Next Article in Journal
Chitosan Oleate Coated PLGA Nanoparticles as siRNA Drug Delivery System
Previous Article in Journal
Manufacturing and Examination of Vaginal Drug Delivery System by FDM 3D Printing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Improving Curcumin Bioavailability: Current Strategies and Future Perspectives

Department of Pharmacy, University of Pisa, Via Bonanno 6, I-56126 Pisa, Italy
*
Author to whom correspondence should be addressed.
Submission received: 23 September 2021 / Revised: 11 October 2021 / Accepted: 14 October 2021 / Published: 17 October 2021
(This article belongs to the Section Pharmacokinetics and Pharmacodynamics)

Abstract

:
Curcumin possesses a plethora of interesting pharmacological effects. Unfortunately, it is also characterized by problematic drug delivery and scarce bioavailability, representing the main problem related to the use of this compound. Poor absorption, fast metabolism, and rapid systemic clearance are the most important factors contributing to low curcumin levels in plasma and tissues. Accordingly, to overcome these issues, numerous strategies have been proposed and are investigated in this article. Due to advances in the drug delivery field, we describe here the most promising strategies for increasing curcumin bioavailability, including the use of adjuvant, complexed/encapsulated curcumin, specific curcumin formulations, and curcumin nanoparticles. We analyze current strategies, already available in the market, and the most advanced technologies that can offer a future perspective for effective curcumin formulations. We focus the attention on the effectiveness of curcumin-based formulations in clinical trials, providing a comprehensive summary. Clinical trial results, employing various delivery methods for curcumin, showed that improved bioavailability corresponds to increased therapeutic efficacy. Furthermore, advances in the field of nanoparticles hold great promise for developing curcumin-based complexes as effective therapeutic agents. Summarizing, suitable delivery methods for this polyphenol will ensure the possibility of using curcumin-derived formulations in clinical practice as preventive and disease-modifying therapeutics.

1. Introduction

Finding appropriate delivery systems is a challenging task to utilize natural products as effective treatments for a given disease. Considering that many natural products have interesting pharmacological effects and that they can represent a valuable source for developing innovative drugs [1,2], currently, natural compounds of interest are under investigation for improving their bioavailability by employing different technologies [3,4,5]. Natural products are not chemically stable and are susceptible to oxidative degradation, affecting the integrity of molecules and leading to the development of free radicals. Additionally, it is difficult to employ natural products in formulations due to unfavorable features, including low solubility, inadequate bioavailability, and weak stability in front of environmental stresses (e.g., exposure to oxygen and light, high temperature, and pH) [6]. Consequently, to maintain the peculiar profile of these compounds, specific methods of delivery offer the possibility to use natural compounds for treating human disorders. In this review article, we examine the most advanced delivery approaches for enhancing curcumin bioavailability, considering the technologies used and clinical trials that prove the effectiveness of the methods. Curcumin holds great promise in clinical practice since it behaves as a multifunctional ligand possessing a plethora of pharmacological effects, targeting several cellular pathways [7,8]. According to the multifactorial and heterogeneous nature of different disorders, including inflammatory, metabolic, neoplastic, neurodegenerative, and central nervous system (CNS)-related disorders, compounds with no side effects, that are non-toxic to patients, and have multiple properties, such as curcumin, are excellent candidates for treating these pathologies [9,10]. This strategy offers a better chance of effective prophylaxis or treatment and supports the ongoing research and development of curcumin as a preventive and disease-modifying agent [11]. These aspects have been reviewed considering clinical trials performed by enrolling healthy people to assess the enhancement of curcumin bioavailability by exploring different formulations. Furthermore, we discuss clinical outcomes using diverse curcumin formulations for treating several disorders (e.g., nonalcoholic fatty liver disease (NAFLD), knee osteoarthritis (OA), moderate hyperlipidemia metabolic syndrome risk, glioblastoma, obesity, impaired glucose tolerance or non-insulin-dependent diabetes mellitus, delayed onset muscle soreness (DOMS) and related muscle damage, osteo-muscular pain, chronic diabetic macular edema, advanced or metastatic pancreatic cancer).

2. From the Kitchen to the Bench

Curcumin (IUPAC: (1E,6E)-1,7-bis(4-hydroxy-3-methoxyphenyl)-1,6-heptadiene-3,5-dione), a polyphenol extracted from turmeric Curcuma longa L., was used from ancient times in traditional and ayurvedic medicine, especially in India and China [12]. Furthermore, for over 2000 years, the rhizome of turmeric has been employed in Asian cuisine, cosmetics, and fabric dyeing [13]. The main component of turmeric, curcumin, has been demonstrated to possess a plethora of interesting pharmacological effects, including anti-inflammatory, antioxidant, neuroprotective, chemopreventive, and chemotherapeutic activity [14,15]. Anti-inflammatory effects of curcumin were observed in the acute carrageenan-induced edema test in mice and rats after oral administration. The oral doses required to produce an anti-inflammatory effect, however, were much higher than the doses that were necessary for intraperitoneal (i.p.) administration, giving a similar effect. Thus, the oral ED50 was 100.2 mg/kg in mice and 48.0 mg/kg in rats [14]. Several studies have been conducted on anti-inflammatory effects of curcumin, concluding that in mice, this polyphenol inhibited edema at doses between 50 and 200 mg/kg. A 50% reduction in edema was achieved with a dose of 48 mg/kg body weight, with curcumin nearly as effective as cortisone and phenylbutazone at similar doses. In rats, a lower dose of 20–80 mg/kg decreased paw edema and inflammation. Curcumin also inhibited formaldehyde-induced arthritis in rats at a dose of 40 mg/kg [16,17]. Regarding its antioxidant profile, a recent meta-analysis highlighted that the effective dose of curcumin to obtain such an effect is 645 mg/die [18]. The anticancer activity of curcumin has been recently reviewed by Tomeh and colleagues, who reported a comprehensive overview of clinical applications of curcumin for treating different tumors (e.g., benign prostatic hypertrophy, 1 g/day for 24 weeks; breast cancer, 0.5–8 g/day for 7 days plus docetaxel in a phase I clinical trial; chronic myeloid leukemia, 5 g three times daily for 6 weeks plus imatinib (400 mg twice daily); pancreatic cancer, in a phase II clinical trial, 8 g/day for 8 weeks; prostate cancer, in a randomized controlled trial, 3 g/day for 3 months as a supplement to radiotherapy) [19]. Finally, curcumin showed neuroprotective effects, considering CNS-related disorders (e.g., depression) and ageing-related diseases (Alzheimer’s and Parkinson’s diseases), at doses ranging from 40 mg/kg i.p. for depression (rat models) to several grams (500 mg twice daily—4 g daily) per day in clinical trials for Alzheimer’s disease [20,21,22].
The beneficial properties of curcumin stem from the capacity to modulate many signaling pathways, such as survival pathways mediated by NF-kB, Akt, and growth factors, as well as cytoprotective pathways involving Nrf2, and metastatic and angiogenic pathways [23,24,25]. NF-kB modulates the expression of several chemokines and cytokines, including interleukins, interferons, lymphokines, and tumor necrosis factors (TNFs) [26]. Specifically, curcumin inhibits cytokine production by suppressing the NF-kB phosphorylation. By this specific targeting, curcumin regulates inflammatory processes and related diseases such as OA. It has been demonstrated that curcumin exerts protective effects on IL-1β-, TNFα-, or TNFβ-stimulated chondrocytes, indicating its chondroprotective properties for treating OA and related osteoarticular problems [27]. Accordingly, curcumin suppressed NF-kB pathway activation by repressing IkBα phosphorylation and degradation, as well as P65 nuclear translocation, attenuating inflammation [28,29,30]. Additionally, curcumin is a free radical scavenger with pro- and antioxidant activity [31]. Furthermore, curcumin binds metals, especially iron and copper, acting as an iron chelator [32]. Interestingly, in general, curcumin is well tolerated and safe, as several studies have confirmed low toxicity in humans (oral doses up to 12 g/die administered to adult subjects) [33,34,35,36]. On the other hand, other reports highlighted possible adverse effects, including DNA damage [37,38], time- and dose-dependent increase in reactive oxygen species (ROS) [39], and inhibition of p53 in colorectal cancer [40]. Moreover, also in clinical studies, some undesirable effects have been reported after employing higher dosage for 4 months (e.g., nausea, diarrhea, headache, rash, and yellow stool) [35,41]. Despite the encouraging pharmacological profile, curcumin suffers from several drawbacks, limiting its therapeutic potential. The problematic drug delivery and poor bioavailability are the main problems related to the use of this compound [42,43]. In fact, curcumin (logP 3.29 [44]) is almost insoluble in pure water (0.6 μg/mL), while an improvement of solubility was observed in organic solvents such as ethanol (1 mg/mL) [45,46]. Poor absorption, fast metabolism, and rapid systemic clearance are the most relevant factors that contribute to the low curcumin levels in plasma and tissues [43]. To overcome these issues, numerous strategies have been recently described, such as the use of adjuvants like piperine (interferes with curcumin metabolism), complexed/encapsulated curcumin, specific curcumin formulations, and curcumin nanoparticles (Figure 1) [47,48,49,50]. Accordingly, in the following sections, we report the main promising strategies for improving curcumin bioavailability that are supported by clinical trials, or in vivo studies, to provide an up-to-date overview of the technologies used and related evidence of success in human trials.

2.1. Ready-to-Go Strategies to Enhance Bioavailability

As a result of the broad spectrum of potential health benefits shown by curcumin, attempts to improve its bioavailability have been described over the years. Despite the high number of formulations tested, only some of them reached the market due to their unsuitability for commercial applications. In fact, many delivery formulations have been discontinued due to being inadequate for commercial purposes since they were too expensive, too difficult to produce, unstable under environmental conditions, or used ingredients not suitable for food applications [54,55]. This section examines commercial formulations, that overcame these concerns, for which human pharmacokinetic data are available. Formulations are presented according to the bioenhancement strategy, addressing the issues regarding the limited curcumin bioavailability. It is important to notice that the great variability within the clinical studies significantly affects the accurate comparison of the results [56]. Moreover, most of the available pharmacokinetic studies, analyzed the total curcumin content instead of free curcumin [57,58,59,60,61,62] since it was reported that blood samples were hydrolyzed with glucuronidase or sulfatase before the analysis. Due to the fact that curcumin undergoes extensive metabolism once orally administered, sulfate and glucuronide conjugates represent the predominant, although physiologically inactive, compounds. Indeed, many authors agreed that plasma-free curcumin represents the bioactive form of curcumin and is currently the best indicator of bioavailability and bioequivalence [63,64,65]. Therefore, enzymatic hydrolysis can lead to a misreading of the results with an over-estimation of the free, bioactive curcumin at least 10-fold higher compared to non-hydrolyzed plasma samples [66]. Thus, in this review, data resulting from sample hydrolysis will be defined as “total curcumin” compared to “free curcumin” from not-hydrolyzed samples.

2.1.1. The Early Factors Impacting Metabolism

In Table 1, the pharmacokinetic profile of curcumin considering various curcumin-based preparations is reported. One of the first adopted strategies was to combine curcumin with the alkaloid piperine (inhibitor of UDP-glucuronosyltransferase and CYP3A4, and P-glycoprotein blocker). This approach showed a clear inhibition of curcumin metabolism [67,68]. The administration of 2 g of curcumin with 20 mg of piperine to ten healthy adult males, in a randomized crossover trial, resulted in a 20-fold increase in AUC compared with the administration of 2 g of pure curcumin alone [69]. Similarly, in a further work, the effect of co-administration of 2 g of curcumin and 5 mg of piperine (Bioperine®) [70] in six healthy subjects in a crossover-designed study was assessed. Three volunteers received only curcumin, while the remaining three received both the drug and the adjuvant. Results showed that the absorption of curcumin co-administered with piperine was approximately doubled [42]. However, more recently, the administration of 12 g of Curcumin C3 Complex® [71] with 5 mg of Bioperine® in an open-label, uncontrolled phase I pilot study, did not detect any trace of curcumin in blood samples of ten volunteers at any endpoint [72]. Recently, the efficacy of the formulation of curcumin + piperine has also been investigated in clinical practice. Panhai and colleagues evaluated the effect of co-administration of 500 mg of curcumin + 5 mg of piperine (Curcumin C3 Complex® and Bioperine®, Sami Labs Ldt) in adult patients suffering from NAFLD. Seventy adults were arbitrarily assigned to the placebo or treatment group and consumed one capsule after meals for 12 weeks. Results showed that the administration of Curcumin C3 Complex® and Bioperine® improved liver and lipid profile without altering hematological parameters, thus reducing NAFLD severity compared to placebo [73]. The same research group also considered the effect of the formulation on serum inflammatory factors, namely, TNFα and IL-6, in a randomized double-blind placebo-controlled trial with 55 subjects with NAFLD. Dietary supplementation for 8 weeks resulted in a decrease in cytokine serum levels, suggesting their reduction can mediate the anti-steatotic effect of curcuminoids [74].
Another strategy that addresses the modulation of curcumin metabolism consists of the mixture of turmeric powder and turmeric essential oil. The latter, being particularly rich in turmerones, which act as p-glycoprotein inhibitors, increases intestinal permeability of curcumin [75]. BCM-95®CG is a patented [76], cost-effective formulation that relies on the synergism between sesquiterpenoids found in turmeric essential oil and curcuminoids. The mechanism hypothesized to underlie its efficacy considers a possible extra-intestinal absorption due to non-curcuminoid components of turmeric and a variation of the activity of phase II enzymes and P-glycoprotein due to its longer presence in blood [77,78]. Evaluation of BCM-95®CG (Biocurcumax™) oral bioavailability was obtained from a pilot crossover study conducted on 11 healthy human subjects. Volunteers, divided into three groups, were administered 2000 mg of BCM-95®CG and an equal dose of curcumin control or a formulation of curcumin–lecithin–piperine, respectively. Drugs were administered and then crossed over after a two-week washout period. Free curcumin AUC appeared to be 6.93 times higher in BCM-95®CG than that of curcumin control; furthermore, results suggested both early absorption and retained release of curcumin from BCM-95®CG compared to unformulated and curcumin-lecithin-piperine preparation [77]. Two clinical trials assessed the effectiveness of a curcumin—essential oil complex in the management of knee OA compared to common drugs, such as diclofenac or paracetamol. Together, these studies revealed that co-administration of 500 mg of BCM-95®CG plus 50 mg of diclofenac twice a day (group 1, 71 patients) compared to administration of 50 mg of diclofenac twice a day alone (group 2, 69 patients), as well as the administration of 500 mg of BCM-95®CG twice a day (73 patients) or 650 mg of paracetamol three times a day (71 patients), not only significantly improved knee OA symptoms, but reduced the gastrointestinal (GI) side effects related to NSAIDs treatment and the need of other analgesics, thus suggesting a good alternative treatment option [79,80].

2.1.2. Solubility Enhancement as a Strategy to Improve Curcumin Bioavailability

Due to the fact that solubility enhancement is one of the major areas of interest in regards to curcumin dissolution, there are two main strategies to consider, the particle size reduction for increasing the surface area and the use of substances to improve curcumin dispersibility. Among them, surfactants, hydrophobic carriers, and the formation of inclusion complexes encompass the most used and successful approaches [56,86].
Novasol® (patented formulation by AQUANOVA AG) [87] is based on micellization: when Tween-80, a common surfactant, exceeds the crucial micellar concentration in aqueous solutions, micelles form and incorporate hydrophobic drugs, such as curcumin [56,86]. Micellar curcumin containing 7% curcumin powder and 93% Tween-80 has been tested on human healthy volunteers to assess its bioavailability by the same research group in at least three small clinical trials. In a single-blind crossover study, curcumin (410 mg) was administered to 23 healthy human volunteers as native powder or micellar curcumin after a 1-week washout period. Data based on total curcumin AUC showed a great increase in formulated curcumin bioavailability that appeared 185-fold higher compared to the unformulated powder [62]. Furthermore, they tested the efficacy of a lower dose of the aforesaid formulation: 80 mg of curcumin was given to 23 healthy subjects in a single-blind crossover trial. Administration of formulated and unformulated curcumin was separated by a 1-week washout. The total curcumin AUC was increased 88-fold compared to native curcumin, suggesting that curcumin bioavailability is affected by the administered dose [81]. Finally, Kocher and colleagues have deepened the efficacy of micellar curcumin in a randomized, double-blind crossover study involving much more moderately overweight and hyperlipidemic patients. Curcuminoids or placebo were administered with each principal meal for 6 weeks (curcumin dose = 241.2 mg/day) and then crossed over to the other regimen after 4 weeks of the washout phase. Although the total plasma curcumin concentration was higher compared to placebo and baseline samples after three and six weeks of treatment, the intervention did not show any alteration in blood lipid levels, inflammatory status, glucose, or iron homeostasis. The lack of effects could be ascribable to an insufficient tissue concentration for biological activity, longer intervention required, or more severely afflicted patients [50]. Novasol® efficacy was studied also on glioblastoma patients scheduled for surgery. Ten out of thirteen completed a study that consisted of drinking juice containing 1 g of micellar curcumin (corresponding to 57.4 mg curcumin) three times a day, for 4 days prior to surgery. Tissue analysis revealed an approximated intratumoral curcumin concentration of 0.15 μmol/L, being the highest ever found in a clinical study, although far from cytotoxic concentrations based on glioma cell culture models [88].
As mentioned, solubility issues might be overcome by curcumin inclusion in cyclodextrin complexes or by taking advantage of hydrophilic carriers such as polyvinyl pyrrolinone (PVP). Cyclodextrins consist of cyclic oligosaccharides formed by non-reducing chiral glucose-building blocks connected with a ring structure. Thanks to the hydrophilic portions of glucose units facing outwards, lipophilic cavity forms inside, allowing fat-soluble drugs to lodge. However, PVP consists of a water-soluble vector that increases curcumin water dispersibility [56,57,58]. Despite the abundance of in vitro and in vivo investigations, there have not been many clinical trials. One of these clinical studies was designed for testing the bioavailability of two innovative curcumin formulations, namely Cavacurmin® [89], a γ-cyclodextrin-based compound, and Curcuwin® [90], a combination of 63–75% PVP, 10–40% cellulosic derivatives, and 1–3% natural antioxidants. These formulations were both examined in two different randomized, double-blind trials and compared with unformulated standard curcumin and two other well-known curcumin associations. A single oral dose of formulated curcuminoids (376 mg) or standard curcumin (unformulated, 1800 mg) was administered to healthy human volunteers with 7 days of washout between doses. In both studies, twelve subjects out of fifteen completed the study. The total curcumin AUC resulted in 85-fold ad 136-fold greater in Cavacurmin® and Curcuwin®, respectively, compared to unformulated curcumin [57,58].
CurQfen® is a water-soluble curcumin formulation made of turmeric powder and fenugreek dietary fiber. The latter are rich in galactomannan units and form a non-digestible gel hydrocolloid that undergoes fermentation in the colon, thus protecting curcumin from degradation. Moreover, the association of curcumin and soluble fiber seems to prolong the drug release [56,66]. A randomized double-blind crossover study was completed on 50 subjects who were administered 1000 mg of CurQfen® (307 mg curcumin) or 411 mg of unformulated curcumin with similar content of curcuminoids. The two administrations were separated by a 10-day washout phase. The same group of subjects was tested also on a low-dose formulation—they were administered 250 mg of fenugreek formulation (76.7 mg of curcumin) or standard unformulated curcumin. Moreover, ten out of the fifty volunteers were selected to participate in an additional study. After the administration of a single oral dose of 1000 mg of CurQfen®, blood samples were withdrawn at specific time points and divided into two batches. Interestingly, for each sample, one half was treated with β-glucuronidase enzyme, while the other half was not. Beyond the dose dependency in the absorption of curcuminoids, CurQfen® significantly increased free curcumin oral bioavailability compared to the standard formulation. From the analysis of free and conjugated metabolites ratio emerged an AUC average enhancement of 17.21%, confirming the hydrolysis-related overestimation, and the ability of fenugreek formulation to mainly provide free unconjugated curcuminoids in plasma, probably due to curcumin protection against enzymatic activity [66]. CurQfen® effects were examined on 22 young obese men to test whether it could improve aortic stiffness and cardiovascular disease-associated blood biomarkers. Subjects were randomly administered 500 mg CurQfen® capsules (about 158 mg of curcumin) or placebo (fenugreek fibers) for 12 weeks. These two randomized double-blind studies confirmed a significant reduction in aortic stiffness related to inflammation markers decrease, while HDL and homocysteine improved significantly [91,92].
The particle size reduction increases the surface area, favoring rapid dissolution of curcumin [86]. Several patented curcumin formulations profit by this strategy to improve bioavailability. Among these, Biocurc® is a patented lipidic formulation [93] composed of curcumin (6.17% w/w), vitamin E (3.29% w/w), Labrasol (5.76% w/w), ethanol (8.23% w/w), Gelucire®44/14 (16.46% w/w), anhydrous citric acid (2.88% w/w), PEG 400 (55.55% w/w), and hydroxyl propyl cellulose (1.64% w/w) [82]. A preliminary open-label, randomized, parallel-design study evaluated the oral bioavailability of a single dose of formulated curcumin (750 mg) on 20 healthy male subjects. Despite the fact that plasma curcumin levels are relatively low, pharmacokinetic modeling revealed that curcumin from the formulation rapidly moves from the central circulation to peripheral tissues, providing an explanation of its therapeutic efficacy [82]. A randomized, double-blind, crossover study was executed considering 12 healthy adult subjects to compare the bioavailability of a single oral dose of Biocurc® (76 mg of total curcuminoids, of which 64.6 mg was curcumin) with unformulated curcumin (380 mg of total curcuminoids, of which 323 mg was curcumin, 95%). The samples were tested for curcumin glucuronide, curcumin sulfate, and free curcumin. While free curcumin from unformulated powder stayed near the baseline during the entire test period, free curcumin from Biocurc® showed at least two peaks, while total curcumin AUC appeared to be 94 times higher for the product than for the unformulated curcumin [63]. Particle size reduction is also a strategy that underlies Theracurmin® technology [94]. Theracurmin® consists of a nanoparticle colloidal dispersion, made of 10 w/w % curcumin, 2% other curcuminoids, 4% gum ghatti, 38% water, and 46% glycerin. Fourteen healthy volunteers were enrolled in a randomized clinical trial to test Theracurmin® oral bioavailability; 30 mg of the formulation or unformulated powder were given as a single oral dose to the subjects. Total curcumin AUC showed a 27-fold increase compared to unformulated curcumin [61]. The same researchers also evaluated the oral bioavailability of commercial curcumin beverages and compared it with a beverage containing Theracurmin®. A double-blind, single-dose, four-way crossover study involving 24 healthy volunteers was conducted. They were separated into four groups, each receiving four curcumin beverages, one every 7 days. Beverages were slightly different in composition, containing 30 or 40 mg of curcumin per 100 mL. For each beverage, pharmacokinetic results indicated that total curcumin was already detectable 30 min after ingestion, but peak plasma concentrations were undetectable for any preparation. This occurrence indicates a substantial involvement of other food ingredients in the beverage, and underling the role of food intake in curcumin oral bioavailability. Total curcumin AUC values from Theracurmin® preparation became about 1.5- to 4-fold higher than those of other formulations [59]. Theracurmin® efficacy was investigated in a double-blind placebo-controlled parallel-group randomized trial in patients with impaired glucose tolerance or non-insulin-dependent diabetes mellitus. Thirty-three patients were separated into two groups, one (n = 18) receiving placebo, the other (n = 15) receiving Theracurmin® (180 mg/day per for 6 months). After six months of screening, subjects in the placebo group displayed a considerable increase in oxidized LDL level, while both triglycerides (TG) and γ-GTP diminished in the Theracurmin® group [95]. In a 6-month open-label perspective study, 45 patients affected by knee OA were administered Theracurmin® (180 mg/day for 6 months) to test its clinical efficacy and safety. Thirteen out of 45 patients were treated with only Theracurmin®, while other patients were allowed other combined therapies, such as NSAIDs, pain relief patches, and hyaluronic acid knee injection treatment. Thirty-four patients were effective cases (75.6%), in which the treatment was effective, and the scores improved in at least one assessment within the 6 months. In contrast, 11 patients were not effective. The Theracurmin®-only group, included 10 effective (76.9%) and 3 not-effective cases [96].

2.1.3. How to Improve Absorption Affecting Intestinal Uptake

Cureit® is a new formulation from Aurea Biolabs [97], established on the recreation of the natural turmeric matrix employing polar—nonpolar sandwich (PNS) technology. The formulation was obtained by combining the hydrophobic and hydrophilic compounds—the three major curcuminoids, once extracted, are combined with turmeric water-soluble constituents, such as dietary fiber, carbohydrates, proteins, and lipophilic turmeric essential oil [84,98]. The synergism that occurs within turmeric matrix constituents enhances curcumin bioavailability thanks to physical stability improvement, protection from degradation, controlled release of curcuminoids, and, consequently, higher absorbability [84,99]. Human bioavailability was assessed by a pilot crossover study involving 12 healthy males and the formulation resulted in a 5.5-fold AUC increase compared to unformulated powder [83]. A single 500 mg oral dose of Cureit® was also compared with the other two commercially available formulations, namely, a volatile oil and a phospholipid formula, in an open-label parallel-arm study. Forty-five healthy males were casually distributed to one of three groups and given 500 mg of the assigned formulation under fasting conditions. The results indicated that free curcumin AUC from Cureit® was approximately 7.3-fold and 5.6-fold more effective, respectively, when compared with the volatile oil formula and phospholipid formula. Cureit® supplementation was also shown to attenuate markers of muscle injury, decrease DOMS-associated pain symptoms, and lower CK blood levels. A randomized, placebo-controlled, double-blind trial was performed on 15 subjects, randomly assigned to the Cureit® or placebo group. They underwent 4 days of intervention, during which they ingested a single oral dose of 500 mg of Cureit® or placebo after a 45 min eccentric muscle injury test running downhill on a treadmill. Evidence showed a significant improvement after Cureit® treatment against DOMS and a positive impact on muscle recovery [100].
Liposomes are part of a phospholipid-based vesicular system, among the most extensively investigated lipid drug carriers. Their success is due to their amphiphilic nature, high biocompatibility, and high affinity for biological membranes, since they can be rapidly taken up through pinocytosis or facilitated diffusion across lipophilic cell membranes [56,60]. Meriva® [101] is a curcumin–phosphatidylcholine complex derived from soy lecithin, microcrystalline cellulose and 18–20% curcuminoids [56]. When nine volunteers, supplemented with a low or high dosage (209 mg and 376 mg of total curcuminoids, respectively) of Meriva®, as well as unformulated curcumin (1799 mg total curcuminoids), in a single oral administration, were randomly assigned one treatment for each 7-day washout period, free curcumin was not observed in any plasma samples but total curcumin appeared to be 18-fold higher in Meriva® compared to the corresponding unformulated powder [60]. The administration of Meriva® (500 mg/die, 4-month treatment) in association with glucosamine (500 mg) was found to alleviate the symptoms and decrease the need for further medications in patients affected by knee OA (n = 63) compared to the administration of chondroitin sulphate + glucosamine (400 mg + 415 mg; n = 61) [102]. Similarly, the administration of Algocur® (containing 1 g of Meriva®) twice a day for 5 or 10 days significantly improved impaired physical function of male rugby players affected by osteo-muscular pain caused by traumatic injuries (n = 25). The effect was superimposable to conventional anti-inflammatory therapy (n = 25), but in the Meriva® group, the adherence to therapy was greater due to the absence of adverse events [103]. The daily assumption of 500 mg of Meriva® was also tested on visual acuity and optical coherence tomography (OCT) retinal thickness in patients with chronic diabetic macular edema. During a 3-month open-label study of 12 eyes, 92% showed a reduction of macula edema, 8% showed stabilization, and 0% showed a clinically meaningful increase [104]. Finally, a prospective phase II, single-center, single arm trial was conducted to investigate the efficacy and safety of curcumin as a complementary therapy to gemcitabine in advanced pancreatic cancer patients (n = 44). Meriva® (2000 mg/day) was administered orally while gemcitabine was infused on days 1, 8, and 15 within a 28-day treatment cycle. Overall cycles were nine and Meriva® was continued after treatment completion (up to 42 months). Results suggested preliminary evidence of the effectiveness of the co-treatment with Meriva® and gemcitabine that could be compared to the combination of paclitaxel and gemcitabine with the advantage of producing less toxicity [105] (Table 2).
Unlike liposomes, solid lipid (nano)-particles (SLPs) are only suitable for hydrophobic drugs. They share with liposomes high biocompatibility, enhanced bioavailability, and high drug loading, but SLPs appear to be more stable and easier to manufacture [86,106]. Longvida® is a patented formulation [107] manufactured by Verdure Sciences, USA, based on solid lipid curcumin particles (SLCPs), in which curcumin content is between 20% and 30% [56]. Pharmacokinetic studies were conducted on both healthy volunteers and osteosarcoma patients. Firstly, 11 subjects with diagnosed metastatic high-grade osteogenic sarcoma were divided into three groups and orally administered 2000 mg, 3000 mg, and 4000 mg of Longvida® (curcumin content ranging from 400 mg to 1200 mg). Free curcumin concentration was improved and AUC values increased in a dose-related manner. Similarly, six healthy subjects were randomly clustered to receive a single oral dose of 650 mg of SLCPs or the same dose of unformulated curcumin. Free curcumin did not show any appreciable plasma level from the non-formulated curcuminoids extract, while SLCP AUC showed a clear increase in free curcumin concentration. The characteristic formulation of turmeric powder and phospholipid, the curcumin/lipid/antioxidant ratio, and globule size are indicated to underlie specific pharmacokinetic behavior and extended absorption of the formulation [85]. In Table 2 the discussed clinical studies are summarized.

3. Nanoformulations: What Is Known and What Is New

The lack of data regarding the protracted administration of high-dose curcumin formulations makes it difficult to foresee long-term effects on the human body. Conversely, epidemiological studies have revealed how the assumption of daily, low-dose, long-term curcumin preparations could produce significant health benefits on curry eaters [9,108]. Consequently, although dietary curcumin belongs to substances generally regarded as safe (GRAS), high doses could not be completely exempted by the onset of side effects [108]. This occurrence could be ascribable to the hormesis phenomenon, a dose–response relationship where low doses of xenobiotics induce stimulation, whereas high doses of the same substance cause inhibition or even toxicity. Among the strategies to enhance permeation and solubilization while providing low doses of drugs, nanoparticles represent one of the most attractive possibilities, particularly for all substances belonging to class IV in the biopharmaceutics classification system (BCS), such as curcumin. Nanoformulations provide substantial benefits due to their fine features that make them extremely versatile. Indeed, because of their tiny size (1–100 nm), they can easily penetrate through the intestinal mucus layer via transcytosis, increasing nutraceutical bioavailability [109]. Moreover, the small dimension prevents nanoparticles from aggregating and settling since they are affected by the brownian motion that overcomes the force of gravity and keeps the particles moving randomly. Furthermore, the reduction of particle size makes their surface area become larger so that any chemical reactions occurring at their surface will be faster. Eventually, the chemical, physical, optical, electronic, and magnetic properties of materials could differ enormously when referred to the nanoscale, giving nanoparticles unique properties [109]. All the features listed above match together the major advantages and disadvantages of nanoparticles since their behavior is often different, or even unpredictable, from the same material in bulk.
As a result of the growing interest in environmental issues from both consumers and producers, another point in favor of nanomaterials is their green synthesis methods that take advantage of green chemistry [109,110]. Two main approaches are reported top-down and bottom-up. While the top-down process consists of breaking down larger particles into small ones using intense mechanical forces, in the bottom-up approach the molecules associate with each other, thanks to the creation of appropriate conditions favoring intra-molecular attraction [109].
Nanotechnology is not new to the food industry and, though still portrayed as dangerous for health concerns, its applications are numerous: from agricultural production, protecting crops, or feeding plants, to enhancing and preserving food quality from the processing to the packaging [109]. Nevertheless, encapsulation technology to control and target the delivery and bioavailability enhancement of healthy nutrients and nutraceuticals remains the most attractive aspect of nanomaterials.

3.1. Nanotech in Bioavailability-Boosting Systems

Colloidal delivery systems and excipient foods are among the latest approaches developed to enhance bioavailability of nutraceuticals like curcumin. On one hand, excipient foods do not have any bioactivity themselves but, once in the gut, form an environment that increases bioaccessibility, stability, or absorption of any substances co-ingested with them [106,109,111]. In vitro studies tested the effects of excipient emulsions on stability and bioaccessibility of powdered curcumin at both 30 °C and 100 °C, to simulate conditions commonly found in foods such as dressings or cooked sauces. Excipient emulsion was prepared by mixing corn oil (10% w/w) and Tween-80 aqueous solution (90% w/w) with a high-shear mixer and the use of a microfluidizer allowed researchers to obtain particles with different diameter sizes—large, medium, or small droplets, respectively. Powdered curcumin was dissolved in the excipient emulsion at a percentage of 3% and the resulting mixtures were then incubated at either 30 °C for 30 min or 100 °C for 10 min. From the analysis of the collected samples, it emerged that no appreciable changes were found in curcumin excipient emulsion mixtures—in droplet characteristics, electrical charge, or droplet microstructures of the samples incubated at 30 °C for 30 min—while a small increase in the mean droplet diameters and the appearance of some larger oil droplets in the systems that firstly contained small droplets occurred after heating at 100 °C for 10 min, indicating that some aggregation and coalescence happened during incubation at high temperatures [112,113]. As for curcumin bioaccessibility, the studies demonstrated that a significantly greater amount of curcumin was found in the emulsion after incubation at 100 °C than 30 °C and it was considerably higher for large and medium droplets than for the small droplets at 30 °C. Moreover, some curcumin crystals were clearly observed in all the mixtures incubated at lower temperatures, suggesting incomplete dissolution. Since the excipient emulsion incubated at 100 °C contained a major amount of curcumin, it was tested for its potential biological fate passing through a three-step GI tract model. This evidence supported the involvement of droplet sizes of the initial excipient emulsion on the curcumin concentration in digesta or mixed micelle phases produced from lipid digestion (large > small ≈ medium), however curcumin bioaccessibility did not result from droplet size, since the same amount of mixed micelles formed for all three droplet sizes studied [113]. Excipient emulsions of curcumin, when compared with curcumin dissolved in corn oil or buffer solution (either at 30 °C for 30 min or 100 °C for 10 min) and exposed to simulated GI tract conditions, revealed a higher curcumin concentration in the mixed micelle phase, highlighting the impact of the nature of the food matrix on the digestion fate [112].
Conversely, colloidal delivery systems enable the encapsulation of low-bioavailable bioactive agents inside nanoparticles. The latter are specifically designed for specific nutraceuticals to be encapsulated in an appropriate food matrix and incorporated into and tailored corresponding to the needed characteristics in the final product [106]. The systematic approach to obtain the most appropriate final-purpose delivery system is the recently developed delivery by design (DbD) principle. It consists of a seven-stage process that aims to design, fabricate, and test colloidal delivery systems suitable for commercial applications. All stages are summarized in Table 3. Briefly, stages 1 and 2 involve defining the physicochemical properties of either the active agent or the final product, highlighting the major obstacles to overcome and the functional attributes, respectively. Therefore, stage 3 entails defining the characteristics that a colloidal delivery system should possess to effectively encapsulate the active ingredient. Once the colloidal delivery system has been defined, it is necessary to qualify the features of the colloidal particles and select the most suitable delivery system, bearing in mind the provided features (stage 4). Stages 5 to 7 involve the optimization of the fabrication method: (i) definition of the manufacturing approaches to ensure that the production is economically reliable for commercial applications, (ii) establishment of testing protocols for evaluating and ensuring the performances of a delivery system in a specific end-product, (iii) monitoring and recording properties of delivery systems and end-products to make it easier to adjust, where needed, the factors that can affect their properties [54,106].
The DbD approach has been tested on curcumin as a case study. In the reported case study, since curcumin is a water-insoluble, low-molecular-weight, and crystalline constituent that cannot be integrated into an aqueous-based functional beverage without being dispersed in colloidal particles, it has been encapsulated to design a fortified nutritional drink like the Ayurvedic “golden milk”. As this beverage appears cloudy, yellowish, and creamy, oil-in-water nanoemulsion was selected as the most appropriate delivery system to reach the goal [54,106].
The range of functional food matrix where curcumin could be incorporated is extremely wide and each food type has its own characteristics (bread, beverages, sauces, dressings, frozen meals, cereal bars), therefore, colloidal delivery systems should fulfill each requirement according to the end-product’s variability. An overview of the major food-grade colloidal delivery systems that efficiently encapsulate curcumin is given below. Moreover, the latter, categorized as simple, delivery systems could be used as building blocks to prepare complex delivery systems through structural design techniques, such as embedding, clustering, coating, or mixing [54].

Lipid-Based Colloidal Delivery Systems

Micelles, emulsions, nano- and microemulsions, liposomes, and solid lipid nanoparticles (SLNPs) or nanostructured lipid carriers belong to this category. As the lipid-based colloidal particles are surely the most widely investigated and used delivery system for curcumin, detailed information from clinical studies involving these carriers have already been given in the previous section and the current paragraph would only deepen their specific use according to the DbD practice.
Micelles are among the smallest colloidal particles used to deliver drugs, since their diameter typically ranges from 5 to 20 nm. Consequently, they hinder the light scatter and appear optically clear, which makes them appropriate for application in beverage products that are supposed to be transparent. Micelles are thermodynamically stable systems formed from natural or, more often, synthetic surfactant molecules that self-assemble when the critical micellar concentration (CMC) is reached. They organize in particles with a hydrophobic core, containing surfactant tails and a hydrophilic shell made up of the surfactant heads. Their fabrication is somewhat simple and consists of heating and/or mixing the surfactant, the bioactive agent, and water together. The most used food-grade surfactants for micelles production include Tween-20 and Tween-80 [81], although micelles could also be formed from natural biosurfactants, such as sophorolipids, synthesized by yeast fermentation. Among their advantages, it is important to consider the high stability and good bioavailability. However, sophorolipids need a high concentration of surfactant to self-assemble, which represents an obstacle in terms of costs and consumer acceptance [54,86,106,111]. Microemulsions and micelles are structurally and functionally similar, except for the presence of some oil in microemulsions that form an additional core located between the surfactant tails, which is responsible for their larger dimensions. Capryol-90, Transcutol P, Cremophor-RH40, soybean oil, soy lecithin, and Tween-80 are some of the food-grade ingredients used to prepare microemulsions [106].
Emulsions and nanoemulsions are usually formed by two immiscible liquids (oil and water) that are stabilized by emulsifiers, thickening, jelly agents, or many other food-grade ingredients. They differ mainly for their dimensions since their diameters typically range from 10 to 100 nm in the case of nanoemulsion or from 100 nm to 100 µm for emulsions. This difference leads to different physicochemical properties and functional attributes—while emulsions appear cloudy or opaque, nanoemulsions may appear clear. Considering that both are thermodynamically unstable, it is necessary to apply mechanical or chemical energy to produce them—high-pressure homogenization, microfluidization, and sonication are processes generally employed by the food industry and are relatively inexpensive [114]. Curcumin-loaded emulsion could be prepared from proteins or polysaccharides, such as soy carbohydrates or casein, whey protein, Arabic gum, lecithin, or Tween-80 [115,116]. Unfortunately, the instability affects their storage as they tend to separate and break down [54,86,106,111]. After comparing curcumin oil-in-water nanoemulsion and emulsion, it emerged that curcumin stability in conventional emulsion was slightly higher, but nanoemulsions had greater stability [117]. However, a curcumin nanoemulsion prepared from organogel in the oil phase with Tween-20 as an emulsifier showed a ninefold increase in curcumin oral bioavailability when compared with unformulated curcumin [118].
Liposomes are self-assembled systems formed by one or more phospholipid bilayers. Their dimensions vary considerably according to the formulation or the fabrication method. Commonly, particles with a diameter smaller than 100 nm are called nanoliposomes, while particles whose diameters are larger than 100 nm are the so-called liposomes. There are a variety of methods for fabricating liposomes, many of which use organic solvents that are inacceptable for food applications; the pH-driven procedure is an organic solvent-free technique based on the deprotonation and dissolution of hydrophobic drugs under alkaline conditions followed by neutralization with an acidic solution of surfactants. Lowering the pH by mixing the solutions forces the hydrophobic molecules of the drug to precipitate and segregate inside the surfactant molecules [119]. According to the desired end-product, liposomes could be designed to have different sizes and, consequently, different appearances, which means they can be integrated in several marketed products. Moreover, liposomal delivery systems could be fabricated from all natural ingredients to meet the compliance of consumers, such as egg lecithin [120] or, more commonly, soy lecithin, milk, rapeseed, canola seed, cottonseed, and sunflower [115]. Despite these benefits, industrial manufacturing is quite expensive, and they are not robust enough to withstand the environmental stresses encountered in foods [54,86,106,111].
SLNPs are comparable to oil-in-water emulsion, except for the lipid phase NPs, which are solid rather than liquid. Indeed, they could be prepared by homogenization of a melted fat with an aqueous emulsifier solution at a temperature higher than the fat melting point, the cooling of the emulsion below the crystallization temperature of the fat leads to the solidification of the fat droplets. SNLPs are particularly suitable for loading hydrophobic drugs that should be locked into the solid interior during the cooling step. As for the other colloidal particles, particle size could range from 20 to 1000 nm, giving the suspension of transparent or cloudy appearances. Food-grade ingredients to synthesize SLNPs include propylene glycol, glycerol stearate, esters of fatty acids, palmitic acid, soy lecithin, and glyceryl behenate in the lipid phase and Tween-80 solutions in the aqueous phase [121,122]. Unfortunately, if not properly formulated, SLNPs tend to expel the bioactive agent and to form particle aggregates. Apart from this, their fabrication profits by common and green processing operations, such as high-pressure homogenization, and can be manufactured on a commercial scale with food-grade ingredients. The overcoming of fabrication issues could be achieved by developing lipid particles that cannot form highly ordered crystalline structures, thanks to the use of a mixture of fat with diverse melting points or lipids that remain in an amorphous phase when cooled [86,106].

3.2. Biopolymer-Based Nanoparticles

The most common food-grade biopolymers used to fabricate colloidal delivery systems are either synthetic or natural [106]. When dispersed in a solvent, the biopolymer-based particles form a 3D network capable of trapping both the active agent and the solvent inside. Biopolymer nanoparticles (d < 100 nm) could be prepared from a variety of methods, among which, the simplest consists of a dispersion of drug-loaded polymeric particles in organic solutions that precipitate in aqueous media. Ultrasonication and microwaves are the latest green methods developed. Fabrication method affects the characteristics of the biopolymers, which allows for the designing of particles with specific delivery properties, such as retention or release of the active agent inside, especially for those preparations where a triggered or sustained release is required [106,111]. Poly-(lactic-co-glycolic acid) (PLGA) is a biodegradable aliphatic polyester commonly utilized for fabricating curcumin nanoparticles [123,124,125]. Several studies have underlined that PLGA has high encapsulation efficiency—curcumin loading is high and the nanoformulation offers great stability and curcumin cellular uptake is enhanced and exhibits pro-apoptosis and anti-proliferative effects on the growth of metastatic cancer (MDA-MB-231 and A2780CP) cells compared to free curcumin, while showing no effects on cell viability of the polymer itself [123]. Moreover, the kinetics of curcumin release display an initial burst release with about 43 ± 3% of the drug released from PLGA within the first hour, followed by a negligible amount of curcumin (>5%) released between 1 h and 24 h. As a matter of fact, this release/retention characteristic is a needed feature for oral administration of BCS class II substances, such as curcumin, as the improvement of the dissolution ratio could increase the bioavailability that is inhibited by their scarce solubility [124]. Despite the fact that PLGA is one of the most extensively studied biopolymer-based nanocarriers for curcumin, polysaccharides and proteins are perceived as healthier by consumers when compared not only to synthetic colloidal particles, but also to the most common lipid-based ones [106]. Many studies have focused on encapsulating curcumin with chitosan nanoparticles and the results generally revealed bioavailability and solubility enhancement in encapsulated curcumin compared to native powder [126,127,128,129]. When orally administered to rats (10 or 50 mg/kg), chitosan–curcumin nanoparticles showed an 11.45-fold increase in bioavailability compared to native curcumin and persistence in blood circulation up to 7 days, probably due to the bioadhesion properties of the polymer itself to the intestinal mucosa [129]. In another study, chitosan–curcumin nanoparticles synthesized using an ionotropic gelation technique indicated an initial burst release of curcumin for 2 h, further followed by a sustained release of the drug up to 96 h. Moreover, the formulation showed good stability at storage temperature for two months (4 °C or 25 °C). When the cytotoxic effect was evaluated, a relatively low concentration was found to be successful at inhibiting HeLa cell proliferation, while the chitosan nanoparticles alone demonstrated no considerable decrease in cell viability and satisfactory biocompatibility [126]. Caseins represent a valuable natural alternative to polysaccharides to encapsulate hydrophobic drugs. Indeed, beta-caseins from camel milk were found to form an efficient self-assembling nanostructured carrier for curcumin. The presence of the micellar structure increases curcumin solubility, bioavailability, and antioxidant activity, while proving a cytotoxic effect on human leukemia cells line K-562 [130]. Interestingly, curcumin microparticles, prepared from whey protein isolated via spray drying method, have been integrated into a model food matrix (yogurt) to test whether the characteristics of retention or release, as well as bioaccessibility or sensory properties, were affected. Findings revealed that even if the presence of yogurt affects the release of curcumin from the complex, the kinetic features, such as burst or prolonged release, are remain, highlighting the role of food components in the bioavailability of nutrients. On the other hand, curcumin microparticles did not affect yogurt properties, except for color and curcumin flavor [131]. It is worth mentioning that curcumin-loaded Pluronic® polymeric micelles gained attention as a novel drug delivery strategy through complex microbial communities into target cells. Indeed, Pluronics® are copolymers of polyethylene oxides (PEO) and lipophilic polypropylene oxide (PPO), organized in a triblock structure (PEO-PPO-PEO), characterized by an amphiphilic nature that enables self-aggregation in aqueous solution. Although further investigation is needed for confirming, improving, and better understanding the early results, curcumin-loaded Pluronic® F-127 micelles showed good chemical stability within 30 days from the fabrication and encouraging results indicate this product as an alternative therapy to inhibit biofilm proliferation processes in oral cavity diseases, such as dental caries [132].

3.3. Naturally Derived Colloidal Particles

In trying to find more natural and healthier alternatives to processed food, it was necessary to investigate the possibility of employing preexisting natural colloidal particles, such as fat globules found in bovine milk or oil bodies in plant-based milks. For instance, curcumin has been successfully encapsulated inside milk fat globules in bovine milk and oil bodies in soybean milk. In both experiments, curcumin was loaded in the oily globules thanks to the PH-shift method without any alteration of the beverages. It also showed good stability under storage in refrigerated conditions, particularly between 5 and 20 °C. Bioavailability was assessed using simulated GI tract conditions and was around 40% for bovine milk fat particles, while 55–59% was available for absorption in the mixed micelle phase from soybean milk [133].

3.4. Nanotech in Targeted Delivery Systems

The drug-targeting approach is an integral component of the overall process of drug development. Indeed, targeting the drug selectively to the cells where the pharmacological action is needed increases the therapeutic efficiency while decreasing undesirable side effects [134,135]. The conjugation of the drug-targeting approach with a nanodelivery systems would allow the nutraceuticals to be at the proper place at the correct time, increasing the beneficial effects, decreasing undesirable toxic effects, limiting nutraceuticals metabolism, and boosting their bioavailability. As many strategies have been taken under consideration by scientists, three of them are discussed below.

3.4.1. Peyer’s Patches

Peryer’s patches (PPs) are lymphoid nodules that aggregate in follicular structures that cover the ileum portion of the small intestine, giving birth to the gut-associated lymphoid tissue (GALT). PP follicles consist of a core containing B lymphocytes, dendritic cells, and macrophages that are detached from the intestinal lumen by a specific layer, the follicle-associated epithelium (FAE), which incorporates enterocytes, goblet cells, and modified epithelial cells termed microfold cells (M cells) [136,137]. M cells are specialized enterocytes whose role is to bind and transfer xenobiotics from the lumen to the underlying immune cells, which then ignore or trigger the immune response depending on the antigen being processed [137]. Although several studies have already taken PPs into consideration for drug-targeting purposes [138,139,140], most of the orally administered drugs still get metabolized before reaching the systemic circulation, unless selective uptake occurs in lymph. Lipid-based and polymeric nanoparticles are the delivery systems known to pass the intestinal membrane and enter the lymphatic circulation, the first being absorbed via chylomicrons/enterocyte uptake, while the second are taken up by M cells [136], indeed, they can transport substances across the intestinal wall through endocytosis, pinocytosis, micropinocytosis, and phagocytosis and deliver it to the basolateral side by exocytosis. Accordingly, M cells represent an intriguing opportunity to target drugs at PPs, though the uptake strongly depends on the particle size, zeta potential, and surface modifications of the particulate system. In consequence, non-ionized, hydrophobic nanoparticles, sized below 1 μm (more precisely below 200 nm) are easily transcytosed by M cells. Furthermore, surface functionalization with specific M cells ligands, such as lecithin, IgA, or antibodies, enables more specific targeting [136,137]. Despite these characteristics, the main issue with M cells uptake remains the integrity of the substance reaching the intestinal wall, thanks to carriers specifically tailored for the purpose.
As already described above, polymeric nanoparticles trap hydrophobic bioactive agents inside a 3D network structure, preventing enzymatic degradation occurring over the GI tract and delivering the intact drug with unchanged PP levels [137,138]. Size, shape, surface properties, and surface charge deeply influence the rate and length of PP absorption. Indeed, the transport rate of hydrophobic particles, whose dimensions are 50–200 nm, rod-shaped, and negatively charged is higher than bigger, sphere- or disc-shaped hydrophilic particles with neutral or positive surface charge. Nevertheless, an in vivo study proposed the encapsulation of curcumin in the lauroyl sulphated chitosan (LSCS-CUR), a hydrophilic and positively charged polymer. LSCS-CUR nanoparticles displayed better aqueous solubility than unformulated curcumin and better mucoadhesion than chitosan alone, which plays a pivotal role in enhancing curcumin cellular uptake by expanding the absorption surface. Similarly, mucoadhesion is involved in the longer persistence of curcumin in blood up to 7 days after oral administration. It is noteworthy that tissue distribution analysis showed a higher concentration of curcumin localized in the intestinal portion, particularly at the duodenum level [129]. As mentioned above, the rate of transport or release of bioactive substances reflects the nature of the polymer matrix. In fact, encapsulating curcumin in PLGA nanoparticles allows biphasic release of the active ingredient—a small amount is instantly issued from the formulation, followed by a slow and steady release of the remaining. This release trend has further been demonstrated in an animal test, where Wistar rats were fed with curcumin encapsulated in PLGA nanoparticles or curcumin alone. The formulation showed a 24 h steady release, as opposed to unformulated curcumin, which achieved the maximum plasma concentration after 30 min from oral intake and then fell following its metabolism [141].
Dendrimers are nanometer-sized treelike structures characterized by an initiator core of monomeric units covered by an outermost exterior layer that provides a multifunctional surface for surface chemistry modification [86]. Once again, dendrimer size, charge, and concentration influence their absorption rate at intestinal level, which can occur for molecules with diameters up to 3 nm, either through the paracellular or transcellular routes. It is noteworthy that dendrimers appeared to express specific tropism for the small rather than the large intestine. Although only few studies have been conducted on curcumin–dendrimer formulation, there is evidence of improved aqueous solubility and bioavailability of dendrosomal curcumin [142,143], cytotoxic activity and tumor target efficacy against several human cancer cell lines [142,143,144,145], reduced cancer cell proliferation, and apoptosis induction with a dose-dependent cytotoxicity showed by polyamidoamine curcumin dendrimer formulation [146].
Self-microemulsifying drug delivery systems (SMEDDSs) are dry microemulsions without an aqueous phase; they can be inserted into capsules or converted into solid dosage forms, such as tablets or pellets, that, upon oral administration, are dispersed in the gut lumen, spontaneously forming a microemulsion [86]. This fact allows not only the active agent to be presented in a solution form in the gut, but also, since SMEDDSs do not need biliary salts and enzymes to be digested, limits possible absorption alterations due to individual variability. Lastly, enhancing drug solubilization and improving lymphatic target efficiency of SMEDDSs increases membrane permeability in the GI tract and lymphatic uptake of drugs, thus lowering the drug amount required to have a clinical effect and reducing the systemic side effects, thereby improving patient compliance and cost of therapy [137]. Several studies conducted on animals have agreed, finding a meaningful increase in absorption and bioavailability when curcumin SMEDDS formulation is compared to unformulated curcumin [147,148,149,150,151].

3.4.2. Inorganic Nanoparticles

Metallic or non-metallic particles, sized between 1 and 100 nm are commonly referred to as inorganic nanomaterials [110]. As for other nanoparticles, their tiny features, small size, large surface area, and higher reactivity mark the major benefits and the deeper concerns associated with their use [109], since the average particle diameter, size distribution, and charge affect the physical stability and the in vivo distribution of nanoparticles. Nevertheless, inorganic nanoparticles offer a wide range of applications in the biomedical field as they are already used in diagnostic imaging, gene vehicles, protein separation and purification, cellular tracking, and more [110,152]. Interestingly, many of these inorganic nanomaterials are routinely used in food and agricultural fields, as pesticides to treat the crops or as additives to provide food and beverages with specific optical characteristics [109]. Inorganic nanoparticles provide high payload loading and payload protection, tunable degradation rates, localized target delivery, enhanced permeation, or accumulation into specific tissues due to the base material properties and chemistry surface functionalization chosen—all these characteristics make them an extremely practical targeted drug delivery system for nutraceuticals.
Among metallic nanoparticles, gold and silver (Au and Ag, respectively) particles have garnered attention because of their high plasma absorption, biocompatibility, versatility, and antimicrobial properties [153]. Gold nanoparticles are mainly synthesized by chemical methods that encompass the reduction in gold salt with the help of reducing agents, although plant-based materials, sunlight irradiation, and green chemistry have been used as green methods. The modulation of thiol/gold ratio during the fabrication process permits control of the size and shape of nanoparticles, which are closely related to nanoparticle distribution and toxicity [152,153]. In a hypertrophy-induced animal model, encapsulating curcumin in gold-loaded PLGA nanoparticles (CAu-PLGA NPs) resulted in the inhibition of cardiac hypertrophy due to the structural modification of curcumin in CAu-PLGA NPs that increased its bioavailability, solubility, and absorption, improved blood circulation and resistance to enzyme degradation; furthermore, CAu-PLGA NPs allow a more controlled and complete drug delivery, increasing therapeutic efficiency [154]. Another in vivo study assessed the potential and specificity of curcumin-encapsulated chitosan-graft-poly (N-vinyl caprolactam) nanoparticles containing gold nanoparticles (Au–curcumin–TRC NPs) in targeted cancer drug delivery. Interestingly, Au–curcumin–TRC NPs were found to facilitate extended persistency of curcumin in the blood for up to a week, compared to curcumin alone, which disappeared from circulation after 6 h, and maximum curcumin accumulation was observed in the tumor compared to other organs [155]. To support these findings, several in vitro studies proved cytotoxicity and increased uptake of curcumin-functionalized gold nanoparticles (curcumin–Au NPs) in human prostate cancer cell culture [156]. Moreover, antiproliferative and apoptotic effects on colon and breast human cancer cell lines [157] and a synergistic effect produced by both the gold core and curcumin that triggers the generation of ROS, which consequently initiates the apoptotic pathway with the release of protein Bax, activation of PARP cleavage, and DNA fragmentation, were observed [158].
Silver nanoparticle synthesis takes advantage of chemical, physical, or green methods based on fungi, bacteria, algae, and plants [153]. As with other nanoparticles, shape, size, concentration, colloidal state, and surface charge influence the antimicrobial properties exhibited by silver nanoparticles that lead to the fabrication of curcumin-loaded silver nanoparticles as antibacterial agents. Curcumin–Ag NPs of 25–35 nm were synthesized and demonstrated efficacy against Gram-positive and Gram-negative bacteria, lowering the minimum effective curcumin concentration fourfold (20–5 mg/L) and toxicity to human keratinocytes. Moreover, skin biocompatibility studies showed an induced anti-inflammatory effect on human macrophages [159]. Although daily approximate silver intake by humans through food and/or water is 0.4–30 μg, which indicates that in the given range its consumption is safe, there are some reports where silver nanoparticle-related toxic effects have been observed in animal models [153] that need further investigation.
Magnetic nanoparticles (MNPs) consist of a metallic oxide core or metal that could be functionalized within an inorganic metal or polymer coating. Among MNPs, magnetic iron oxide nanoparticles, otherwise known as magnetite or Fe3O4, have been approved by US Food and Drug Administration (FDA) for clinical imaging and drug delivery applications as they are considered safe and biocompatible [153]. Among the variety of existing fabrication methods, green biosynthesis is usually selected over other techniques due to higher biocompatibility, biodegradability, and non-toxic surface coating of green materials [153]. Curcumin-loaded MNPs were evaluated both in vitro and in vivo in pancreatic cancer for assessing therapeutic efficacy. The MNPs–curcumin formulation appeared efficiently dose-dependently internalized by human pancreatic cancer cells, whereas the in vivo study showed the ability of the MNPs–curcumin formulation to suppress pancreatic tumor growth. Furthermore, the MNPs–curcumin formulation improved serum bioavailability of curcumin in mice up to 2.5-fold compared with free curcumin, without any sign of hemotoxicity [160]. Another in vivo study tested the hemocompatibility and bioavailability of polyethyleneimine and glutathione functionalized iron oxide nanoparticles for targeted curcumin, resulting in up to 2.5-fold enhanced serum bioavailability for curcumin-loaded MNPs compared with free curcumin [161].
Mesoporous silica nanoparticles encountered the favor of the biomedical industry thanks to their high chemical inactivity, biocompatibility, thermal stability, high hydrophilicity, high loading capacity, and resistance to microbial attack. As many synthesis methods are employed for fabrication, the green process of silica nanoparticles fabrication is gaining more and more resonance since it relies on the use of agricultural waste. Amorphous silica, to which mesoporous silica belongs, was considered secure by the FDA at an oral intake of up to 1500 mg/die for foods and as a delivery vehicle [162,163], contrary to its crystalline form, involved in emphysema, pulmonary tuberculosis, lung cancer, and fibrotic lung disease (silicosis) [153]. Curcumin-loaded mesoporous silica nanoparticles were prepared and coated with polyethylene glycol (PEG) and ultimately conjugated with the targeting moiety transferrin (Tf) for targeting pancreatic cancer cells. Data showed enhanced uptake and, consequently, higher cell killing in vitro, while in vivo results on animal models clearly demonstrated inhibition of tumor growth and prevention from distant metastasis spreading [164].

3.5. Microbiota

It is largely acknowledged that the gut microbiota is pivotal for maintaining human health and conditions recognized as dysbiosis are correlated with the onset of chronic pathologies, including allergies, autoimmunity, GI disorders, metabolic and cardiovascular disorders, cancer, and CNS dysfunctions [165,166]. Gut microbiota composition tightly depends on host genetics and interpersonal variance and it is profoundly influenced by age, diet, lifestyle, and environmental factors [166], so that any improper diets could determine alterations in the gut microbiota composition, leading to intestinal permeability changes and low-grade gut inflammation [167].
Recently, a growing number of studies have focused on the correlation between curcumin and gut microbiota after oral administration as a high concentration of polyphenols is found in the GI tract [168,169,170]. Taken together, these results highlight a bidirectional interaction [166,167]: on one hand gut microbiota participate actively to curcumin metabolism, as it acts as a metabolic bioreactor, producing various enzymes that contribute to curcumin biotransformation in metabolites, which display similar or superior activity compared to curcumin itself [171,172,173]; in contrast, the administration of curcumin promotes the growth of beneficial bacterial strains, inducing a shift in the ratio between beneficial and pathogenic species [168], improving intestinal barrier functions [174,175] and counteracts the expression of biomarkers of inflammation and oxidative stress [169].
Though these studies were executed under different conditions, using various doses or formulations, thus being hardly comparable, they give the path to novel microbiota-targeting curcumin-based therapies.

3.6. A Brief Overview of Fabrication Issues

In order to deepen commercial pros and cons of nanoformulations used as strategies to improve drug bioavailability, it is firstly necessary to establish their intended use. Ingredients, fabrication processes, and research and development investments would vary considerably depending on food and food-related production or dietary supplement production, inevitably affecting the cost to both manufacturers and consumers.
When referring to the food industry, it is important to consider, next to the ease of fabrication and related costs, consumer acceptance, which influences consumer choice. All the lipid-based colloidal delivery systems, as well as the naturally derived biopolymers, could be produced from natural food-grade ingredients commonly perceived as healthier by consumers, although only emulsions, nanoemulsions, and SLNPs could be fabricated in a large-scale process, using common and inexpensive processing operations that are already used in the food industry, like high-pressure homogenization, microfluidization, or sonication [106,176]. According to curcumin bioavailability enhancements, data suggest that nanoemulsions are the most suitable delivery system since they could be fabricated from a variety of cheap, plant-based emulsifiers (soy, pea, lentil, bean proteins) and lipids (corn, sunflower or coconut oil) on a commercial scale [106,177]; if properly formulated, they have relatively good stability over the stresses encountered by foods during production and storage [178], as well as high bioavailability [179]. Moreover, they are extremely versatile since they could be incorporated into a beverage or dried into a powdered form [180].
On the side of clinical or pharmaceutical applications, the latest approaches investigated above, such as inorganic nanoparticles, nanopolymers, or dendrimers, suffer from a deficiency of data in terms of human pharmacokinetics and pharmacodynamics, which limit predictable evaluation of costs related to raw materials. Nevertheless, fabrication methods have been well investigated. Concerning polymer nanoparticles, among many production approaches [181,182,183], nanoprecipitation (emulsion–solvent evaporation method) and the emulsions–diffusion method are the easiest to scale up [183] and the simplest to perform [86], although green methods based on ultrasonication and microwaves are emerging in order to reduce toxic reagents and simplify the procedure to allow economic scale-up [86,182].
Inorganic nanoparticles could be fabricated using numerous methods, including chemical, physical, and bio-based green methods (sunlight irradiation, from bacteria, fungi, yeast, algae, and plants) [110]. As already mentioned above, top-down and bottom-up approaches are among the easiest. Due to the fact that inorganic nanoparticles are already used in diagnostic medicine and imaging, scaling up their production is a reality, although not free from costs.
As they consist of a repetition of monomeric interior units encircled by the outermost region, dendrimers are generally synthesized by a sequential repeat of reaction steps. Indeed, they are formed of a core, from which a number of spacers depart, whose ends combine with a high degree of branching that allows surface chemistry functionalization with proteins, amino acids, fatty acids, and polymers in order to enhance solubility and reduce toxicity [184,185]. The use of dendrimers as drug carriers is still the subject of study, though recent applications include nanomedicines, gene-delivery vectors, biological adhesives, and imaging agents [137,185,186,187,188]. Finally, SMEDDSs belong to the emulsion class since they are dry microemulsions without an aqueous phase and they have the great advantage to spontaneously form emulsions in vivo without any energy input [86,137]. According to their high targeting efficiency, it is possible to reduce the drug amount required to have a clinical effect, thereby improving patient compliance and the cost of therapy [137].
Taking these evaluations into account and considering the absence of data regarding some aspects of the newest drug delivery approaches, it is appropriate to affirm that nanoemulsions are, at present, the most suitable, well-investigated, and low-cost approach to enhance bioavailability of curcumin, either as a drug or a dietary supplement to enrich food and beverages.

4. Conclusions

Currently, the polyphenolic compound curcumin is one of the most investigated natural products and its beneficial effects in several human disorders have been established, along with a low toxic profile. During the years, numerous curcumin drug targets have been described, providing the molecular basis for the pharmacological action. Unfortunately, the bioavailability and delivery of curcumin are the main obstacles for its effectiveness, precluding the use in medicine. In this review, we have explored the most promising and modern strategies for improving curcumin bioavailability, offering a comprehensive overview regarding this field. Surely, the progress in the nanoparticle field could provide, in the near future, further curcumin-based complexes with significant pharmacological profiles that can be developed as effective therapeutic agents for treating human diseases. The advancement in this field holds great promise to employ this polyphenol in medicine. Moreover, due to the multifunctional profile of curcumin, an efficacious delivery method may offer a better possibility of effective prophylaxis or treatment in complicated multifactorial disorders, such as systemic inflammatory diseases and cancer, supporting the curcumin-based agents as preventive and disease-modifying therapeutics.

Author Contributions

Conceptualization, R.T., S.B. and V.C.; investigation, R.T. and S.B.; data curation, R.T., S.B. and V.C.; writing—original draft preparation, R.T. and S.B.; writing—review and editing, R.T., S.B. and V.C.; supervision, S.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

3DThree dimensional
γ-GTPγ-glutamyl transpeptidase
AktProtein kinase B
ALBAlbumin
ALPAlkaline phosphatase
ALTAlanine aminotransferase
ASTAspartate aminotransferase
AT-LDLA1-antitrypsin-LDL
AUCArea under the curve
BaxBcl-2-like protein 4
BCSBiopharmaceutics classification system
BCVABest-corrected visual acuity
BMIBody mass index
BSBlood sugar
CAu-PLGA NPsGold-loaded PLGA nanoparticles
CCDCCambridge Crystallographic Data Centre
cfPWVCarotid-femoral pulse wave velocity
CKCreatine kinase
CMCCritical micellar concentration
CNScentral nervous system
CRPC-reactive protein
CYP3A4Cytochrome P450 3A4
DbDDelivery by Design
DOMSDelayed onset muscle soreness
EGFEpithelial growth factor
FAEFollicle-associated epithelium
FDAUS Food and Drug Administration
GALTGut-associated lymphoid tissue
GBGlioblastoma
GEMGemcitabine
GIGastrointestinal
GRASGenerally regarded as safe
HDLHigh-density lipoprotein
HDL-CHigh-density lipoprotein cholesterol
IκBαnuclear factor of kappa light polypeptide gene enhancer in B-cells inhibitor α
IL-1βInterleukin-1β
IL-6Interleukin-6
i.p.Intraperitoneal
JKOMJapanese knee osteoarthritis measure
JOAJapanese orthopedic association
LDHLactate dehydrogenase
LDLLow-density lipoprotein
LSCS-CURLauroyl sulphated chitosan–curcumin
M cellsMicrofold cells
MCTMedium-chain triglyceride
MNPsMagnetic nanoparticles
mPa·sMillipascal second
NAFLDNonalcoholic fatty liver disease
NF-kBNuclear factor kappa-light-chain-enhancer of activated B cells
Nrf2Nuclear factor erythroid 2–related factor 2
NSAIDsNon-steroidal anti-inflammatory drugs
OCTOptical coherence tomography
PARPPoly (ADP-ribose) polymerase
PEGPolyethylene glycol
PEG 400Polyethylene glycol 400
PEOPolyethylene oxides
PNSPolar–nonpolar sandwich
PPOPolypropylene oxide
PPsPeryer’s patches
PLGAPoly-(lactic-co-glycolic acid)
PVPPolyvinyl pyrrolinone
QoLQuality of life
OAOsteoarthritis
ROSReactive Oxygen Species
sCD40LSoluble CD40 ligand
SMEDDSsSelf-microemulsifying drug delivery systems
SLCPsSolid lipid curcumin particles
SLNPsSolid lipid nanoparticles
SLPsSolid lipid particles
TAGTriacylglycerols
TCTotal cholesterol
TfTransferrin
TGTriglycerides
TNFαTumor necrosis factor α
TNFβTumor necrosis factor β
UDP-glucuronosyltransferaseUridine 5′-diphospho-glucuronosyltransferase
UMIN Clinical TrialsUniversity hospital medical information network Clinical Trials
UV lightUltraviolet light
VASVisual analog scale
WOMACWestern Ontario and McMaster Universities’ arthritis index

References

  1. Atanasov, A.G.; Zotchev, S.B.; Dirsch, V.M.; Supuran, C.T. Natural products in drug discovery: Advances and opportunities. Nat. Rev. Drug Discov. 2021, 20, 200–216. [Google Scholar] [CrossRef] [PubMed]
  2. Rodrigues, T.; Reker, D.; Schneider, P.; Schneider, G. Counting on natural products for drug design. Nat. Chem. 2016, 8, 531–541. [Google Scholar] [CrossRef] [PubMed]
  3. Gupta, S.; Kesarla, R.; Omri, A. Formulation strategies to improve the bioavailability of poorly absorbed drugs with special emphasis on self-emulsifying systems. ISRN Pharm. 2013, 2013, 848043. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Kumar Vivekanandhan, D.; Ranjan Prasad Verma, P.; Kumar Singh, S. Emerging Technologies for Improving Bioavailability of Polyphenols. Curr. Nutr. Food Sci. 2016, 12, 12–22. [Google Scholar] [CrossRef]
  5. Zhao, J.; Yang, J.; Xie, Y. Improvement strategies for the oral bioavailability of poorly water-soluble flavonoids: An overview. Int. J. Pharm. 2019, 570, 118642. [Google Scholar] [CrossRef]
  6. Shoji, Y.; Nakashima, H. Nutraceutics and delivery systems. J. Drug Target. 2004, 12, 385–391. [Google Scholar] [CrossRef]
  7. Zhou, H.; Beevers, C.S.; Huang, S. The targets of curcumin. Curr. Drug Targets 2011, 12, 332–347. [Google Scholar] [CrossRef]
  8. Lin, J.K. Molecular targets of curcumin. Adv. Exp. Med. Biol. 2007, 595, 227–243. [Google Scholar] [CrossRef]
  9. Kunnumakkara, A.B.; Bordoloi, D.; Padmavathi, G.; Monisha, J.; Roy, N.K.; Prasad, S.; Aggarwal, B.B. Curcumin, the golden nutraceutical: Multitargeting for multiple chronic diseases. Br. J. Pharmacol. 2017, 174, 1325–1348. [Google Scholar] [CrossRef] [Green Version]
  10. Youdim, M.B.H.; Buccafusco, J.J. CNS Targets for multi-functional drugs in the treatment of Alzheimer?s and Parkinson?s diseases. J. Neural Transm. 2005, 112, 519–537. [Google Scholar] [CrossRef]
  11. Epstein, J.; Sanderson, I.R.; Macdonald, T.T. Curcumin as a therapeutic agent: The evidence from in vitro, animal and human studies. Br. J. Nutr. 2010, 103, 1545–1557. [Google Scholar] [CrossRef] [Green Version]
  12. Hatcher, H.; Planalp, R.; Cho, J.; Torti, F.M.; Torti, S.V. Curcumin: From ancient medicine to current clinical trials. Cell. Mol. Life Sci. 2008, 65, 1631–1652. [Google Scholar] [CrossRef]
  13. Aggarwal, B.B.; Sundaram, C.; Malani, N.; Ichikawa, H. Curcumin: The Indian solid gold. Adv. Exp. Med. Biol. 2007, 595, 1–75. [Google Scholar] [CrossRef] [PubMed]
  14. Ammon, H.P.; Wahl, M.A. Pharmacology of Curcuma longa. Planta Med. 1991, 57, 1–7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Abe, Y.; Hashimoto, S.; Horie, T. Curcumin inhibition of inflammatory cytokine production by human peripheral blood monocytes and alveolar macrophages. Pharmacol. Res. 1999, 39, 41–47. [Google Scholar] [CrossRef] [PubMed]
  16. Srimal, R.C.; Dhawan, B.N. Pharmacology of diferuloyl methane (curcumin), a non-steroidal anti-inflammatory agent. J. Pharm. Pharmacol. 1973, 25, 447–452. [Google Scholar] [CrossRef]
  17. Jurenka, J.S. Anti-inflammatory properties of curcumin, a major constituent of Curcuma longa: A review of preclinical and clinical research. Altern. Med. Rev. 2009, 14, 141–153. [Google Scholar]
  18. Jakubczyk, K.; Druzga, A.; Katarzyna, J.; Skonieczna-Zydecka, K. Antioxidant Potential of Curcumin-A Meta-Analysis of Randomized Clinical Trials. Antioxidants 2020, 9, 1092. [Google Scholar] [CrossRef]
  19. Tomeh, M.A.; Hadianamrei, R.; Zhao, X. A Review of Curcumin and Its Derivatives as Anticancer Agents. Int. J. Mol. Sci. 2019, 20, 1033. [Google Scholar] [CrossRef] [Green Version]
  20. Goozee, K.G.; Shah, T.M.; Sohrabi, H.R.; Rainey-Smith, S.R.; Brown, B.; Verdile, G.; Martins, R.N. Examining the potential clinical value of curcumin in the prevention and diagnosis of Alzheimer’s disease. Br. J. Nutr. 2016, 115, 449–465. [Google Scholar] [CrossRef]
  21. Fan, C.; Song, Q.; Wang, P.; Li, Y.; Yang, M.; Yu, S.Y. Neuroprotective Effects of Curcumin on IL-1beta-Induced Neuronal Apoptosis and Depression-Like Behaviors Caused by Chronic Stress in Rats. Front. Cell. Neurosci. 2018, 12, 516. [Google Scholar] [CrossRef] [PubMed]
  22. Cole, G.M.; Teter, B.; Frautschy, S.A. Neuroprotective effects of curcumin. Adv. Exp. Med. Biol. 2007, 595, 197–212. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Jobin, C.; Bradham, C.A.; Russo, M.P.; Juma, B.; Narula, A.S.; Brenner, D.A.; Sartor, R.B. Curcumin blocks cytokine-mediated NF-kappa B activation and proinflammatory gene expression by inhibiting inhibitory factor I-kappa B kinase activity. J. Immunol. 1999, 163, 3474–3483. [Google Scholar] [PubMed]
  24. Gururaj, A.E.; Belakavadi, M.; Venkatesh, D.A.; Marme, D.; Salimath, B.P. Molecular mechanisms of anti-angiogenic effect of curcumin. Biochem. Biophys. Res. Commun. 2002, 297, 934–942. [Google Scholar] [CrossRef]
  25. Bhandarkar, S.S.; Arbiser, J.L. Curcumin as an inhibitor of angiogenesis. Adv. Exp. Med. Biol. 2007, 595, 185–195. [Google Scholar] [CrossRef] [PubMed]
  26. Karunaweera, N.; Raju, R.; Gyengesi, E.; Munch, G. Plant polyphenols as inhibitors of NF-kappaB induced cytokine production-a potential anti-inflammatory treatment for Alzheimer’s disease? Front. Mol. Neurosci. 2015, 8, 24. [Google Scholar] [CrossRef] [Green Version]
  27. Shakibaei, M.; John, T.; Schulze-Tanzil, G.; Lehmann, I.; Mobasheri, A. Suppression of NF-kappaB activation by curcumin leads to inhibition of expression of cyclo-oxygenase-2 and matrix metalloproteinase-9 in human articular chondrocytes: Implications for the treatment of osteoarthritis. Biochem. Pharmacol. 2007, 73, 1434–1445. [Google Scholar] [CrossRef]
  28. Jhun, J.; Min, H.K.; Na, H.S.; Kwon, J.Y.; Ryu, J.; Cho, K.H.; Choi, J.; Jung, K.; Lee, S.Y.; Kim, S.J.; et al. Combinatmarion treatment with Lactobacillus acidophilus LA-1, vitamin B, and curcumin ameliorates the progression of osteoarthritis by inhibiting the pro-inflammatory mediators. Immunol. Lett. 2020, 228, 112–121. [Google Scholar] [CrossRef] [PubMed]
  29. Wang, J.; Wang, X.; Cao, Y.; Huang, T.; Song, D.X.; Tao, H.R. Therapeutic potential of hyaluronic acid/chitosan nanoparticles for the delivery of curcuminoid in knee osteoarthritis and an in vitro evaluation in chondrocytes. Int. J. Mol. Med. 2018, 42, 2604–2614. [Google Scholar] [CrossRef] [Green Version]
  30. Buhrmann, C.; Brockmueller, A.; Mueller, A.L.; Shayan, P.; Shakibaei, M. Curcumin Attenuates Environment-Derived Osteoarthritis by Sox9/NF-kB Signaling Axis. Int. J. Mol. Sci. 2021, 22, 7645. [Google Scholar] [CrossRef]
  31. Ahsan, H.; Parveen, N.; Khan, N.U.; Hadi, S.M. Pro-oxidant, anti-oxidant and cleavage activities on DNA of curcumin and its derivatives demethoxycurcumin and bisdemethoxycurcumin. Chem.-Biol. Interact. 1999, 121, 161–175. [Google Scholar] [CrossRef]
  32. Jiao, Y.; Wilkinson, J.T.; Christine Pietsch, E.; Buss, J.L.; Wang, W.; Planalp, R.; Torti, F.M.; Torti, S.V. Iron chelation in the biological activity of curcumin. Free Radic. Biol. Med. 2006, 40, 1152–1160. [Google Scholar] [CrossRef] [PubMed]
  33. Soni, K.B.; Kuttan, R. Effect of oral curcumin administration on serum peroxides and cholesterol levels in human volunteers. Indian J. Physiol. Pharmacol. 1992, 36, 273–275. [Google Scholar]
  34. Sharma, R.A.; McLelland, H.R.; Hill, K.A.; Ireson, C.R.; Euden, S.A.; Manson, M.M.; Pirmohamed, M.; Marnett, L.J.; Gescher, A.J.; Steward, W.P. Pharmacodynamic and pharmacokinetic study of oral Curcuma extract in patients with colorectal cancer. Clin. Cancer Res. 2001, 7, 1894–1900. [Google Scholar] [PubMed]
  35. Lao, C.D.; Ruffin, M.T.T.; Normolle, D.; Heath, D.D.; Murray, S.I.; Bailey, J.M.; Boggs, M.E.; Crowell, J.; Rock, C.L.; Brenner, D.E. Dose escalation of a curcuminoid formulation. BMC Complement. Altern. Med. 2006, 6, 10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Kocaadam, B.; Sanlier, N. Curcumin, an active component of turmeric (Curcuma longa), and its effects on health. Crit. Rev. Food Sci. Nutr. 2017, 57, 2889–2895. [Google Scholar] [CrossRef]
  37. Sakano, K.; Kawanishi, S. Metal-mediated DNA damage induced by curcumin in the presence of human cytochrome P450 isozymes. Arch. Biochem. Biophys. 2002, 405, 223–230. [Google Scholar] [CrossRef]
  38. Urbina-Cano, P.; Bobadilla-Morales, L.; Ramirez-Herrera, M.A.; Corona-Rivera, J.R.; Mendoza-Magana, M.L.; Troyo-Sanroman, R.; Corona-Rivera, A. DNA damage in mouse lymphocytes exposed to curcumin and copper. J. Appl. Genet. 2006, 47, 377–382. [Google Scholar] [CrossRef]
  39. Strasser, E.M.; Wessner, B.; Manhart, N.; Roth, E. The relationship between the anti-inflammatory effects of curcumin and cellular glutathione content in myelomonocytic cells. Biochem. Pharmacol. 2005, 70, 552–559. [Google Scholar] [CrossRef] [PubMed]
  40. Moos, P.J.; Edes, K.; Mullally, J.E.; Fitzpatrick, F.A. Curcumin impairs tumor suppressor p53 function in colon cancer cells. Carcinogenesis 2004, 25, 1611–1617. [Google Scholar] [CrossRef] [Green Version]
  41. Sharma, R.A.; Euden, S.A.; Platton, S.L.; Cooke, D.N.; Shafayat, A.; Hewitt, H.R.; Marczylo, T.H.; Morgan, B.; Hemingway, D.; Plummer, S.M.; et al. Phase I clinical trial of oral curcumin: Biomarkers of systemic activity and compliance. Clin. Cancer Res. 2004, 10, 6847–6854. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Anand, P.; Kunnumakkara, A.B.; Newman, R.A.; Aggarwal, B.B. Bioavailability of curcumin: Problems and promises. Mol. Pharm. 2007, 4, 807–818. [Google Scholar] [CrossRef] [PubMed]
  43. Sharma, R.A.; Steward, W.P.; Gescher, A.J. Pharmacokinetics and pharmacodynamics of curcumin. Adv. Exp. Med. Biol. 2007, 595, 453–470. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. de Souza Ferreira, S.B.; Bruschi, M.L. Improving the bioavailability of curcumin: Is micro/nanoencapsulation the key? Ther. Deliv. 2019, 10, 83–86. [Google Scholar] [CrossRef] [PubMed]
  45. Kurien, B.T.; Singh, A.; Matsumoto, H.; Scofield, R.H. Improving the solubility and pharmacological efficacy of curcumin by heat treatment. Assay Drug Dev. Technol. 2007, 5, 567–576. [Google Scholar] [CrossRef]
  46. Priyadarsini, K.I. The chemistry of curcumin: From extraction to therapeutic agent. Molecules 2014, 19, 20091–20112. [Google Scholar] [CrossRef] [Green Version]
  47. Bisht, S.; Feldmann, G.; Soni, S.; Ravi, R.; Karikar, C.; Maitra, A.; Maitra, A. Polymeric nanoparticle-encapsulated curcumin (“nanocurcumin”): A novel strategy for human cancer therapy. J. Nanobiotechnol. 2007, 5, 3. [Google Scholar] [CrossRef] [Green Version]
  48. Li, L.; Braiteh, F.S.; Kurzrock, R. Liposome-encapsulated curcumin: In vitro and in vivo effects on proliferation, apoptosis, signaling, and angiogenesis. Cancer 2005, 104, 1322–1331. [Google Scholar] [CrossRef]
  49. Suresh, D.; Srinivasan, K. Studies on the in vitro absorption of spice principles--curcumin, capsaicin and piperine in rat intestines. Food Chem. Toxicol. 2007, 45, 1437–1442. [Google Scholar] [CrossRef]
  50. Kocher, A.; Bohnert, L.; Schiborr, C.; Frank, J. Highly bioavailable micellar curcuminoids accumulate in blood, are safe and do not reduce blood lipids and inflammation markers in moderately hyperlipidemic individuals. Mol. Nutr. Food Res. 2016, 60, 1555–1563. [Google Scholar] [CrossRef]
  51. Samuels, E.R.; Sevrioukova, I.F. Rational Design of CYP3A4 Inhibitors: A One-Atom Linker Elongation in Ritonavir-Like Compounds Leads to a Marked Improvement in the Binding Strength. Int. J. Mol. Sci. 2021, 22, 852. [Google Scholar] [CrossRef]
  52. Kim, Y.; Chen, J. Molecular structure of human P-glycoprotein in the ATP-bound, outward-facing conformation. Science 2018, 359, 915–919. [Google Scholar] [CrossRef] [Green Version]
  53. Ramos, A.I.; Braga, T.M.; Silva, P.; Fernandes, J.A.; Ribeiro-Claro, P.; Lopes, M.d.F.S.; Paz, F.A.A.; Braga, S.S. Chloramphenicol·cyclodextrin inclusion compounds: Co-dissolution and mechanochemical preparations and antibacterial action. CrystEngComm 2013, 15, 2822. [Google Scholar] [CrossRef]
  54. McClements, D.J. Delivery by Design (DbD): A Standardized Approach to the Development of Efficacious Nanoparticle- and Microparticle-Based Delivery Systems. Compr. Rev. Food Sci. Food Saf. 2018, 17, 200–219. [Google Scholar] [CrossRef] [Green Version]
  55. Douglass, B.J.; Clouatre, D.L. Beyond Yellow Curry: Assessing Commercial Curcumin Absorption Technologies. J. Am. Coll. Nutr. 2015, 34, 347–358. [Google Scholar] [CrossRef] [Green Version]
  56. Jamwal, R. Bioavailable curcumin formulations: A review of pharmacokinetic studies in healthy volunteers. J. Integr. Med. 2018, 16, 367–374. [Google Scholar] [CrossRef] [PubMed]
  57. Purpura, M.; Lowery, R.P.; Wilson, J.M.; Mannan, H.; Munch, G.; Razmovski-Naumovski, V. Analysis of different innovative formulations of curcumin for improved relative oral bioavailability in human subjects. Eur. J. Nutr. 2018, 57, 929–938. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Jager, R.; Lowery, R.P.; Calvanese, A.V.; Joy, J.M.; Purpura, M.; Wilson, J.M. Comparative absorption of curcumin formulations. Nutr. J. 2014, 13, 11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Morimoto, T.; Sunagawa, Y.; Katanasaka, Y.; Hirano, S.; Namiki, M.; Watanabe, Y.; Suzuki, H.; Doi, O.; Suzuki, K.; Yamauchi, M.; et al. Drinkable preparation of Theracurmin exhibits high absorption efficiency--a single-dose, double-blind, 4-way crossover study. Biol. Pharm. Bull. 2013, 36, 1708–1714. [Google Scholar] [CrossRef] [Green Version]
  60. Cuomo, J.; Appendino, G.; Dern, A.S.; Schneider, E.; McKinnon, T.P.; Brown, M.J.; Togni, S.; Dixon, B.M. Comparative absorption of a standardized curcuminoid mixture and its lecithin formulation. J. Nat. Prod. 2011, 74, 664–669. [Google Scholar] [CrossRef] [PubMed]
  61. Sasaki, H.; Sunagawa, Y.; Takahashi, K.; Imaizumi, A.; Fukuda, H.; Hashimoto, T.; Wada, H.; Katanasaka, Y.; Kakeya, H.; Fujita, M.; et al. Innovative preparation of curcumin for improved oral bioavailability. Biol. Pharm. Bull. 2011, 34, 660–665. [Google Scholar] [CrossRef] [Green Version]
  62. Schiborr, C.; Kocher, A.; Behnam, D.; Jandasek, J.; Toelstede, S.; Frank, J. The oral bioavailability of curcumin from micronized powder and liquid micelles is significantly increased in healthy humans and differs between sexes. Mol. Nutr. Food Res. 2014, 58, 516–527. [Google Scholar] [CrossRef] [PubMed]
  63. Stohs, S.J.; Ji, J.; Bucci, L.R.; Preuss, H.G. A Comparative Pharmacokinetic Assessment of a Novel Highly Bioavailable Curcumin Formulation with 95% Curcumin: A Randomized, Double-Blind, Crossover Study. J. Am. Coll. Nutr. 2018, 37, 51–59. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Stohs, S.J.; Chen, C.Y.O.; Preuss, H.G.; Ray, S.D.; Bucci, L.R.; Ji, J.; Ruff, K.J. The fallacy of enzymatic hydrolysis for the determination of bioactive curcumin in plasma samples as an indication of bioavailability: A comparative study. BMC Complement. Altern. Med. 2019, 19, 293. [Google Scholar] [CrossRef]
  65. Stohs, S.J.; Chen, O.; Ray, S.D.; Ji, J.; Bucci, L.R.; Preuss, H.G. Highly Bioavailable Forms of Curcumin and Promising Avenues for Curcumin-Based Research and Application: A Review. Molecules 2020, 25, 1397. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Kumar, D.; Jacob, D.; Ps, S.; Maliakkal, A.; Nm, J.; Kuttan, R.; Maliakel, B.; Konda, V.; Im, K. Enhanced bioavailability and relative distribution of free (unconjugated) curcuminoids following the oral administration of a food-grade formulation with fenugreek dietary fibre: A randomised double-blind crossover study. J. Funct. Foods 2016, 22, 578–587. [Google Scholar] [CrossRef]
  67. Bhardwaj, R.K.; Glaeser, H.; Becquemont, L.; Klotz, U.; Gupta, S.K.; Fromm, M.F. Piperine, a major constituent of black pepper, inhibits human P-glycoprotein and CYP3A4. J. Pharmacol. Exp. Ther. 2002, 302, 645–650. [Google Scholar] [CrossRef] [Green Version]
  68. Srinivasan, K. Black pepper and its pungent principle-piperine: A review of diverse physiological effects. Crit. Rev. Food Sci. Nutr. 2007, 47, 735–748. [Google Scholar] [CrossRef]
  69. Shoba, G.; Joy, D.; Joseph, T.; Majeed, M.; Rajendran, R.; Srinivas, P.S. Influence of piperine on the pharmacokinetics of curcumin in animals and human volunteers. Planta Med. 1998, 64, 353–356. [Google Scholar] [CrossRef] [Green Version]
  70. Majeed, M.; Badmaev, V.; Rajendran, R. Use of Piperine to Increase the Bioavailability of Nutritional Compounds. U.S. Patent 5,536,506, 16 July 1996. [Google Scholar]
  71. Vijaya, N. Dietary Nutritional Supplements for Healthcare. U.S. Patent 2009/0004334 A1, 1 January 2009. [Google Scholar]
  72. Klickovic, U.; Doberer, D.; Gouya, G.; Aschauer, S.; Weisshaar, S.; Storka, A.; Bilban, M.; Wolzt, M. Human pharmacokinetics of high dose oral curcumin and its effect on heme oxygenase-1 expression in healthy male subjects. BioMed Res. Int. 2014, 2014, 458592. [Google Scholar] [CrossRef]
  73. Panahi, Y.; Valizadegan, G.; Ahamdi, N.; Ganjali, S.; Majeed, M.; Sahebkar, A. Curcuminoids plus piperine improve nonalcoholic fatty liver disease: A clinical trial. J. Cell. Biochem. 2019, 120, 15989–15996. [Google Scholar] [CrossRef]
  74. Saberi-Karimian, M.; Keshvari, M.; Ghayour-Mobarhan, M.; Salehizadeh, L.; Rahmani, S.; Behnam, B.; Jamialahmadi, T.; Asgary, S.; Sahebkar, A. Effects of curcuminoids on inflammatory status in patients with non-alcoholic fatty liver disease: A randomized controlled trial. Complement. Ther. Med. 2020, 49, 102322. [Google Scholar] [CrossRef]
  75. Yue, G.G.; Cheng, S.W.; Yu, H.; Xu, Z.S.; Lee, J.K.; Hon, P.M.; Lee, M.Y.; Kennelly, E.J.; Deng, G.; Yeung, S.K.; et al. The role of turmerones on curcumin transportation and P-glycoprotein activities in intestinal Caco-2 cells. J. Med. Food 2012, 15, 242–252. [Google Scholar] [CrossRef] [PubMed]
  76. Antony, B. Composition to Enhance the Bioavailability of Curcumin. U.S. Patent 7,879,373 B2, 1 February 2011. [Google Scholar]
  77. Antony, B.; Merina, B.; Iyer, V.S.; Judy, N.; Lennertz, K.; Joyal, S. A Pilot Cross-Over Study to Evaluate Human Oral Bioavailability of BCM-95CG (Biocurcumax), A Novel Bioenhanced Preparation of Curcumin. Indian J. Pharm. Sci. 2008, 70, 445–449. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Dei Cas, M.; Ghidoni, R. Dietary Curcumin: Correlation between Bioavailability and Health Potential. Nutrients 2019, 11, 2147. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Singhal, S.; Hasan, N.; Nirmal, K.; Chawla, R.; Chawla, S.; Kalra, B.S.; Dhal, A. Bioavailable turmeric extract for knee osteoarthritis: A randomized, non-inferiority trial versus paracetamol. Trials 2021, 22, 105. [Google Scholar] [CrossRef] [PubMed]
  80. Shep, D.; Khanwelkar, C.; Gade, P.; Karad, S. Efficacy and safety of combination of curcuminoid complex and diclofenac versus diclofenac in knee osteoarthritis: A randomized trial. Medicine 2020, 99, e19723. [Google Scholar] [CrossRef] [PubMed]
  81. Kocher, A.; Schiborr, C.; Behnam, D.; Frank, J. The oral bioavailability of curcuminoids in healthy humans is markedly enhanced by micellar solubilisation but not further improved by simultaneous ingestion of sesamin, ferulic acid, naringenin and xanthohumol. J. Funct. Foods 2015, 14, 183–191. [Google Scholar] [CrossRef] [Green Version]
  82. Pawar, Y.B.; Munjal, B.; Arora, S.; Karwa, M.; Kohli, G.; Paliwal, J.K.; Bansal, A.K. Bioavailability of a lipidic formulation of curcumin in healthy human volunteers. Pharmaceutics 2012, 4, 517–530. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Gopi, S.; George, R.; Thomas, M.; Jude, S. A pilot cross-over study to assess the human bioavailability of“Cureit” a bioavailable curcumin in complete natural matrix. Asian J. Pharm. Technol. Innov. 2015, 3, 92–96. [Google Scholar]
  84. Gopi, S.; Jacob, J.; Varma, K.; Jude, S.; Amalraj, A.; Arundhathy, C.A.; George, R.; Sreeraj, T.R.; Divya, C.; Kunnumakkara, A.B.; et al. Comparative Oral Absorption of Curcumin in a Natural Turmeric Matrix with Two Other Curcumin Formulations: An Open-label Parallel-arm Study. Phytother. Res. 2017, 31, 1883–1891. [Google Scholar] [CrossRef] [PubMed]
  85. Gota, V.S.; Maru, G.B.; Soni, T.G.; Gandhi, T.R.; Kochar, N.; Agarwal, M.G. Safety and pharmacokinetics of a solid lipid curcumin particle formulation in osteosarcoma patients and healthy volunteers. J. Agric. Food Chem. 2010, 58, 2095–2099. [Google Scholar] [CrossRef]
  86. Ipar, V.S.; Dsouza, A.; Devarajan, P.V. Enhancing Curcumin Oral Bioavailability Through Nanoformulations. Eur. J. Drug Metab. Pharmacokinet. 2019, 44, 459–480. [Google Scholar] [CrossRef] [PubMed]
  87. Behnam, D. Curcumin Solubilsate. U.S. Patent Application No. US2015/0342881 A1, 3 December 2015. [Google Scholar]
  88. Dutzmann, S.; Schiborr, C.; Kocher, A.; Pilatus, U.; Hattingen, E.; Weissenberger, J.; Gessler, F.; Quick-Weller, J.; Franz, K.; Seifert, V.; et al. Intratumoral Concentrations and Effects of Orally Administered Micellar Curcuminoids in Glioblastoma Patients. Nutr. Cancer 2016, 68, 943–948. [Google Scholar] [CrossRef] [PubMed]
  89. Simsek, S.; Simsek, S.A. A Pharmaceutical Composition with Enhanced Bioavailability Containing a Medical Herbal Extract. European Patent EP3581172A1, 18 December 2019. [Google Scholar]
  90. Jayantven Katesh, D.; Shrinivas Krishnarao, K. A Water Soluble Composition Comprising Curcumin Having Enhanced Bioavailability and Process Thereof. WO Patent WO2012156979A1, 22 November 2012. [Google Scholar]
  91. Campbell, M.S.; Ouyang, A.; Krishnakumar, I.M.; Charnigo, R.J.; Westgate, P.M.; Fleenor, B.S. Influence of enhanced bioavailable curcumin on obesity-associated cardiovascular disease risk factors and arterial function: A double-blinded, randomized, controlled trial. Nutrition 2019, 62, 135–139. [Google Scholar] [CrossRef]
  92. Campbell, M.S.; Berrones, A.J.; Krishnakumar, I.M.; Charnigo, R.J.; Westgate, P.M.; Fleenor, B.S. Responsiveness to curcumin intervention is associated with reduced aortic stiffness in young, obese men with higher initial stiffness. J. Funct. Foods 2017, 29, 154–160. [Google Scholar] [CrossRef]
  93. Fahrner, R.L. Compositions and Methods for Increasing the Bioavailability of One or More Compounds. WO Patent WO2016057839A1, 14 April 2016. [Google Scholar]
  94. Takeshi, M.; Takashi, M.; Masayuki, N.; Taiji, S.; Yoshiharu, T. Turmeric Dye Composition and Preparation Method Thereof. JP Patent JP5448511B2, 19 March 2014. [Google Scholar]
  95. Funamoto, M.; Shimizu, K.; Sunagawa, Y.; Katanasaka, Y.; Miyazaki, Y.; Kakeya, H.; Yamakage, H.; Satoh-Asahara, N.; Wada, H.; Hasegawa, K.; et al. Effects of Highly Absorbable Curcumin in Patients with Impaired Glucose Tolerance and Non-Insulin-Dependent Diabetes Mellitus. J. Diabetes Res. 2019, 2019, 8208237. [Google Scholar] [CrossRef] [PubMed]
  96. Nakagawa, Y.; Mukai, S.; Yamada, S.; Murata, S.; Yabumoto, H.; Maeda, T.; Akamatsu, S. The Efficacy and Safety of Highly-Bioavailable Curcumin for Treating Knee Osteoarthritis: A 6-Month Open-Labeled Prospective Study. Clin. Med. Insights Arthritis Musculoskelet. Disord. 2020, 13, 1179544120948471. [Google Scholar] [CrossRef] [PubMed]
  97. Gopi, S. A Novel Composition of Curcumin with Enhanced Bioavailability. U.S. Patent US20160151440 A1, 2 June 2016. [Google Scholar]
  98. Amalraj, A.; Jude, S.; Varma, K.; Jacob, J.; Gopi, S.; Oluwafemi, O.S.; Thomas, S. Preparation of a novel bioavailable curcuminoid formulation (Cureit) using Polar-Nonpolar-Sandwich (PNS) technology and its characterization and applications. Mater. Sci. Eng. C Mater. Biol. Appl. 2017, 75, 359–367. [Google Scholar] [CrossRef]
  99. Ahmed Nasef, N.; Loveday, S.M.; Golding, M.; Martins, R.N.; Shah, T.M.; Clarke, M.; Coad, J.; Moughan, P.J.; Garg, M.L.; Singh, H. Food matrix and co-presence of turmeric compounds influence bioavailability of curcumin in healthy humans. Food Funct. 2019, 10, 4584–4592. [Google Scholar] [CrossRef]
  100. Amalraj, A.; Divya, C.; Gopi, S. The Effects of Bioavailable Curcumin (Cureit) on Delayed Onset Muscle Soreness Induced By Eccentric Continuous Exercise: A Randomized, Placebo-Controlled, Double-Blind Clinical Study. J. Med. Food 2020, 23, 545–553. [Google Scholar] [CrossRef] [PubMed]
  101. Giori, A.; Franceschi, F. Phospholipid Complexes of Curcumin Having Improved Bioavailability. WO Patent WO2007101551A2, 13 September 2007. [Google Scholar]
  102. Belcaro, G.; Dugall, M.; Luzzi, R.; Ledda, A.; Pellegrini, L.; Cesarone, M.R.; Hosoi, M.; Errichi, M. Meriva(R)+Glucosamine versus Condroitin+Glucosamine in patients with knee osteoarthritis: An observational study. Eur. Rev. Med. Pharmacol. Sci. 2014, 18, 3959–3963. [Google Scholar] [PubMed]
  103. Di Pierro, F.; Zacconi, P.; Bertuccioli, A.; Togni, S.; Eggenhoffner, R.; Giacomelli, L.; Scaltrini, S. A naturally-inspired, curcumin-based lecithin formulation (Meriva(R) formulated as the finished product Algocur(R)) alleviates the osteo-muscular pain conditions in rugby players. Eur. Rev. Med. Pharmacol. Sci. 2017, 21, 4935–4940. [Google Scholar]
  104. Mazzolani, F.; Togni, S.; Giacomelli, L.; Eggenhoffner, R.; Franceschi, F. Oral administration of a curcumin-phospholipid formulation (Meriva(R)) for treatment of chronic diabetic macular edema: A pilot study. Eur. Rev. Med. Pharmacol. Sci. 2018, 22, 3617–3625. [Google Scholar] [CrossRef] [PubMed]
  105. Pastorelli, D.; Fabricio, A.S.C.; Giovanis, P.; D’Ippolito, S.; Fiduccia, P.; Solda, C.; Buda, A.; Sperti, C.; Bardini, R.; Da Dalt, G.; et al. Phytosome complex of curcumin as complementary therapy of advanced pancreatic cancer improves safety and efficacy of gemcitabine: Results of a prospective phase II trial. Pharmacol. Res. 2018, 132, 72–79. [Google Scholar] [CrossRef] [PubMed]
  106. Kharat, M.; McClements, D.J. Recent advances in colloidal delivery systems for nutraceuticals: A case study—Delivery by Design of curcumin. J. Colloid Interface Sci. 2019, 557, 506–518. [Google Scholar] [CrossRef] [PubMed]
  107. Frautschy, S.A.; Cole, G.M. Bioavailable Curcuminoid Formulations for Treating Alzheimer’s Disease and Other Age-Related Disorders. WO Patent WO2007103435A2, 25 October 2007. [Google Scholar]
  108. Esatbeyoglu, T.; Huebbe, P.; Ernst, I.M.; Chin, D.; Wagner, A.E.; Rimbach, G. Curcumin—From molecule to biological function. Angew. Chem. Int. Ed. Engl. 2012, 51, 5308–5332. [Google Scholar] [CrossRef] [PubMed]
  109. McClements, D.J. Food Nanotechnology: Harnessing the Power of the Miniature World Inside Our Foods. In Future Foods; Copernicus: Cham, Switzerland, 2019; pp. 287–321. [Google Scholar] [CrossRef]
  110. Pandey, P.; Dahiya, M. A brief review on inorganic nanoparticles. J. Crit. Rev. 2016, 3, 18–26. [Google Scholar]
  111. Zheng, B.; McClements, D.J. Formulation of More Efficacious Curcumin Delivery Systems Using Colloid Science: Enhanced Solubility, Stability, and Bioavailability. Molecules 2020, 25, 2791. [Google Scholar] [CrossRef]
  112. Zou, L.; Liu, W.; Liu, C.; Xiao, H.; McClements, D.J. Utilizing food matrix effects to enhance nutraceutical bioavailability: Increase of curcumin bioaccessibility using excipient emulsions. J. Agric. Food Chem. 2015, 63, 2052–2062. [Google Scholar] [CrossRef]
  113. Zou, L.; Zheng, B.; Liu, W.; Liu, C.; Xiao, H.; McClements, D.J. Enhancing nutraceutical bioavailability using excipient emulsions: Influence of lipid droplet size on solubility and bioaccessibility of powdered curcumin. J. Funct. Foods 2015, 15, 72–83. [Google Scholar] [CrossRef]
  114. Li, G.; Zhang, Z.; Liu, H.; Hu, L. Nanoemulsion-based delivery approaches for nutraceuticals: Fabrication, application, characterization, biological fate, potential toxicity and future trends. Food Funct. 2021, 12, 1933–1953. [Google Scholar] [CrossRef] [PubMed]
  115. McClements, D.J.; Gumus, C.E. Natural emulsifiers—Biosurfactants, phospholipids, biopolymers, and colloidal particles: Molecular and physicochemical basis of functional performance. Adv. Colloid Interface Sci. 2016, 234, 3–26. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Ma, P.; Zeng, Q.; Tai, K.; He, X.; Yao, Y.; Hong, X.; Yuan, F. Preparation of curcumin-loaded emulsion using high pressure homogenization: Impact of oil phase and concentration on physicochemical stability. LWT 2017, 84, 34–46. [Google Scholar] [CrossRef]
  117. Ahmed, K.; Li, Y.; McClements, D.J.; Xiao, H. Nanoemulsion- and emulsion-based delivery systems for curcumin: Encapsulation and release properties. Food Chem. 2012, 132, 799–807. [Google Scholar] [CrossRef]
  118. Yu, H.; Huang, Q. Improving the oral bioavailability of curcumin using novel organogel-based nanoemulsions. J. Agric. Food Chem. 2012, 60, 5373–5379. [Google Scholar] [CrossRef] [PubMed]
  119. Cheng, C.; Peng, S.; Li, Z.; Zou, L.; Liu, W.; Liu, C. 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]
  120. Matloob, A.H.; Mourtas, S.; Klepetsanis, P.; Antimisiaris, S.G. Increasing the stability of curcumin in serum with liposomes or hybrid drug-in-cyclodextrin-in-liposome systems: A comparative study. Int. J. Pharm. 2014, 476, 108–115. [Google Scholar] [CrossRef]
  121. 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]
  122. Kakkar, V.; Singh, S.; Singla, D.; Kaur, I.P. Exploring solid lipid nanoparticles to enhance the oral bioavailability of curcumin. Mol. Nutr. Food Res. 2011, 55, 495–503. [Google Scholar] [CrossRef]
  123. Yallapu, M.M.; Gupta, B.K.; Jaggi, M.; Chauhan, S.C. Fabrication of curcumin encapsulated PLGA nanoparticles for improved therapeutic effects in metastatic cancer cells. J. Colloid Interface Sci. 2010, 351, 19–29. [Google Scholar] [CrossRef]
  124. Umerska, A.; Gaucher, C.; Oyarzun-Ampuero, F.; Fries-Raeth, I.; Colin, F.; Villamizar-Sarmiento, M.G.; Maincent, P.; Sapin-Minet, A. Polymeric Nanoparticles for Increasing Oral Bioavailability of Curcumin. Antioxidants 2018, 7, 46. [Google Scholar] [CrossRef] [Green Version]
  125. Tajbakhsh, A.; Hasanzadeh, M.; Rezaee, M.; Khedri, M.; Khazaei, M.; ShahidSales, S.; Ferns, G.A.; Hassanian, S.M.; Avan, A. Therapeutic potential of novel formulated forms of curcumin in the treatment of breast cancer by the targeting of cellular and physiological dysregulated pathways. J. Cell. Physiol. 2018, 233, 2183–2192. [Google Scholar] [CrossRef]
  126. Yadav, P.; Bandyopadhyay, A.; Chakraborty, A.; Sarkar, K. Enhancement of anticancer activity and drug delivery of chitosan-curcumin nanoparticle via molecular docking and simulation analysis. Carbohydr. Polym. 2018, 182, 188–198. [Google Scholar] [CrossRef] [PubMed]
  127. Chauhan, P.; Tamrakar, A.K.; Mahajan, S.; Prasad, G. Chitosan encapsulated nanocurcumin induces GLUT-4 translocation and exhibits enhanced anti-hyperglycemic function. Life Sci. 2018, 213, 226–235. [Google Scholar] [CrossRef] [PubMed]
  128. Khan, M.A.; Zafaryab, M.; Mehdi, S.H.; Ahmad, I.; Rizvi, M.M.A. Physicochemical Characterization of Curcumin Loaded Chitosan Nanoparticles: Implications in Cervical Cancer. Anticancer Agents Med. Chem. 2018, 18, 1131–1137. [Google Scholar] [CrossRef] [PubMed]
  129. Shelma, R.; Sharma, C.P. In vitro and in vivo evaluation of curcumin loaded lauroyl sulphated chitosan for enhancing oral bioavailability. Carbohydr. Polym. 2013, 95, 441–448. [Google Scholar] [CrossRef]
  130. 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 2011, 44, 2166–2172. [Google Scholar] [CrossRef]
  131. Ye, Q.; Ge, F.; Wang, Y.; Woo, M.W.; Wu, P.; Chen, X.D.; Selomulya, C. On improving bioaccessibility and targeted release of curcumin-whey protein complex microparticles in food. Food Chem. 2021, 346, 128900. [Google Scholar] [CrossRef]
  132. Dantas Lopes Dos Santos, D.; Besegato, J.F.; de Melo, P.B.G.; Oshiro Junior, J.A.; Chorilli, M.; Deng, D.; Bagnato, V.S.; Rastelli, A.N.S. Curcumin-loaded Pluronic((R)) F-127 Micelles as a Drug Delivery System for Curcumin-mediated Photodynamic Therapy for Oral Application. Photochem. Photobiol. 2021. [Google Scholar] [CrossRef]
  133. Zheng, B.; Zhang, X.; Lin, H.; McClements, D.J. Loading natural emulsions with nutraceuticals using the pH-driven method: Formation & stability of curcumin-loaded soybean oil bodies. Food Funct. 2019, 10, 5473–5484. [Google Scholar] [CrossRef]
  134. Mills, J.K.; Needham, D. Targeted drug delivery. Expert Opin. Ther. Pat. 2005, 9, 1499–1513. [Google Scholar] [CrossRef]
  135. Pandey, A.; Mishra, S.; Tiwari, A.; Misra, K. Targeted drug delivery (Site specific drug delivery). J. Sci. Ind. Res. 2004, 63, 230–247. [Google Scholar]
  136. Managuli, R.S.; Raut, S.Y.; Reddy, M.S.; Mutalik, S. Targeting the intestinal lymphatic system: A versatile path for enhanced oral bioavailability of drugs. Expert Opin. Drug Deliv. 2018, 15, 787–804. [Google Scholar] [CrossRef]
  137. Patel, R.P.; Shah, P.; Barve, K.; Patel, N.; Gandhi, J. Peyer’s Patch: Targeted Drug Delivery for Therapeutics Benefits. In Novel Drug Delivery Technologies; Springer: New York, NY, USA, 2019; pp. 121–149. [Google Scholar] [CrossRef]
  138. Bachhav, S.S.; Dighe, V.D.; Devarajan, P.V. Exploring Peyer’s Patch Uptake as a Strategy for Targeted Lung Delivery of Polymeric Rifampicin Nanoparticles. Mol. Pharm. 2018, 15, 4434–4445. [Google Scholar] [CrossRef]
  139. Joshi, G.; Kumar, A.; Sawant, K. Enhanced bioavailability and intestinal uptake of Gemcitabine HCl loaded PLGA nanoparticles after oral delivery. Eur. J. Pharm. Sci. 2014, 60, 80–89. [Google Scholar] [CrossRef]
  140. van der Lubben, I. Chitosan microparticles for oral vaccination: Preparation, characterization and preliminary in vivo uptake studies in murine Peyer’s patches. Biomaterials 2001, 22, 687–694. [Google Scholar] [CrossRef]
  141. Khalil, N.M.; do Nascimento, T.C.; Casa, D.M.; Dalmolin, L.F.; de Mattos, A.C.; Hoss, I.; Romano, M.A.; Mainardes, R.M. Pharmacokinetics of curcumin-loaded PLGA and PLGA-PEG blend nanoparticles after oral administration in rats. Colloids Surf. B Biointerfaces 2013, 101, 353–360. [Google Scholar] [CrossRef]
  142. Gamage, N.H.; Jing, L.; Worsham, M.J.; Ali, M.M. Targeted Theranostic Approach for Glioma Using Dendrimer-Based Curcumin Nanoparticle. J. Nanomed. Nanotechnol. 2016, 7, 393. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Debnath, S.; Saloum, D.; Dolai, S.; Sun, C.; Averick, S.; Raja, K.; Fata, J.E. Dendrimer-curcumin conjugate: A water soluble and effective cytotoxic agent against breast cancer cell lines. Anticancer Agents Med. Chem. 2013, 13, 1531–1539. [Google Scholar] [CrossRef]
  144. Esmatabadi, M.J.D.; Motamedrad, M.; Sadeghizadeh, M. Down-regulation of lncRNA, GAS5 decreases chemotherapeutic effect of dendrosomal curcumin (DNC) in breast cancer cells. Phytomedicine 2018, 42, 56–65. [Google Scholar] [CrossRef]
  145. Mollazade, M.; Nejati-Koshki, K.; Akbarzadeh, A.; Zarghami, N.; Nasiri, M.; Jahanban-Esfahlan, R.; Alibakhshi, A. PAMAM dendrimers augment inhibitory effects of curcumin on cancer cell proliferation: Possible inhibition of telomerase. Asian Pac. J. Cancer Prev. 2013, 14, 6925–6928. [Google Scholar] [CrossRef] [Green Version]
  146. Mollazade, M.; Zarghami, N.; Nasiri, M.; Nejati, K.; Rahmati, M.; Pourhasan, M. Polyamidoamine (PAMAM) encapsulated curcumin inhibits telomerase activity in breast cancer cell line. Clin. Biochem. 2011, 44, S217. [Google Scholar] [CrossRef]
  147. Mande, P.P.; Bachhav, S.S.; Devarajan, P.V. Solid Dispersion of Curcumin as Polymeric Films for Bioenhancement and Improved Therapy of Rheumatoid Arthritis. Pharm. Res. 2016, 33, 1972–1987. [Google Scholar] [CrossRef] [PubMed]
  148. Cui, J.; Yu, B.; Zhao, Y.; Zhu, W.; Li, H.; Lou, H.; Zhai, G. Enhancement of oral absorption of curcumin by self-microemulsifying drug delivery systems. Int. J. Pharm. 2009, 371, 148–155. [Google Scholar] [CrossRef] [PubMed]
  149. Setthacheewakul, S.; Mahattanadul, S.; Phadoongsombut, N.; Pichayakorn, W.; Wiwattanapatapee, R. Development and evaluation of self-microemulsifying liquid and pellet formulations of curcumin, and absorption studies in rats. Eur. J. Pharm. Biopharm. 2010, 76, 475–485. [Google Scholar] [CrossRef]
  150. Wu, X.; Xu, J.; Huang, X.; Wen, C. Self-microemulsifying drug delivery system improves curcumin dissolution and bioavailability. Drug Dev. Ind. Pharm. 2011, 37, 15–23. [Google Scholar] [CrossRef] [PubMed]
  151. Yan, Y.-D.; Kim, J.A.; Kwak, M.K.; Yoo, B.K.; Yong, C.S.; Choi, H.-G. Enhanced Oral Bioavailability of Curcumin via a Solid Lipid-Based Self-Emulsifying Drug Delivery System Using a Spray-Drying Technique. Biol. Pharm. Bull. 2011, 34, 1179–1186. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Zhang, J.; Chen, L.; Tse, W.H.; Bi, R.; Chen, L. Inorganic Nanoparticles: Engineering for Biomedical Applications. IEEE Nanotechnol. Mag. 2014, 8, 21–28. [Google Scholar] [CrossRef]
  153. Meena, J.; Gupta, A.; Ahuja, R.; Singh, M.; Bhaskar, S.; Panda, A.K. Inorganic nanoparticles for natural product delivery: A review. Environ. Chem. Lett. 2020, 18, 2107–2118. [Google Scholar] [CrossRef]
  154. Liu, Y.; Liu, Q.; Liu, Y.; Ju, F.; Ma, Q.; He, Q. In vivo evaluation of enhanced drug carrier efficiency and cardiac anti-hypertrophy therapeutic potential of nano-curcumin encapsulated photo-plasmonic nanoparticles combined polymerized nano-vesicles: A novel strategy. J. Photochem. Photobiol. B 2019, 199, 111619. [Google Scholar] [CrossRef]
  155. Sanoj Rejinold, N.; Thomas, R.G.; Muthiah, M.; Chennazhi, K.P.; Manzoor, K.; Park, I.K.; Jeong, Y.Y.; Jayakumar, R. Anti-cancer, pharmacokinetics and tumor localization studies of pH-, RF- and thermo-responsive nanoparticles. Int. J. Biol. Macromol. 2015, 74, 249–262. [Google Scholar] [CrossRef]
  156. Nambiar, S.; Osei, E.; Fleck, A.; Darko, J.; Mutsaers, A.J.; Wettig, S. Synthesis of curcumin-functionalized gold nanoparticles and cytotoxicity studies in human prostate cancer cell line. Appl. Nanosci. 2018, 8, 347–357. [Google Scholar] [CrossRef]
  157. 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]
  158. Kondath, S.; Rajaram, R.; Anantanarayanan, R. Curcumin reduced gold nanoparticles synergistically induces ROS mediated apoptosis in MCF-7 cancer cells. Inorg. Nano-Met. Chem. 2021, 51, 601–613. [Google Scholar] [CrossRef]
  159. Jaiswal, S.; Mishra, P. Antimicrobial and antibiofilm activity of curcumin-silver nanoparticles with improved stability and selective toxicity to bacteria over mammalian cells. Med. Microbiol. Immunol. 2018, 207, 39–53. [Google Scholar] [CrossRef]
  160. Yallapu, M.M.; Ebeling, M.C.; Khan, S.; Sundram, V.; Chauhan, N.; Gupta, B.K.; Puumala, S.E.; Jaggi, M.; Chauhan, S.C. Novel curcumin-loaded magnetic nanoparticles for pancreatic cancer treatment. Mol. Cancer Ther. 2013, 12, 1471–1480. [Google Scholar] [CrossRef] [Green Version]
  161. Aeineh, N.; Salehi, F.; Akrami, M.; Nemati, F.; Alipour, M.; Ghorbani, M.; Nikfar, B.; Salehian, F.; Riyahi Alam, N.; Sadat Ebrahimi, S.E.; et al. Glutathione conjugated polyethylenimine on the surface of Fe3O4 magnetic nanoparticles as a theranostic agent for targeted and controlled curcumin delivery. J. Biomater. Sci. Polym. Ed. 2018, 29, 1109–1125. [Google Scholar] [CrossRef]
  162. Goncalves, M.C. Sol-gel Silica Nanoparticles in Medicine: A Natural Choice. Design, Synthesis and Products. Molecules 2018, 23, 2021. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Kasaai, M.R. Nanosized Particles of Silica and Its Derivatives for Applications in Various Branches of Food and Nutrition Sectors. J. Nanotechnol. 2015, 2015, 852394. [Google Scholar] [CrossRef] [Green Version]
  164. Prabhuraj, R.S.; Mal, A.; Valvi, S.K.; Srivastava, R.; De, A.; Bandyopadhyaya, R. Noninvasive Preclinical Evaluation of Targeted Nanoparticles for the Delivery of Curcumin in Treating Pancreatic Cancer. ACS Appl. Bio Mater. 2020, 3, 4643–4654. [Google Scholar] [CrossRef]
  165. Flandroy, L.; Poutahidis, T.; Berg, G.; Clarke, G.; Dao, M.C.; Decaestecker, E.; Furman, E.; Haahtela, T.; Massart, S.; Plovier, H.; et al. The impact of human activities and lifestyles on the interlinked microbiota and health of humans and of ecosystems. Sci. Total Environ. 2018, 627, 1018–1038. [Google Scholar] [CrossRef]
  166. Scazzocchio, B.; Minghetti, L.; D’Archivio, M. Interaction between Gut Microbiota and Curcumin: A New Key of Understanding for the Health Effects of Curcumin. Nutrients 2020, 12, 2499. [Google Scholar] [CrossRef] [PubMed]
  167. Di Meo, F.; Margarucci, S.; Galderisi, U.; Crispi, S.; Peluso, G. Curcumin, Gut Microbiota, and Neuroprotection. Nutrients 2019, 11, 2426. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Peterson, C.T.; Vaughn, A.R.; Sharma, V.; Chopra, D.; Mills, P.J.; Peterson, S.N.; Sivamani, R.K. Effects of Turmeric and Curcumin Dietary Supplementation on Human Gut Microbiota: A Double-Blind, Randomized, Placebo-Controlled Pilot Study. J. Evid. Based Integr. Med. 2018, 23, 2515690X18790725. [Google Scholar] [CrossRef] [PubMed]
  169. Chashmniam, S.; Mirhafez, S.R.; Dehabeh, M.; Hariri, M.; Azimi Nezhad, M.; Nobakht, M.G.B.F. A pilot study of the effect of phospholipid curcumin on serum metabolomic profile in patients with non-alcoholic fatty liver disease: A randomized, double-blind, placebo-controlled trial. Eur. J. Clin. Nutr. 2019, 73, 1224–1235. [Google Scholar] [CrossRef]
  170. Peron, G.; Sut, S.; Dal Ben, S.; Voinovich, D.; Dall’Acqua, S. Untargeted UPLC-MS metabolomics reveals multiple changes of urine composition in healthy adult volunteers after consumption of curcuma longa L. extract. Food Res. Int. 2020, 127, 108730. [Google Scholar] [CrossRef]
  171. Edwards, R.L.; Luis, P.B.; Varuzza, P.V.; Joseph, A.I.; Presley, S.H.; Chaturvedi, R.; Schneider, C. The anti-inflammatory activity of curcumin is mediated by its oxidative metabolites. J. Biol. Chem. 2017, 292, 21243–21252. [Google Scholar] [CrossRef] [Green Version]
  172. Hassaninasab, A.; Hashimoto, Y.; Tomita-Yokotani, K.; Kobayashi, M. Discovery of the curcumin metabolic pathway involving a unique enzyme in an intestinal microorganism. Proc. Natl. Acad. Sci. USA 2011, 108, 6615–6620. [Google Scholar] [CrossRef] [Green Version]
  173. Tsuda, T. Curcumin as a functional food-derived factor: Degradation products, metabolites, bioactivity, and future perspectives. Food Funct. 2018, 9, 705–714. [Google Scholar] [CrossRef]
  174. Hou, H.T.; Qiu, Y.M.; Zhao, H.W.; Li, D.H.; Liu, Y.T.; Wang, Y.Z.; Su, S.H. Effect of curcumin on intestinal mucosal mechanical barrier in rats with non-alcoholic fatty liver disease. Zhonghua Gan Zang Bing Za Zhi 2017, 25, 134–138. [Google Scholar] [CrossRef] [PubMed]
  175. Wang, J.; Ghosh, S.S.; Ghosh, S. Curcumin improves intestinal barrier function: Modulation of intracellular signaling, and organization of tight junctions. Am. J. Physiol. Cell Physiol. 2017, 312, C438–C445. [Google Scholar] [CrossRef] [PubMed]
  176. McClements, D.J.; Rao, J. Food-grade nanoemulsions: Formulation, fabrication, properties, performance, biological fate, and potential toxicity. Crit. Rev. Food Sci. Nutr. 2011, 51, 285–330. [Google Scholar] [CrossRef]
  177. Xu, G.; Wang, C.; Yao, P. Stable emulsion produced from casein and soy polysaccharide compacted complex for protection and oral delivery of curcumin. Food Hydrocoll. 2017, 71, 108–117. [Google Scholar] [CrossRef]
  178. Wu, M.-H.; Yan, H.H.; Chen, Z.-Q.; He, M. Effects of emulsifier type and environmental stress on the stability of curcumin emulsion. J. Dispers. Sci. Technol. 2016, 38, 1375–1380. [Google Scholar] [CrossRef]
  179. Zheng, B.; Peng, S.; Zhang, X.; McClements, D.J. Impact of Delivery System Type on Curcumin Bioaccessibility: Comparison of Curcumin-Loaded Nanoemulsions with Commercial Curcumin Supplements. J. Agric. Food Chem. 2018, 66, 10816–10826. [Google Scholar] [CrossRef] [PubMed]
  180. Wang, Y.; Lu, Z.; Lv, F.; Bie, X. Study on microencapsulation of curcumin pigments by spray drying. Eur. Food Res. Technol. 2009, 229, 391–396. [Google Scholar] [CrossRef]
  181. Tiruwa, R. A review on nanoparticles—Preparation and evaluation parameters. Indian J. Pharm. Biol. Res. 2016, 4, 27–31. [Google Scholar] [CrossRef]
  182. Nagavarma, B.; Yadav, H.; Ayaz, A.; Vasudha, L.; Shivakumar, H. Different techniques for preparation of polymeric nanoparticles a review. Asian J. Pharm. Clin. Res. 2012, 5, 16–23. [Google Scholar]
  183. Mallakpour, S.; Behranvand, V. Polymeric nanoparticles: Recent development in synthesis and application. Express Polym. Lett. 2016, 10, 895–913. [Google Scholar] [CrossRef]
  184. Tsai, H.C.; Imae, T. Fabrication of dendrimers toward biological application. Prog. Mol. Biol. Transl. Sci. 2011, 104, 101–140. [Google Scholar] [CrossRef] [PubMed]
  185. Svenson, S. Dendrimers as versatile platform in drug delivery applications. Eur. J. Pharm. Biopharm. 2009, 71, 445–462. [Google Scholar] [CrossRef]
  186. Dufes, C.; Uchegbu, I.F.; Schatzlein, A.G. Dendrimers in gene delivery. Adv. Drug Deliv. Rev. 2005, 57, 2177–2202. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Joshi, N.; Grinstaff, M. Applications of dendrimers in tissue engineering. Curr. Top. Med. Chem. 2008, 8, 1225–1236. [Google Scholar] [CrossRef]
  188. Kobayashi, H.; Brechbiel, M.W. Dendrimer-based nanosized MRI contrast agents. Curr. Pharm. Biotechnol. 2004, 5, 539–549. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic representation of the main strategies for improving the bioavailability of curcumin. CYP3A4 and P-glycoprotein were downloaded from protein data bank (PDB https://www.rcsb.org/; PDB IDs 7KVS [51] and 6C0V [52], respectively). Cyclodextrin structure was downloaded from The Cambridge Crystallographic Data Centre (CCDC https://www.ccdc.cam.ac.uk/; ID 762697; WEWTOJ) [53]. Structures were manipulated by PyMOL software (The PyMOL Molecular Graphics System, v1.8; Schrödinger, LLC, New York, 2015). Au–curcumin nanoparticle was designed according to Priyadarsini [46].
Figure 1. Schematic representation of the main strategies for improving the bioavailability of curcumin. CYP3A4 and P-glycoprotein were downloaded from protein data bank (PDB https://www.rcsb.org/; PDB IDs 7KVS [51] and 6C0V [52], respectively). Cyclodextrin structure was downloaded from The Cambridge Crystallographic Data Centre (CCDC https://www.ccdc.cam.ac.uk/; ID 762697; WEWTOJ) [53]. Structures were manipulated by PyMOL software (The PyMOL Molecular Graphics System, v1.8; Schrödinger, LLC, New York, 2015). Au–curcumin nanoparticle was designed according to Priyadarsini [46].
Pharmaceutics 13 01715 g001
Table 1. The pharmacokinetic profile of curcumin considering various commercially available, curcumin-based formulations.
Table 1. The pharmacokinetic profile of curcumin considering various commercially available, curcumin-based formulations.
FormulationCommercial ProductStudy DesignSubjectsSample HydrolysisFood AdministrationInterventionDoseAUC(0−t)
(ng*h/mL)
Refs.
Curcumin + piperineCurcumin C3
Complex® +
Bioperine®
Randomized crossover8 (10) Asian
males
20–26 yo
NoOvernight fasting administration with 150 mL of water and standard mealsFormulation
Control
2000 mg curcumin + 20 mg piperine
2000 mg curcumin
80 ± 10 a,2
4 a,2
[69]
Curcumin + piperineCurcumin C3
Complex® +
Bioperine®
Randomized crossover6 malesYes-Formulation
Control
2000 mg curcumin + 5 mg piperine
2000 mg curcumin
15.55 d,1
8.44 d,1
[42]
Curcumin + piperineCurcumin C3
Complex® +
Bioperine®
Open-label, uncontrolled phase I pilot study10 European males
23–33 yo
NoOvernight
fasting
Formulation12,000 mg curcumin + 5 mg piperineND 6[72]
Curcuminoids + turmeric essential oilBCM-95®CG
Biocurcumax™
Pilot study crossover11
(Gender not reported)
28–50 yo
NoOvernight
fasting
Formulation
Control
2000 mg curcuminoids
2000 mg curcuminoids
3201.28 (μg*h/g) b,7
461.86 b,7
[77]
Micellar
curcumin
Novasol®Single-blind crossover23
13 females 19–26 yo
10 males
20–28 yo
Yes12 h overnight fasting; drugs mixed with 50 g woodruff syrup; meals with neither food nor water restrictionFormulation
Control
410 mg curcumin
410 mg curcumin
4474.97 ± 1675.02 c,4
24.16 ± 42.58 c,4
[62]
Micellar
curcumin
Novasol®Single-blind crossover23
6 females
23 ± 2 yo
5 females 67 ± 3 yo
6 males
26 ± 2 yo
6 males
62 ± 3 yo
Yes12 h fasting
administration with water
Formulation
Control
80 mg curcumin
80 mg curcumin
212.48 ± 76.36 c,5
2.54 ± 4.20 c,5
[81]
Curcumin + cyclodextrinCavacurmin®Randomized double-blind crossover12
11 males
1 female 23.0 ± 2.4 yo 1 African
-American and 11 Caucasians
Yes10 h overnight fast; at 4 and 8 h after administration curcumin-free meals were served (first mealtime: 40 g chocolate whey protein isolate and 80 g instant oatmeal dissolved in 30 mL of water + 473 mL of water to drink; second mealtime: 230 g turkey breast, 2 slices of whole wheat bread, 15 g light miracle whip, 170 g of fat free Greek yogurt, and 473 mL of water)Formulation
Control
376 mg curcuminoids
1800 mg total curcuminoids
327.7 ± 58.1 a,4
3.9 ± 0.5 a,4
[57]
Curcumin + hydrophilic carrierCurcuwin®Randomized double-blind crossover12
11 males
1 female 23.0 ± 2.4 yo
1 African-American and 11 Caucasians
YesFormulation
Control
376 mg curcuminoids
1800 mg total curcuminoids
307.6 ± 44.6 a,4
10.8 ± 1.7 a,4
[58]
Curcuminoids + fenugreek dietary fibersCurQfen®Randomized double-blind crossover50
43 males
7 females
24–46 yo
NoOvernight fasting + 200 mL of water with the administration;
standardized food and beverages were provided: breakfast at 1 h time point, lunch after 5 h time point, and dinner after 8 h time point. No further food allowed except water
Formulation
Control
307 mg curcumin
307 mg curcumin
2,274,000 c,4
49.87 c,4
[66]
Curcuminoids + fenugreek
dietary fibers
CurQfen®Randomized double-blind crossover50
43 males
7 females
24-46 yo
NoFormulation
Control
76.7 mg curcumin
76.7 mg curcumin
963 c,4
38.8 c,4
Lipid droplet micromicellar techniqueBiocurc®Randomized open label
parallel
20 males 26.45 ± 5.02 yoNo10 h overnight fasting + 4 h fasting post administration. Curcumin- and garlic-free meals were provided Formulation750 mg curcumin231.31 ± 24.45 a,5[82]
Lipid droplet micromicellar techniqueBiocurc®Randomized double-blind crossover12
6 males
6 females
18–45 yo
No-Formulation
Control
64.6 mg curcumin
323 mg 95% curcumin
351 b,6
4.16 b,6
[63]
Nano-particle colloidal
dispersion
Theracurmin®Randomized14
8 males
6 females
30–59 yo
Japanese
YesAdministration with 100 mL of waterFormulation
Control
30 mg curcumin
30 mg curcumin
113 ± 61 c,2
4.1 ± 7 c,2
[61]
Nano-particle colloidal
dispersion
Theracurmin®Single-dose, double-blind, 4-way crossover24
13 males
11 females
23–32 yo; Japanese
YesOvernight fasting and light lunch after assumption.
Administration such as nutritional beverage
Formul. A
Formul. B
Formul. C
Formul. D
30 mg curcumin
30 mg curcumin
40 mg curcumin
30 mg curcumin
121 ± 65.6 c,3
79.5 ± 31.4 c,3
40.9 ± 24.9 c,3
30.1 ± 14.4 c,3
[59]
Natural turmeric matrix obtained by polar-nonpolar sandwich techniqueCureit®Pilot, crossover12 males (18–45 y)No Overnight fasting for at least 8 hFormulation
Control
190 mg curcumin
500 mg curcumin
980 ± 508.104 c,7
192.8 ± 63.886 c,7
[83]
Natural turmeric matrix obtained by polar-nonpolar sandwich techniqueCureit®Randomized open-label three treatment groups, and parallel comparative design45 males
18–45 yo
No10 h overnight fasting, 4 h fasting post administration with 250 mL of water. Lunch, snacks, and dinner at 4, 8, and 12 h from the time of dosingFormulation
Volatile oil formulation
Lecithin
formulation
180 mg curcumin
351 mg curcumin
80.5 mg curcumin
824.9 ± 466.5 c,5
117.3 ± 56.8 c,5
187.3 ± 190.9 c,5
[84]
Lecithin
formulation
Meriva®Randomized double-blind crossover9
8 males
1 female
35 ± 10 yo
YesOvernight fasting before administration followed by a standardized breakfast and lunch after 4 h of blood draw (one plain bagel with cream cheese).
Subjects were released after the 8 h blood draw, resuming their normal dietary intake but restricting foods with curcumin
Formulation high
Formulation low
Control
297 mg curcumin
165 mg curcumin
1295 mg curcumin
538.0 ± 130.75 e
272.6 ± 68.52 e
122.5 ± 29.35 e
[60]
Solid lipid particle
formulation
Longvida®Single-dose double-blind
crossover
6 males
18–40 yo
Indians
No10 h fasting prior to administration,
1 h fasting post administration. Administration with 240 mL of water
Formulation
Control
130-–195 mg
Curcumin
>390 mg curcumin
95.26 ± 4.62 a,5
NA
[85]
Control: unformulated curcumin; ND: not detected; AUC: area under the drug concentration–time curve; NA: not applicable; a mean ± standard error of mean; b mean; c mean ± standard deviation; d no reported standard deviation or error; e results in mg/mL; 1 AUC0-3; 2 AUC0-6; 3 AUC0-8; 4 AUC0-12; 5 AUC0-24; 6 AUC0-48; 7 AUC0-∞.
Table 2. Clinical studies of curcumin considering various curcumin-based formulations.
Table 2. Clinical studies of curcumin considering various curcumin-based formulations.
Commercial ProductStudy DesignClinical Trial NumberSubjectsDiseaseDose/InterventionTimeClinical Trial ResultsRefs.
Curcumin C3
Complex® +
Bioperine®
Randomized controlled
parallel-group
UMIN
Clinical Trials No. UMIN000033774
70
(treatment n = 35
46.63 ± 2.2 yo;
placebo
n = 35
47.51 ± 2.4 yo
Asians)
NAFLD500 mg Curcumin C3 Complex® +
5 mg Bioperine®
after meal
12 weeks
  • ↓ ALT (p = 0.035), AST (p = 0.042), ALP (p = 0.004) in the curcumin + piperine group
  • ↑ ALB and LDH (p = 0.001) in the placebo group
  • ↓ cholesterol (p < 0.016) and LDL-C (p < 0.017) in the curcumin + piperine group
  • ↓ cholesterol (p = 0.035) and LDL-C (p = 0.000) in the placebo group
  • ↓ NAFLD severity (p < 0.001) in the curcumin + piperine group
No adverse effects on hematological parameters
[73]
Curcumin C3
Complex® +
Bioperine®
Randomized double-blind placebo-controlled-49 (18–70 yo Asians)NAFLD500 mg Curcumin C3 Complex® +
5 mg Bioperine®
after breakfast
8
weeks
  • Weight changes at baseline and after intervention in the curcumin + piperine compared to placebo (p = 0.016)
  • ↓ serum concentrations TNFα (p = 0.001), EGF (p = 0.0001), MCP-1 (p = 0.008) in the curcumin + piperine group
NAFLD improvement by curcumin + piperine supplementation
[74]
BCM-95®CG
Biocurcumax™
Randomized non-inferiority controlledClinical Trials Registry India CTRI/2017/02/007962144
37 males
107 females; BCM-95®CG group n = 73 53.1 yo;
paracetamol group n = 71 50.8 yo)
Knee OA500 mg BCM-95®CG caps twice a day or 650 mg paracetamol tablet thrice a day6
weeks
  • ↓ TNFα in turmeric compared to paracetamol (p = 0.0095)
  • ↓ WOMAC pain score and WOMAC pain and function/stiffness score
All patients with knee OA acheived ≥ 20% score decreases; 18% of patients with knee OA in the turmeric group got ≥ 50% increase in WOMAC pain and function/stiffness score; 3% acheived ≥ 70% improvement; no patients in the paracetamol group acheived ≥ 50% improvement (18% vs 0%; p = 0.0002)
[79]
BCM-95®CG
Biocurcumax™
Prospective randomized open-label parallel-groupISRCTN
registry
ISRCTN
10074826
140
98 males
42 females BCM-95®CG + diclofenac group n = 71; diclofenac group n = 69 38–65 yo
Knee OA50 mg of diclofenac twice daily or 500mg BCM-95®CG + 50 mg diclofenac twice a day
Patients were provided with paracetamol 500 mg tab and Ranitidine 150 mg tab as rescue medication
4
weeks
  • Patients receiving BCM-95®CG + diclofenac showed larger improvement in symptoms, pain, and QoL than those treated only with diclofenac (p < 0.001)
  • ↓ pain intensity in patients treated with BCM-95®CG + diclofenac than those treated with diclofenac alone (p < 0.001)
  • Lower number of patients who required H2 blockers in the BCM-95®CG + diclofenac group compared to the diclofenac group (6% vs 28%, respectively; p < 0.001)
  • Lower need for rescue medication (p < 0.005) in the BCM-95®CG + diclofenac group (3%) compared to the diclofenac group (17%)
13% of patients treated by curcuminoid complex + diclofenac and 38% of patients treated with diclofenac showed undesirable effects (p < 0.001)
[80]
Novasol®Randomized double-blind crossoverRegistered at clinicaltrials.gov (NCT01925547)42
17 males
50 ± 20 yo;
25 females
52 ± 16 yo
Moderate hyperlipidemia - metabolic syndrome risk80.4 mg of curcumin thrice a day with each principal meal
(241.2 mg curcumin/day) or placebo
6
weeks
  • TC, TAG, HDL-C, LDL-C, and LDL/HDL-C concentrations neither diverged among treatments nor showed a difference over time
  • CRP and IL-6 concentrations were neither influenced by treatment nor time
  • A time effect (p < 0.001), but no treatment effect or treatment × time interaction, was detected for fasting glucose concentration
  • Free iron, transferrin, and transferrin saturation were neither modified by treatment nor time; no treatment or time effects were found for ALT and AST activities
[50]
Novasol®ProspectiveRegistered at clinicaltrials.gov (NCT01712542)10
4 females
6 males
67 ± 8 yo
GB1 g of micellar curcumin (57.4 mg curcumin) mixed in 200 mL of pear juice 3 times/day, after meals4 days
  • Mean intratumoral concentrations of curcumin, 56 pg/mg tissue (range 9–151)
  • The concentration of curcumin within the tumor correlated with the cumulative dose that patients received (p = 0.025).
  • Mean serum concentration of total curcumin was 253 ng/mL (range 129–364)
[88]
CurQfen®Double-blinded randomized placebo-controlled-22 males
18–35 yo
Obesity500 mg CurQfen® or placebo12 weeks
  • ↓ plasma homocysteine concentrations in the curcumin group compared with the placebo group (p = 0.04)
  • ↑ HDL serum concentrations in the curcumin group compared with the placebo group (p = 0.04)
No relevant variation was detected in endothelial function between curcumin and control groups
[91]
CurQfen®Randomized placebo-controlled pilot study-22 males
18–35 yo
Obesity500 mg CurQfen® or placebo12 weeks
  • ↓ brachial pulse pressure in the curcumin group compared with the placebo group (p = 0.040)
  • ↑ anti-inflammatory cytokine IL-10 with curcumin treatment (p = 0.071)
  • Different response among the curcumin group detected in cfPWV: responders were subjects with reductions in cfPWV after 12-week curcumin treatment, while the non-responders showed no decreases in cfPWV from baseline rates
The global variation in cfPWV among the curcumin responders shows a larger curcumin-induced destiffening effect for those with increased baseline cfPWV (p = 0.006)
[92]
Theracurmin®Randomized double-blind placebo-controlled parallel-groupRegistered at UMIN
UMIN000007361
33
placebo group n = 18 14 males
4 females
69 ± 7 yo;
treatment group n=15 9 males
6 females
70 ± 6 yo
Impaired glucose tolerance or non-insulin-dependent diabetes mellitus180 mg of Theracurmin® or placebo6 months
  • ↑ BS and AT-LDL (p = 0.017, p = 0.024, respectively) in patients treated by placebo; no significant change in Theracurmin® patients
  • ↓ TG and γ-GTP (p= 0.015, p = 0.007, respectively) in patients receiving Theracurmin®
  • Time-dependent increase in AT-LDL levels detected in the placebo group (p = 0.017)
  • No time-dependent modifications in AT-LDL levels in the Theracurmin® group compared with the placebo group,
Theracurmin® patients showed a higher percentage variation in their BMI. The modification in adiponectin, was negative in placebo patients, while it was positive in Theracurmin® patients
[95]
Theracurmin®Open-label Prospective-45
11 males
34 females 42–85 yo
Knee OA180 mg of Theracurmin® + combined therapies
(NSAIDs, pain relief patches, and hyaluronic acid knee injection treatment) or 180 mg of Theracurmin®
(n = 13 patients)
6 months
  • VAS, JKOM, and JOA scores were significantly better after 6-month treatment (p < 0.0001, p = 0.0003, p < 0.0001 respectively)
  • JOA score of 13 Theracurmin® patients considerably improved after the therapy (p = 0.0203)
Of the 45 patients, 34 were effective (75.6%) and 11 not effective cases, 8 of which did not feel the efficacy of Theracurmin®. The Theracurmin®-only group was comprised of 10 effective cases (76.9%) and 3 not-effective cases, 2 of which did not feel the efficacy of Theracurmin®
[96]
Cureit®Randomized, placebo-controlled, double-blindRegistered with Clinical Trials
Registry India (CTRI/2018/05/014174)
30
12 males
18 females 36 ± 11 yo
DOMS and related muscle damage500 mg of Cureit® or placebo12 weeks
  • ↓ VAS score in subjects treated with Cureit® (p < 0.0001)
  • Substantial changes (p < 0.001) in the pain score of Cureit® and placebo groups
[100]
Meriva®Observational study-124
61 males
63 females 56.4 ± 5.2 yo
n = 63 Meriva® +
glucosamine
n = 61
chondroitin sulphate +
glucosamine
Knee OA1 tablet/day containing 500 mg Meriva® + 500 mg Regenasure® (vegetarian glucosamine HCl) or 2 capsules/day, containing 400 mg chondroitin sulphate and 415 mg glucosamine HCl4 months
  • ↑ Karnofsky Index and WOMAC scores in the Meriva® + glucosamine group compared to the chondroitin + glucosamine group
  • ↑ of walking distance on the treadmill in the Meriva® + glucosamine group compared with the baseline and chondroitin + glucosamine group already at 1 month; this benefit was continued until the end of the investigation
  • ↓ need for associated drugs and medical care in both groups: the usage of Meriva® + glucosamine was related to a diminished need for medicines and medical care compared to chondroitin + glucosamine
[102]
Meriva®Pilot study-50 males
n = 25
common analgesic drugs group
18–40 yo
n = 25
Meriva® group
17–52 yo
osteo-muscular pain1 tablet of Algocur® every 12 h (each tablet contains 1 g of Meriva®) or conventional analgesic drugs10
days
  • ↓ perceived pain was scored by VAS compared to baseline in both treated groups, starting from the third day of therapy
  • Improvement of impaired physical function assessed in subjects treated with the Meriva®-based formulation and in subjects cured with common analgesic drugs, compared to the condition at baseline
  • Adherence to therapy was higher in the Meriva® group (24.96%) compared to the conventional analgesic drug group (15.6%) (p = 0.005)
Tolerability scored as outstanding by 24 (96%) subjects treated by the Meriva®-based formulation and by 14 (56%) subjects treated with standard therapy (p =0.002)
[103]
Meriva®Pilot study-12 eyes from 11 patients
7 males
4 females
Chronic
diabetic macular edema
500 mg of Meriva®
twice a day
3 months
  • BCVA improved after 3 months of treatment
  • 0% of eyes exhibited a decrease in visual acuity, 17% revealed stabilization, and 83% demonstrated improvement (p = 0.0072)
  • 92% of eyes revealed a decrease in macular edema, 8% displayed stabilization, and 0% demonstrated a clinically meaningful increase
  • Mean macular thickness decreased from baseline after 3-month treatment (p = 0.0090)
58% of eyes showed an improvement for near best-corrected visual acuity and 42% showed stability for near best-adjusted visual acuity
[104]
Meriva®Prospective
single arm
single center phase II trial
-44
29 males
15 females 42–87 yo
Advanced or metastatic pancreatic cancer2000 mg/day of Meriva® + GEM was infused on days 1, 8, and 15, within a 28-day treatment cycle42 months
  • Limited response in 27.3% of patients, stable disorder in 34.1%, progression in 17 patients (38.6%)
  • ↑ basal levels of IL-6 (p = 0.03), sCD40L (p = 0.05), and CRP (p = 0.03) in patients not responding to therapy
  • In responsive patients, no relevant changes in biomarkers were detected among baseline levels and those receiving 1 or 3 cycles of chemotherapy
A total number of 35 patients (79.5%) achieved the baseline QoL questionnaire and 30 (68.2%) filled it in for at least three cycles
[105]
ALB: albumin; ALP: alkaline phosphatase; ALT: alanine aminotransferase; AST: aspartate aminotransferase; AT-LDL: a1-antitrypsin-LDL; BCVA: best-corrected visual acuity; BMI: body mass index; BS: blood sugar; cfPWV: carotid-femoral pulse wave velocity; CRP: C-reactive protein; DOMS: delayed onset muscle soreness; EGF: epithelial growth factor; GB: glioblastoma; HDL-C: high-density lipoprotein cholesterol; IL-6: interleukin-6; JKOM: Japanese Knee Osteoarthritis Measure; JOA: Japanese Orthopedic Association; Knee OA: knee osteoarthritis; LDH: lactate dehydrogenase; LDL-C: low-density lipoprotein cholesterol; NAFLD: nonalcoholic fatty liver disease; NSAIDs: non-steroidal anti-inflammatory drugs; QoL: quality of life; sCD40L: soluble CD40 ligand; TAG: triacylglycerols; TC: total cholesterol; TG: triglycerides; TNFα: tumor necrosis factor alpha; VAS: visual analog scale; WOMAC: Western Ontario and McMaster Universities’ Arthritis Index; γ-GTP: γ-glutamyl transpeptidase; GEM: gemcitabine.
Table 3. A detailed step-by-step description of the DbD approach with an overview of curcumin-fortified beverages fabrication.
Table 3. A detailed step-by-step description of the DbD approach with an overview of curcumin-fortified beverages fabrication.
Stage ProcessPropertyRequirementsCurcumin-Fortified Functional Beverage
Stage 1 
Active agent definition
  • Molecular and physico-chemical features
  • Solubility features
  • Stability features
Chemical formula and structure, molar mass, molecular dimensions, density, refractive index, melting and boiling points, partitioning, diffusion coefficient, molar volume, surface tension, solubility, pKa values, hydrogen bond acceptors/donors
Molecular weight, number of hydrogen bond donors/acceptors, oil-water partition coefficient, ionic strength, pH, matrix composition, temperature
pH, temperature, oxygen levels, pro-oxidant levels, light exposure
Polyphenol, molecular formula: C21H20O6
Low-molecular-weight substance (368.380 Da)
At room temperature, it exits as a yellow-orange crystallized powder (melting point 180 °C, boiling point 521.3 °C). Density 1.3 g/cm, refractive index 1.643
Water solubility is highly pH-dependent (24 mg/L at pH < 8, while it becomes hydrophilic at pH > 11) and its surface and interfacial tension (54.3 dyne/cm)
Oil solubility tends to increase as the molecular weight of the carrier oil decreases—olive oil (1.18 mg/mL), corn oil (3 mg/g), tributyrin (29.8 mg/g), medium-chain triglycerides (7.9 mg/g), sunflower oil (1.08 mg/mL)
The partition coefficient oil/water is expressed as LogP value (4.12), influenced by pH. pH values could influence electrical charge thanks to the presence of ionizable functional groups (pKa values of the three readily dissociable hydrogen atoms: 8.1, 9.5, and 10.1)
6 O2 atoms act as hydrogen bond acceptors, while 2 H2 atoms can donate hydrogen bonds
Highly stable at acidic condition (half-life around 175 days at pH 5.97) but the degradation rate increases at the increase in pH values
Photochemical degradation occurs more frequently under visible light than UV light
Thermal degradation occurs at high temperature
Dissolution in oils prevents degradation compared to dissolution in water
Stage 2 
End-product definition
  • Compositional analysi
  • Environmental stress analysis
  • Physical state and rheological properties
  • Optical properties
pH, ionic strength, polymer type, transition metals, surface-active substances
Temperatures, oxygen levels, light conditions
Solid, semi-solid, liquid, viscosity, elastic modules, breaking stress
Transparency, color
The product pH should be 6.3-–6.8 with total carbohydrates ≤10%, including dietary fibers ≥0.25%, total fat ≤5%, protein ≥3%, minerals (ionic strength of 90–110 mM)
Curcumin content should be at least 150 mg/100 g of product and the quantity reaching the blood should exceed 50 ng/mL
It would be desirable to use plant-based ingredients to meet the favor of health-conscious consumers
The product should be physically stable and avoid phase separation or increase in mean particle size as well as chemically stable: curcumin content should not decrease by >15% and color should not reduce during manufacturing, storage, and consumption (by >5% at 20 °C or by >10% at 30 °C for 6 months, in light exposure)
The product should stay stable after heating (72 °C for 15 s) and at temperatures 0–45 °C for 12 months of storage, and after exposure to high-shear mixing conditions (500 rev/min for 5 min)
Ideally, the product should be like the “golden milk”, a creamy, cloudy, and yellow beverage (viscosity of 2–4 mPa·s) with an earthy spicy flavor profile
Stage 3 
Delivery system specification
  • Physical form, rheological properties
  • Optical properties
  • Stability characteristics
  • Functional attributes
Solid, semi-solid, liquid, viscosity, elastic modules, breaking stress
Clear, turbid, or opaque
pH values, temperatures, ionic strengths, light exposure, oxygen levels, ingredients
Loading, active agent protection, retention, and release profile, dispersibility
Fluid or powdered delivery systems would readily disperse into an aqueous solution at room temperature (pH 6.5)
The delivery system should match the appearance of the end product that is a creamy yellow aspect
Colloidal particles should remain stable to aggregation, growth (increase >20% of particle diameter), dissolution, or separation during storage
Physical and chemical stability should be maintained after thermal processing at 75 °C for 15 s, when stored at temperatures of 0–45 °C (12 months), or at pH values ranging from 6.0 to 7.0 for 12 months (30 °C)
Curcumin not degraded (<20%) during production, storage, and utilization; concentration of active curcumin should be >1000 mg when stored for 12 months (30 °C, pH 6.5)
The delivery system should prevent curcumin degradation during passage through certain parts of the GI tract, while allowing release in others
Stage 4 
Particle specification & delivery system selection
  • Particle characteristics
  • Loading characteristics
  • Retention and release characteristics
  • Environmental responsiveness
  • Delivery system selection
Particles, composition, concentration, morphology, dimensions, physico-chemical properties, charge interactions
Loading capacity, encapsulation efficiency
Prolonged, burst, or triggered release
Ease of manufacture, cost, legal status, and consumer acceptability (plant-based/ vegetarian, animal-based, Kosher, or synthetic ingredients)
Micelles, liposomes, emulsions, microemulsion, nanoemulsion, solid lipid nanoparticles, or polymer particles
To obtain a creamy beverage stable to gravitational separation, colloidal particles should be relatively small (d < 300 nm) and/or have great density, have a smooth mouth feel, and the diameter of the colloidal particles should be smaller than 50 μm
Since curcumin is a hydrophobic molecule, it would be advantageous to use colloidal particles with a hydrophobic interior; moreover, curcumin has higher solubility in medium-chain triglycerides (e.g., coconut oil) and improved bioaccessibility in MCT-nanoemulsions
The delivery system should be designed to selectively break down in the small intestine and release free fatty acids for increasing the curcumin solubilization capability within the intestinal liquids
The ideal delivery system should be the simplest and cheapest (lower than 2 cents per unit) to manufacture; ingredients should preferably be plant-based or derived from natural food-grade components
Among the variety of the delivery systems available, oil-in-water emulsions and nanoemulsions are the most appropriate delivery system for curcumin, thanks to their ease of fabrication with plant-based emulsifiers and lipids, their optically opaque appearance, and their low viscosity. They have great stability and a superior bioavailability
Stage 5 
Process specification
  • Ingredient quality: Critical ingredient attributes
  • Manufacturing steps: Critical processing attributes
Composition, quality; ingredient properties/colloidal delivery system performance interaction
Resources, equipment, and facilities available; cost, simplicity, reliability, scale up; dissolution times, temperatures, mixing conditions, appropriate particle size distribution.
According to the equipment, resources, and available facilities, different approaches to produce colloidal particles required for the end-product are available
To prepare nanoemulsions, it is possible to dissolve curcumin in the oil phase and then homogenize with an aqueous protein mixture, employing a high-pressure homogenizer or microfluidizer. Experiments should be performed for optimizing the system composition and homogenization conditions (minimum mixing time and mixing rate, optimum solubilization temperature, protein concentration, and operating pressure/number of passes through the homogenizer for obtaining nanoemulsions with the desirable droplet size)
Otherwise, nanoemulsions could be produced thanks to the pH-driven method, which consists of loading curcumin into a preexisting nanoemulsion or milk-like product starting from an alkaline curcumin solution mixed with an acidic colloidal dispersion at room temperature
Stage 6 
Performance testing
  • Analytical tools
  • Testing protocols
Static and dynamic light scattering techniques, microelectrophoresis, optical, confocal fluorescence, electron microscopy, colorimetry, spectrophotometry, chromatography, mass spectrometry, sensory tests
In-product tests, accelerated screening tests, environmental stress tests
The particle size distribution of the nanoemulsions and their electrical characteristics (ζ potential) could be measured by dynamic light scattering and particle electrophoresis, respectively
Environmental stress could be affected by exposing it to ionic strength, different pH and temperature, and mechanical conditions
Measuring the optical properties (colorimeter), rheology (rheometer), and phase separation (visual observation) of the functional beverage allows researchers to monitor the impact of the delivery system on the appearance, texture, and stability of the end-product
Stage 7
System optimization
Monitoring and tabulating the characteristics of the delivery system and end-product throughout its life cycle so as to optimize their properties and make appropriate adjustments where possible, if needed
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Tabanelli, R.; Brogi, S.; Calderone, V. Improving Curcumin Bioavailability: Current Strategies and Future Perspectives. Pharmaceutics 2021, 13, 1715. https://0-doi-org.brum.beds.ac.uk/10.3390/pharmaceutics13101715

AMA Style

Tabanelli R, Brogi S, Calderone V. Improving Curcumin Bioavailability: Current Strategies and Future Perspectives. Pharmaceutics. 2021; 13(10):1715. https://0-doi-org.brum.beds.ac.uk/10.3390/pharmaceutics13101715

Chicago/Turabian Style

Tabanelli, Rita, Simone Brogi, and Vincenzo Calderone. 2021. "Improving Curcumin Bioavailability: Current Strategies and Future Perspectives" Pharmaceutics 13, no. 10: 1715. https://0-doi-org.brum.beds.ac.uk/10.3390/pharmaceutics13101715

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

Article Metrics

Back to TopTop