saedeleer, C.J.; Fruytier, A.-C.; Schleich, N.; Rieux, A.D.; Sonveaux, P.; Gallez, B.; Préat, V. Paclitaxel-loaded micelles enhance transvascular permeability and retention of nanomedicines in tumors. Int. J. Pharm. 2015, 479, 399–407. [Google Scholar] [CrossRef]
  204. Ruttala, H.B.; Ko, Y.T. Liposomal co-delivery of curcumin and albumin/paclitaxel nanoparticle for enhanced synergistic antitumor efficacy. Colloids Surf. B Biointerfaces 2015, 128, 419–426. [Google Scholar] [CrossRef]
  205. Li, Q.; Xiong, W.; Peng, L.; Chen, H. Surface modification of MPEG-b-PCL-based nanoparticles via oxidative self-polymerization of dopamine for malignant melanoma therapy. Int. J. Nanomed. 2015, 10, 2985–2996. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  206. Danhier, F.; Lecouturier, N.; Vroman, B.; Jérôme, C.; Marchand-Brynaert, J.; Feron, O.; Préat, V. Paclitaxel-loaded PEGylated PLGA-based nanoparticles: In vitro and in vivo evaluation. J. Control. Release 2009, 133, 11–17. [Google Scholar] [CrossRef] [PubMed]
  207. Fonseca, C.; Simões, S.; Gaspar, R. Paclitaxel-loaded PLGA nanoparticles: Preparation, physicochemical characterization and in vitro anti-tumoral activity. J. Control. Release 2002, 83, 273–286. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  208. Su, Y.; Hu, J.; Huang, Z.; Huang, Y.; Peng, B.; Xie, N.; Liu, H. Paclitaxel-loaded star-shaped copolymer nanoparticles for enhanced malignant melanoma chemotherapy against multidrug resistance. Drug Des. Dev. Ther. 2017, 11, 659–668. [Google Scholar] [CrossRef] [Green Version]
  209. Cui, Y.; Zhang, M.; Zeng, F.; Jin, H.; Xu, Q.; Huang, Y. Dual-Targeting Magnetic PLGA Nanoparticles for Codelivery of Paclitaxel and Curcumin for Brain Tumor Therapy. ACS Appl. Mater. Interfaces 2016, 8, 32159–32169. [Google Scholar] [CrossRef]
  210. Abriata, J.P.; Turatti, R.C.; Luiz, M.T.; Raspantini, G.L.; Tofani, L.B.; Amaral, R.L.F.D.; Swiech, K.; Marcato, P.D.; Marchetti, J.M. Development, characterization and biological in vitro assays of paclitaxel-loaded PCL polymeric nanoparticles. Mater. Sci. Eng. C 2018, 96, 347–355. [Google Scholar] [CrossRef]
  211. Bernabeu, E.; Helguera, G.; Legaspi, M.J.; González, L.; Hocht, C.; Taira, C.; Chiappetta, D.A. Paclitaxel-loaded PCL–TPGS nanoparticles: In vitro and in vivo performance compared with Abraxane®. Colloids Surf. B Biointerfaces 2014, 113, 43–50. [Google Scholar] [CrossRef]
  212. Gupta, U.; Sharma, S.; Khan, I.; Gothwal, A.; Sharma, A.K.; Singh, Y.; Chourasia, M.K.; Kumar, V. Enhanced apoptotic and anticancer potential of paclitaxel loaded biodegradable nanoparticles based on chitosan. Int. J. Biol. Macromol. 2017, 98, 810–819. [Google Scholar] [CrossRef]
  213. Hans, M.L.; Lowman, A.M. Biodegradable nanoparticles for drug delivery and targeting. Curr. Opin. Solid State Mater. Sci. 2002, 6, 319–327. [Google Scholar] [CrossRef]
  214. Das, P.; Jana, N.R. Dopamine functionalized polymeric nanoparticle for targeted drug delivery. RSC Adv. 2015, 5, 33586–33594. [Google Scholar] [CrossRef]
  215. Amgoth, C.; Phan, C.; Banavoth, M.; Rompivalasa, S.; Tang, G. Polymer Properties: Functionalization and Surface Modified Nanoparticles. In Role of Novel Drug Delivery Vehicles in Nanobiomedicine; IntechOpen: London, UK, 2019. [Google Scholar]
  216. Roux, S.; Garcia, B.; Bridot, J.-L.; Salomé, M.; Marquette, C.; Lemelle, L.; Gillet, P.; Blum, L.; Perriat, P.; Tillement, O. Synthesis, Characterization of Dihydrolipoic Acid Capped Gold Nanoparticles, and Functionalization by the Electroluminescent Luminol. Langmuir 2005, 21, 2526–2536. [Google Scholar] [CrossRef] [PubMed]
  217. Zayats, M.; Katz, E.; Baron, R.; Willner, I. Reconstitution of Apo-Glucose Dehydrogenase on Pyrroloquinoline Quinone-Functionalized Au Nanoparticles Yields an Electrically Contacted Biocatalyst. J. Am. Chem. Soc. 2005, 127, 12400–12406. [Google Scholar] [CrossRef] [PubMed]
  218. Ulman, A. Formation and Structure of Self-Assembled Monolayers. Chem. Rev. 1996, 96, 1533–1554. [Google Scholar] [CrossRef]
  219. Coester, C.; Kreuter, J.; von Briesen, H.; Langer, K. Preparation of avidin-labelled gelatin nanoparticles as carriers for biotinylated peptide nucleic acid (PNA). Int. J. Pharm. 2000, 196, 147–149. [Google Scholar] [CrossRef] [PubMed]
  220. Rodrigues, J.; Santos-Magalhães, N.; Coelho, L.; Couvreur, P.; Ponchel, G.; Gref, R. Novel core(polyester)-shell(polysaccharide) nanoparticles: Protein loading and surface modification with lectins. J. Control. Release 2003, 92, 103–112. [Google Scholar] [CrossRef] [PubMed]
  221. Feng, S.-S. Nanoparticles of biodegradable polymers for new-concept chemotherapy. Expert Rev. Med. Devices 2004, 1, 115–125. [Google Scholar] [CrossRef]
  222. Owens, D.E., III; Peppas, N.A. Opsonization, biodistribution, and pharmacokinetics of polymeric nanoparticles. Int. J. Pharm. 2006, 307, 93–102. [Google Scholar] [CrossRef]
  223. Torchilin, V.P.; Trubetskoy, V.S. Which polymers can make nanoparticulate drug carriers long-circulating? Adv. Drug Deliv. Rev. 1995, 16, 141–155. [Google Scholar] [CrossRef]
  224. Storm, G.; Belliot, S.O.; Daemen, T.; Lasic, D.D. Surface modification of nanoparticles to oppose uptake by the mononuclear phagocyte system. Adv. Drug Deliv. Rev. 1995, 17, 31–48. [Google Scholar] [CrossRef]
  225. Stolnik, S.; Illum, L.; Davis, S. Long circulating microparticulate drug carriers. Adv. Drug Deliv. Rev. 2012, 64, 290–301. [Google Scholar] [CrossRef]
  226. Tobío, M.; Sánchez, A.; Vila, A.; Soriano, I.; Evora, C.; Vila-Jato, J.; Alonso, M. The role of PEG on the stability in digestive fluids and in vivo fate of PEG-PLA nanoparticles following oral administration. Colloids Surf. B Biointerfaces 2000, 18, 315–323. [Google Scholar] [CrossRef] [PubMed]
  227. Kim, S.Y.; Lee, Y.M. Taxol-loaded block copolymer nanospheres composed of methoxy poly(ethylene glycol) and poly(ε-caprolactone) as novel anticancer drug carriers. Biomaterials 2001, 22, 1697–1704. [Google Scholar] [CrossRef] [PubMed]
  228. Calvo, P.; Gouritin, B.; Brigger, I.; Lasmezas, C.; Deslys, J.-P.; Williams, A.; Andreux, J.P.; Dormont, D.; Couvreur, P. PEGylated polycyanoacrylate nanoparticles as vector for drug delivery in prion diseases. J. Neurosci. Methods 2001, 111, 151–155. [Google Scholar] [CrossRef]
  229. Perrault, S.D.; Chan, W.C.W. Synthesis and surface modification of highly monodispersed, spherical gold nanoparticles of 50–200 nm. J. Am. Chem. Soc. 2009, 131, 17042–17043. [Google Scholar] [CrossRef] [PubMed]
  230. Fang, J.-Y. Nano- or submicron-sized liposomes as carriers for drug delivery. Chang. Gung Med. J. 2006, 29, 358. [Google Scholar] [PubMed]
  231. Gref, R.; Lück, M.; Quellec, P.; Marchand, M.; Dellacherie, E.; Harnisch, S.; Blunk, T.; Müller, R.H. ‘Stealth’ corona-core nanoparticles surface modified by polyethylene glycol (PEG): Influences of the corona (PEG chain length and surface density) and of the core composition on phagocytic uptake and plasma protein adsorption. Colloids Surf. B Biointerfaces 2000, 18, 301–313. [Google Scholar] [CrossRef] [PubMed]
  232. Devalpally, H.; Shenoy, D.; Little, S.; Langer, R.; Amiji, M. Poly (ethylene oxide)-modified poly (epsiloncaprolactone) nanoparticles for targeted delivery of tamoxifen in breast cancer. Cancer Chemother. Pharmacol. 2007, 59, 477. [Google Scholar]
  233. Gulyaev, A.E.; Gelperina, S.E.; Skidan, I.N.; Antropov, A.S.; Kivman, G.Y.; Kreuter, J. Significant Transport of Doxorubicin into the Brain with Polysorbate 80-Coated Nanoparticles. Pharm. Res. 1999, 16, 1564–1569. [Google Scholar] [CrossRef]
  234. Alyautdin, R.N.; Tezikov, E.B.; Ramge, P.; Kharkevich, D.A.; Begley, D.J.; Kreuter, J. Significant entry of tubocurarine into the brain of rats by adsorption to polysorbate 80–coated polybutylcyanoacrylate nanoparticles: An in situ brain perfusion study. J. Microencapsul. 1998, 15, 67–74. [Google Scholar] [CrossRef]
  235. Borchard, G.; Audus, K.L.; Shi, F.; Kreuter, J. Uptake of surfactant-coated poly(methyl methacrylate)-nanoparticles by bovine brain microvessel endothelial cell monolayers. Int. J. Pharm. 1994, 110, 29–35. [Google Scholar] [CrossRef]
  236. Lemarchand, C.; Gref, R.; Lesieur, S.; Hommel, H.; Vacher, B.; Besheer, A.; Maeder, K.; Couvreur, P. Physico-chemical characterization of polysaccharide-coated nanoparticles. J. Control. Release 2005, 108, 97–111. [Google Scholar] [CrossRef] [PubMed]
  237. Lee, S.H.; Zhang, Z.; Feng, S.-S. Nanoparticles of poly(lactide)—Tocopheryl polyethylene glycol succinate (PLA-TPGS) copolymers for protein drug delivery. Biomaterials 2007, 28, 2041–2050. [Google Scholar] [CrossRef] [PubMed]
  238. Gavas, S.; Quazi, S.; Karpiński, T.M. Nanoparticles for Cancer Therapy: Current Progress and Challenges. Nanoscale Res. Lett. 2021, 16, 1–21. [Google Scholar] [CrossRef] [PubMed]
  239. Wicki, A.; Witzigmann, D.; Balasubramanian, V.; Huwyler, J. Nanomedicine in cancer therapy: Challenges, opportunities, and clinical applications. J. Control. Release 2015, 200, 138–157. [Google Scholar] [CrossRef]
  240. Ai, J.; Biazar, E.; Jafarpour, M.; Montazeri, M.; Majdi, A.; Aminifard, S.; Zafari, M.; Akbari, H.R.; Rad, H.G. Nanotoxicology and nanoparticle safety in biomedical designs. Int. J. Nanomed. 2011, 6, 1117–1127. [Google Scholar] [CrossRef] [Green Version]
  241. Gatoo, M.A.; Naseem, S.; Arfat, M.Y.; Dar, A.M.; Qasim, K.; Zubair, S. Physicochemical Properties of Nanomaterials: Implication in Associated Toxic Manifestations. BioMed Res. Int. 2014, 2014, 1–8. [Google Scholar] [CrossRef] [Green Version]
  242. Singh, N.; Joshi, A.; Toor, A.P.; Verma, G. Drug delivery: Advancements and challenges. In Nanostructures for Drug Delivery; Elsevier: Amsterdam, The Netherlands, 2017; pp. 865–886. [Google Scholar]
  243. Prabha, A.S.; Dorothy, R.; Jancirani, S.; Rajendran, S.; Singh, G.; Kumaran, S.S. Recent advances in the study of toxicity of polymer-based nanomaterials. Nanotoxicity 2020, 7, 143–165. [Google Scholar] [CrossRef]
  244. Krishnaswamy, K.; Orsat, V. Sustainable delivery systems through green nanotechnology. In Nano-and Microscale Drug Delivery Systems; Elsevier: Amsterdam, The Netherlands, 2017; pp. 17–32. [Google Scholar]
  245. Jena, S.K.; Sangamwar, A.T. Polymeric micelles of amphiphilic graft copolymer of a-tocopherol succinate-g-carboxymethyl chitosan for tamoxifen delivery: Synthesis, characterization and in vivo pharmacokinetic study (vol 151, pg 1162, 2016). Carbohydr. Polym. 2017, 157, 904. [Google Scholar] [CrossRef]
  246. Liu, Y.; Kong, M.; Feng, C.; Yang, K.K.; Li, Y.; Su, J.; Cheng, X.J.; Park, H.J.; Chen, X.G. Biocompatibility, cellular uptake and biodistribution of the polymeric amphiphilic nanoparticles as oral drug carriers. Colloids Surf. B Biointerfaces 2013, 103, 345–353. [Google Scholar] [CrossRef]
  247. Radwan, S.E.-S.; Sokar, M.S.; Abdelmonsif, D.A.; El-Kamel, A.H. Mucopenetrating nanoparticles for enhancement of oral bioavailability of furosemide: In vitro and in vivo evaluation/sub-acute toxicity study. Int. J. Pharm. 2017, 526, 366–379. [Google Scholar] [CrossRef]
  248. Maity, S.; Mukhopadhyay, P.; Kundu, P.P.; Chakraborti, A.S. Alginate coated chitosan core-shell nanoparticles for efficient oral delivery of naringenin in diabetic animals—An in vitro and in vivo approach. Carbohydr. Polym. 2017, 170, 124–132. [Google Scholar] [CrossRef]
  249. Sharma, M.; Sharma, S.; Sharma, V.; Sharma, K.; Yadav, S.K.; Dwivedi, P.; Agrawal, S.; Paliwal, S.K.; Dwivedi, A.K.; Maikhuri, J.P.; et al. Oleanolic–bioenhancer coloaded chitosan modified nanocarriers attenuate breast cancer cells by multimode mechanism and preserve female fertility. Int. J. Biol. Macromol. 2017, 104, 1345–1358. [Google Scholar] [CrossRef]
  250. Lanone, S.; Boczkowski, J. Biomedical applications and potential health risks of nanomaterials: Molecular mechanisms. Curr. Mol. Med. 2006, 6, 651–663. [Google Scholar] [CrossRef]
  251. Rejman, J.; Oberle, V.; Zuhorn, I.S.; Hoekstra, D. Size-dependent internalization of particles via the pathways of clathrin- and caveolae-mediated endocytosis. Biochem. J. 2004, 377, 159–169. [Google Scholar] [CrossRef] [PubMed]
  252. Oberdörster, G.; Oberdörster, E.; Oberdörster, J. Nanotoxicology: An emerging discipline evolving from studies of ultrafine particles. Environ. Health Perspect. 2005, 113, 823–839. [Google Scholar] [CrossRef]
  253. Nel, A.; Xia, T.; Mädler, L.; Li, N. Toxic Potential of Materials at the Nanolevel. Science 2006, 311, 622–627. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  254. Verma, A.; Stellacci, F. Effect of surface properties on nanoparticle–cell interactions. Small 2010, 6, 12–21. [Google Scholar] [CrossRef] [PubMed]
  255. Sahoo, S.K.; Panyam, J.; Prabha, S.; Labhasetwar, V. Residual polyvinyl alcohol associated with poly (d,l-lactide-co-glycolide) nanoparticles affects their physical properties and cellular uptake. J. Control. Release 2002, 82, 105–114. [Google Scholar] [CrossRef] [PubMed]
  256. Grabowski, N.; Hillaireau, H.; Vergnaud, J.; Tsapis, N.; Pallardy, M.; Kerdine-Römer, S.; Fattal, E. Surface coating mediates the toxicity of polymeric nanoparticles towards human-like macrophages. Int. J. Pharm. 2014, 482, 75–83. [Google Scholar] [CrossRef] [PubMed]
  257. He, C.; Hu, Y.; Yin, L.; Tang, C.; Yin, C. Effects of particle size and surface charge on cellular uptake and biodistribution of polymeric nanoparticles. Biomaterials 2010, 31, 3657–3666. [Google Scholar] [CrossRef]
  258. Das, S.C.; Khadka, P.; Shah, R.; McGill, S.; Smyth, H.D.C. Nanomedicine in pulmonary delivery. Theory Appl. Nonparenteral Nanomed. 2021, 4, 319–354. [Google Scholar]
  259. Bancos, S.; Stevens, D.L.; Tyner, K.M. Effect of silica and gold nanoparticles on macrophage proliferation, activation markers, cytokine production, and phagocytosis in vitro. Int. J. Nanomed. 2015, 10, 183. [Google Scholar]
  260. Gonzalez, L.; Lukamowicz-Rajska, M.; Thomassen, L.; Kirschhock, C.; Leyns, L.; Lison, D.; Martens, J.; Elhajouji, A.; Kirsch-Volders, M. Co-assessment of cell cycle and micronucleus frequencies demonstrates the influence of serum on the in vitro genotoxic response to amorphous monodisperse silica nanoparticles of varying sizes. Nanotoxicology 2014, 8, 876–884. [Google Scholar] [CrossRef]
  261. Maser, E.; Schulz, M.; Sauer, U.G.; Wiemann, M.; Ma-Hock, L.; Wohlleben, W.; Hartwig, A.; Landsiedel, R. In vitro and in vivo genotoxicity investigations of differently sized amorphous SiO2 nanomaterials. Mutat. Res. Toxicol. Environ. Mutagen. 2015, 794, 57–74. [Google Scholar] [CrossRef]
  262. Singh, R.; RamaRao, P. Accumulated Polymer Degradation Products as Effector Molecules in Cytotoxicity of Polymeric Nanoparticles. Toxicol. Sci. 2013, 136, 131–143. [Google Scholar] [CrossRef]
  263. Platel, A.; Carpentier, R.; Becart, E.; Mordacq, G.; Betbeder, D.; Nesslany, F. Influence of the surface charge of PLGA nanoparticles on their in vitro genotoxicity, cytotoxicity, ROS production and endocytosis. J. Appl. Toxicol. 2016, 36, 434–444. [Google Scholar] [CrossRef]
  264. Yang, X.; He, C.; Li, J.; Chen, H.; Ma, Q.; Sui, X.; Tian, S.; Ying, M.; Zhang, Q.; Luo, Y.; et al. Uptake of silica nanoparticles: Neurotoxicity and Alzheimer-like pathology in human SK-N-SH and mouse neuro2a neuroblastoma cells. Toxicol. Lett. 2014, 229, 240–249. [Google Scholar] [CrossRef] [PubMed]
  265. Guadagnini, R.; Kenzaoui, B.H.; Walker, L.; Pojana, G.; Magdolenova, Z.; Bilanicova, D.; Saunders, M.; Juillerat-Jeanneret, L.; Marcomini, A.; Huk, A.; et al. Toxicity screenings of nanomaterials: Challenges due to interference with assay processes and components of classic in vitro tests. Nanotoxicology 2015, 9, 13–24. [Google Scholar] [CrossRef] [PubMed]
  266. Thomas, O.S.; Weber, W. Overcoming Physiological Barriers to Nanoparticle Delivery—Are We There Yet? Front. Bioeng. Biotechnol. 2019, 7, 415. [Google Scholar] [CrossRef] [PubMed]
  267. Garay, R.P.; El-Gewely, R.; Armstrong, J.K.; Garratty, G.; Richette, P. Antibodies against polyethylene glycol in healthy subjects and in patients treated with PEG-conjugated agents. Expert Opin. Drug Deliv. 2012, 9, 1319–1323. [Google Scholar] [CrossRef] [PubMed]
  268. Suk, J.S.; Xu, Q.; Kim, N.; Hanes, J.; Ensign, L.M. PEGylation as a strategy for improving nanoparticle-based drug and gene delivery. Adv. Drug Deliv. Rev. 2016, 99, 28–51. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  269. Colby, A.H.; Liu, R.; Doyle, R.P.; Merting, A.; Zhang, H.; Savage, N.; Chu, N.-Q.; Hollister, B.A.; McCulloch, W.; Burdette, J.E.; et al. Pilot-scale production of expansile nanoparticles: Practical methods for clinical scale-up. J. Control. Release 2021, 337, 144–154. [Google Scholar] [CrossRef] [PubMed]
  270. Date, T.; Nimbalkar, V.; Kamat, J.; Mittal, A.; Mahato, R.I.; Chitkara, D. Lipid-polymer hybrid nanocarriers for delivering cancer therapeutics. J. Control. Release 2018, 271, 60–73. [Google Scholar] [CrossRef] [PubMed]
  271. Mandal, B.; Bhattacharjee, H.; Mittal, N.; Sah, H.; Balabathula, P.; Thoma, L.; Wood, G. Core–shell-type lipid–polymer hybrid nanoparticles as a drug delivery platform. Nanomed. Nanotechnol. Biol. Med. 2013, 9, 474–491. [Google Scholar] [CrossRef] [PubMed]
  272. Salvador-Morales, C.; Zhang, L.; Langer, R.; Farokhzad, O.C. Immunocompatibility properties of lipid–polymer hybrid nanoparticles with heterogeneous surface functional groups. Biomaterials 2009, 30, 2231–2240. [Google Scholar] [CrossRef] [Green Version]
  273. Zhao, P.; Wang, H.; Yu, M.; Liao, Z.; Wang, X.; Zhang, F.; Ji, W.; Wu, B.; Han, J.; Zhang, H.; et al. Paclitaxel loaded folic acid targeted nanoparticles of mixed lipid-shell and polymer-core: In vitro and in vivo evaluation. Eur. J. Pharm. Biopharm. 2012, 81, 248–256. [Google Scholar] [CrossRef]
  274. Quirós-Fallas, M.I.; Wilhelm-Romero, K.; Quesada-Mora, S.; Azofeifa-Cordero, G.; Vargas-Huertas, L.F.; Alvarado-Corella, D.; Mora-Román, J.J.; Vega-Baudrit, J.R.; Navarro-Hoyos, M.; Araya-Sibaja, A.M. Curcumin Hybrid Lipid Polymeric Nanoparticles: Antioxidant Activity, Immune Cellular Response, and Cytotoxicity Evaluation. Biomedicines 2022, 10, 2431. [Google Scholar] [CrossRef]
  275. Naziris, N.; Sekowski, S.; Olchowik-Grabarek, E.; Buczkowski, A.; Balcerzak, Ł.; Chrysostomou, V.; Pispas, S.; Małecka, M.; Bryszewska, M.; Ionov, M. Biophysical interactions of mixed lipid-polymer nanoparticles incorporating curcumin: Potential as antibacterial agent. Biomater. Adv. 2023, 144, 213200. [Google Scholar] [CrossRef]
  276. Abosabaa, S.A.; Arafa, M.G.; ElMeshad, A.N. Hybrid chitosan-lipid nanoparticles of green tea extract as natural anti-cellulite agent with superior in vivo potency: Full synthesis and analysis. Drug Deliv. 2021, 28, 2160–2176. [Google Scholar] [CrossRef]
  277. Patel, G.; Thakur, N.S.; Kushwah, V.; Patil, M.D.; Nile, S.H.; Jain, S.; Kai, G.; Banerjee, U.C. Mycophenolate co-administration with quercetin via lipid-polymer hybrid nanoparticles for enhanced breast cancer management. Nanomed. Nanotechnol. Biol. Med. 2019, 24, 102147. [Google Scholar] [CrossRef]
  278. El-Sawy, H.S.; Al-Abd, A.M.; Ahmed, T.A.; El-Say, K.M.; Torchilin, V.P. Stimuli-responsive nano-architecture drug-delivery systems to solid tumor micromilieu: Past, present, and future perspectives. ACS Nano 2018, 12, 10636–10664. [Google Scholar] [CrossRef] [PubMed]
  279. Ganta, S.; Devalapally, H.; Shahiwala, A.; Amiji, M. A review of stimuli-responsive nanocarriers for drug and gene delivery. J. Control. Release 2008, 126, 187–204. [Google Scholar] [CrossRef] [PubMed]
  280. Deirram, N.; Zhang, C.; Kermaniyan, S.S.; Johnston, A.P.R.; Such, G.K. pH-responsive polymer nanoparticles for drug delivery. Macromol. Rapid Commun. 2019, 40, 1800917. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  281. Hsu, P.-H.; Almutairi, A. Recent progress of redox-responsive polymeric nanomaterials for controlled release. J. Mater. Chem. B 2021, 9, 2179–2188. [Google Scholar] [CrossRef]
  282. Zhang, X.-X.; Eden, H.S.; Chen, X. Peptides in cancer nanomedicine: Drug carriers, targeting ligands and protease substrates. J. Control. Release 2012, 159, 2–13. [Google Scholar] [CrossRef] [Green Version]
  283. Shum, P.; Kim, J.-M.; Thompson, D.H. Phototriggering of liposomal drug delivery systems. Adv. Drug Deliv. Rev. 2001, 53, 273–284. [Google Scholar] [CrossRef]
  284. Chilkoti, A.; Dreher, M.R.; E Meyer, D.; Raucher, D. Targeted drug delivery by thermally responsive polymers. Adv. Drug Deliv. Rev. 2002, 54, 613–630. [Google Scholar] [CrossRef]
  285. Huang, S.-L. Liposomes in ultrasonic drug and gene delivery. Adv. Drug Deliv. Rev. 2008, 60, 1167–1176. [Google Scholar] [CrossRef]
  286. Sun, Q.; Bi, H.; Wang, Z.; Li, C.; Wang, X.; Xu, J.; Zhu, H.; Zhao, R.; He, F.; Gai, S.; et al. Hyaluronic acid-targeted and pH-responsive drug delivery system based on metal-organic frameworks for efficient antitumor therapy. Biomaterials 2019, 223, 119473. [Google Scholar] [CrossRef]
  287. Shen, W.; Liu, W.; Yang, H.; Zhang, P.; Xiao, C.; Chen, X. A glutathione-responsive sulfur dioxide polymer prodrug as a nanocarrier for combating drug-resistance in cancer chemotherapy. Biomaterials 2018, 178, 706–719. [Google Scholar] [CrossRef]
  288. Turk, B. Targeting proteases: Successes, failures and future prospects. Nat. Rev. Drug Discov. 2006, 5, 785–799. [Google Scholar] [CrossRef] [PubMed]
  289. Zhou, Y.; Chen, R.; Yang, H.; Bao, C.; Fan, J.; Wang, C.; Lin, Q.; Zhu, L. Light-responsive polymersomes with a charge-switch for targeted drug delivery. J. Mater. Chem. B 2019, 8, 727–735. [Google Scholar] [CrossRef]
  290. Al-Jawadi, S.; Thakur, S.S. Ultrasound-responsive lipid microbubbles for drug delivery: A review of preparation techniques to optimise formulation size, stability and drug loading. Int. J. Pharm. 2020, 585, 119559. [Google Scholar] [CrossRef] [PubMed]
  291. Chatterjee, S.; Hui, P.C.-L. Review of Stimuli-Responsive Polymers in Drug Delivery and Textile Application. Molecules 2019, 24, 2547. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  292. Chen, D.; Zhang, G.; Li, R.; Guan, M.; Wang, X.; Zou, T.; Zhang, Y.; Wang, C.; Shu, C.; Hong, H.; et al. Biodegradable, Hydrogen Peroxide, and Glutathione Dual Responsive Nanoparticles for Potential Programmable Paclitaxel Release. J. Am. Chem. Soc. 2018, 140, 7373–7376. [Google Scholar] [CrossRef]
  293. Li, D.; van Nostrum, C.F.; Mastrobattista, E.; Vermonden, T.; Hennink, W.E. Nanogels for intracellular delivery of biotherapeutics. J. Control. Release 2017, 259, 16–28. [Google Scholar] [CrossRef] [PubMed]
  294. Ye, Y.; Yu, J.; Gu, Z. Versatile Protein Nanogels Prepared by In Situ Polymerization. Macromol. Chem. Phys. 2015, 217, 333–343. [Google Scholar] [CrossRef]
  295. Howaili, F.; Özliseli, E.; Küçüktürkmen, B.; Razavi, S.M.; Sadeghizadeh, M.; Rosenholm, J.M. Stimuli-Responsive, Plasmonic Nanogel for Dual Delivery of Curcumin and Photothermal Therapy for Cancer Treatment. Front. Chem. 2021, 8, 602941. [Google Scholar] [CrossRef]
Figure 1. Different polymeric-based nanocarrier systems.
Figure 1. Different polymeric-based nanocarrier systems.
Polymers 15 01123 g001
Figure 2. Advantages and disadvantages of polymeric nanoparticles.
Figure 2. Advantages and disadvantages of polymeric nanoparticles.
Polymers 15 01123 g002
Figure 3. Classification of polymeric nanoparticles according to structural organization.
Figure 3. Classification of polymeric nanoparticles according to structural organization.
Polymers 15 01123 g003
Figure 4. Chemical structures of different polymeric materials used in fabrication of polymeric nanoparticles.
Figure 4. Chemical structures of different polymeric materials used in fabrication of polymeric nanoparticles.
Polymers 15 01123 g004
Figure 5. Schematic representation of frequently used techniques used to fabricate synthetic-based polymeric nanoparticles.
Figure 5. Schematic representation of frequently used techniques used to fabricate synthetic-based polymeric nanoparticles.
Polymers 15 01123 g005
Figure 6. Schematic representation showing the mechanisms by which natural bioactive agents can act as anticancer agents after release from polymeric nanoparticles.
Figure 6. Schematic representation showing the mechanisms by which natural bioactive agents can act as anticancer agents after release from polymeric nanoparticles.
Polymers 15 01123 g006
Table 1. Examples of recent studies utilizing natural polymer-based polymeric nanoparticles for loading natural bioactive agents.
Table 1. Examples of recent studies utilizing natural polymer-based polymeric nanoparticles for loading natural bioactive agents.
Polymeric CompositionLoaded Natural Bioactive AgentResearch OutcomesReference
Natural polymersChitosan/chitosan derivativesQuercetin
  • Efficient uptake by HaCaT cells.
  • Improvement of percutaneous absorption and skin retention.
  • Enhancement of anti-UVB effect.
[11]
Effective antimicrobial and antiadhesion effects in multidrug-resistant isolates (E. coli and S. aureus).[12]
Significant cytotoxic effect against human breast tumor (MCF-7) and human lung tumor (A549).[13]
Improvement of wound healing by modulation of cytokines and growth factors involved in inflammatory and proliferative phases of wound healing.[14]
CurcuminImproved antimicrobial activities against C. glabrata and A. niger.[15]
Improved anticancer activity against Vero cell line.[16]
Improved neuroprotective effect of curcumin and curcumin-loaded chitosan nanoparticles against stress-induced neurobehavioral and neurochemical deficits and protection against stress-associated gastric ulcer. [17]
  • Improved mucoadhesion, cellular uptake, and cytotoxicity against Caco-2 cells.
  • Improved antioxidant and anti-inflammatory effects in activated RAW264.7 cells.
[18]
Higher suppression of hepatocellular carcinoma growth in murine xenograft models and inhibited tumor angiogenesis when compared to free curcumin.[19]
ResveratrolLarger AUC0-6 (6.44-fold) and higher mean residence time in plasma (2.458-fold) of the optimized formulation when compared to of resveratrol solution.[20]
Higher solubility and stability of resveratrol-loaded nanoparticles.[21]
  • Negligible in vitro release of resveratrol in simulated gastric.
  • Sustained release in simulated intestinal conditions.
  • Higher relative bioavailability of resveratrol.
[22]
Improved anticancer activity against HepG2 cells.[23]
  • Significant improvement in the in vivo permeation of resveratrol.
  • Improvement in the human stratum corneum penetration of resveratrol after topical application.
[24]
Silymarin
  • Significantly improved depressive-like behaviors.
  • Significantly reduced levels of malondialdehyde (MDA) and expression of interleukin-6 (IL-6) and tumor necrosis factor-α (TNF-α).
  • Significant increase in the activities of superoxide dismutase (SOD), catalase (CAT), glutathione peroxidase, glutathione reductase, and glutathione (GSH) levels in I/R brain.
[25]
SaponinNanosaponin showed higher uptake and specific toxicity on PC3 and KB cell lines. [26]
Paclitaxel Safe and High loading efficiency and encapsulation efficiency.[27]
Essential oilsImproved antifungal activity against Aspergillus versicolor, A. niger, and Fusarium oxysporum. [28]
  • Improved antioxidant activity
  • Improved antibacterial activity against L. monocytogenes and S. aureus.
[29]
AlginateQuercetinEffective down-regulating the inflammation-related gene expression of tumor necrosis factor-α, interleukin-6, inducible nitric oxide synthase, and monocyte chemotactic protein-1. [30]
Curcumin
  • Prolonged release of curcumin.
  • Enhanced collagenesis and increased number of fibroblasts.
  • Improved wound healing efficacy.
[31]
  • Amelioration of colonic inflammation and tissue damage by inhibiting the TLR4 expression in colonic epithelial cells.
  • Reducing the transcription and expression of the pro-inflammation cytokines downstream.
[32]
GelatinResveratrol
  • Greater ROS generation, DNA damage, and apoptotic incidence.
  • Higher bioavailability and a longer half-life than free resveratrol.
[33]
High loading efficiency and superior efficacy in NCI-H460 cells.[34]
AlbuminSilymarin
  • Enhancement in the viability of hepatocytes in vitro against both APAP and LPS/D-GalN-induced hepatocyte damage.
  • Improvement of antioxidant effects against intracellular oxidative stress.
[35]
Curcumin Enhanced cytotoxicity on triple-negative human breast cancer cells (MDA-MB-231) compared to free curcumin. [36]
Resveratrol Potent activity against oxidative stress-based diseases.[37]
Paclitaxel/resveratrol
  • High encapsulation efficiency.
  • Sustained drug release profiles.
[38]
Table 2. Examples of recent studies utilizing synthetic polymer-based polymeric nanoparticles for loading natural bioactive agents.
Table 2. Examples of recent studies utilizing synthetic polymer-based polymeric nanoparticles for loading natural bioactive agents.
Polymeric CompositionLoaded Natural Bioactive AgentResearch OutcomesReference
Synthetic polymersPLAPaclitaxel
  • Suppression of the tumor growth in mice.
  • Better tumor suppression trend with conjugated resveratrol nanoparticles with no side effects.
[61]
Significant therapeutic improvement in arthritis proved by measuring rats’ knee diameter as well as the tumor necrosis factor-alpha (TNF-α).[62]
Effective targeting of folate-decorated paclitaxel-loaded copolymer nanoparticles on cancer cells both in vitro and in vivo.[63]
Improved anti-tumoral activity of paclitaxel when compared to the commercial paclitaxel formulation Taxol®.[64]
  • Enhanced cellular uptake in both human umbilical vein endothelial cells and rat C6 glioma cells.
  • Increased cytotoxicity and improved penetration and growth inhibition in avascular C6 glioma spheroids.
[65]
QuercetinEnhanced anticancer efficacy in terms of its sustained release kinetics revealing novel vehicle for the treatment of cancer.[66]
  • Initial burst release followed by slow and sustained release.
  • High antioxidant activity.
[67]
Good antibacterial effects against Staphylococcus aureus (S. aureus), Escherichia coli (E. coli), and Klebsiella pneumoniae (K. pneumoniae).[68]
CurcuminAmelioration of the negative changes in diabetes with a more pronounced treated effect than free curcumin. [69]
Cytocompatibility and suppressed production of TNF-α.[70]
CarvacrolEnhanced in vitro antimicrobial activity against Escherichia coli, Listeria monocytogenes, Salmonella enterica, and Staphylococcus aureus. [71]
PLGAGrape extractsSustained release with good stability in GIT fluids.[72]
Curcumin and docetaxel
  • Significant increase in cytotoxic activity.
  • Significantly improved brain penetration.
[73]
Curcumin Higher percentage of curcumin internalization in cells, leading to enhanced cancer cell killing with targeted nanoparticles. [74]
  • Increased curcumin means half-life and high Cmax.
  • Improved oral bioavailability.
[75]
QuercetinProduction of thermosensitive gel as cost-effective option for burn wound therapy.[76]
Improved anti-oxidant activity.[77]
Significant blockage of UVB irradiation-induced COX-2 up-expression and NF-kB activation in Hacat cell line. [78]
Inhibition of the neurotoxicity of Zn2+-Aβ42 system and enhancement of the viability of neuron cells. [79]
Plant extract
  • Decreased level in the inflammation-related gene expression and DNA fragmentation level.
  • High survival rate of post-damaged corneal epithelial cells.
  • Enhanced corneal wound healing and decreased inflammation.
[80]
PCLEssential oilPowerful antibacterial and antifungal activities that may be utilized as dermal patches for acne treatment.[81]
  • Prompt antioxidant activity.
  • Inhibition of Staphylococcus aureus and Escherichia coli growth.
[82]
Curcumin
  • Increase in the cell population of Sub G1 and apoptosis.
  • Decreased tumor cell proliferation and angiogenesis.
[83]
  • Enhanced antibacterial efficacy against both Gram-positive and Gram-negative strains.
  • Proficient scar-less wound healing activity within 21 days with marked re-epithelization of tissue, collagen deposition, and formation of granulation tissue.
[84]
PAMAMPaclitaxel and curcumin
  • High encapsulation efficiency.
  • Improved bioavailability and bioactivity against skin cancer.
[85]
MethacrylateEpigallocatechin gallateEnhanced antimicrobial activity of EGCG-attached polymer that can be used as an alternative strategy to preclude microbial colonization.[86]
Curcumin
  • Enhanced tumor cell regression activity when compared to free curcumin.
  • Significant reduction in G0/G1 cells.
  • High biocompatibility.
[87]
Silymarin
  • High capability to reverse the induced liver fibrosis and decrease serum TNF-α, serum TGF-β1, and hepatic hydroxyproline.
  • Down-regulation in the expression of different hepatic biomarkers.
  • Very little collagen of livers in extracellular matrix and restored hepatic architecture.
[88]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Elmowafy, M.; Shalaby, K.; Elkomy, M.H.; Alsaidan, O.A.; Gomaa, H.A.M.; Abdelgawad, M.A.; Mostafa, E.M. Polymeric Nanoparticles for Delivery of Natural Bioactive Agents: Recent Advances and Challenges. Polymers 2023, 15, 1123. https://doi.org/10.3390/polym15051123

AMA Style

Elmowafy M, Shalaby K, Elkomy MH, Alsaidan OA, Gomaa HAM, Abdelgawad MA, Mostafa EM. Polymeric Nanoparticles for Delivery of Natural Bioactive Agents: Recent Advances and Challenges. Polymers. 2023; 15(5):1123. https://doi.org/10.3390/polym15051123

Chicago/Turabian Style

Elmowafy, Mohammed, Khaled Shalaby, Mohammed H. Elkomy, Omar Awad Alsaidan, Hesham A. M. Gomaa, Mohamed A. Abdelgawad, and Ehab M. Mostafa. 2023. "Polymeric Nanoparticles for Delivery of Natural Bioactive Agents: Recent Advances and Challenges" Polymers 15, no. 5: 1123. https://doi.org/10.3390/polym15051123

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop