Next Article in Journal
Synthesis, Docking, 3-D-Qsar, and Biological Assays of Novel Indole Derivatives Targeting Serotonin Transporter, Dopamine D2 Receptor, and Mao-A Enzyme: In the Pursuit for Potential Multitarget Directed Ligands
Next Article in Special Issue
The Influence of UV Radiation on the Degradation of Pharmaceutical Formulations Containing Quercetin
Previous Article in Journal
The Chemical Composition of Brazilian Green Propolis and Its Protective Effects on Mouse Aortic Endothelial Cells against Inflammatory Injury
Previous Article in Special Issue
Investigating the Modes of Action of the Antimicrobial Chalcones BC1 and T9A
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Nanoformulations to Enhance the Bioavailability and Physiological Functions of Polyphenols

Jiangsu Provincial Key Lab for the Chemistry and Utilization of Agro-Forest Biomass, Jiangsu Key Lab of Biomass-Based Green Fuels and Chemicals, Jiangsu Co-Innovation Center of Efficient Processing and Utilization of Forest Resources, College of Chemical Engineering, Nanjing Forestry University, Nanjing 210037, China
*
Author to whom correspondence should be addressed.
Molecules 2020, 25(20), 4613; https://doi.org/10.3390/molecules25204613
Submission received: 16 September 2020 / Revised: 4 October 2020 / Accepted: 6 October 2020 / Published: 10 October 2020

Abstract

:
Polyphenols are micronutrients that are widely present in human daily diets. Numerous studies have demonstrated their potential as antioxidants and anti-inflammatory agents, and for cancer prevention, heart protection and the treatment of neurodegenerative diseases. However, due to their vulnerability to environmental conditions and low bioavailability, their application in the food and medical fields is greatly limited. Nanoformulations, as excellent drug delivery systems, can overcome these limitations and maximize the pharmacological effects of polyphenols. In this review, we summarize the biological activities of polyphenols, together with systems for their delivery, including phospholipid complexes, lipid-based nanoparticles, protein-based nanoparticles, niosomes, polymers, micelles, emulsions and metal nanoparticles. The application of polyphenol nanoparticles in food and medicine is also discussed. Although loading into nanoparticles solves the main limitation to application of polyphenolic compounds, there are some concerns about their toxicological safety after entry into the human body. It is therefore necessary to conduct toxicity studies and residue analysis on the carrier.

1. Introduction

Many effective medical treatments have originated from plant extracts, which are important sources of materials for the treatment of many diseases [1]. Phenolic compounds are widely present as secondary metabolites in all vascular plants [2,3,4,5]. They play important roles in the growth and development of plants, and are involved in defense against ultraviolet light and pathogens [6]. In recent years, because of their potential positive role in human metabolism, they have attracted increasing attention and research. These compounds have biological properties that include antioxidant, anti-inflammatory, antibacterial, anticancer and cardiovascular protection activities [7,8,9].
However, the use of phenolic compounds in humans is limited by many factors, such as low solubility, poor permeability, instability, rapid release, susceptibility to environmental influences and low bioavailability [10,11]. In order to overcome these limitations, polyphenols are often loaded into various carriers to enhance their bioavailability. This can increase biocompatibility, prevent degradation caused by the external environment, and prevent interaction with other components in the human body. Nanocarriers have been demonstrated to be excellent materials for encapsulating phenolic compounds and improving their bioavailability.
With the rapid development of nanotechnology in the food and pharmaceutical industries, many advanced nanoparticles have been developed to protect and control/target the release of biologically active ingredients, including various polyphenols [12,13,14,15]. The size of nanoparticles and nanocarriers is in the range of 1–100 nm [16]. By loading phenolic compounds into nanoparticles, not only can their bioavailability be improved, but also the controlled/targeted release and protection of active substances can be achieved [17]. In recent years, many nanoparticles have been developed for the delivery of polyphenolic compounds, including liposomes, phospholipid complexes, niosomes, protein-based nanoparticles, micelles, emulsions and metal nanoparticles.
This review first summarizes the biological activities of polyphenolic compounds, and then introduces the application of a variety of nanoparticles to improve their bioavailability (Figure 1).

2. Bioactivities of Polyphenols

With the continuous emergence of the benefits of polyphenols, there is growing interest in the study of its biological activities, such as antioxidant, heart protection, cancer prevention and nerve protection. Polyphenols play an irreplaceable role in medicine, food, health products and other fields.

2.1. Antioxidant Activity

Among all biological activities of phenolic compounds, antioxidant activity is one of its most important activities, and has been extensively studied, including scavenging free radicals, inhibiting the oxidation of lipids and reducing the formation of hydrogen peroxide, etc. Phenolic compounds have significant antioxidant properties, as the structure contains a large number of hydroxyl groups, which has a great influence on the ability to scavenge free radicals and chelate metal ions [18,19]. Polyphenols can neutralize free radicals by providing electrons or hydrogen atoms to a variety of reactive oxygen species (ROS). Acting as metal ion chelating agents, polyphenols can transform peroxides or metal oxides into stable substances and thus disrupt the proliferation stage of lipid autotrophic chain reactions [20,21]. Previous studies have shown that some flavonoids can directly scavenge peroxides, while others have the ability to scavenge a strong oxygen-derived free radical (such as peroxynitrite) to inhibit low-density lipoprotein (LDL) oxidation [22]. It is true that phenolic compounds are good antioxidants, however, when losing electrons or acting as reducing agents, the molecules themselves will be converted into free radicals, and the interactions with transition metal ions will also lead to the formation of pro-oxidants [23]. Therefore, it is necessary to understand correctly its activity and control its dosage reasonably.
The balance of the oxidative defense mechanism in human cells is maintained by the enzymatic redox system, which mainly includes catalase (CAT), hydrogen peroxide dismutase (SOD), glutathione peroxidase (GPx), glutathione reductase (GR) and peroxidase (PRXs) [24]. However, under the oxidative stress condition of excessive reactive oxygen, the defense ability will be greatly reduced [25]. It has been proved that polyphenols can restore equilibrium by increasing the activity of SOD, CAT, GPx, GR and PRXs, thus preventing systemic or local inflammation [26]. The antioxidant activity of plant polyphenol extracts has been extensively studied in different biological systems (Table 1).

2.2. Cardioprotection Activity

Cardiovascular disease (CVD), also known as circulatory disease, is a range of diseases that involve the circulatory system, including coronary artery disease, stroke, heart failure and high blood pressure. Numerous studies have shown that polyphenols have beneficial effects on cardiovascular health, and consuming foods rich in polyphenols can reduce the risk of these diseases [54,55,56,57].
Wang et al. [58] reported that anthocyanins, proanthocyanidins, flavonoids, and flavan-3-ols in polyphenols are closely related to the incidence of CVD. Increasing the daily intake of flavonols by 10 mg can reduce the risk of cardiovascular disease by 5%. The meta-analysis by Menezes et al. [59] also showed that the intake of flavonols reduced the risk of CVD, but the impact of flavonols intake on lipid levels may vary depending on the country and health status of individuals.
Tang et al. [60] demonstrated in a study on the risk of stroke that ingesting 100 mg of flavonoids per day reduced the stroke rate by approximately 10%. A similar study conducted by Wang et al. [61] demonstrated that an intake of 20 mg of flavonols per day was associated with a 14% reduction in stroke risk. Studies have shown that the main polyphenolic compounds that can significantly reduce the risk of hypertension are flavonoids [62,63,64]. Through a meta-analysis of randomized controlled trials, Huang and coworkers concluded that participants who took quercetin for 8 weeks or longer showed significant changes in HDL cholesterol and triglyceride levels [65]. Important evidence from epidemiological investigations suggests that dietary polyphenols may treat and prevent type 2 diabetes. Cao and coworkers have demonstrated that resveratrol can improve blood glucose control in subjects with insulin resistance, and anthocyanins can lower blood glucose or optimize insulin secretion and resistance [66]. It is reported that genistein could significantly improve glycemic control and sensitivity to insulin in postmenopausal women [67].
Atherosclerosis is a chronic inflammatory disease that occurs in vulnerable areas of middle arteries. It may have existed for many years before it develops and produces physiological states, such as acute myocardial infarction, unstable angina, or sudden cardiac death [68]. Anthocyanins can increase blood lipids, inhibit inflammation, and improve endothelium-dependent vasodilation by activating NO-cGMP signaling pathway [69]. It is also demonstrated that, by inhibiting the activity of phosphodiesterase-5, the blood vessels can be relaxed, thus reducing the risk of cardiovascular disease [70].
The formation of thrombosis is one of the main causes of myocardial infarction, ischemic heart disease and other diseases. Plenty of evidence has proven that dietary polyphenols can reduce the formation of thrombus [71,72]. Red wine, rich in polyphenols (such as resveratrol, proanthocyanidins, etc.), is known to have the function of protecting the heart. Extensive data have shown that resveratrol can prevent thrombosis by reducing oxidative stress and platelet aggregation, thereby protecting blood vessels [73,74].

2.3. Anticancer Activities

Cancer is a cell proliferative disease caused and controlled by genetic mutations, resulting from various factors, such as diet, radiation, unhealthy life habits (such as regular drinking and smoking) and infectious microorganisms [75]. Considerable evidence has proven that it is the plant secondary metabolites, polyphenols, that make a significant contribution to anticancer activity [76,77,78,79]. Polyphenols are mainly used to prevent cancer by inhibiting cancer cell activation, promotion, angiogenesis, and strengthening the immune system [80,81,82]. Polyphenols have preventive and therapeutic effects on many cancers, such as esophageal, gastric, colon, liver, lung, breast, ovarian and skin cancer (Table 2) [83,84,85].
Grosso et al. [86] conducted a systematic evaluation of the role of polyphenols in cancer prevention, and the analysis showed that isoflavones significantly reduced the risk of gastric, lung, breast and colorectal cancers, by directly inhibiting oxidative stress, oxidative damage, anti-angiogenesis and anti-metastasis, while total flavones had little effect on the risk of breast cancer. Another study reviewed the role of flavonoids in colon cancer risk and found that the sustained consumption of flavonoids appeared to be an effective complementary treatment that reduced colon cancer risk from cellular antioxidants and anti-inflammatory effects [87]. The study also revealed that the flavonoid subclasses had a significant synergistic effect on preventing tumorigenesis, tumor growth and promoting apoptosis of cancer cells [87]. Similar studies by Chang and coworkers proved that the intake of quercetin and other flavonol compounds could reduce the risk of colon cancer, while the intake of apigenin and other flavonoids could reduce the risk of rectal cancer [88].
Table 2. Anti-cancer effects of some polyphenols.
Table 2. Anti-cancer effects of some polyphenols.
PolyphenolCancer Type/
Cell Line
Major OutcomesReferences
CurcuminMBA-MB-231cells, MCF-7 cellsDown-regulate the mRNA expression of Vimentin, Fibronectin, and β-catenin; up-regulate E-cadherin mRNA expression levels[89]
HCT-116 cellsReduce the expression of SIRT1 protein, suppress the oncogenicity of human-colon cancer cells [90]
T98G, U87MG, T67 cells, HCT-116 cellsInhibit AP-1 and NF-κB signaling pathways, suppress JNK activation induced by carcinogens[91]
Resveratrol LNCaP cells Induce the expression of COX-2, promoting ERK1/2 activation, and facilitate p53-dependent anti-proliferation gene expression[92]
NSCLC cellsPrevent tumorigenesis and progression, and down-regulate EGFR/Akt/ERK1/2 signaling pathway[93]
Apc10.1 cellsShow superior efficacy than high doses due to the pro-oxidant activity and AMPK signaling upregulation[94]
Hela cellsInhibit the expression of PLSCR1, leading to the growth inhibition of HeLa cells[95]
SGC7901, BGC823 cellsInhibit the invasion and migration of human gastric cancer cells [96]
QuercetinMDA-MB-231 cellsIncrease FasL mRNA expression and p51, p21, and GADD45 signaling activities, induce protein level, transcriptional activity, and nuclear translocation of Foxo3a[97]
AsPC-1, CRL-4023, PANC-1 cellsReduce the expression levels of cellular FLICE-like inhibitory protein, activate c-Jun N-terminal kinase (JNK)[98]
A549 cellsTrigger BCL2/BAX-mediated apoptosis, as well as necrosis and mitotic catastrophe[99]
PC-3 cellsDecrease tumor improvement, down-regulate Ki67, and enhance caspase 7[100]
EGCGBreast T47DUp-regulate PTEN, CASP3, CASP9, down-regulate AKT[101]
GenisteinPancreatic Mia-PaCa2 Induce mitochondrial apoptosis, block cell cycle and regulate STAT3[102]
Colorectal HCT 116Inhibite cell proliferation, induce apoptosis of colorectal cancer cells[103]
DaidzeinColorectal HT-29, MIA PaCa-2Cytotoxic effects on both MIA PaCa-2 and HT-29 cell lines [104]
Ovarian SKOV3Up-regulate B-cell lymphoma 2-associated X protein, cytochrome c, down-regulate pCdc25c, Cdc25c[105]
BEL-7402Increased the levels of reactive oxygen species (ROS) and induce a decrease in mitochondrial membrane potential[106]
ChrysinHCT-116;
HepG2; Hep 3B
The combination of chrysin and cisplatin promoted apoptosis of HepG2 cells in both dose- and time- dependent manners[107]
A549Reinforce the therapeutic efficacy of DTX and mitigate
edema
[108]

2.4. Neuroprotective Activity

With the continuous improvement of medical standards, the average life span of humans has been significantly extended, but correspondingly, diseases related to brain aging caused by an aging population have also increased significantly, such as cognitive and neurodegenerative diseases [109]. Neurodegenerative diseases, characterized by the progressive loss of functions of a large number of neurons and neural stem cells, leading to sensory deficits and cognitive impairments, are a type of progressive, disabling and severely fatal complex disease [110]. Studies have found that all neurodegenerative diseases share a common cellular and molecular mechanism, that is, oxidative stress accumulation, inflammation, protein misfolding and aggregation, neurotoxicity, etc. [111,112,113].
In addition to genetic and environmental factors, the increase of oxidative stress in cells is considered to be the main cause of neurodegenerative diseases [114,115]. A large number of studies have shown that polyphenols can inhibit the increase of oxidative stress through many mechanisms. For example, polyphenols can enhance the activity of detoxification and antioxidant enzymes by activating the Nrf2 pathway [116,117,118]. They can also regulate the activity of reactive oxygen generation enzymes and modify the structural integrity and metabolic efficiency of mitochondria [119,120]. Inflammation is also the cause of such diseases. Polyphenols can regulate the expression of pro-inflammatory genes such as nitric oxide, lipoxygenase, cyclooxygenase (COX) and chemokines [115,121]. In addition, the neuroprotective effect of polyphenol compounds is also attributed to the reduction of amyloid aggregation and/or the formation of precursors. Curcumin has been shown to have anti-amyloidosis activity, and can not only inhibit the formation of new β-aggregates, but also decompose already formed aggregates [122]. Pomegranate polyphenols, myricetin, luteolin and honokiol variably altered the morphology of Aβ aggregation, the flavonoids all bound in a similar hydrophobic region of the amyloid pentamer and exhibited the most obvious inhibitory effect on Aβ1-42 aggregation [123].

3. Nanoformulations for Loading and Delivery of Polyphenols

As mentioned earlier, polyphenols have been widely concerned and applied in many fields due to various beneficial functions, but some of their restrictive factors have greatly hindered their applications, in vivo and in clinic. These factors mainly include low solubility, permeability, and bioavailability. In order to overcome the limitations, nanocarriers have been developed extensively [124]. The unique physicochemical properties of nanocarriers, such as high loading, drug protection and tumor cell penetration, provide preconditions for the delivery of polyphenols and other drugs [125,126]. At present, the delivery systems like phytosomes, liposomes, niosomes, protein-based nanoparticles, polymer nanoparticles, microspheres and emulsions have emerged as attractive options for controlled bioactive systems [127,128].

3.1. Phytosome

Phytosome is a relatively stable complex formed by electrostatic interaction between phospholipids, (mainly phosphatidylcholine), and plant extracts (mainly polyphenols) (Figure 2) [129]. This electrostatic effect mainly includes ion-dipole, dipole-dipole and hydrogen bonding [130]. Phospholipid complexes are more bioavailable than purified extracts because the presence of phosphatidyl cholines, a major component of cell membranes, enhances the ability of plant extracts to circulate in the body [131].
Phytosomes were developed in the late 1980s by a company called Indena, which developed a way to increase the bioavailability of drugs by complexing them to phospholipids. Many of the drug extracts currently on the market are in the form of phospholipid complexes. Phytosome technology has greatly improved the bioavailability of plant extracts, especially polyphenols. Yang et al. [132] used a simple method to prepare the rosmarinic acid-phospholipid complex (RA-PLC). The study showed that compared with natural RA, the membrane permeability and antioxidant properties of RA-PLC are significantly improved, and the biological utilization is 1.2 times higher than RA. Ravarotto et al. [133] reported that the silibinin-phospholipid complex has better anti-hepatotoxic activity than silibinin alone, and it can reduce the toxic effect of aflatoxin B1 on broilers. Similarly, there have been numerous reports that curcumin-phospholipid complexes can significantly increase bioavailability, improve pharmacokinetics, and enhance liver protection [134,135]. In the study of Marczylo et al. [136] rats given the same oral dose of free curcumin and curcumin–phospholipid complex (340 mg/kg) were found to have 5 times more drug content in plasma than the other group after 2 h. Pharmacokinetic studies in rats in another study showed that oral curcumin–phospholipid complexes had a longer half-life than oral free drugs [137]. Quercetin and polyphenol extracts from Moringa oleifera leaf have the function of improving bioavailability after complexing with phospholipids [138,139].
The application of polyphenol–phospholipid complexes in anti-cancer has gradually emerged. Narges Mahmoodi found that both silibinin and silibinin phosphatidylcholine can down-regulate the expression of HER2 on SKBR3 breast cancer cells, but silibinin–phospholipid complex had much greater inhibitory effect on cancer cells than natural silibinin (approximately 2 ~ 2.5 times) [140]. A list of polyphenols complexed with phospholipids is shown in Table 3.

3.2. Lipid-Based Nanoparticles

Liposome-based nanoparticles are spherical lipid particles widely used for drug delivery. The ability to encapsulate water-soluble, lipid-soluble and amphiphilic substances makes them ideal carriers for many kinds of drugs. These kinds of nanoparticles mainly include nanoliposomes and solid lipid nanoparticles [14].

3.2.1. Liposomes

The term liposome is composed of two Greek words, lipos (fat) and soma (body or structure), meaning a membrane of fat (mainly phospholipids and cholesterol) that surrounds a water-soluble carrier or compartment [152]. Liposomes are self-assembled amphiphilic spherical vesicles with at least one phospholipid bilayer, similar to the bilayer shape on the cell membrane, which can separate the internal and external aqueous medium [14,153] (Figure 3). Liposomes vary in size, ranging from 50 nm to 1 μm, depending on composition and preparation methods. The presence of such an amphiphilic substance allows water-soluble, fat-soluble and amphiphilic compounds to be encapsulated, transferred and released [154].
There are many methods for preparing liposomes. The traditional methods include thin film hydration, reverse evaporation, injection, and heating [155]. The biggest drawback of these technologies is that the liposomes formed are large in size. The methods currently used include membrane contactor-based method, freeze drying of double emulsion method, proliposome method [156].
With the ability to improve biocompatibility, prevent drug degradation, deliver low-solubility drugs, and improve drug targeting [157,158,159], liposomes are undoubtedly a kind of promising and flexible polyphenol delivery system. In addition, it has been recognized in clinical and biological research for its role in human health, including protecting the liver, enhancing memory and reducing cholesterol intake [152]. With the deep understanding of liposomes and polyphenols, the combination of liposomes and polyphenols has attracted much attention.
Cheng reported that nano-liposomes greatly improved the bioavailability and water solubility of curcumin [160]. Vanaja et al. [161] used a thin-film method to load resveratrol into liposomes. Compared with free resveratrol, resveratrol packed in liposomes was more active in cells in vivo and had more significant antioxidant effect. A novel remote loading approach using chemically modified β-cyclodextrin was applied to incorporate curcumin into liposomes [162]. The results proved that the encapsulation of the nanoparticles significantly improved the bioavailability of curcumin, and the complexation of cyclodextrin further increased the encapsulation efficiency of curcumin.
Huang and coworkers co-loaded curcumin and resveratrol in liposomes [163]. Infrared spectroscopy and fluorescence techniques demonstrated that curcumin was connected to the hydrophobic acyl chain region of liposomes, while resveratrol was located in the polar hydrophilic region. The study also investigated the physical and chemical properties of the liposomes. The results showed that, when the ratio of curcumin to resveratrol was 5:1, the encapsulation rate and the antioxidant activity was the highest. Similarly, co-loading quercetin and resveratrol, using liposome as the nanocarrier, enhanced the cellular uptake and ROS scavenging ability than single drug loading [164].
However, the limitation of liposome is that it has a short half-life and can be easily oxidized and hydrolyzed. It was previously reported that the intravenous administration of resveratrol resulted in a short t1/2, ranging from 7.8 to 33 min [165]. Therefore, many modified liposomes came into being. In order to increase the half-life and extend the blood circulation time, Caddeo et al. [166] grafted polyethylene glycol chains (PEGylated liposomes) at the ends of liposomes to deliver resveratrol. Drug release results indicated that the half-life of polyethylene glycol modified liposomes was increased by about nine times, while the inherent antioxidant activity was still maintained. By mixing with various biopolymers, the storage stability of liposomes was improved [167]. The biopolymers include anionic (such as arabinose and whey protein) and cationic (such as chitosan). It was also proven that the encapsulation rate of polyphenols in liposomes and their antioxidant activity will increase with the addition of biopolymers.

3.2.2. Solid Lipid Nanoparticles (SLNs)

Solid lipid nanoparticle is a solid colloidal drug delivery system which is made by wrapping or inserting drugs in lipid nucleus with natural or synthetic lipids, such as lecithin and triglyceride. The emergence and development of SLNs successfully make up for the limitations of traditional nanocarriers (such as nanoemulsion and liposomes), with advantages including controlling drug release and targeting, increasing drug stability, high drug loading, low toxicity, and loading of hydrophilic lipophilic drugs.
SLNs coated with resveratrol were prepared by Pandita and coworkers using solvent diffusion - solvent evaporation method [168]. The encapsulation rate of resveratrol in nanoparticles was 88.9±3.1%, and the release time in vitro could be extended to 120 h. It was also found that the use of solid lipid formulations increased the bioavailability of oral resveratrol by eight times compared to drug suspensions. Similarly, glyceryl behenate-based SLNs were used for the encapsulation of resveratrol to explore its brain targeting ability [169]. Cytotoxicity experiments demonstrated that SLNs had the same antitumor effect as free resveratrol, while the drug biodistribution in mice indicated that SLN greatly increased the concentration of resveratrol in brain cells.
Curcumin was also incorporated within chitosan coated SLNs by homogenization and ultra-sonication technique [170]. It showed that the oral bioavailability of nanoparticles was more improved than that of curcumin suspension, and the reason was not only that the encapsulation prevented the drug from being degraded by enzymes, but also that the chitosan could be easily absorbed. The production of such formulations could further expand the use of curcumin in food and nutraceuticals.
SLN was also used to improve the bioavailability of quercetin by carrying transferrin that enabled it to be delivered to the brain at designated points to study the role in Alzheimer’s disease [171]. It was found that the nanoparticles could inhibit the formation of fibrils and reduce the amyloid accumulation of peptides.

3.3. Niosomes

Liposomes have the disadvantages of high production cost, poor chemical stability, and impure phospholipid content. With similar structures to liposomes, niosomes can perfectly avoid those disadvantages and become promising polyphenol delivery systems that have been used for continuous, controllable and targeted drug delivery of polyphenols and other drugs [172,173,174]. Most of the niosomes are composed of non-toxic non-ionic surfactants (mainly alkylamides, alkyl esters, and fatty acid esters), and contain cholesterol or its derivatives and charged molecules. The presence of cholesterol increases the rigidity of nanoparticles, while the presence of charged molecules contributes to the stability during the preparation (Figure 4) [174,175,176].
The preparation methods of niosomes include the ethanol injection method, transmembrane pH gradient method, reverse phase evaporation, microfluidization, lipid layer hydration, etc. Due to the unique structure, niosomes can be used to load and deliver hydrophilic and hydrophobic substances. Lu et al. designed Span60-Rh40-based niosomes for loading flavonoids to improve the solubility, stability and penetration [177]. It was demonstrated that niosomes significantly improved the solubility and light stability of quercetin. The skin water-locking effect of quercetin-niosomes was almost three times higher than that of free quercetin solution. Another group extracted polyphenol-rich propolis with ethanol and prepared propolis-loaded niosomes (PLNs) with different concentrations of Span 60 and cholesterol to enhance the local antibacterial effect of propolis. The use of niosomes significantly enhanced the antibacterial activity of propolis against Staphylococcus aureus and Candida albicans. The enhancement of antibacterial activity attributes to the fact that niosomes can directly interact with the bacterial envelope, thus facilitating the entry of antibacterial components in propolis into cells [178]. To increase chemotherapeutic efficacy while reducing toxic effects, a rational design for synergy-based drug regimens is essential. Alemi et al. found that the combination therapy of paclitaxel with curcumin using PEGylated noisome delivery enhanced cytotoxicity against MCF-7 cells [179].

3.4. Protein-Based Nanoformulations

Protein-based drug delivery system has high nutritional value and good functional characteristics, making it continue to develop in the food industry.

3.4.1. Casein-Based Nanoparticles

Casein is the predominant protein in milk (approximately 80%), composed of αs1-, αs2-, β-, and κ-caseins, with unique hydrophilic and hydrophobic domains. These proteins self-assemble in the presence of calcium phosphate to form a spherical colloid with a diameter of 50–500 nm (average diameter 150 nm). Because the casein structure is rich in proline and can adapt to changes in its environment, it is defined as a rheological protein [180,181].
Casein has many properties that favor its use in drug delivery systems, including exceptional surface activity, excellent emulsification and self-assembly properties, and excellent water binding ability [182]. In addition, casein can interact with many molecules to form stable complexes and conjugates. These unique properties make it an excellent choice for drug delivery systems.
Studies have shown that β-casein encapsulation increases the solubility of curcumin by at least 2500-fold, and the presence of this casein micellar protein enhances the cytotoxicity of the drug to human leukemia cells K-562. Furthermore, curcumin–casein showed significantly higher antioxidant activity than free curcumin [183]. Luo et al. [184] prepared rutin-sodium caseinate/pectin complex nanoparticles by acidification and heat treatment. It was found that heating not only improved the rate of nanoparticles formation, but also significantly improved the rate of rutin encapsulation. The presence of pectin delayed the hydrolysis of sodium caseinate by pepsin and allowed the controlled release of rutin in the gastrointestinal tract. In another study, rutin was coated with sodium caseinate and trehalose complex, and an analysis revealed that rutin was amorphous after addition of trehalose and pH adjustment. The powder produced by co-precipitation with sodium caseinate contained large amounts of rutin [185]. Ghayour et al. [186] encapsulated curcumin and quercetin using the coating method. The encapsulation efficiency of both compounds was greater than 90%, and their solubilities in nanoparticles were higher than those of the free polyphenol molecules. In addition, polyphenol-casein nanoparticles were more cytotoxic towards MCF-7 human breast cancer cells than unloaded molecules.

3.4.2. Gelatin Nanoparticles

Gelatin is a water-soluble protein, that is prepared from collagen by acid or alkali hydrolysis (Figure 5) [187]. The US Food and Drug Administration generally believes that it is safe for use in medicine, cosmetics and food [188]. If sulfuric acid or hydrochloric acid is used to hydrolyze egg collagen, the gelatin obtained is of type A, with an isoelectric point of about 9. Correspondingly, if an alkaline solution is used to hydrolyze collagen, the gelatin produced is of type B with an isoelectric point of about 5 [189]. Gelatin is a polyamphoteric electrolyte with hydrophobic groups. Apart from being cheap and easily obtained, gelatins have good biocompatibility and biodegradability. First, because it is a deformable protein, the antigenicity of gelatin is relatively low compared with that of collagen. Second, gelatin produces no harmful by-products after enzymatic hydrolysis. In addition, the intrinsic protein structure of gelatin and the presence of many functional groups enable it to be coupled with many crosslinking agents and ligands. This has far-reaching significance for the development of targeting vectors [190,191]. Methods of preparing gelatin nanoparticles include desolvation, coacervation-phase separation, emulsification-solvent evaporation and nanoprecipitation [192,193,194].
Gelatin nanoparticles have been used for the effective delivery of a variety of drugs, including polyphenols. Shutava et al. [196] encapsulated several polyphenolic compounds—epigallocatechin gallate (EGCG), curcumin, tannic acid and catechin—in gelatin nanoparticles, and modified the nanoparticle surface with a layer-by-layer polyelectrolyte shell to increase the stability of the gel and control the release of polyphenols. The release of EGCG from the gel was found to be up to 8 h, much longer than the minutes when in the free state. Furthermore, nanoparticle-coated EGCG retained its biological activity in MB-MD-231 breast cancer cells. Karthikeyan et al. [197] prepared resveratrol-gelatin nanoparticles by agglomeration. We understand that the nanoparticles may induce the apoptosis of cancer cells by affecting the expression of p53, p21, caspase-3, Bax, Bcl-2 and NF-κB. To investigate the release of free tea polyphenols and nanoparticulate tea polyphenols in fatty foods from gelatin films, chitosan nanoparticles were prepared using the ionic gel method and bonded to gelatin films. It was found that the amount of tea polyphenols released from the gum film was related to the type of fatty food used and the encapsulation rate of the tea polyphenols. The presence of chitosan hydrochloride increased the diffusion time of tea polyphenols in the simulant, and the association was positively correlated [198].

3.4.3. Whey Protein (Mainly β-lactoglobulin) Nanoparticles

Whey protein is extracted from whey, a by-product of cheese production, and is composed of many proteins, including α-lactalbumin (α-la), β-lactoglobulin (β-lg), bovine serum albumin (BSA) and immunoglobulins, and lactoferrin [199]. Whey protein is considered to be ideal for encapsulating and delivering compounds such as polyphenols, due to its high safety, low cost, high nutritional value, and diverse functions [200]. Among them, β-lactoglobulin is the most widely used. Beta-lactoglobulin is the main whey protein and gelling agent in milk. It is present in most mammalian milk, but not in human milk, and is a small globular protein, consisting of only 162 amino acids, with a molecular weight of 18.3 kDa. Whey proteins have several transfer-friendly functions, such as binding to hydrophobic active substances, gelation and emulsification. In addition, whey protein is resistant to pepsin, so it is beneficial for the oral transport of polyphenols and other substances [201].
Shpigelman et al. [202] delivered EGCG with thermally modified lactoglobulin, and found that the correlation constant of EGCG with preheated protein was about 3.5 times higher than that of the natural protein. Because the size of EGCG-lactoglobulin nanoparticles is relatively small, they can maintain good transparency for the processing and preparation of transparent drinks. In addition, lactoglobulin encapsulation greatly protected the antioxidant activity of EGCG, and the degradation of EGCG in nanoparticles was 3.2 times slower than that of free EGCG within eight days. Li et al. [203] delivered curcumin with β-lactoglobulin and nanoemulsion as the carrier. The results showed that the water solubility, pH stability and permeability of curcumin were significantly improved by binding to β-lactoglobulin. The curcumin-β-lactoglobulin complex was resistant to pepsin, but sensitive to trypsin. In another study, whey protein-safflower complex was prepared under different pH conditions. The emulsification, thermal stability and oxidation resistance of the complex were improved compared with whey protein alone [204].
A large number of studies have shown that the complexation of polyphenolic compounds with whey protein can not only improve the bioavailability of polyphenolic compounds, but also improve the functional properties of whey protein. Chen et al. [205] studied the effect of lotus heart proanthocyanidin (LSPC) on the stability of carotene whey protein nanoemulsion. The results proved that the addition of LSPC significantly improved the chemical stability of the nanoemulsion, and the use of whey protein enhanced the antioxidant activity of LSPC. Morais has reported on the interactions of whey protein with (−)-EGCG and caffeic acid (CA) at different pH values and the antioxidant capacities of the complexes [206]. The study showed that the complexation of the two polyphenols with whey protein altered the protein structure under both acidic and neutral conditions, and the effect of CA was more significant. The complexation of whey protein with CA had a synergistic effect on its reducing power and antioxidant capacity. In contrast, the complexation of whey protein with EGCG inhibited its reducing power and antioxidant capacity. As previously mentioned, whey protein complexation prevents CA and EGCG from being degraded during digestion, increasing the oral availability of the phenolic compounds.

3.5. Polymeric Nanoparticles

Polymer nanoparticles are spherical or irregularly shaped colloidal particles formed by polymer materials [207]. They have high stability, uniform particle size, high drug loading rate, high biocompatibility, good drug release control and are easily produced in a factory. These advantages make them widely used for the encapsulation of natural extracts, including polyphenols [208,209,210]. Polymer systems are mainly divided into two categories: natural polymers (proteins and polysaccharides) and synthetic polymers.

3.5.1. PLA/PLGA

Polylactic acid (PLA) and polylactic acid-glycolic acid copolymer (PLGA) (Figure 6) have been widely used in drug delivery systems, due to their excellent biocompatibility, biodegradability and particle size control [211,212].
The use of PLA/PLGA nanoparticles improves the water solubility and poor stability of polyphenols. The release kinetics of polyphenol-PLA/PLGA nanoparticles in vitro show rapid release are at first followed by slow release, which demonstrates the controlled release of polyphenol compounds from PLA/PLGA nanoparticles [214]. Similar to other nanoparticles, PLA/PLGA also enhances the functional activity of polyphenols, such as anti-inflammatory, antioxidant and anti-cancer properties [215,216,217].
Although PLA has many advantages, it has certain limitations as a drug delivery system due to its hydrophobicity and low chemical stability. These limitations are particularly obvious when used for oral administration, because the nanopariticles are very likely to be intercepted and quickly cleared by mucus and cilia [218]. To avoid these limitations, PLA/PLGA nanoparticles are often modified by other polymers (such as polyethylene glycol (PEG) or chitosan) or are used in combination with other polymers. Studies have shown that a PEG coating on the nanoparticle surface can ensure rapid passage through the mucus layer, and significantly improve hydrophilicity and stability [219]. The toxicity of resveratrol-loaded PEG-PLA polymer nanoparticles to CT26 colon cancer cells was the same as that of free resveratrol, but encapsulation significantly increased its stability and cycling time, conferring significant anti-tumor effects [220]. Mixed micelles prepared from mPEG-PLA/TPGS enhanced the absorption of curcumin from the gastrointestinal tract, and its oral availability was much higher compared to curcumin suspension [221]. New targeted nanoparticles coated with EGCG exhibited excellent anti-proliferative activity in vitro, and tumor inhibition by the nanoparticles was significantly enhanced compared with the natural compounds [222]. In addition to PEG modification of PLGA, other polymers such as chitosan are often used as modifiers to increase the stability of nanoparticles. Chitosan oleic acid was applied to PLGA containing curcumin by emulsification and solvent evaporation, resulting in PLGA nanoparticles that were more stable than polymer micellar ones [223].

3.5.2. Chitosan

Chitosan is a semi-synthetic polysaccharide obtained by the deacetylation of chitin, which is widely present in nature, and is the only alkaline polysaccharide. Chitosan has received increasing attention, due to its good biocompatibility, degradability, non-antigenicity, high permeability, non-toxicity and good film-forming properties [224,225]. In addition, it has adhesive properties, and can reversibly open the tight junctions between epithelial cells, thereby promoting paracellular transport between cells [226]. Moreover, chitosan readily interacts with negatively charged polymers, conferring targeting properties to nanoparticles.
Chitosan has been widely used as a carrier in drug delivery systems. By increasing the release time and adhesion of polyphenols, chitosan nanoparticles significantly improved the functional activity and oral availability of the active substances. Encapsulation by chitosan enables the available concentration of green tea polyphenols to meet physiological requirements during treatment. The EGCG nanoformulation has an 8-fold dosage advantage compared with natural polyphenols. In addition, EGCG nanoparticles significantly induced apoptosis in human melanoma cells [227]. Similarly, the nanoparticles inhibited the growth of gastric cancer cells and reduced the expression of vascular endothelial growth factor protein by controlling the release of EGCG in gastric acid [228]. Studies have shown that the stability of chitosan-gum arabic polysaccharide nanoparticles loaded with curcumin in a simulated gastrointestinal environment is significantly improved. The release of curcumin is delayed, and the antioxidant activity of the active substance is also significantly enhanced [229]. Compared with free curcumin, curcumin-chitosan nanoparticles are more easily absorbed by colon cancer cells. The adhesion also prolongs contact time, resulting in a decrease in cell viability [230].
Although chitosan has many excellent features, properties required for some specific applications can be obtained by modifying it. Curcumin diethyl ester encapsulated with chitosan-alginate nanoparticles can be stored stably at 4 °C for three months, and the nanoparticles significantly improve cellular uptake by Caco-2 cells [231]. Studies have shown that folate-modified chitosan nanoparticles are able to target breast cancer cells, which complements curcumin-induced apoptosis of breast cancer cells. In addition, folic acid-modified chitosan nanoparticles are also responsive to pH changes. Compared with natural resveratrol, resveratrol loaded in carboxymethyl chitosan exhibits prolonged absorption and duration of action. Its antioxidant activity and bioavailability are also significantly increased [232].
In addition to the simple encapsulation of polyphenols, chitosan is increasingly popular in the food packaging industry, due to its film-forming ability and biodegradability [233]. It has been reported that pure chitosan has good mechanical strength and air permeability, but its antioxidant performance is not ideal. Addition of polyphenolic compounds with antioxidant properties can solve such problems [234]. A large number of studies have shown that the addition of polyphenols greatly improves resistance to oxidation, thermal stability, mechanical strength and pH-responsiveness of chitosan membranes [235,236,237].

3.5.3. Cyclodextrins

Cyclodextrin (CD) is the general term for a series of circular oligosaccharides produced by amylose under the action of cyclodextrin glucosyl transferase. As a natural carrier, it has been widely used in the pharmaceutical and food industries for encapsulation of bioactive compounds to improve their bioavailability and stability. Cyclodextrins are circular oligosaccharides with hollow pyramidal structures connected by 6, 7 or 8 glucose residues via 1,4 glycosidic bonds. They are produced by enzymatic hydrolysis of starch [238]. In nature, they exist as α-, β- and γ-CDs (Figure 7). The hydrophobic inner cavities of CDs can be used to encapsulate phenolic compounds to protect them from the external environment, such as pH, light, temperature and oxygen.
Studies have confirmed that the encapsulation of resveratrol into a cyclodextrin-based metal-organic framework greatly improves its stability [239]. It is controversial whether the complexation of polyphenols by cyclodextrin will affect their biological activity. It has been found that catechins and cyclodextrins interact through intermolecular O–H...O hydrogen bonds. The establishment of this host-guest relationship helps to improve the antioxidant capacity of polyphenols [240]. Similarly, when β-cyclodextrin is used to extract and protect phenolic compounds in olive oil, it does not affect their in vitro bioavailability [241]. In contrast, it was found that the oxygen radical absorbance capacity-fluorescein (ORAC-FL) value of the (+) catechin-cyclodextrin complex was lower than that of free catechin, indicating that complexation reduced the antioxidant activity of polyphenols [242]. This difference may be due to a variety of factors, such as the method, temperature, and matrix used in the complexation. Ho et al. found that catechin complexes were more stable in a solid matrix compared to semi-solid or liquid matrixes. The food matrix will affect the stability and recovery rate of catechins from the inclusion compound [243].

3.5.4. Hydrogels

A hydrogel is a three-dimensional, porous, shape-retaining, chemically or physically cross-linked highly water-soluble polymer [12]. The characteristics of hydrogels include mechanical resistance, swelling potential and moisture retention capacity [245,246].
When anti-inflammatory polyphenols are added to hydrogels that mimic the natural extracellular matrix and hydration environment, they form an ideal material for skin wound healing. Resveratrol-polypeptide-hydrogel nanoparticles are not cytotoxic and inhibit the macrophage release of pro-inflammatory factors to accelerate wound healing [247]. Animal experiments and molecular mechanism studies have shown that PVA/alginate hydrogel particles coated with tea polyphenols can promote wound healing in diabetic rats by regulating the PI3K/AKT signaling pathway [248]. Similarly, a newly developed hydrogel based on plant polyphenols, tannins and polypyrrole chains can stimulate tissue repair after bone marrow injury [249]. Of course, other biological activities, such as the antibacterial effects of polyphenols, are also fully compatible with hydrogel encapsulation. One study showed that PEG-lysozyme (LZM) polyphenolic hydrogels were more flexible than the original PEG-LZM, and the addition of polyphenols significantly improved the antibacterial and anti-inflammatory properties of the hydrogels [250]. The edible film prepared from tea polyphenols and calcium alginate gel has anti-oxidant and anti-inflammatory functions. The ductility, fracture strain, and air permeability of the film increases as the polyphenol content is increased [251]. Another study confirmed that injectable hydrogels based on curcumin, thiochitosan and polyethylene glycol diacrylate can promote the apoptosis of cancer cells and effectively delay tumor growth (Figure 8) [252].

3.5.5. Dendrimers

A dendrimer is a monodisperse, three-dimensional, hyper-branched radial symmetric polymer with host-guest capabilities [253]. Its core structure is a cavity that can contains biologically active components, and the branches can be modified or complexed with other compounds [254]. In addition, dendrimers can simply pass through biological barriers. Among the known dendrimers, PAMAM is the most widely used in drug delivery systems. Using silicon-PAMAM hybrid nanoparticles to load procyanidins, proanthocyanidins were completely released after six days, and the cytotoxicity reached at 87.9% after 134 h. Moreover, the mixed nanoparticles were found to have no toxicity towards normal cells [255]. Studies have shown that resveratrol, genistein and curcumin combine with PAMAM-G3 and PAMAM-G4 through hydrophilic, hydrophobic and hydrogen bonds to form stable complexes. The larger the nanoparticles, the higher the loading and stability of polyphenols, increasing their bioavailability [256,257].
In addition to PAMAM, other dendrimers have also been used to protect and deliver polyphenols, thus improving their bioavailability and stability. Gallic acid (GA) was loaded into the fifth-generation polyester dendrimer, and the large surface area of the nanoparticles significantly improved GA retention time and biological efficacy. Moreover, the antioxidant capacity of encapsulated gallic acid was four times higher than that of free GA [258]. The binding of dendritic plant glycogen (PG) significantly increased the solubility, in vitro permeability (Caco-2 monolayer permeability) and anticancer effects (HeLa cells) of curcumin compared with natural curcumin [259].

3.6. Micelles

Polymer micelles composed of amphiphilic polymer molecules are a new type of polyphenol carrier. Their hydrophobic core can be used to encapsulate water-insoluble substances, and the hydrophilic corona that protects the core can escape removal by the reticuloendothelial system (RES), increase the circulation time, and avoid interaction with blood components [260,261,262]. The micelles have diameters of 20–100 nm, which enables their movement from the tumor blood vessel wall into cancer cells [263]. Micelles are often used to encapsulate doxorubicin and polyphenolic compounds to increase the therapeutic effect while protecting the heart [264]. A study loaded resveratrol and quercetin Pluronic® F127 micelles (mRQ) with doxorubicin hydrochloride and found that the heart was protected by the combination with the two polyphenolic compounds. In addition, the presence of mRQ did not affect the caspase activity of human ovarian cancer cells (SKOV-3), but significantly reduced the caspase activity of rat cardiomyocytes (H9C2) [265]. Replacing the quercetin in the nanoparticles with curcumin gave similar results. The combined use of polyphenols and doxorubicin can reduce cardiac toxicity by reducing apoptosis and ROS production, and increase the efficacy of doxorubicin against ovarian cancer cells [266]. A large number of research results have shown that many biological activities of polyphenols are fully reflected by encapsulation in micelles. Curcumin encapsulated in monomethoxy-polyethylene glycol-chitosan-S-S-hexadecyl micelles effectively promoted the accumulation of active substances in cells and significantly down-regulated the expression of tumor necrosis factor. In addition, the nanoparticles also showed good anti-inflammatory effects in the tumor microenvironment [267]. Compared with free doxorubicin, doxorubicin/curcumin colloidal nanoparticles in human liver cancer SMMC 7721 cells exhibit prolonged release, a higher rate of cell apoptosis, and a stronger anti-angiogenesis effect [268].
In another study, resveratrol was loaded into cholesterol-polyamide micelles by solvent evaporation. The micelles reduced the production of pro-inflammatory factors in the lungs by the inhibiting nuclear transposition of NF-κB, demonstrating the anti-inflammatory effect of resveratrol (Figure 9) [269]. Washington et al. [270] used poly(ethylene glycol)-b-poly(ε-caprolactone) (PEG-b-PCL) and poly(ethylene glycol)-b-poly(γ-benzyl-ε-caprolactone) (PEG-b-PBCL) amphiphilic copolymer micelles loaded with doxorubicin and resveratrol. The encapsulation efficiency of doxorubicin in PEG-b-PBCL micelles was only 31%, while co-loading increased the encapsulation rate to 87.7%. In addition, the composite drug-loaded micelles were more cytotoxic to HeLa cells than micelles containing only DOX.

3.7. Nanoemulsion

Nanoemulsion refers to a system composed of two immiscible liquids, that is divided into an internal phase (or dispersed phase) and an external phase (or continuous phase). The internal phase is dispersed in the external phase [271]. Emulsions can be divided into micro- (10–100 nm), mini (nano)- (100–1000nm), and macro-emulsions (0.5–100mm) according to size [272], and many properties of the emulsion, such as stability, color and stability, are closely related to droplet size. Nanoemulsion is located between normal emulsion and microemulsionQ, and the average diameter is mostly less than 100 nm.
Nanoemulsions have been used to encapsulate a variety of polyphenolic compounds, such as curcumin, due to their small size, large surface area, high optical clarity, good stability, and ability to improve drug bioavailability. Yu et al. [273] prepared a nanoemulsion with a curcumin organogel. Pharmacokinetic analysis in mice showed that the oral availability of curcumin from the nanoemulsion was nine times higher than that of unformulated curcumin, and the digestion of the nanoemulsion in the gastrointestinal tract was significantly faster than that of the organogel. Another team used ultrasound to prepare a curcumin-nanoemulsion. In a gastrointestinal simulation experiment, the release rate of curcumin from the nanoemulsion was slower, because the nanoemulsion was not easily hydrolyzed by pepsin, and pancreatin can cause its release [274]. Similarly, Zou et al. [275] studied the potential of three nano-drug delivery systems, nanoemulsion, zein nanosuspension and nanoliposomes, to improve curcumin bioavailability. The study found that curcumin loaded into nanoemulsion was most effective for gastrointestinal absorption. In one study, EGCG encapsulated in an oil/water nanoemulsion exhibited significantly increased anticancer activity in vitro compared with free EGCG, and the droplet size of the nanoemulsion hardly changed over 14 days [276]. In another study, the authors investigated the effects of nanoemulsion encapsulation on the physical and chemical properties, biological activity, and epithelial permeability of EGCG. Compared with unencapsulated catechins, the bioaccessibility of EGCG-nanoemulsion increased by 2.78 times. In addition, the intestinal permeability of EGCG was significantly increased. These results showed that a soybean protein nanoemulsion could improve the stability, bioaccessibility and permeability of green tea catechin [277].
Although nanoemulsions have significant advantages, they are unstable at low pH, and their small size and liquid nature make drug release difficult to control. In addition, the preparation of nanoemulsions requires specialized equipment, so there is a perception in the food industry that nanoemulsions are unprofitable [14].

3.8. Metal Nanoparticles

Compared with organic nanoparticles, inorganic nanoparticles have unique characteristics, such as good controllability of size and shape, large specific surface area, and imaging potential. They also enable targeted drug delivery and synergistic therapy. These characteristics make them well suited for drug delivery. Among all inorganic nanoparticles, metal nanoparticles are the most widely used. Metal nanoparticles are composed of pure metals such as gold, silver or platinum, with a size of 1–100 nm. In recent years, they have attracted increasing attention due to their huge potential in drug delivery systems [278,279].

3.8.1. Gold Nanoparticles (AuNPs)

Gold nanoparticles are crystal structures composed of metal gold atoms, and are the most widely used metal nanoparticles, with many properties that make them the most promising nanomaterials in biomedical fields such as biosensors, molecular imaging and drug carriers. They are 1–100 nm in size, and form a variety of shapes such as ball, bar and cage. In addition, they are non-toxic, biocompatible, are negatively charged and easily functionalized by other biomolecules [280,281,282]. Traditional gold nanoparticles are synthesized using chemicals that are harmful to human health and the environment [283]. In recent years, numerous gold nanoparticles have been prepared using plant active compounds that are friendly to the human body and the environment, such as polyphenols [284]. Curcumin (Cur) is the most widely used plant active compound in the preparation of gold nanoparticles (Figure 10).
Compared with free curcumin, Cur-AuNPs showed significantly increased cytotoxicity to colon and breast cancer cells. The conjugated nanoparticles showed no toxicity to normal kidney cells, exhibiting excellent biocompatibility [285,286]. Compared with other nanoparticles, gold nanoparticles are composed of a small number of atoms, and fluoresce under visible light (550 nm) irradiation. This feature makes them useful for biological imaging. Compared with the control group, curcumin-conjugated gold clusters (Cur-AuNCs) significantly inhibited the migration of HeLa cells and exhibited significant cytotoxicity [284]. Other polyphenol compounds can also participate in the synthesis of gold nanoparticles. Compared with free EGCG and citrate-gold nanoparticles, AuNPs prepared with EGCG as a reducing agent more potently inhibit the growth of cancer cells such as PC3 and MDA-MB-231 and induce apoptosis. Resveratrol-gold nanoparticles have good stability, photothermal performance and antioxidant capacity. Moreover, under laser irradiation, the nanoparticles block the cancer cell cycle and inhibit cell division, leading to apoptosis [287]. Studies have also evaluated the effects of quercetin-gold nanoparticles in breast cancer cell lines. It was found that the nanoparticles inhibited the angiogenesis and metastasis of breast cancer cells by targeting the EGFR/VEGFR-2 signaling pathway. Compared with free quercetin, nanoparticles induce more apoptosis of cancer cells [288,289].

3.8.2. Silver NPs (Ag NPs)

Similar to gold nanoparticles, silver nanoparticles also have anti-inflammatory, antibacterial, anti-cancer and other biological activities [290]. In recent years, researchers have been interested in the use of plant extracts, such as polyphenols, as reducing agents and stabilizers for green synthesis of silver nanoparticles [291]. Green synthetic methods embody the advantages of cost-effectiveness, eco-friendliness and biocompatibility [292].
The antimicrobial activity of AgNPs has attracted much attention. AgNPs synthesized using hydrolyzate rich in polyphenols have good activity against bacteria and fungi [293]. Silver nanoparticles synthesized from the aqueous extract of laurel stem had significant activity against both Gram-positive and Gram-negative bacteria [294].
Other biological properties of AgNPs synthesized from polyphenolic plant extracts, such as anti-inflammatory, antioxidant and anticancer activities, have also attracted increasing attention. Polyphenols (extracted from Cornus officinalis)-silver nanoparticles inhibit the production of pro-inflammatory cytokines by inhibiting the activation of NF-κB in macrophages, thus showing good anti-inflammatory activity in the treatment of psoriasis [295]. AgNPs prepared using the water extract of Lintong leaf is a good analgesic and muscle relaxant, and can be used for pain care [296]. Although silver nanoparticles themselves have antibacterial and anticancer activities, the role of polyphenols and other plant extracts in AgNPs should not be ignored. Studies have shown that these active substances can bind to the final AgNPs. In fact, polyphenols and silver have a synergistic relationship in respect of antioxidant, antibacterial, anti-inflammatory and other properties [297,298].

4. Conclusions and Future Trends

As some of the most widely distributed plant active compounds, phenolic compounds have many functions that are beneficial to human health. However, the low solubility, poor stability and low bioavailability of these compounds greatly limit their applications in food and medicine. Encapsulation in nanoparticles can overcome these limitations and can control/target their release under specific conditions. Therefore, nanotechnology provides an ideal carrier system to improve the pharmacokinetics and bioavailability of polyphenols.
Although nanoparticles are nearly perfect as carriers, their toxicity and side effects still need to be considered and minimized before clinical application. Because polyphenols are natural compounds that need to be taken for long periods of time to prevent and treat disease, it is important to understand the toxic side effects of nanoparticles when they accumulate in the body, especially if the nanoparticles have a low encapsulation rate. It is therefore necessary to establish standardized in vitro and in vivo models and conduct safety tests to promote the development and application of new nanoparticles beneficial to human health.

Author Contributions

B.Y. and Y.D. drafted the manuscript. F.W. contributed to acquisition of literature data. Y.Z. outlined and edited the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

The work was supported by the National Key Research and Development Program (2016YFD0600801), the Natural Science Foundation of Jiangsu Province (BK20181401) and the National Natural Science Foundation of China (31200564).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hu, B.; Liu, X.X.; Zhang, C.L.; Zeng, X.X. Food macromolecule based nanodelivery systems for enhancing the bioavailability of polyphenols. J. Food Drug Anal. 2017, 25, 3–15. [Google Scholar] [CrossRef] [Green Version]
  2. Cvetanović, A.; Švarc-Gajić, J.; Gašić, U.; Tešić, Ž.; Zengin, G.; Zeković, Z.; Đurović, S. Isolation of apigenin from subcritical water extracts: Optimization of the process. J. Supercrit. Fluids 2017, 120, 32–42. [Google Scholar] [CrossRef] [Green Version]
  3. Kostić, A.Ž.; Gašić, U.M.; Pešić, M.B.; Stanojević, S.P.; Barać, M.B.; Mačukanović-Jocić, M.P.; Avramov, S.N.; Tešić, Ž.L. Phytochemical analysis and total antioxidant capacity of rhizome, above-ground vegetative parts and flower of three Iris species. Chem. Biodivers. 2019, 16, e1800565. [Google Scholar] [CrossRef] [Green Version]
  4. Kostić, A.Ž.; Milinčić, D.D.; Gašić, U.M.; Nedić, N.; Stanojević, S.P.; Tešić, Ž.L.; Pešić, M.B. Polyphenolic profile and antioxidant properties of bee-collected pollen from sunflower (Helianthus annuus L.) plant. LWT Food Sci. Technol. 2019, 112, 108244. [Google Scholar] [CrossRef]
  5. Pavlović, A.V.; Papetti, A.; Zagorac, D.Č.D.; Gašić, U.M.; Mišić, D.M.; Tešić, Ž.L.; Natić, M.M. Phenolics composition of leaf extracts of raspberry and blackberry cultivars grown in Serbia. Ind. Crop. Prod. 2016, 87, 304–314. [Google Scholar] [CrossRef]
  6. Khalid, M.; Saeed-ur, R.; Bilal, M.; Huang, D.F. Role of flavonoids in plant interactions with the environment and against human pathogens—A review. J. Integr. Agric. 2019, 18, 211–230. [Google Scholar] [CrossRef]
  7. Daza, L.D.; Fujita, A.; Granato, D.; Favaro-Trindade, C.S.; Genovese, M.I. Functional properties of encapsulated Cagaita (Eugenia dysenterica DC.) fruit extract. Food Biosci. 2017, 18, 15–21. [Google Scholar] [CrossRef]
  8. Stanisavljević, N.S.; Ilić, M.D.; Matić, I.Z.; Jovanović, Ž.S.; Čupić, T.; Dabić, D.Č.; Natić, M.M.; Tešić, Ž.L. Identification of phenolic compounds from seed coats of differently colored european varieties of pea (Pisum sativum L.) and characterization of their antioxidant and in vitro anticancer activities. Nutr. Cancer 2016, 68, 988–1000. [Google Scholar] [CrossRef] [PubMed]
  9. Murador, D.; Braga, A.R.; Da Cunha, D.; De Rosso, V. Alterations in phenolic compound levels and antioxidant activity in response to cooking technique effects: A meta-analytic investigation. Crit. Rev. Food Sci. Nutr. 2018, 58, 169–177. [Google Scholar] [CrossRef] [PubMed]
  10. Rostami, M.; Yousefi, M.; Khezerlou, A.; Aman Mohammadi, M.; Jafari, S.M. Application of different biopolymers for nanoencapsulation of antioxidants via electrohydrodynamic processes. Food Hydrocolloids 2019, 97, 105170. [Google Scholar] [CrossRef]
  11. Li, Z.; Jiang, H.; Xu, C.; Gu, L. A review: Using nanoparticles to enhance absorption and bioavailability of phenolic phytochemicals. Food Hydrocoll. 2015, 43, 153–164. [Google Scholar] [CrossRef]
  12. Abaee, A.; Mohammadian, M.; Jafari, S.M. Whey and soy protein-based hydrogels and nano-hydrogels as bioactive delivery systems. Trends Food Sci. Technol. 2017, 70, 69–81. [Google Scholar] [CrossRef]
  13. Esfanjani, A.F.; Assadpour, E.; Jafari, S.M. Improving the bioavailability of phenolic compounds by loading them within lipid-based nanocarriers. Trends Food Sci. Technol. 2018, 76, 56–66. [Google Scholar] [CrossRef]
  14. Akhavan, S.; Assadpour, E.; Katouzian, I.; Jafari, S.M. Lipid nano scale cargos for the protection and delivery of food bioactive ingredients and nutraceuticals. Trends Food Sci. Technol. 2018, 74, 132–146. [Google Scholar] [CrossRef]
  15. Miraftab, R.; Xiao, H. Feasibility and potential of graphene and its hybrids with cellulose as drug carriers: A commentary. J. Bioresour. Bioprod. 2019, 4, 200–201. [Google Scholar]
  16. Kassem, M.A.; El-Sawy, H.S.; Abd-Allah, F.I.; Abdelghany, T.M.; El-Say, K.M. Maximizing the therapeutic efficacy of imatinib mesylate-loaded niosomes on human colon adenocarcinoma using box-behnken design. J. Pharm. Sci. 2017, 106, 111–122. [Google Scholar] [CrossRef] [Green Version]
  17. Esfanjani, A.F.; Jafari, S.M. Biopolymer nano-particles and natural nano-carriers for nano-encapsulation of phenolic compounds. Colloids Surf. B Biointerfaces 2016, 146, 532–543. [Google Scholar] [CrossRef]
  18. Heim, K.E.; Tagliaferro, A.R.; Bobilya, D.J. Flavonoid antioxidants: Chemistry, metabolism and structure-activity relationships. J. Nutr. Biochem. 2002, 13, 572–584. [Google Scholar] [CrossRef]
  19. Li, D.W.; Zhou, B.; Lv, B. Antibacterial therapeutic agents composed of functional biological molecules. J. Chem. 2020, 2020, 13. [Google Scholar] [CrossRef] [Green Version]
  20. Rong, T. Antioxidant properties in vitro and in vivo: Realistic assessments of efficacy of plant extracts. Plant. Sci. Rev. 2012, 7. [Google Scholar] [CrossRef]
  21. Perron, N.R.; Brumaghim, J.L. A review of the antioxidant mechanisms of polyphenol compounds related to iron binding. Cell Biochem. Biophys. 2009, 53, 75–100. [Google Scholar] [CrossRef] [PubMed]
  22. Heinrich, J.; Valentova, K.; Vacek, J.; Palikova, I.; Zatloukalova, M.; Kosina, P.; Ulrichova, J.; Vrbkova, J.; Simanek, V. Metabolic profiling of phenolic acids and oxidative stress markers after consumption of Lonicera caerulea L. fruit. J. Agric. Food Chem. 2013, 61, 4526–4532. [Google Scholar] [CrossRef] [PubMed]
  23. Bouayed, J.; Bohn, T. Exogenous antioxidants—Double-edged swords in cellular redox state: Health beneficial effects at physiologic doses versus deleterious effects at high doses. Oxid. Med. Cell. Longev. 2010, 3, 267025. [Google Scholar] [CrossRef] [PubMed]
  24. Bhattacharyya, A.; Chattopadhyay, R.; Mitra, S.; Crowe, S.E. Oxidative stress: An essential factor in the pathogenesis of gastrointestinal mucosal diseases. Physiol. Rev. 2014, 94, 329–354. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Pham-Huy, L.A.; He, H.; Pham-Huy, C. Free radicals, antioxidants in disease and health. Int. J. Biomed. Sci. 2008, 4, 89–96. [Google Scholar] [PubMed]
  26. Zhang, H.; Tsao, R. Dietary polyphenols, oxidative stress and antioxidant and anti-inflammatory effects. Curr. Opin. Food Sci. 2016, 8. [Google Scholar] [CrossRef]
  27. Ji, M.Y.; Gong, X.; Li, X.; Wang, C.C.; Li, M.H. Advanced research on the antioxidant activity and mechanism of polyphenols from hippophae species—A review. Molecules 2020, 25, 917. [Google Scholar] [CrossRef] [Green Version]
  28. Middha, S.K.; Usha, T.; Basistha, B.C.; Goyal, A.K. Amelioration of antioxidant potential, toxicity, and antihyperglycemic activity of Hippophae salicifolia D. Don leaf extracts in alloxan-induced diabetic rats. 3 Biotech. 2019, 9, 8. [Google Scholar] [CrossRef]
  29. Sun, H.N.; Mu, B.N.; Song, Z.; Ma, Z.M.; Mu, T.H. The in vitro antioxidant activity and inhibition of intracellular reactive oxygen species of sweet potato leaf polyphenols. Oxid. Med. Cell. Longev. 2018, 2018, 11. [Google Scholar] [CrossRef] [Green Version]
  30. Xi, L.; Mu, T.; Sun, H. Preparative purification of polyphenols from sweet potato (Ipomoea batatas L.) leaves by AB-8 macroporous resins. Food Chem. 2015, 172, 166–174. [Google Scholar] [CrossRef]
  31. Myint, K.Z.; Wu, K.; Xia, Y.M.; Fan, Y.; Shen, J.; Zhang, P.; Gu, J.X. Polyphenols from Stevia rebaudiana (Bertoni) leaves and their functional properties. J. Food Sci. 2020, 85, 240–248. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Baltas, N.; Can, Z. Bioactivity and Enzyme inhibition properties of Stevia rebaudiana. Curr. Enzym. Inhib. 2016, 12. [Google Scholar] [CrossRef]
  33. Guo, Y.L.; Li, X.Z.; Kuang, C.T. Antioxidant pathways and chemical mechanism of curcumin. In Application of Chemical Engineering; Parts 1–3; Cao, Z., He, Y.H., Sun, L., Cao, X.Q., Eds.; Trans Tech Publications Ltd.: Zurich, Switzerland, 2011; Volume 236–238, pp. 2311–2314. [Google Scholar]
  34. O’Toole, M.G.; Soucy, P.A.; Chauhan, R.; Raju, M.V.R.; Patel, D.N.; Nunn, B.M.; Keynton, M.A.; Ehringer, W.D.; Nantz, M.H.; Keynton, R.S.; et al. Release-Modulated antioxidant activity of a composite curcumin-chitosan polymer. Biomacromolecules 2016, 17, 1253–1260. [Google Scholar] [CrossRef] [PubMed]
  35. Reddy, D.N.K.; Huang, F.-Y.; Wang, S.-P.; Kumar, R. Synergistic antioxidant and antibacterial activity of Curcumin-C3 encapsulated chitosan nanoparticles. Curr. Pharm. Des. 2020. [Google Scholar] [CrossRef] [PubMed]
  36. Bajpai, V.K.; Alam, M.B.; Ju, M.K.; Kwon, K.R.; Huh, Y.S.; Han, Y.K.; Lee, S.H. Antioxidant mechanism of polyphenol-rich Nymphaea nouchali leaf extract protecting DNA damage and attenuating oxidative stress-induced cell death via Nrf2-mediated heme-oxygenase-1 induction coupled with ERK/p38 signaling pathway. Biomed. Pharm. 2018, 103, 1397–1407. [Google Scholar] [CrossRef] [PubMed]
  37. Zou, B.; Xiao, G.S.; Xu, Y.J.; Wu, J.J.; Yu, Y.S.; Fu, M.Q. Persimmon vinegar polyphenols protect against hydrogen peroxide-induced cellular oxidative stress via Nrf2 signalling pathway. Food Chem. 2018, 255, 23–30. [Google Scholar] [CrossRef]
  38. Cho, J.H.; Kim, I.D.; Dhungana, S.K.; Do, H.M.; Shin, D.H. Persimmon fruit enhanced quality characteristics and antioxidant potential of beer. Food Sci. Biotechnol. 2018, 27, 1067–1073. [Google Scholar] [CrossRef]
  39. Zou, B.; Wu, J.J.; Yu, Y.S.; Xiao, G.S.; Xu, Y.J. Evolution of the antioxidant capacity and phenolic contents of persimmon during fermentation. Food Sci. Biotechnol. 2017, 26, 563–571. [Google Scholar] [CrossRef]
  40. Deng, G.F.; Xu, X.R.; Zhang, Y.; Li, D.; Gan, R.Y.; Li, H.B. Phenolic compounds and bioactivities of pigmented rice. Crit. Rev. Food Sci. Nutr. 2013, 53, 296–306. [Google Scholar] [CrossRef]
  41. Meng, L.S.; Zhu, J.Y.; Ma, Y.; Sun, X.Y.; Li, D.N.; Li, L.; Bai, H.Q.; Xin, G.; Meng, X.J. Composition and antioxidant activity of anthocyanins from Aronia melanocarpa cultivated in Haicheng, Liaoning, China. Food Biosci. 2019, 30, 10. [Google Scholar] [CrossRef]
  42. Jiang, W.; Zhou, X.H. Hydrolysis of radish anthocyanins to enhance the antioxidant and antiproliferative capacities. Food Chem. 2019, 294, 477–485. [Google Scholar] [CrossRef] [PubMed]
  43. Sano, A.; Uchida, R.; Saito, M.; Shioya, N.; Komori, Y.; Tho, Y.; Hashizume, N. Beneficial effects of grape seed extract on malondialdehyde-modified LDL. J. Nutr. Sci. Vitam. 2007, 53, 174–182. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Mansouri, E.; Khorsandi, L.; Fard, A.A. Protective role of grape seed proanthocyanidin antioxidant properties on heart of streptozotocin-induced diabetic rats. Vet. Res. Forum 2015, 6, 119–124. [Google Scholar] [PubMed]
  45. Ky, I.; Lorrain, B.; Kolbas, N.; Crozier, A.; Teissedre, P.L. Wine by-products: Phenolic characterization and antioxidant activity evaluation of grapes and grape pomaces from six different French grape varieties. Molecules 2014, 19, 482–506. [Google Scholar] [CrossRef] [Green Version]
  46. Sun, J.; Jiang, Y.M.; Shi, J.; Wei, X.Y.; Xue, S.J.; Shi, J.Y.; Yi, C. Antioxidant activities and contents of polyphenol oxidase substrates from pericarp tissues of litchi fruit. Food Chem. 2010, 119, 753–757. [Google Scholar] [CrossRef]
  47. Xu, J.L.; Shin, J.S.; Park, S.K.; Kang, S.; Jeong, S.C.; Moon, J.K.; Choi, Y. Differences in the metabolic profiles and antioxidant activities of wild and cultivated black soybeans evaluated by correlation analysis. Food Res. Int. 2017, 100, 166–174. [Google Scholar] [CrossRef]
  48. Kou, X.H.; Han, L.H.; Li, X.Y.; Xue, Z.H.; Zhou, F.J. Antioxidant and antitumor effects and immunomodulatory activities of crude and purified polyphenol extract from blueberries. Front. Chem. Sci. Eng. 2016, 10, 108–119. [Google Scholar] [CrossRef]
  49. Su, X.M.; Zhang, J.; Wang, H.Q.; Xu, J.; He, J.M.; Liu, L.Y.; Zhang, T.; Chen, R.Y.; Kang, J. Phenolic acid profiling, antioxidant, and anti-inflammatory activities, and miRNA regulation in the polyphenols of 16 blueberry samples from China. Molecules 2017, 22, 312. [Google Scholar] [CrossRef]
  50. Li, H.; Wang, Z.Y. Comparison in antioxidant and antitumor activities of pine polyphenols and its seven biotransformation extracts by fungi. PeerJ 2017, 5, 21. [Google Scholar] [CrossRef] [Green Version]
  51. Ferreira-Santos, P.; Zanuso, E.; Genisheva, Z.; Rocha, C.M.R.; Teixeira, J.A. Green and sustainable valorization of bioactive phenolic compounds from Pinus by-products. Molecules 2020, 25, 2931. [Google Scholar] [CrossRef]
  52. Yan, Z.M.; Zhong, Y.Z.; Duan, Y.H.; Chen, Q.H.; Li, F.N. Antioxidant mechanism of tea polyphenols and its impact on health benefits. Anim. Nutr. 2020, 6, 115–123. [Google Scholar] [CrossRef] [PubMed]
  53. Zhao, C.N.; Tang, G.Y.; Cao, S.Y.; Xu, X.Y.; Gan, R.Y.; Liu, Q.; Mao, Q.Q.; Shang, A.; Li, H.B. Phenolic profiles and antioxidant activities of 30 tea infusions from green, black, oolong, white, yellow and dark teas. Antioxidants 2019, 8, 215. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Allawadhi, P.; Khurana, A.; Sayed, N.; Kumari, P.; Godugu, C. Isoproterenol-induced cardiac ischemia and fibrosis: Plant-based approaches for intervention. Phytother. Res. 2018, 32, 1908–1932. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Tapsell, L.C.; Neale, E.P.; Probst, Y. Dietary patterns and cardiovascular disease: Insights and challenges for considering food groups and nutrient sources. Curr. Atheroscler. Rep. 2019, 21, 8. [Google Scholar] [CrossRef] [Green Version]
  56. Marventano, S.; Salomone, F.; Godos, J.; Pluchinotta, F.; Del Rio, D.; Mistretta, A.; Grosso, G. Coffee and tea consumption in relation with non-alcoholic fatty liver and metabolic syndrome: A systematic review and meta-analysis of observational studies. Clin. Nutr. 2016, 35, 1269–1281. [Google Scholar] [CrossRef]
  57. Shin, J.Y.; Kim, J.Y.; Kang, H.T.; Han, K.H.; Shim, J.Y. Effect of fruits and vegetables on metabolic syndrome: A systematic review and meta-analysis of randomized controlled trials. Int. J. Food Sci. Nutr. 2015, 66, 416–425. [Google Scholar] [CrossRef]
  58. Wang, X.; Ouyang, Y.; Liu, J.; Zhao, G. Flavonoid intake and risk of CVD: A systematic review and meta-analysis of prospective cohort studies. Br. J. Nutr. 2013, 111, 1–11. [Google Scholar] [CrossRef] [Green Version]
  59. Menezes, R.; Rodriguez-Mateos, A.; Kaltsatou, A.; Gonzalez-Sarrias, A.; Greyling, A.; Giannaki, C.; Andres-Lacueva, C.; Milenkovic, D.; Gibney, E.R.; Dumont, J.; et al. Impact of flavonols on cardiometabolic biomarkers: A meta-analysis of randomized controlled human trials to explore the role of inter-individual variability. Nutrients 2017, 9, 117. [Google Scholar] [CrossRef]
  60. Tang, Z.; Li, M.; Zhang, X.; Hou, W. Dietary flavonoid intake and the risk of stroke: A dose-response meta-analysis of prospective cohort studies. BMJ Open 2016, 6, e008680. [Google Scholar] [CrossRef]
  61. Wang, Z.-M.; Zhao, D.; Nie, Z.-L.; Zhao, H.; Zhou, B.; Gao, W.; Wang, L.-S.; Yang, Z.-J. Flavonol intake and stroke risk: A meta-analysis of cohort studies. Nutrition 2014, 30, 518–523. [Google Scholar] [CrossRef]
  62. Cassidy, A.; O’Reilly, É.J.; Kay, C.; Sampson, L.; Franz, M.; Forman, J.P.; Curhan, G.; Rimm, E.B. Habitual intake of flavonoid subclasses and incident hypertension in adults. Am. J. Clin. Nutr. 2010, 93, 338–347. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Grosso, G.; Stepaniak, U.; Micek, A.; Kozela, M.; Stefler, D.; Bobak, M.; Pajak, A. Dietary polyphenol intake and risk of hypertension in the Polish arm of the HAPIEE study. Eur. J. Nutr. 2017, 57. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Lajous, M.; Rossignol, E.; Fagherazzi, G.; Perquier, F.; Scalbert, A.; Clavel-Chapelon, F.; Boutron-Ruault, M.-C. Flavonoid intake and incident hypertension in women. Am. J. Clin. Nutr. 2016, 103, 1091–1098. [Google Scholar] [CrossRef] [Green Version]
  65. Huang, H.; Liao, D.; Dong, Y.; Pu, R. Effect of quercetin supplementation on plasma lipid profiles, blood pressure, and glucose levels: A systematic review and meta-analysis. Nutr. Rev. 2020. [Google Scholar] [CrossRef] [Green Version]
  66. Cao, H.; Ou, J.Y.; Chen, L.; Zhang, Y.B.; Szkudelski, T.; Delmas, D.; Daglia, M.; Xiao, J.B. Dietary polyphenols and type 2 diabetes: Human study and clinical trial. Crit. Rev. Food Sci. Nutr. 2019, 59, 3371–3379. [Google Scholar] [CrossRef]
  67. Liu, Y.; Li, J.P.; Wang, T.S.; Wang, Y.T.; Zhao, L.B.; Fang, Y. The effect of genistein on glucose control and insulin sensitivity in postmenopausal women: A meta-analysis. Maturitas 2017, 97, 44–52. [Google Scholar] [CrossRef] [PubMed]
  68. Vita, J.A. Polyphenols and cardiovascular disease: Effects on endothelial and platelet function. Am. J. Clin. Nutr. 2005, 81, 292S–297S. [Google Scholar] [CrossRef] [PubMed]
  69. Zhu, Y.N.; Xia, M.; Yang, Y.; Liu, F.Q.; Li, Z.X.; Hao, Y.T.; Mi, M.T.; Jin, T.R.; Ling, W.H. Purified anthocyanin supplementation improves endothelial function via NO-cGMP activation in hypercholesterolemic individuals. Clin. Chem. 2011, 57, 1524–1533. [Google Scholar] [CrossRef] [Green Version]
  70. Dell’Agli, M.; Galli, G.V.; Vrhovsek, U.; Mattivi, F.; Bosisio, E. In vitro inhibition of human cGMP-specific phosphodiesterase-5 by polyphenols from red grapes. J. Agric. Food Chem. 2005, 53, 1960–1965. [Google Scholar] [CrossRef]
  71. Santhakumar, A.B.; Kundur, A.R.; Fanning, K.; Netzel, M.; Stanley, R.; Singh, I. Consumption of anthocyanin-rich Queen Garnet plum juice reduces platelet activation related thrombogenesis in healthy volunteers. J. Funct. Foods 2015, 12, 11–22. [Google Scholar] [CrossRef]
  72. Santhakumar, A.B.; Bulmer, A.C.; Singh, I. A review of the mechanisms and effectiveness of dietary polyphenols in reducing oxidative stress and thrombotic risk. J. Hum. Nutr. Diet. 2014, 27, 1–21. [Google Scholar] [CrossRef]
  73. Du, G.H.; Sun, L.; Zhao, R.; Du, L.D.; Song, J.K.; Zhang, L.; He, G.R.; Zhang, Y.X.; Zhang, J.T. Polyphenols: Potential source of drugs for the treatment of ischaemic heart disease. Pharm. 2016, 162, 23–34. [Google Scholar] [CrossRef] [Green Version]
  74. Slevin, M.; Ahmed, N.; Wang, Q.; McDowell, G.; Badimon, L. Unique vascular protective properties of natural products: Supplements or future main-line drugs with significant anti-atherosclerotic potential? Vasc. Cell 2012, 4, 9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Wu, S.; Zhu, W.; Thompson, P.; Hannun, Y.A. Evaluating intrinsic and non-intrinsic cancer risk factors. Nat. Commun. 2018, 9, 12. [Google Scholar] [CrossRef] [PubMed]
  76. Seca, A.M.L.; Pinto, D. Plant secondary metabolites as anticancer agents: Successes in clinical trials and therapeutic application. Int. J. Mol. Sci. 2018, 19, 263. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Oyenihi, A.B.; Smith, C. Are polyphenol antioxidants at the root of medicinal plant anti-cancer success? J. Ethnopharmacol. 2019, 229, 54–72. [Google Scholar] [CrossRef] [PubMed]
  78. Tariq, A.; Sadia, S.; Pan, K.W.; Ullah, I.; Mussarat, S.; Sun, F.; Abiodun, O.O.; Batbaatar, A.; Li, Z.L.; Song, D.G.; et al. A systematic review on ethnomedicines of anticancer plants. Phytother. Res. 2017, 31, 202–264. [Google Scholar] [CrossRef] [PubMed]
  79. Li, D.W.; Wang, Q.; Zhou, B.; Zhuge, Q.; Lv, B. Small DNA circles as bacterial topoisomerase I inhibitors. RSC Adv. 2019, 9, 18415–18419. [Google Scholar] [CrossRef] [Green Version]
  80. Rengasamy, K.R.R.; Khan, H.; Gowrishankar, S.; Lagoa, R.J.L.; Mahomoodally, F.M.; Khan, Z.; Suroowan, S.; Tewari, D.; Zengin, G.; Hassan, S.T.S.; et al. The role of flavonoids in autoimmune diseases: Therapeutic updates. Pharm. 2019, 194, 107–131. [Google Scholar] [CrossRef]
  81. Salehi, B.; Zucca, P.; Sharifi-Rad, M.; Pezzani, R.; Rajabi, S.; Setzer, W.N.; Varoni, E.M.; Iriti, M.; Kobarfard, F.; Sharifi-Rad, J. Phytotherapeutics in cancer invasion and metastasis. Phytother. Res. 2018, 32, 1425–1449. [Google Scholar] [CrossRef]
  82. Xing, L.J.; Zhang, H.; Qi, R.L.; Tsao, R.; Mine, Y. Recent advances in the understanding of the health benefits and molecular mechanisms associated with green tea polyphenols. J. Agric. Food Chem. 2019, 67, 1029–1043. [Google Scholar] [CrossRef] [PubMed]
  83. Bo, Y.C.; Sun, J.F.; Wang, M.M.; Ding, J.Z.; Lu, Q.J.; Yuan, L. Dietary flavonoid intake and the risk of digestive tract cancers: A systematic review and meta-analysis. Sci. Rep. 2016, 6, 8. [Google Scholar] [CrossRef] [PubMed]
  84. Hua, X.L.; Yu, L.L.; You, R.X.; Yang, Y.; Liao, J.; Chen, D.S.; Yu, L.X. Association among dietary flavonoids, flavonoid subclasses and ovarian cancer risk: A meta-analysis. PLoS ONE 2016, 11, 15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Sajadimajd, S.; Bahramsoltani, R.; Iranpanah, A.; Patra, J.K.; Das, G.; Gouda, S.; Rahimi, R.; Rezaeiamiri, E.; Cao, H.; Giampieri, F.; et al. Advances on natural polyphenols as anticancer agents for skin cancer. Pharm. Res. 2020, 151, 14. [Google Scholar] [CrossRef] [PubMed]
  86. Grosso, G.; Godos, J.; Lamuela-Raventos, R.; Ray, S.; Micek, A.; Pajak, A.; Sciacca, S.; D’Orazio, N.; Del Rio, D.; Galvano, F. A comprehensive meta-analysis on dietary flavonoid and lignan intake and cancer risk: Level of evidence and limitations. Mol. Nutr. Food Res. 2017, 61, 10. [Google Scholar] [CrossRef]
  87. Kopustinskiene, D.M.; Jakstas, V.; Savickas, A.; Bernatoniene, J. Flavonoids as anticancer agents. Nutrients 2020, 12, 457. [Google Scholar] [CrossRef] [Green Version]
  88. Chang, H.; Lei, L.; Zhou, Y.; Ye, F.Y.; Zhao, G.H. Dietary flavonoids and the risk of colorectal cancer: An updated meta-analysis of epidemiological studies. Nutrients 2018, 10, 950. [Google Scholar] [CrossRef] [Green Version]
  89. Hu, C.X.; Li, M.J.; Guo, T.T.; Wang, S.X.; Huang, W.P.; Yang, K.; Liao, Z.W.; Wang, J.; Zhang, F.X.; Wang, H.Q. Anti-metastasis activity of curcumin against breast cancer via the inhibition of stem cell-like properties and EMT. Phytomedicine 2019, 58, 11. [Google Scholar] [CrossRef]
  90. Lee, Y.H.; Song, N.Y.; Suh, J.; Kim, D.H.; Kim, W.; Ann, J.; Lee, J.; Baek, J.H.; Na, H.K.; Surh, Y.J. Curcumin suppresses oncogenicity of human colon cancer cells by covalently modifying the cysteine 67 residue of SIRT1. Cancer Lett. 2018, 431, 219–229. [Google Scholar] [CrossRef]
  91. Shanmugam, M.K.; Rane, G.; Kanchi, M.M.; Arfuso, F.; Sethi, G. The multifaceted role of curcumin in cancer prevention and treatment. Molecules 2015, 20, 2728–2769. [Google Scholar] [CrossRef]
  92. Cheng, T.M.; Chin, Y.T.; Ho, Y.; Chen, Y.R.; Yang, Y.N.; Yang, Y.C.; Shih, Y.J.; Lin, T.I.; Lin, H.Y.; Davis, P.J. Resveratrol induces sumoylated COX-2-dependent anti-proliferation in human prostate cancer LNCaP cells. Food Chem. Toxicol. 2018, 112, 67–75. [Google Scholar] [CrossRef]
  93. Li, W.; Ma, X.Q.; Li, N.; Liu, H.S.; Dong, Q.; Zhang, J.; Yang, C.J.; Liu, Y.; Liang, Q.; Zhang, S.W.; et al. Resveratrol inhibits Hexokinases II mediated glycolysis in non-small cell lung cancer via targeting Akt signaling pathway. Exp. Cell Res. 2016, 349, 320–327. [Google Scholar] [CrossRef]
  94. Cai, H.; Scott, E.; Kholghi, A.; Andreadi, C.; Rufini, A.; Karmokar, A.; Britton, R.G.; Horner-Glister, E.; Greaves, P.; Jawad, D.; et al. Cancer chemoprevention: Evidence of a nonlinear dose response for the protective effects of resveratrol in humans and mice. Sci. Transl. Med. 2015, 7, 12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Zhao, Y.H.; Yuan, X.Y.; Li, X.; Zhang, Y. Resveratrol significantly inhibits the occurrence and development of cervical cancer by regulatingphospholipid scramblase 1. J. Cell. Biochem. 2019, 120, 1527–1531. [Google Scholar] [CrossRef] [PubMed]
  96. Yang, Z.Y.; Xie, Q.G.; Chen, Z.L.; Ni, H.B.; Xia, L.; Zhao, Q.F.; Chen, Z.Y.; Chen, P.F. Resveratrol suppresses the invasion and migration of human gastric cancer cells via inhibition of MALAT1-mediated epithelial-to-mesenchymal transition. Exp. Med. 2019, 17, 1569–1578. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Nguyen, L.T.; Lee, Y.H.; Sharma, A.R.; Park, J.B.; Nam, J.S. Quercetin induces apoptosis and cell cycle arrest in triple-negative breast cancer cells through modulation of Foxo3a activity. Korean J. Physiol. Pharm. Off. J. Korean Physiol. Soc. Korean Soc. Pharm. 2017, 21, 205–213. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Nwaeburu, C.C.; Bauer, N.; Zhao, Z.; Abukiwan, A.; Gladkich, J.; Benner, A.; Herr, I. Up-regulation of microRNA Let-7c by quercetin inhibits pancreatic cancer progression by activation of Numbl. Oncotarget 2016, 7, 58367–58380. [Google Scholar] [CrossRef] [Green Version]
  99. Klimaszewska-Wisniewska, A.; Halas-Wisniewska, M.; Izdebska, M.; Gagat, M.; Grzanka, A.; Grzanka, D. Antiproliferative and antimetastatic action of quercetin on A549 non-small cell lung cancer cells through its effect on the cytoskeleton. Acta Histochem. 2017, 119, 99–112. [Google Scholar] [CrossRef]
  100. Baruah, M.M.; Khandwekar, A.P.; Sharma, N. Quercetin modulates Wnt signaling components in prostate cancer cell line by inhibiting cell viability, migration, and metastases. Tumor Biol. 2016, 37, 14025–14034. [Google Scholar] [CrossRef]
  101. Moradzadeh, M.; Hosseini, A.; Erfanian, S.; Rezaei, H. Epigallocatechin-3-gallate promotes apoptosis in human breast cancer T47D cells through down-regulation of PI3K/AKT and Telomerase. Pharm. Rep. 2017, 69, 924–928. [Google Scholar] [CrossRef]
  102. Bi, Y.L.; Min, M.; Shen, W.; Liu, Y. Genistein induced anticancer effects on pancreatic cancer cell lines involves mitochondrial apoptosis, G0/G1cell cycle arrest and regulation of STAT3 signalling pathway. Phytomedicine 2018, 39, 10–16. [Google Scholar] [CrossRef] [PubMed]
  103. Qin, J.; Teng, J.A.; Zhu, Z.; Chen, J.X.; Huang, W.J. Genistein induces activation of the mitochondrial apoptosis pathway by inhibiting phosphorylation of Akt in colorectal cancer cells. Pharm. Biol. 2016, 54, 74–79. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Gundogdu, G.; Dodurga, Y.; Cetin, M.; Secme, M.; Cicek, B. The cytotoxic and genotoxic effects of daidzein on MIA PaCa-2 human pancreatic carcinoma cells and HT-29 human colon cancer cells. Drug Chem. Toxicol. 2020, 43, 581–587. [Google Scholar] [CrossRef] [PubMed]
  105. Hua, F.; Li, C.H.; Chen, X.G.; Liu, X.P. Daidzein exerts anticancer activity towards SKOV3 human ovarian cancer cells by inducing apoptosis and cell cycle arrest, and inhibiting the Raf/MEK/ERK cascade. Int. J. Mol. Med. 2018, 41, 3485–3492. [Google Scholar] [CrossRef] [PubMed]
  106. Han, B.J.; Li, W.; Jiang, G.B.; Lai, S.H.; Zhang, C.; Zeng, C.C.; Liu, Y.J. Effects of daidzein in regards to cytotoxicity in vitro, apoptosis, reactive oxygen species level, cell cycle arrest and the expression of caspase and Bcl-2 family proteins. Oncol. Rep. 2015, 34, 1115–1120. [Google Scholar] [CrossRef] [PubMed]
  107. Li, X.; Huang, J.M.; Wang, J.N.; Xiong, X.K.; Yang, X.F.; Zou, F. Combination of chrysin and cisplatin promotes the apoptosis of Hep G2 cells by up-regulating p53. Chem. -Biol. Interact. 2015, 232, 12–20. [Google Scholar] [CrossRef] [PubMed]
  108. Lim, H.K.; Kim, K.M.; Jeong, S.Y.; Choi, E.K.; Jung, J. Chrysin increases the therapeutic efficacy of docetaxel and mitigates docetaxel-induced edema. Integr. Cancer 2017, 16, 496–504. [Google Scholar] [CrossRef]
  109. Jaul, E.; Barron, J. Age-related diseases and clinical and public health implications for the 85 years old and over population. Front. Public Health 2017, 5, 7. [Google Scholar] [CrossRef] [Green Version]
  110. Di Meo, F.; Valentino, A.; Petillo, O.; Peluso, G.; Filosa, S.; Crispi, S. Bioactive polyphenols and neuromodulation: Molecular mechanisms in neurodegeneration. Int. J. Mol. Sci. 2020, 21, 2564. [Google Scholar] [CrossRef] [Green Version]
  111. Ciechanover, A.; Kwon, Y.T. Degradation of misfolded proteins in neurodegenerative diseases: Therapeutic targets and strategies. Exp. Mol. Med. 2015, 47, e147. [Google Scholar] [CrossRef] [Green Version]
  112. Soto, C.; Pritzkow, S. Protein misfolding, aggregation, and conformational strains in neurodegenerative diseases. Nat. Neurosci. 2018, 21, 1332–1340. [Google Scholar] [CrossRef] [PubMed]
  113. Vidoni, C.; Castiglioni, A.; Seca, C.; Secomandi, E.; Melone, M.A.B.; Isidoro, C. Dopamine exacerbates mutant Huntingtin toxicity via oxidative-mediated inhibition of autophagy in SH-SY5Y neuroblastoma cells: Beneficial effects of anti-oxidant therapeutics. Neurochem. Int. 2016, 101, 132–143. [Google Scholar] [CrossRef] [PubMed]
  114. Kim, G.H.; Kim, J.E.; Rhie, S.J.; Yoon, S. The role of oxidative stress in neurodegenerative diseases. Exp. Neurobiol 2015, 24, 325–340. [Google Scholar] [CrossRef] [PubMed]
  115. Carmela, S.; Marianna, N.; Idolo, T.; Stefania, M.; Alfonsina, M.; Gian Luigi, R. Neuroprotective role of natural polyphenols. Curr. Top. Med. Chem. 2016, 16, 1943–1950. [Google Scholar] [CrossRef]
  116. Davinelli, S.; Scapagnini, G.; Koverech, G.; Luca, M.; Calandra, C.; Calabrese, V. Neuroprotective mechanisms of dietary phytochemicals: Implications for successful brain aging. In Introduction to the Molecular Basis of Nutrition and Aging; Elsevier: Amsterdam, The Netherlands, 2016; Chapter 19; pp. 251–261. [Google Scholar]
  117. Moosavi, F.; Hosseini, R.; Saso, L.; Firuzi, O. Modulation of neurotrophic signaling pathways by polyphenols. Drug Des. Dev. 2016, 10, 23–42. [Google Scholar] [CrossRef] [Green Version]
  118. Zhao, D.Y.; Shah, N.P. Concomitant ingestion of lactic acid bacteria and black tea synergistically enhances flavonoid bioavailability and attenuates D-galactose-induced oxidative stress in mice via modulating glutathione antioxidant system. J. Nutr. Biochem. 2016, 38, 116–124. [Google Scholar] [CrossRef]
  119. Sandoval-Acuna, C.; Ferreira, J.; Speisky, H. Polyphenols and mitochondria: An update on their increasingly emerging ROS-scavenging independent actions. Arch. Biochem. Biophys. 2014, 559, 75–90. [Google Scholar] [CrossRef]
  120. Petersen, K.S.; Smith, C. Ageing-associated oxidative stress and inflammation are alleviated by products from grapes. Oxidative Med. Cell. Longev. 2016, 2016, 12. [Google Scholar] [CrossRef] [Green Version]
  121. Jayasena, T.; Poljak, A.; Smythe, G.; Braidy, N.; Münch, G.; Sachdev, P. The role of polyphenols in the modulation of sirtuins and other pathways involved in Alzheimer’s disease. Ageing Res. Rev. 2013, 12, 867–883. [Google Scholar] [CrossRef]
  122. Malvajerd, S.S.; Izadi, Z.; Azadi, A.; Kurd, M.; Derakhshankhah, H.; Zadeh, M.S.; Javar, H.A.; Hamidi, M. Neuroprotective potential of curcumin-loaded nanostructured lipid carrier in an animal model of Alzheimer’s disease: Behavioral and biochemical evidence. J. Alzheimers Dis. 2019, 69, 671–686. [Google Scholar] [CrossRef]
  123. Das, S.; Stark, L.; Musgrave, I.F.; Pukala, T.; Smid, S.D. Bioactive polyphenol interactions with beta amyloid: A comparison of binding modelling, effects on fibril and aggregate formation and neuroprotective capacity. Food Funct. 2016, 7, 1138–1146. [Google Scholar] [CrossRef]
  124. Neethirajan, S.; Jayas, D.S. Nanotechnology for the food and bioprocessing industries. Food Bioprocess. Technol. 2011, 4, 39–47. [Google Scholar] [CrossRef] [PubMed]
  125. Akhtartavan, S.; Karimi, M.; Karimian, K.; Azarpira, N.; Khatami, M.; Heli, H. Evaluation of a self-nanoemulsifying docetaxel delivery system. Biomed. Pharm. 2019, 109, 2427–2433. [Google Scholar] [CrossRef] [PubMed]
  126. Ranjbar, M.; Pardakhty, A.; Amanatfard, A.; Asadipour, A. Efficient drug delivery of β-estradiol encapsulated in Zn-metal–organic framework nanostructures by microwave-assisted coprecipitation method. Drug Des. Dev. 2018, 12, 2635–2643. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Alexander, A.; Ajazuddin; Patel, R.J.; Saraf, S.; Saraf, S. Recent expansion of pharmaceutical nanotechnologies and targeting strategies in the field of phytopharmaceuticals for the delivery of herbal extracts and bioactives. J. Control. Release 2016, 241, 110–124. [Google Scholar] [CrossRef]
  128. Prasad, S.; Tyagi, A.K.; Aggarwal, B.B. Recent developments in delivery, bioavailability, absorption and metabolism of curcumin: The golden pigment from golden spice. Cancer Res. Treat. 2014, 46, 2–18. [Google Scholar] [CrossRef] [Green Version]
  129. Bhingardeve, D.; Patil, S.; Patil, R. Phytosome-valuable phytophospholipid carriers. Curr. Pharm. Res. 2014, 1, 1386–1391. [Google Scholar]
  130. Amin, T.; Bhat, S.V. A Review on phytosome technology as a novel approach to improve the bioavailability of nutraceuticals. Int. J. Adv. Res. Technol. 2012, 1, 2278–7763. [Google Scholar]
  131. Rossi, R.; Basilico, F.; Rossoni, G.; Riva, A.; Morazzoni, P.; Mauri, P.L. Liquid chromatography/atmospheric pressure chemical ionization ion trap mass spectrometry of bilobalide in plasma and brain of rats after oral administration of its phospholipidic complex. J. Pharm. Biomed. Anal. 2009, 50, 224–227. [Google Scholar] [CrossRef]
  132. Yang, J.H.; Zhang, L.; Li, J.S.; Chen, L.H.; Zheng, Q.; Chen, T.; Chen, Z.P.; Fu, T.M.; Di, L.Q. Enhanced oral bioavailability and prophylactic effects on oxidative stress and hepatic damage of an oil solution containing a rosmarinic acid-phospholipid complex. J. Funct. Food. 2015, 19, 63–73. [Google Scholar] [CrossRef]
  133. Tedesco, D.; Steidler, S.; Galletti, S.; Tameni, M.; Sonzogni, O.; Ravarotto, L. Efficacy of silymarin-phospholipid complex in reducing the toxicity of aflatoxin B1 in broiler chicks. Poult. Sci. 2004, 83, 1839–1843. [Google Scholar] [CrossRef] [PubMed]
  134. Mirzaei, H.; Shakeri, A.; Rashidi, B.; Jalili, A.; Banikazemi, Z.; Sahebkar, A. Phytosomal curcumin: A review of pharmacokinetic, experimental and clinical studies. Biomed. Pharm. 2017, 85, 102–112. [Google Scholar] [CrossRef] [PubMed]
  135. Allam, A.N.; Komeil, I.A.; Fouda, M.A.; Abdallah, O.Y. Preparation, characterization and in vivo evaluation of curcumin self-nano phospholipid dispersion as an approach to enhance oral bioavailability. Int. J. Pharm. 2015, 489, 117–123. [Google Scholar] [CrossRef] [PubMed]
  136. Marczylo, T.H.; Verschoyle, R.D.; Cooke, D.N.; Morazzoni, P.; Steward, W.P.; Gescher, A.J. Comparison of systemic availability of curcumin with that of curcumin formulated with phosphatidylcholine. Cancer Chemother. Pharm. 2007, 60, 171–177. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Zhang, J.; Tang, Q.; Xu, X.; Li, N. Development and evaluation of a novel phytosome-loaded chitosan microsphere system for curcumin delivery. Int. J. Pharm. 2013, 448, 168–174. [Google Scholar] [CrossRef] [PubMed]
  138. Lim, A.W.; Ng, P.Y.; Chieng, N.; Ng, S.F. Moringa oleifera leaf extract-loaded phytophospholipid complex for potential application as wound dressing. J. Drug Deliv. Sci. Technol. 2019, 54, 9. [Google Scholar] [CrossRef]
  139. Riva, A.; Ronchi, M.; Petrangolini, G.; Bosisio, S.; Allegrini, P. Improved oral absorption of quercetin from quercetin phytosome (R), a new delivery system based on food grade lecithin. Eur. J. Drug Metab. Pharm. 2019, 44, 169–177. [Google Scholar] [CrossRef] [Green Version]
  140. Mahmoodi, N.; Motamed, N.; Paylakhi, S.H.; Mahmoodi, N.O. Comparing the effect of silybin and silybin advanced (TM) on viability and HER2 expression on the human breast cancer SKBR3 cell line by no serum starvation. Iran. J. Pharm. Res. 2015, 14, 521–530. [Google Scholar]
  141. Wei, L.; Li, X.; Guo, F.; Liu, X.; Wang, Z. Structural properties, in vitro release and radical scavenging activity of lecithin based curcumin-encapsulated inverse hexagonal (HII) liquid crystals. Colloids Surf. A 2018, 539, 124–131. [Google Scholar] [CrossRef]
  142. Baradaran, S.; Moghaddam, A.H.; Jelodar, S.K.; Moradi-kor, N. Protective effects of curcumin and its nano-phytosome on carrageenan-induced inflammation in mice model: Behavioral and biochemical responses. J. Inflamm. Res. 2020, 13, 45–51. [Google Scholar] [CrossRef] [Green Version]
  143. Bernardo, J.; Videira, R.A.; Valentão, P.; Veiga, F.; Andrade, P.B. Extraction of phospholipid-rich fractions from egg yolk and development of liposomes entrapping a dietary polyphenol with neuroactive potential. Food Chem. Toxicol. 2019, 133, 110749. [Google Scholar] [CrossRef]
  144. Semalty, A.; Semalty, M.; Singh, D.; Rawat, M.S.M. Phyto-phospholipid complex of catechin in value added herbal drug delivery. J. Incl. Phenom. Macrocycl. Chem. 2012, 73, 377–386. [Google Scholar] [CrossRef]
  145. Khan, J.; Saraf, S.; Saraf, S. Preparation and evaluation of luteolin-phospholipid complex as an effective drug delivery tool against GalN/LPS induced liver damage. Pharm. Dev. Technol. 2016, 21, 475–486. [Google Scholar] [CrossRef] [PubMed]
  146. Chi, C.; Zhang, C.S.; Liu, Y.; Nie, H.C.; Zhou, J.P.; Ding, Y. Phytosome-nanosuspensions for silybin-phospholipid complex with increased bioavailability and hepatoprotection efficacy. Eur. J. Pharm. Sci. 2020, 144, 12. [Google Scholar] [CrossRef] [PubMed]
  147. Lazzeroni, M.; Guerrieri-Gonzaga, A.; Gandini, S.; Johansson, H.; Serrano, D.; Cazzaniga, M.; Aristarco, V.; Macis, D.; Mora, S.; Caldarella, P.; et al. A Presurgical study of lecithin formulation of green tea extract in women with early breast cancer. Cancer Prev. Res. 2017, 10, 363–369. [Google Scholar] [CrossRef] [Green Version]
  148. Mao, J.T.; Xue, B.Y.; Smoake, J.; Lu, Q.Y.; Park, H.; Henning, S.M.; Burns, W.; Bernabei, A.; Elashoff, D.; Serio, K.J.; et al. MicroRNA-19a/b mediates grape seed procyanidin extract-induced anti-neoplastic effects against lung cancer. J. Nutr. Biochem. 2016, 34, 118–125. [Google Scholar] [CrossRef]
  149. Abd El-Fattah, A.I.; Fathy, M.M.; Ali, Z.Y.; El-Garawany, A.E.-R.A.; Mohamed, E.K. Enhanced therapeutic benefit of quercetin-loaded phytosome nanoparticles in ovariectomized rats. Chem. Biol. Interact. 2017, 271, 30–38. [Google Scholar] [CrossRef]
  150. Angelico, R.; Ceglie, A.; Sacco, P.; Colafemmina, G.; Ripoli, M.; Mangia, A. Phyto-liposomes as nanoshuttles for water-insoluble silybin–phospholipid complex. Int. J. Pharm. 2014, 471, 173–181. [Google Scholar] [CrossRef]
  151. Direito, R.; Reis, C.; Roque, L.; Goncalves, M.; Sanches-Silva, A.; Gaspar, M.M.; Pinto, R.; Rocha, J.; Sepodes, B.; Bronze, M.R.; et al. Phytosomes with persimmon (Diospyros kaki L.) Extract: Preparation and preliminary demonstration of in vivo tolerability. Pharmaceutics 2019, 11, 296. [Google Scholar] [CrossRef] [Green Version]
  152. Mozafari, M.R.; Johnson, C.; Hatziantoniou, S.; Demetzos, C. Nanoliposomes and their applications in food nanotechnology. J. Liposome Res. 2008, 18, 309–327. [Google Scholar] [CrossRef]
  153. Ghorbanzade, T.; Jafari, S.M.; Akhavan, S.; Hadavi, R. Nano-encapsulation of fish oil in nano-liposomes and its application in fortification of yogurt. Food Chem. 2017, 216, 146–152. [Google Scholar] [CrossRef] [PubMed]
  154. Fang, J.-Y.; Fang, C.-L.; Liu, C.-H.; Su, Y.-H. Lipid nanoparticles as vehicles for topical psoralen delivery: Solid lipid nanoparticles (SLN) versus nanostructured lipid carriers (NLC). Eur. J. Pharm. Biopharm. 2008, 70, 633–640. [Google Scholar] [CrossRef] [PubMed]
  155. Mozafari, M.R. Liposomes: An overview of manufacturing techniques. Cell. Mol. Biol. Lett. 2005, 10, 711–719. [Google Scholar] [PubMed]
  156. Huang, Z.; Li, X.; Zhang, T.; Song, Y.; She, Z.; Li, J.; Deng, Y. Progress involving new techniques for liposome preparation. Asian J. Pharm. Sci. 2014, 9, 176–182. [Google Scholar] [CrossRef] [Green Version]
  157. Pattni, B.S.; Chupin, V.V.; Torchilin, V.P. New developments in liposomal drug delivery. Chem. Rev. 2015, 115, 10938–10966. [Google Scholar] [CrossRef]
  158. Askarizadeh, A.; Butler, A.E.; Badiee, A.; Sahebkar, A. Liposomal nanocarriers for statins: A pharmacokinetic and pharmacodynamics appraisal. J. Cell. Physiol. 2019, 234, 1219–1229. [Google Scholar] [CrossRef]
  159. Shishir, M.R.I.; Karim, N.; Gowd, V.; Zheng, X.D.; Chen, W. Liposomal delivery of natural product: A promising approach in health research. Trends Food Sci. Technol. 2019, 85, 177–200. [Google Scholar] [CrossRef]
  160. Cheng, C.; Peng, S.F.; Li, Z.L.; Zou, L.Q.; Liu, W.; Liu, C.M. Improved bioavailability of curcumin in liposomes prepared using a pH-driven, organic solvent-free, easily scalable process. RSC Adv. 2017, 7, 25978–25986. [Google Scholar] [CrossRef] [Green Version]
  161. Vanaja, K.; Wahl, M.A.; Bukarica, L.; Heinle, H. Liposomes as carriers of the lipid soluble antioxidant resveratrol: Evaluation of amelioration of oxidative stress by additional antioxidant vitamin. Life Sci. 2013, 93, 917–923. [Google Scholar] [CrossRef]
  162. Odeh, F.; Nsairat, H.; Alshaer, W.; Alsotari, S.; Bawab, A.A. Remote loading of curcumin-in-modified β-cyclodextrins into liposomes using a transmembrane pH gradient. RSC Adv. 2019, 9, 37148–37161. [Google Scholar] [CrossRef] [Green Version]
  163. Huang, M.G.; Liang, C.P.; Tan, C.; Huang, S.; Ying, R.F.; Wang, Y.S.; Wang, Z.J.; Zhang, Y.F. Liposome co-encapsulation as a strategy for the delivery of curcumin and resveratrol. Food Funct. 2019, 10, 6447–6458. [Google Scholar] [CrossRef] [PubMed]
  164. Caddeo, C.; Nacher, A.; Vassallo, A.; Armentano, M.F.; Pons, R.; Fernàndez-Busquets, X.; Carbone, C.; Valenti, D.; Fadda, A.M.; Manconi, M. Effect of quercetin and resveratrol co-incorporated in liposomes against inflammatory/oxidative response associated with skin cancer. Int. J. Pharm. 2016, 513, 153–163. [Google Scholar] [CrossRef] [PubMed]
  165. Colom, H.; Alfaras, I.; Maijo, M.; Juan, M.E.; Planas, J.M. Population pharmacokinetic modeling of trans-resveratrol and its glucuronide and sulfate conjugates after oral and intravenous administration in rats. Pharm. Res. 2011, 28, 1606–1621. [Google Scholar] [CrossRef] [PubMed]
  166. Caddeo, C.; Pucci, L.; Gabriele, M.; Carbone, C.; Fernàndez-Busquets, X.; Valenti, D.; Pons, R.; Vassallo, A.; Fadda, A.M.; Manconi, M. Stability, biocompatibility and antioxidant activity of PEG-modified liposomes containing resveratrol. Int. J. Pharm. 2018, 538, 40–47. [Google Scholar] [CrossRef] [PubMed]
  167. Dag, D.; Guner, S.; Oztop, M.H. Physicochemical mechanisms of different biopolymers’ (lysozyme, gum arabic, whey protein, chitosan) adsorption on green tea extract loaded liposomes. Int. J. Biol. Macromol. 2019, 138, 473–482. [Google Scholar] [CrossRef] [PubMed]
  168. Pandita, D.; Kumar, S.; Poonia, N.; Lather, V. Solid lipid nanoparticles enhance oral bioavailability of resveratrol, a natural polyphenol. Food Res. Int. 2014, 62, 1165–1174. [Google Scholar] [CrossRef]
  169. Jose, S.; Anju, S.S.; Cinu, T.A.; Aleykutty, N.A.; Thomas, S.; Souto, E.B. In vivo pharmacokinetics and biodistribution of resveratrol-loaded solid lipid nanoparticles for brain delivery. Int. J. Pharm. 2014, 474, 6–13. [Google Scholar] [CrossRef]
  170. Ramalingam, P.; Yoo, S.W.; Ko, Y.T. Nanodelivery systems based on mucoadhesive polymer coated solid lipid nanoparticles to improve the oral intake of food curcumin. Food Res. Int. 2016, 84, 113–119. [Google Scholar] [CrossRef]
  171. Pinheiro, R.G.R.; Granja, A.; Loureiro, J.A.; Pereira, M.C.; Pinheiro, M.; Neves, A.R.; Reis, S. Quercetin lipid nanoparticles functionalized with transferrin for Alzheimer’s disease. Eur. J. Pharm. Sci. 2020, 148, 11. [Google Scholar] [CrossRef]
  172. Rajera, R.; Nagpal, K.; Singh, S.K.; Mishra, D.N. Niosomes: A controlled and novel drug delivery system. Biol. Pharm. Bull. 2011, 34, 945–953. [Google Scholar] [CrossRef] [Green Version]
  173. Biswas, G.R.; Majee, S.B. Niosomes in ocular drug delivery. Eur. J. Pharm. Med. Res. 2017, 4, 813–819. [Google Scholar]
  174. Mahale, N.B.; Thakkar, P.D.; Mali, R.G.; Walunj, D.R.; Chaudhari, S.R. Niosomes: Novel sustained release nonionic stable vesicular systems—An overview. Adv. Colloid Interface Sci. 2012, 183, 46–54. [Google Scholar] [CrossRef] [PubMed]
  175. Marianecci, C.; Di Marzio, L.; Rinaldi, F.; Celia, C.; Paolino, D.; Alhaique, F.; Esposito, S.; Carafa, M. Niosomes from 80s to present: The state of the art. Adv. Colloid Interface Sci. 2014, 205, 187–206. [Google Scholar] [CrossRef] [PubMed]
  176. Moghassemi, S.; Hadjizadeh, A. Nano-niosomes as nanoscale drug delivery systems: An illustrated review. J. Control. Release 2014, 185, 22–36. [Google Scholar] [CrossRef]
  177. Lu, B.Y.; Huang, Y.T.; Chen, Z.Y.; Ye, J.Y.; Xu, H.Y.; Chen, W.R.; Long, X.Y. Niosomal nanocarriers for enhanced skin delivery of quercetin with functions of anti-tyrosinase and antioxidant. Molecules 2019, 24, 2322. [Google Scholar] [CrossRef] [Green Version]
  178. Patel, J.; Ketkar, S.; Patil, S.; Fearnley, J.; Mahadik, K.R.; Paradkar, A.R. Potentiating antimicrobial efficacy of propolis through niosomal-based system for administration. Integr. Med. Res. 2015, 4, 94–101. [Google Scholar] [CrossRef] [Green Version]
  179. Alemi, A.; Reza, J.Z.; Haghiralsadat, F.; Jaliani, H.Z.; Karamallah, M.H.; Hosseini, S.A.; Karamallah, S.H. Paclitaxel and curcumin coadministration in novel cationic PEGylated niosomal formulations exhibit enhanced synergistic antitumor efficacy. J. Nanobiotechnol. 2018, 16, 20. [Google Scholar] [CrossRef]
  180. Pereira, P.C. Milk nutritional composition and its role in human health. Nutrition 2014, 30, 619–627. [Google Scholar] [CrossRef]
  181. Fox, P.F.; Brodkorb, A. The casein micelle: Historical aspects, current concepts and significance. Int. Dairy J. 2008, 18, 677–684. [Google Scholar] [CrossRef]
  182. Livney, Y.D. Milk proteins as vehicles for bioactives. Curr. Opin. Colloid Interface Sci. 2010, 15, 73–83. [Google Scholar] [CrossRef]
  183. Esmaili, M.; Ghaffari, S.-M.; Moosavi-Movahedi, Z.; Atri, M.S.; Sharifizadeh, A.; Farhadi, M.; Yousefi, R.; Chobert, J.M.; Haertlé, T.; Moosavi-Movahedi, A.A. Beta casein-micelle as a nano vehicle for solubility enhancement of curcumin; food industry application. LWT Food Sci. Technol. 2011, 44, 2166–2172. [Google Scholar] [CrossRef]
  184. Luo, Y.C.; Pan, K.; Zhong, Q.X. Casein/pectin nanocomplexes as potential oral delivery vehicles. Int. J. Pharm. 2015, 486, 59–68. [Google Scholar] [CrossRef] [PubMed]
  185. Rashidinejad, A.; Loveday, S.M.; Jameson, G.B.; Hindmarsh, J.P.; Singh, H. Rutin-casein co-precipitates as potential delivery vehicles for flavonoid rutin. Food Hydrocoll. 2019, 96, 451–462. [Google Scholar] [CrossRef]
  186. Ghayour, N.; Hosseini, S.M.H.; Eskandari, M.H.; Esteghlal, S.; Nekoei, A.R.; Gahruie, H.H.; Tatar, M.; Naghibalhossaini, F. Nanoencapsulation of quercetin and curcumin in casein-based delivery systems. Food Hydrocoll. 2019, 87, 394–403. [Google Scholar] [CrossRef]
  187. Foox, M.; Zilberman, M. Drug delivery from gelatin-based systems. Expert Opin. Drug Deliv. 2015, 12, 1–17. [Google Scholar] [CrossRef] [PubMed]
  188. Elzoghby, A.O.; Samy, W.M.; Elgindy, N.A. Protein-based nanocarriers as promising drug and gene delivery systems. J. Control. Release 2012, 161, 38–49. [Google Scholar] [CrossRef] [PubMed]
  189. Sahoo, N.; Sahoo, R.K.; Biswas, N.; Guha, A.; Kuotsu, K. Recent advancement of gelatin nanoparticles in drug and vaccine delivery. Int. J. Biol. Macromol. 2015, 81, 317–331. [Google Scholar] [CrossRef] [PubMed]
  190. Kommareddy, S.; Shenoy, D.B.; Amiji, M.M. Gelatin Nanoparticles and Their Biofunctionalization; Wiley: New York, NY, USA, 2007. [Google Scholar]
  191. Wang, H.; Boerman, O.C.; Sariibrahimoglu, K.; Li, Y.; Jansen, J.A.; Leeuwenburgh, S.C.G. Comparison of micro vs. nanostructured colloidal gelatin gels for sustained delivery of osteogenic proteins: Bone morphogenetic protein-2 and alkaline phosphatase. Biomaterials 2012, 33, 8695–8703. [Google Scholar] [CrossRef] [PubMed]
  192. Ofokansi, K.; Winter, G.; Fricker, G.; Coester, C. Matrix-loaded biodegradable gelatin nanoparticles as new approach to improve drug loading and delivery. Eur. J. Pharm. Biopharm. 2010, 76, 1–9. [Google Scholar] [CrossRef]
  193. Zhao, Y.-Z.; Li, X.; Lu, C.-T.; Xu, Y.-Y.; Lv, H.-F.; Dai, D.-D.; Zhang, L.; Sun, C.-Z.; Yang, W.; Li, X.-K.; et al. Experiment on the feasibility of using modified gelatin nanoparticles as insulin pulmonary administration system for diabetes therapy. Acta Diabetol. 2012, 49, 315–325. [Google Scholar] [CrossRef]
  194. Khan, S.A.; Schneider, M. Improvement of nanoprecipitation technique for preparation of gelatin nanoparticles and potential macromolecular drug loading. Macromol. Biosci. 2013, 13, 455–463. [Google Scholar] [CrossRef] [PubMed]
  195. Thein-Han, W.W.; Saikhun, J.; Pholpramoo, C.; Misra, R.D.K.; Kitiyanant, Y. Chitosan-gelatin scaffolds for tissue engineering: Physico-chemical properties and biological response of buffalo embryonic stem cells and transfectant of GFP-buffalo embryonic stem cells. Acta Biomater. 2009, 5, 3453–3466. [Google Scholar] [CrossRef] [PubMed]
  196. Shutava, T.G.; Balkundi, S.S.; Vangala, P.; Steffan, J.J.; Bigelow, R.L.; Cardelli, J.A.; O’Neal, D.P.; Lvov, Y.M. Layer-by-layer-coated gelatin nanoparticles as a vehicle for delivery of natural polyphenols. ACS Nano 2009, 3, 1877–1885. [Google Scholar] [CrossRef] [PubMed]
  197. 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. Pharm. 2015, 70, 274–282. [Google Scholar] [CrossRef]
  198. Liu, F.; Avena-Bustillos, R.J.; Chiou, B.S.; Li, Y.; Ma, Y.; Williams, T.G.; Wood, D.F.; McHugh, T.H.; Zhong, F. Controlled-release of tea polyphenol from gelatin films incorporated with different ratios of free/nanoencapsulated tea polyphenols into fatty food simulants. Food Hydrocoll. 2017, 62, 212–221. [Google Scholar] [CrossRef]
  199. Fox, P.F.; Mcsweeney, P. Advanced dairy chemistry. Adv. Dairy Chem. 2013, 1, 337–385. [Google Scholar]
  200. Gunasekaran, S.; Ko, S.; Xiao, L. Use of whey proteins for encapsulation and controlled delivery applications. J. Food Eng. 2007, 83, 31–40. [Google Scholar] [CrossRef]
  201. Chen, L.; Subirade, M. Chitosan/β-lactoglobulin core–shell nanoparticles as nutraceutical carriers. Biomaterials 2005, 26, 6041–6053. [Google Scholar] [CrossRef]
  202. Shpigelman, A.; Israeli, G.; Livney, Y.D. Thermally-induced protein-polyphenol co-assemblies: Beta lactoglobulin-based nanocomplexes as protective nanovehicles for EGCG. Food Hydrocoll. 2010, 24, 735–743. [Google Scholar] [CrossRef]
  203. Li, M.; Cui, J.; Ngadi, M.O.; Ma, Y. Absorption mechanism of whey-protein-delivered curcumin using Caco-2 cell monolayers. Food Chem. 2015, 180, 48–54. [Google Scholar] [CrossRef]
  204. Li, X.Y.; Gao, Z.L.; Li, T.; Sarker, S.K.; Chowdhury, S.; Jiang, Z.M.; Mu, Z.S. Effects of pH Values on physicochemical properties and antioxidant potential of whey protein isolate-safflower yellow complexes. Food Sci. Technol. Res. 2018, 24, 475–484. [Google Scholar] [CrossRef] [Green Version]
  205. Chen, Y.; Zhang, R.; Xie, B.; Sun, Z.; McClements, D.J. Lotus seedpod proanthocyanidin-whey protein complexes: Impact on physical and chemical stability of β-carotene-nanoemulsions. Food Res. Int. 2020, 127, 108738. [Google Scholar] [CrossRef] [PubMed]
  206. De Morais, F.P.R.; Pessato, T.B.; Rodrigues, E.; Mallmann, L.P.; Mariutti, L.R.B.; Netto, F.M. Whey protein and phenolic compound complexation: Effects on antioxidant capacity before and after in vitro digestion. Food Res. Int. 2020, 133, 11. [Google Scholar] [CrossRef]
  207. Pérez-Herrero, E.; Fernández-Medarde, A. Advanced targeted therapies in cancer: Drug nanocarriers, the future of chemotherapy. Eur. J. Pharm. Biopharm. 2015, 93, 52–79. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  208. Wu, W.; Chen, M.; Luo, T.R.; Fan, Y.; Zhang, J.Q.; Zhang, Y.; Zhang, Q.Y.; Sapin-Minet, A.; Gaucher, C.; Xia, X.F. ROS and GSH-responsive S-nitrosoglutathione functionalized polymeric nanoparticles to overcome multidrug resistance in cancer. Acta Biomater. 2020, 103, 259–271. [Google Scholar] [CrossRef]
  209. Kamaly, N.; Yameen, B.; Wu, J.; Farokhzad, O.C. Degradable controlled-release polymers and polymeric nanoparticles: Mechanisms of controlling drug release. Chem. Rev. 2016, 116, 2602–2663. [Google Scholar] [CrossRef] [Green Version]
  210. El-Say, K.M.; El-Sawy, H.S. Polymeric nanoparticles: Promising platform for drug delivery. Int. J. Pharm. 2017, 528, 675–691. [Google Scholar] [CrossRef]
  211. Li, J.Y.; Sabliov, C. PLA/PLGA nanoparticles for delivery of drugs across the blood-brain barrier. Nanotechnol. Rev. 2013, 2, 241–257. [Google Scholar] [CrossRef]
  212. Pan, Q.; Li, W.; Yuan, X.; Rakhmanov, Y.; Wang, P.; Lu, R.; Mao, Z.; Shang, X.; You, H. Chondrogenic effect of cell-based scaffold of self-assembling peptides/PLGA-PLL loading the hTGFβ3 plasmid DNA. J. Mater. Sci. Mater. Med. 2016, 27, 19. [Google Scholar] [CrossRef]
  213. 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, 41573. [Google Scholar] [CrossRef] [Green Version]
  214. Cheng, C.Y.; Quoc-Hue, P.; Wu, X.Y.; Ting-Yu, C.; Chen, C.M.; Fang, P.H.; Lin, Y.C.; Ming-Fa, H. PLGA microspheres loaded with β-cyclodextrin complexes of epigallocatechin-3-gallate for the anti-inflammatory properties in activated microglial cells. Polymers 2018, 10, 519. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Martins, C.; Vilarinho, F.; Silva, A.S.; Andrade, M.; Machado, A.V.; Castilho, M.C.; Sa, A.; Cunha, A.; Vaz, M.F.; Ramos, F. Active polylactic acid film incorporated with green tea extract: Development, characterization and effectiveness. Ind. Crop. Prod. 2018, 123, 100–110. [Google Scholar] [CrossRef] [Green Version]
  216. Nassir, A.M.; Shahzad, N.; Ibrahim, I.A.A.; Ahmad, I.; Md, S.; Ain, M.R. Resveratrol-loaded PLGA nanoparticles mediated programmed cell death in prostate cancer cells. Saudi Pharm. J. 2018, 26, 876–885. [Google Scholar] [CrossRef] [PubMed]
  217. Yallapu, M.M.; Khan, S.; Maher, D.M.; Ebeling, M.C.; Sundram, V.; Chauhan, N.; Ganju, A.; Balakrishna, S.; Gupta, B.K.; Zafar, N.; et al. Anti-cancer activity of curcumin loaded nanoparticles in prostate cancer. Biomaterials 2014, 35, 8635–8648. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  218. Khutoryanskiy, V.V. advances in mucoadhesion and mucoadhesive polymers. Macromol. Biosci. 2011, 11, 748–764. [Google Scholar] [CrossRef] [PubMed]
  219. Palacio, J.; Agudelo, N.A.; Lopez, B.L. PEGylation of PLA nanoparticles to improve mucus-penetration and colloidal stability for oral delivery systems. Curr. Opin. Chem. Eng. 2016, 11, 14–19. [Google Scholar] [CrossRef]
  220. 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]
  221. Duan, Y.; Zhang, B.; Chu, L.; Tong, H.H.Y.; Liu, W.; Zhai, G. Evaluation in vitro and in vivo of curcumin-loaded mPEG-PLA/TPGS mixed micelles for oral administration. Colloids Surf. B Biointerfaces 2016, 141, 345–354. [Google Scholar] [CrossRef]
  222. Radhakrishnan, R.; Kulhari, H.; Pooja, D.; Gudem, S.; Bhargava, S.; Shukla, R.; Sistla, R. Encapsulation of biophenolic phytochemical EGCG within lipid nanoparticles enhances its stability and cytotoxicity against cancer. Chemistry and Physics of Lipids 2016, 198, 51–60. [Google Scholar] [CrossRef]
  223. Miele, D.; Catenacci, L.; Sorrenti, M.; Rossi, S.; Sandri, G.; Malavasi, L.; Dacarro, G.; Ferrari, F.; Bonferoni, M.C. Chitosan oleate coated poly lactic-glycolic acid (PLGA) nanoparticles versus chitosan oleate self-assembled polymeric micelles, loaded with resveratrol. Mar. Drugs 2019, 17, 16. [Google Scholar] [CrossRef] [Green Version]
  224. Elgadir, M.A.; Uddin, M.S.; Ferdosh, S.; Adam, A.; Chowdhury, A.J.K.; Sarker, M.Z.I. Impact of chitosan composites and chitosan nanoparticle composites on various drug delivery systems: A review. J. Food Drug Anal. 2015, 23, 619–629. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Zhang, X.J.; He, C.C.; Yan, R.C.; Chen, Y.; Zhao, P.X.; Li, M.S.; Fan, T.; Yang, T.; Lu, Y.; Luo, J.; et al. HIF-1 dependent reversal of cisplatin resistance via anti-oxidative nano selenium for effective cancer therapy. Chem. Eng. J. 2020, 380, 12. [Google Scholar] [CrossRef]
  226. Ways, T.M.M.; Lau, W.M.; Khutoryanskiy, V.V. Chitosan and its derivatives for application in mucoadhesive drug delivery systems. Polymers 2018, 10, 267. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  227. Siddiqui, I.A.; Bharali, D.J.; Nihal, M.; Adhami, V.M.; Khan, N.; Chamcheu, J.C.; Khan, M.I.; Shabana, S.; Mousa, S.A.; Mukhtar, H. Excellent anti-proliferative and pro-apoptotic effects of (−)-epigallocatechin-3-gallate encapsulated in chitosan nanoparticles on human melanoma cell growth both in vitro and in vivo. Nanomed. Nanotechnol. Biol. Med. 2014, 10, 1619–1626. [Google Scholar] [CrossRef] [PubMed]
  228. Lin, Y.H.; Chen, Z.R.; Lai, C.H.; Hsieh, C.H.; Feng, C.L. Active targeted nanoparticles for oral administration of gastric cancer therapy. Biomacromolecules 2015, 16, 3021–3032. [Google Scholar] [CrossRef] [PubMed]
  229. Tan, C.; Xie, J.; Zhang, X.; Cai, J.; Xia, S. Polysaccharide-based nanoparticles by chitosan and gum Arabic polyelectrolyte complexation as carriers for curcumin. Food Hydrocoll. 2016, 57, 236–245. [Google Scholar] [CrossRef]
  230. Chuah, L.H.; Roberts, C.J.; Billa, N.; Abdullah, S.; Rosli, R. Cellular uptake and anticancer effects of mucoadhesive curcumin-containing chitosan nanoparticles. Colloid Surf. B Biointerfaces 2014, 116, 228–236. [Google Scholar] [CrossRef]
  231. Bhunchu, S.; Rojsitthisak, P.; Rojsitthisak, P. Effects of preparation parameters on the characteristics of chitosan–alginate nanoparticles containing curcumin diethyl disuccinate. J. Drug Deliv. Sci. Technol. 2015, 28, 64–72. [Google Scholar] [CrossRef]
  232. Zu, Y.G.; Zhang, Y.; Wang, W.G.; Zhao, X.H.; Han, X.; Wang, K.L.; Ge, Y.L. Preparation and in vitro/in vivo evaluation of resveratrol-loaded carboxymethyl chitosan nanoparticles. Drug Deliv. 2016, 23, 981–991. [Google Scholar] [CrossRef]
  233. Wang, H.X.; Qan, J.; Ding, F.Y. Emerging chitosan-based films for food packaging applications. J. Agric. Food Chem. 2018, 66, 395–413. [Google Scholar] [CrossRef]
  234. Liu, J.; Meng, C.-g.; Liu, S.; Kan, J.; Jin, C.-h. Preparation and characterization of protocatechuic acid grafted chitosan films with antioxidant activity. Food Hydrocoll. 2017, 63, 457–466. [Google Scholar] [CrossRef]
  235. Liu, J.; Liu, S.; Wu, Q.; Gu, Y.; Kan, J.; Jin, C. Effect of protocatechuic acid incorporation on the physical, mechanical, structural and antioxidant properties of chitosan film. Food Hydrocoll. 2017, 73, 90–100. [Google Scholar] [CrossRef]
  236. Wang, X.C.; Yong, H.M.; Gao, L.; Li, L.L.; Jin, M.J.; Liu, J. Preparation and characterization of antioxidant and pH-sensitive films based on chitosan and black soybean seed coat extract. Food Hydrocoll. 2019, 89, 56–66. [Google Scholar] [CrossRef]
  237. Kurek, M.; Garofulic, I.E.; Bakic, M.T.; Scetar, M.; Uzelac, V.D.; Galic, K. Development and evaluation of a novel antioxidant and pH indicator film based on chitosan and food waste sources of antioxidants. Food Hydrocoll. 2018, 84, 238–246. [Google Scholar] [CrossRef]
  238. Kurkov, S.V.; Loftsson, T. Cyclodextrins. Int. J. Pharm. 2013, 453, 167–180. [Google Scholar] [CrossRef]
  239. Qiu, C.; Julian McClements, D.; Jin, Z.; Qin, Y.; Hu, Y.; Xu, X.; Wang, J. Resveratrol-loaded core-shell nanostructured delivery systems: Cyclodextrin-based metal-organic nanocapsules prepared by ionic gelation. Food Chem. 2020, 317, 126328. [Google Scholar] [CrossRef]
  240. Aree, T.; Jongrungruangchok, S. Structure–antioxidant activity relationship of β-cyclodextrin inclusion complexes with olive tyrosol, hydroxytyrosol and oleuropein: Deep insights from X-ray analysis, DFT calculation and DPPH assay. Carbohydr. Polym. 2018, 199, 661–669. [Google Scholar] [CrossRef]
  241. Malapert, A.; Tomao, V.; Margier, M.; Nowicki, M.; Gleize, B. β-Cyclodextrin does not alter the bioaccessibility and the uptake by caco-2 cells of olive by-product phenolic compounds. Nutrients 2018, 10, 1653. [Google Scholar] [CrossRef] [Green Version]
  242. Żyżelewicz, D.; Oracz, J.; Kaczmarska, M.; Budryn, G.; Grzelczyk, J. Preparation and characterization of inclusion complex of (+)-catechin with β-cyclodextrin. Food Res. Int. 2018, 113, 263–268. [Google Scholar] [CrossRef]
  243. Ho, S.; Thoo, Y.Y.; Young, D.J.; Siow, L.F. Stability and recovery of cyclodextrin encapsulated catechin in various food matrices. Food Chem. 2019, 275, 594–599. [Google Scholar] [CrossRef]
  244. Chao, Q.; Jinpeng, W.; Huang, Z.; Yang, Q.; Xueming, X.; Zhengyu, J. A novel approach with controlled nucleation and growth for green synthesis of size-controlled cyclodextrin-based metal—Organic frameworks based on short-chain starch nanoparticles. J. Agric. Food Chem. 2018, 66, 9785–9793. [Google Scholar]
  245. Chen, Y.; Tang, H.L.; Liu, Y.; Tan, H.M. Preparation and study on the volume phase transition properties of novel carboxymethyl chitosan grafted polyampholyte superabsorbent polymers. J. Taiwan Inst. Chem. Eng. 2016, 59, 569–577. [Google Scholar] [CrossRef]
  246. Ng, W.L.; Yeong, W.Y.; Naing, M.W. Development of polyelectrolyte chitosan-gelatin hydrogels for skin bioprinting. In Second Cirp Conference on Biomanufacturing; Bartolo, P., Ed.; Elsevier Science Bv: Amsterdam, The Netherlands, 2016; Volume 49, pp. 105–112. [Google Scholar]
  247. Zhao, C.-C.; Zhu, L.; Wu, Z.; Yang, R.; Xu, N.; Liang, L. Resveratrol-loaded peptide-hydrogels inhibit scar formation in wound healing through suppressing inflammation. Regen. Biomater. 2020, 7, 99–107. [Google Scholar] [CrossRef] [PubMed]
  248. Chen, G.Y.; He, L.B.; Zhang, P.; Zhang, J.; Mei, X.F.; Wang, D.H.; Zhang, Y.Y.; Ren, X.L.; Chen, Z.H. Encapsulation of green tea polyphenol nanospheres in PVA/alginate hydrogel for promoting wound healing of diabetic rats by regulating PI3K/AKT pathway. Mater. Sci. Eng. C Mater. Biol. Appl. 2020, 110, 10. [Google Scholar] [CrossRef]
  249. Zhou, L.; Fan, L.; Yi, X.; Zhou, Z.N.; Liu, C.; Fu, R.M.; Dai, C.; Wang, Z.G.; Chen, X.X.; Yu, P.; et al. Soft conducting polymer hydrogels cross-linked and doped by tannic acid for spinal cord injury repair. ACS Nano 2018, 12, 10957–10967. [Google Scholar] [CrossRef]
  250. Tan, H.; Sun, J.; Jin, D.; Song, J.; Lei, M.; Antoshin, A.; Chen, X.; Yin, M.; Qu, X.; Liu, C. Coupling PEG-LZM polymer networks with polyphenols yields suturable biohydrogels for tissue patching. Biomater. Sci. 2020, 8, 3334–3347. [Google Scholar] [CrossRef]
  251. Yuan, B.; Cao, Y.X.; Tang, Q.; Yuan, Z.Q.; Zhou, Y.R.; McClements, D.J.; Cao, C.J. Enhanced performance and functionality of active edible films by incorporating tea polyphenols into thin calcium alginate hydrogels. Food Hydrocoll. 2019, 97, 8. [Google Scholar] [CrossRef]
  252. Ning, P.A.; Lu, S.Y.; Bai, X.; Wu, X.; Gao, C.M.; Wen, N.; Liu, M.Z. High encapsulation and localized delivery of curcumin from an injectable hydrogel. Mater. Sci. Eng. C Mater. Biol. Appl. 2018, 83, 121–129. [Google Scholar] [CrossRef]
  253. Abbasi, E.; Aval, S.F.; Akbarzadeh, A. Dendrimers: Synthesis, applications, and properties. Nanoscale Res. Lett. 2014, 9, 247. [Google Scholar] [CrossRef] [Green Version]
  254. Tripathy, S.; Das, M.K. Dendrimers and their applications as novel drug delivery carriers. J. Appl. Pharm. Sci. 2013, 3, 142–149. [Google Scholar]
  255. Yesil-Celiktas, O.; Pala, C.; Cetin-Uyanikgil, E.O.; Sevimli-Gur, C. Synthesis of silica-PAMAM dendrimer nanoparticles as promising carriers in Neuro blastoma cells. Anal. Biochem. 2017, 519, 1–7. [Google Scholar] [CrossRef] [PubMed]
  256. Chanphai, P.; Tajmir-Riahi, H.A. Binding analysis of antioxidant polyphenols with PAMAM nanoparticles. J. Biomol. Struct. Dyn. 2018, 36, 3487–3495. [Google Scholar] [CrossRef] [PubMed]
  257. Chanphai, P.; Tajmir-Riahi, H.A. Encapsulation of micronutrients resveratrol, genistein, and curcumin by folic acid-PAMAM nanoparticles. Mol. Cell. Biochem. 2018, 449, 157–166. [Google Scholar] [CrossRef] [PubMed]
  258. Alfei, S.; Catena, S.; Turrini, F. Biodegradable and biocompatible spherical dendrimer nanoparticles with a gallic acid shell and a double-acting strong antioxidant activity as potential device to fight diseases from “oxidative stress”. Drug Deliv. Transl. Res. 2020, 10, 259–270. [Google Scholar] [CrossRef] [PubMed]
  259. Rodriguez-Rosales, R.J.; Yao, Y. Phytoglycogen, a natural dendrimer-like glucan, improves the soluble amount and Caco-2 monolayer permeation of curcumin and enhances its efficacy to reduce HeLa cell viability. Food Hydrocoll. 2020, 100, 8. [Google Scholar] [CrossRef]
  260. Amirmahani, N.; Mahmoodi, N.O.; Mohammadi Galangash, M.; Ghavidast, A. Advances in nanomicelles for sustained drug delivery. J. Ind. Eng. Chem. 2017, 55, 21–34. [Google Scholar] [CrossRef]
  261. Movassaghian, S.; Merkel, O.M.; Torchilin, V.P. Applications of polymer micelles for imaging and drug delivery. Wires Nanomed. Nanobiotechnology 2015, 7, 691–707. [Google Scholar] [CrossRef]
  262. Biswas, S.; Kumari, P.; Lakhani, P.M.; Ghosh, B. Recent advances in polymeric micelles for anti-cancer drug delivery. Eur. J. Pharm. Sci. 2016, 83, 184–202. [Google Scholar] [CrossRef]
  263. Chen, S.Z.; Yang, K.N.; Tuguntaev, R.G.; Mozhi, A.; Zhang, J.C.; Wang, P.C.; Liang, X.J. Targeting tumor microenvironment with PEG-based amphiphilic nanoparticles to overcome chemoresistance. Nanomed. Nanotechnol. Biol. Med. 2016, 12, 269–286. [Google Scholar] [CrossRef] [Green Version]
  264. Al Fatease, A.; Shah, V.; Nguyen, D.X.; Cote, B.; LeBlanc, N.; Rao, D.A.; Alani, A.W.G. Chemosensitization and mitigation of Adriamycin-induced cardiotoxicity using combinational polymeric micelles for co-delivery of quercetin/resveratrol and resveratrol/curcumin in ovarian cancer. Nanomed. Nanotechnol. Biol. Med. 2019, 19, 39–48. [Google Scholar] [CrossRef]
  265. Cote, B.; Carlson, L.J.; Rao, D.A.; Alani, A.W.G. Combinatorial resveratrol and quercetin polymeric micelles mitigate doxorubicin induced cardiotoxicity in vitro and in vivo. J. Control. Release 2015, 213, 128–133. [Google Scholar] [CrossRef] [PubMed]
  266. Carlson, L.J.; Cote, B.; Alani, A.W.G.; Rao, D.A. Polymeric micellar co-delivery of resveratrol and curcumin to mitigate in vitro doxorubicin-induced cardiotoxicity. J. Pharm. Sci. 2014, 103, 2315–2322. [Google Scholar] [CrossRef] [PubMed]
  267. Zhao, S.; Ma, L.T.; Cao, C.W.; Yu, Q.Q.; Chen, L.M.; Liu, J. Curcumin-loaded redox response of self-assembled micelles for enhanced antitumor and anti-inflammation efficacy. Int. J. Nanomed. 2017, 12, 2489–2504. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  268. Zhang, J.M.; Li, J.J.; Shi, Z.; Yang, Y.; Xie, X.; Lee, S.M.; Wang, Y.T.; Leong, K.W.; Chen, M.W. pH-sensitive polymeric nanoparticles for co-delivery of doxorubicin and curcumin to treat cancer via enhanced pro-apoptotic and anti-angiogenic activities. Acta Biomater. 2017, 58, 349–364. [Google Scholar] [CrossRef]
  269. Kim, G.; Piao, C.; Oh, J.; Lee, M. Self-assembled polymeric micelles for combined delivery of anti-inflammatory gene and drug to the lungs by inhalation. Nanoscale 2018, 10, 8503–8514. [Google Scholar] [CrossRef]
  270. Washington, K.E.; Kularatne, R.N.; Biewer, M.C.; Stefan, M.C. Combination loading of doxorubicin and resveratrol in polymeric micelles for increased loading efficiency and efficacy. ACS Biomater. Sci. Eng. 2018, 4, 997–1004. [Google Scholar] [CrossRef]
  271. Assadpour, E.; Maghsoudlou, Y.; Jafari, S.M.; Ghorbani, M.; Aalami, M. Optimization of folic acid nano-emulsification and encapsulation by maltodextrin-whey protein double emulsions. Int. J. Biol. Macromol. 2016, 86, 197–207. [Google Scholar] [CrossRef]
  272. Kenar, J. International News on Fats Oils. In Food Emulsions: Principles, Practices, and Techniques; CRC Press: Boca Raton, FL, USA, 2005. [Google Scholar]
  273. Yu, H.L.; Huang, Q.R. Improving the oral bioavailability of curcumin using novel organogel-based nanoemulsions. J. Agric. Food Chem. 2012, 60, 5373–5379. [Google Scholar] [CrossRef]
  274. Sari, T.P.; Mann, B.; Kumar, R.; Singh, R.R.B.; Sharma, R.; Bhardwaj, M.; Athira, S. Preparation and characterization of nanoemulsion encapsulating curcumin. Food Hydrocoll. 2015, 43, 540–546. [Google Scholar] [CrossRef]
  275. Zou, L.Q.; Zheng, B.J.; Zhang, R.J.; Zhang, Z.P.; Liu, W.; Liu, C.M.; Xiao, H.; McClements, D.J. Food-grade nanoparticles for encapsulation, protection and delivery of curcumin: Comparison of lipid, protein, and phospholipid nanoparticles under simulated gastrointestinal conditions. RSC Adv. 2016, 6, 3126–3136. [Google Scholar] [CrossRef]
  276. Ru, Q.; Yu, H.; Huang, Q. Encapsulation of epigallocatechin-3-gallate (EGCG) using oil-in-water (O/W) submicrometer emulsions stabilized by ι-carrageenan and β-lactoglobulin. J. Agric. Food Chem. 2010, 58, 10373–10381. [Google Scholar] [CrossRef] [PubMed]
  277. Bhushani, J.A.; Karthik, P.; Anandharamakrishnan, C. Nanoemulsion based delivery system for improved bioaccessibility and Caco-2 cell monolayer permeability of green tea catechins. Food Hydrocoll. 2016, 56, 372–382. [Google Scholar] [CrossRef]
  278. Cordeiro, M.; Ferreira Carlos, F.; Pedrosa, P.; Lopez, A.; Baptista, P.V. Gold nanoparticles for diagnostics: Advances towards points of care. Diagnostics (Basel, Switzerland) 2016, 6, 43. [Google Scholar] [CrossRef] [PubMed]
  279. Chowdhury, A.; Kunjiappan, S.; Panneerselvam, T.; Somasundaram, B.; Bhattacharjee, C. Nanotechnology and nanocarrier-based approaches on treatment of degenerative diseases. Int. Nano Lett. 2017, 7, 91–122. [Google Scholar] [CrossRef] [Green Version]
  280. Verissimo, T.V.; Santos, N.T.; Silva, J.R.; Azevedo, R.B.; Gomes, A.J.; Lunardi, C.N. In vitro cytotoxicity and phototoxicity of surface-modified gold nanoparticles associated with neutral red as a potential drug delivery system in phototherapy. Mater. Sci. Eng. C Materials Biol. Appl. 2016, 65, 199–204. [Google Scholar] [CrossRef]
  281. Wang, C.; Zhang, H.; Zeng, D.; San, L.; Mi, X. DNA Nanotechnology mediated gold nanoparticle conjugates and their applications in biomedicine. Chin. J. Chem. 2016, 34, 299–307. [Google Scholar] [CrossRef]
  282. Fratoddi, I.; Venditti, I.; Cametti, C.; Russo, M.V. How toxic are gold nanoparticles? The state-of-the-art. Nano Res. 2015, 8, 1771–1799. [Google Scholar] [CrossRef]
  283. Singh, P.; Kim, Y.J.; Wang, C.; Mathiyalagan, R.; El-Agamy Farh, M.; Yang, D.C. Biogenic silver and gold nanoparticles synthesized using red ginseng root extract, and their applications. Artif. Cells Nanomed. Biotechnol. 2016, 44, 811–816. [Google Scholar] [CrossRef]
  284. Govindaraju, S.; Rengaraj, A.; Arivazhagan, R.; Huh, Y.-S.; Yun, K. Curcumin-conjugated gold clusters for bioimaging and anticancer applications. Bioconjugate Chem. 2018, 29, 363–370. [Google Scholar] [CrossRef]
  285. Elbialy, N.S.; Abdelfatah, E.A.; Khalil, W.A. Antitumor Activity of curcumin-green synthesized gold nanoparticles: In vitro study. BioNanoScience 2019, 9, 813–820. [Google Scholar] [CrossRef]
  286. Vemuri, S.K.; Banala, R.R.; Mukherjee, S.; Uppula, P.; GPV, S.; A.V., G.R.; T., M. Novel biosynthesized gold nanoparticles as anti-cancer agents against breast cancer: Synthesis, biological evaluation, molecular modelling studies. Mater. Sci. Eng. C 2019, 99, 417–429. [Google Scholar] [CrossRef] [PubMed]
  287. Wang, W.J.; Tang, Q.; Yu, T.R.; Li, X.; Gao, Y.; Li, J.; Liu, Y.; Rong, L.; Wang, Z.G.; Sun, H.C.; et al. Surfactant-free preparation of Au@resveratrol hollow nanoparticles with photothermal performance and antioxidant activity. ACS Appl. Mater. Interfaces 2017, 9, 3376–3387. [Google Scholar] [CrossRef] [PubMed]
  288. Balakrishnan, S.; Mukherjee, S.; Das, S.; Bhat, F.A.; Raja Singh, P.; Patra, C.R.; Arunakaran, J. Gold nanoparticles–conjugated quercetin induces apoptosis via inhibition of EGFR/PI3K/Akt–mediated pathway in breast cancer cell lines (MCF-7 and MDA-MB-231). Cell Biochem. Funct. 2017, 35, 217–231. [Google Scholar] [CrossRef]
  289. Balakrishnan, S.; Bhat, F.A.; Raja Singh, P.; Mukherjee, S.; Elumalai, P.; Das, S.; Patra, C.R.; Arunakaran, J. Gold nanoparticle–conjugated quercetin inhibits epithelial–mesenchymal transition, angiogenesis and invasiveness via EGFR/VEGFR-2-mediated pathway in breast cancer. Cell Prolif. 2016, 49, 678–697. [Google Scholar] [CrossRef] [PubMed]
  290. Mousavi, S.M.; Hashemi, S.A.; Ghasemi, Y.; Atapour, A.; Amani, A.M.; Savar Dashtaki, A.; Babapoor, A.; Arjmand, O. Green synthesis of silver nanoparticles toward bio and medical applications: Review study. Artif. Cells Nanomed. Biotechnol. 2018, 46, S855–S872. [Google Scholar] [CrossRef] [Green Version]
  291. Rafique, M.; Sadaf, I.; Rafique, M.S.; Tahir, M.B. A review on green synthesis of silver nanoparticles and their applications. Artif. Cells Nanomed. Biotechnol. 2017, 45, 1272–1291. [Google Scholar] [CrossRef]
  292. Chung, I.M.; Park, I.; Seung-Hyun, K.; Thiruvengadam, M.; Rajakumar, G. Plant-mediated synthesis of silver nanoparticles: Their characteristic properties and therapeutic applications. Nanoscale Res. Lett. 2016, 11, 14. [Google Scholar] [CrossRef] [Green Version]
  293. Shivakumar, M.; Nagashree, K.L.; Yallappa, S.; Manjappa, S.; Manjunath, K.S.; Dharmaprakash, M.S. Biosynthesis of silver nanoparticles using pre-hydrolysis liquor of Eucalyptus wood and its effective antimicrobial activity. Enzym. Microb. Technol. 2017, 97, 55–62. [Google Scholar] [CrossRef]
  294. Ahmed, M.J.; Murtaza, G.; Mehmood, A.; Bhatti, T.M. Silver nanoparticles, green synthesis: Characterization, in vitro antioxidant and antimicrobial study. Inorg. Nano Met. Chem. 2019, 49, 240–248. [Google Scholar] [CrossRef]
  295. Crisan, D.; Scharffetter-Kochanek, K.; Crisan, M.; Schatz, S.; Hainzl, A.; Olenic, L.; Filip, A.; Schneider, L.A.; Sindrilaru, A. Topical silver and gold nanoparticles complexed with Cornus mas suppress inflammation in human psoriasis plaques by inhibiting NF-B activity. Exp. Derm. 2018, 27, 1166–1169. [Google Scholar] [CrossRef] [Green Version]
  296. Hao, D.Y.; Xu, Y.Y.; Zhao, M.H.; Ma, J.X.; Wei, Y.J.; Wang, X.L. Biosynthesis of Clinacanthus nutans Lindau leaf extract mediated ag NPs, au NPs and their comparative strong muscle relaxant, analgesic activities for pain management in nursing care for using in intensive nursing care unit. J. Photochem. Photobiol. B Biol. 2020, 202, 5. [Google Scholar] [CrossRef] [PubMed]
  297. Singh, A.; Dar, M.Y.; Joshi, B.; Sharma, B.; Shrivastava, S.; Shukla, S. Phytofabrication of silver nanoparticles: Novel drug to overcome hepatocellular ailments. Toxicol. Rep. 2018, 5, 333–342. [Google Scholar] [CrossRef] [PubMed]
  298. Soto-Quintero, A.; Guarrotxena, N.; Garcia, O.; Quijada-Garrido, I. Curcumin to promote the synthesis of silver nps and their self-assembly with a thermoresponsive polymer in core-shell nanohybrids. Sci. Rep. 2019, 9, 14. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic representation of nanoformulations to enhance the bioavailability and physiological functions of polyphenols.
Figure 1. Schematic representation of nanoformulations to enhance the bioavailability and physiological functions of polyphenols.
Molecules 25 04613 g001
Figure 2. Main structure of the phosphatidylcholine, representative phosphatidylcholine groups and main fatty acid residues.
Figure 2. Main structure of the phosphatidylcholine, representative phosphatidylcholine groups and main fatty acid residues.
Molecules 25 04613 g002
Figure 3. Liposome structure and drug loading diagram: (A) Cross-section structure of liposome, made of phospholipid and cholesterol, showing the magnified molecular structure of a phospholipid that consists of a polar head and a non-polar tail. Phospholipid head is hydrophobic and comprises choline, phosphate and glycerol, while the tail is a hydrocarbon chain that shows lipophilicity. (B) A schematic representation of the structure and preparation of phytosomal curcumin. Reprinted from reference [134,159]. Copyright 2016 Elsevier Masson SAS, 2019 Elsevier Ltd.
Figure 3. Liposome structure and drug loading diagram: (A) Cross-section structure of liposome, made of phospholipid and cholesterol, showing the magnified molecular structure of a phospholipid that consists of a polar head and a non-polar tail. Phospholipid head is hydrophobic and comprises choline, phosphate and glycerol, while the tail is a hydrocarbon chain that shows lipophilicity. (B) A schematic representation of the structure and preparation of phytosomal curcumin. Reprinted from reference [134,159]. Copyright 2016 Elsevier Masson SAS, 2019 Elsevier Ltd.
Molecules 25 04613 g003
Figure 4. Schematic representation of niosome prepared by sorbitan monostearate (Span-60). Redrawn from reference [176]. Copyright 2014 Elsevier B.V.
Figure 4. Schematic representation of niosome prepared by sorbitan monostearate (Span-60). Redrawn from reference [176]. Copyright 2014 Elsevier B.V.
Molecules 25 04613 g004
Figure 5. (A) The chemical structure of the gelatin. (B) Scanning electron microscopy (SEM) images of pure chitosan scaffold and (C) chitosan–gelatin scaffold. Redrawn from reference [195]. Copyright 2009 Elsevier Ltd.
Figure 5. (A) The chemical structure of the gelatin. (B) Scanning electron microscopy (SEM) images of pure chitosan scaffold and (C) chitosan–gelatin scaffold. Redrawn from reference [195]. Copyright 2009 Elsevier Ltd.
Molecules 25 04613 g005
Figure 6. Schematic representation of the designed targeted EGCG NPs. Chemical structure of (-)- epigallocatechin-3-gallate (EGCG), the PEGylated PLGA polymers (PLGA-PEG), and the targeting ligands DCL and AG. Adapted from reference [213].
Figure 6. Schematic representation of the designed targeted EGCG NPs. Chemical structure of (-)- epigallocatechin-3-gallate (EGCG), the PEGylated PLGA polymers (PLGA-PEG), and the targeting ligands DCL and AG. Adapted from reference [213].
Molecules 25 04613 g006
Figure 7. Schematic representation of the size-controlled synthesis of γ-CD-MOFs through facile and green seed-mediated method. Adapted from reference [244]. Copyright 2018 American Chemical Society.
Figure 7. Schematic representation of the size-controlled synthesis of γ-CD-MOFs through facile and green seed-mediated method. Adapted from reference [244]. Copyright 2018 American Chemical Society.
Molecules 25 04613 g007
Figure 8. (A) Scheme of thiolated chitosan/poly(ethylene glycol) diacrylate (TCS/PEGDA) injectable hydrogel for localized intratumoral delivery of anti-cancer drugs. (B) Curcumin release behavior from TCS/PEGDA injectable hydrogel in PBS buffer at 37 °C with shaking (100 rpm). TP0 is the gel with micro porous starch but without lysozyme; TP1, TP2 and TP3 are the gels with micro porous starch adsorbed 0.46, 0.60 and 0.75 g/g lysozyme, respectively. (C) HepG2 cells viability determined using MTT assay when incubated with free curcumin, and curcumin loaded TCS/PEGDA injectable hydrogels with (TP3) or without (TP0) lysozyme, respectively. Adapted from reference [252]. Copyright 2017 Elsevier B.V.
Figure 8. (A) Scheme of thiolated chitosan/poly(ethylene glycol) diacrylate (TCS/PEGDA) injectable hydrogel for localized intratumoral delivery of anti-cancer drugs. (B) Curcumin release behavior from TCS/PEGDA injectable hydrogel in PBS buffer at 37 °C with shaking (100 rpm). TP0 is the gel with micro porous starch but without lysozyme; TP1, TP2 and TP3 are the gels with micro porous starch adsorbed 0.46, 0.60 and 0.75 g/g lysozyme, respectively. (C) HepG2 cells viability determined using MTT assay when incubated with free curcumin, and curcumin loaded TCS/PEGDA injectable hydrogels with (TP3) or without (TP0) lysozyme, respectively. Adapted from reference [252]. Copyright 2017 Elsevier B.V.
Molecules 25 04613 g008
Figure 9. Synthesis of cholesterol-conjugated PAMAM (PAM-Chol) (A) and preparation of the pHO-1/PAM-Chol/Res complex (B). The heme oxygenase-1 gene (pHO-1). Adapted from reference [269]. Copyright 2018 Royal Society of Chemistry.
Figure 9. Synthesis of cholesterol-conjugated PAMAM (PAM-Chol) (A) and preparation of the pHO-1/PAM-Chol/Res complex (B). The heme oxygenase-1 gene (pHO-1). Adapted from reference [269]. Copyright 2018 Royal Society of Chemistry.
Molecules 25 04613 g009
Figure 10. Morphology characterization of CUR-AuNCs: (A) TEM of CUR-AuNCs; the inset image is SAED pattern. (B) HR-TEM; arrow indicates the collections of atoms to form a cluster. (C) Bio-AFM height image of CUR-AuNCs. Adapted from reference [284]. Copyright 2018 American Chemical Society.
Figure 10. Morphology characterization of CUR-AuNCs: (A) TEM of CUR-AuNCs; the inset image is SAED pattern. (B) HR-TEM; arrow indicates the collections of atoms to form a cluster. (C) Bio-AFM height image of CUR-AuNCs. Adapted from reference [284]. Copyright 2018 American Chemical Society.
Molecules 25 04613 g010
Table 1. Antioxidant activities of some extracts/compounds from plants.
Table 1. Antioxidant activities of some extracts/compounds from plants.
PolyphenolsAntioxidant ActivityDetection Method *References
Extracts of Hippophae speciesRegulate enzyme activity, affect the antioxidant reaction of cellsDPPH assay[27,28]
Extracts of sweet potato leavesDecrease the level of intracellular ROSPhotochemiluminescence assay, ORAC assay[29,30]
Polyphenols from stevia rebaudiana Radical scavenging, regulate enzyme activityDPPH assay, ABTS assay[31,32]
CurcuminDirect reaction with free radicals, regulation of antioxidant-related enzyme activity and gene expressionDPPH assay, ABTS assay, total phenolic content assays[33,34,35]
Extracts of Nymphaea nouchali leafReducing DNA damage and attenuating oxidative stress-induced cell deathFRAP assay, ORAC assay, DPPH assay[36]
Persimmon vinegar polyphenolsActivate of the Nrf2 antioxidative pathwayFluorescent probe method, DPPH assay, total phenolic content assays[37,38,39]
AnthocyaninsRadical scavenging, reduce the catalytic effect of metal ionsDPPH assay, T-AOC assay, ABTS assay, FRAP assay[40,41,42]
Grape seed extractDecrease the oxidized LDL in plasma, regulate enzyme activityAntioxidant enzyme activity, DPPH assay, ORAC assay, ABTS assay[43,44,45]
(−)-Epicatechin and procyanidinPreservatives for fruit, radical scavengingDPPH assay, hydroxyl radical scavenging capacity method, superoxide anion radical method[46,47]
Extracts of blueberriesRegulate enzyme activity, chelate trace metals, regulate miRNAFRAP assay, DPPH assay, ABTS assay[48,49]
Extracts of pineRadical scavenging, the skin against oxygen reactive speciesDPPH assay, superoxide anion radical method, hydroxyl radical scavenging capacity method[50,51]
Extracts of teaIncrease antioxidant enzyme activity, inhibit lipid peroxidation, radical scavengingDPPH assay, FRAP assay, TEAC assay[52,53]
* DPPH, 2,2-diphenyl-1-picrylhydrazyl; ORAC, oxygen radical absorbance capacity; ABTS, 2,2’-azino-bis(3-ethylbenzothiazoline-6-sulfonic acid; FRAP, fluorescence recovery after photobleaching; T-AOC, total antioxidant capacity; TEAC, trolox equivalent antioxidant capacity.
Table 3. Polyphenols complexed with phospholipids.
Table 3. Polyphenols complexed with phospholipids.
Phytosomal
Formulations
Biological ActivityRoute of AdministrationReference
Moringa oleifera leaf phytophospholipid complexWound healingIn vitro[138]
Quercetin phytosomeAntimicrobial, anti-infammatory, anticancerOral[139]
Curcumin phytosomeAntioxidantIn vitro, oral [141,142]
Rutin-phospholipidAnticytotoxicity, neuroprotectionIn vitro[143]
Catechin phyto-phospholipidAntioxidantIn vitro[144]
Luteolin phytosome HepatoprotectiveOral[145]
Silybin phospholipidHepatoprotective, antioxidant, anticancerIn vivo[140,146]
EGCG phytosome AnticancerOral[147]
Grape seed phytosome AnticancerOral[148]
Quercetin phytosome Antioxidant, anti-inflammatoryOral[149]
Silybin phytosome HepatoprotectiveOral[150]
Persimmon phytosomeAntioxidantOral[151]

Share and Cite

MDPI and ACS Style

Yang, B.; Dong, Y.; Wang, F.; Zhang, Y. Nanoformulations to Enhance the Bioavailability and Physiological Functions of Polyphenols. Molecules 2020, 25, 4613. https://doi.org/10.3390/molecules25204613

AMA Style

Yang B, Dong Y, Wang F, Zhang Y. Nanoformulations to Enhance the Bioavailability and Physiological Functions of Polyphenols. Molecules. 2020; 25(20):4613. https://doi.org/10.3390/molecules25204613

Chicago/Turabian Style

Yang, Bingyan, Yixin Dong, Fei Wang, and Yu Zhang. 2020. "Nanoformulations to Enhance the Bioavailability and Physiological Functions of Polyphenols" Molecules 25, no. 20: 4613. https://doi.org/10.3390/molecules25204613

Article Metrics

Back to TopTop