Next Article in Journal
Prevalence and Antimicrobial Resistance of Escherichia coli, Salmonella and Vibrio Derived from Farm-Raised Red Hybrid Tilapia (Oreochromis spp.) and Asian Sea Bass (Lates calcarifer, Bloch 1970) on the West Coast of Peninsular Malaysia
Next Article in Special Issue
Cyclic Tetrapeptides with Synergistic Antifungal Activity from the Fungus Aspergillus westerdijkiae Using LC-MS/MS-Based Molecular Networking
Previous Article in Journal
Groin Surgical Site Infection in Vascular Surgery: Systemic Review on Peri-Operative Antibiotic Prophylaxis
Previous Article in Special Issue
Antimicrobial Bacillus: Metabolites and Their Mode of Action
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Curcumin: Biological Activities and Modern Pharmaceutical Forms

Faculty of Technology, University of Niš, Bulevar Oslobođenja 124, 16000 Leskovac, Serbia
*
Authors to whom correspondence should be addressed.
Antibiotics 2022, 11(2), 135; https://doi.org/10.3390/antibiotics11020135
Submission received: 10 December 2021 / Revised: 6 January 2022 / Accepted: 10 January 2022 / Published: 20 January 2022
(This article belongs to the Special Issue Antimicrobial Natural Products)

Abstract

:
Curcumin (1,7-bis-(4-hydroxy-3-methoxyphenyl)-hepta-1,6-diene-3,5-dione) is a natural lipophilic polyphenol that exhibits significant pharmacological effects in vitro and in vivo through various mechanisms of action. Numerous studies have identified and characterised the pharmacokinetic, pharmacodynamic, and clinical properties of curcumin. Curcumin has an anti-inflammatory, antioxidative, antinociceptive, antiparasitic, antimalarial effect, and it is used as a wound-healing agent. However, poor curcumin absorption in the small intestine, fast metabolism, and fast systemic elimination cause poor bioavailability of curcumin in human beings. In order to overcome these problems, a number of curcumin formulations have been developed. The aim of this paper is to provide an overview of recent research in biological and pharmaceutical aspects of curcumin, methods of sample preparation for its isolation (Soxhlet extraction, ultrasound extraction, pressurised fluid extraction, microwave extraction, enzyme-assisted aided extraction), analytical methods (FTIR, NIR, FT-Raman, UV-VIS, NMR, XRD, DSC, TLC, HPLC, HPTLC, LC-MS, UPLC/Q-TOF-MS) for identification and quantification of curcumin in different matrices, and different techniques for developing formulations. The optimal sample preparation and use of an appropriate analytical method will significantly improve the evaluation of formulations and the biological activity of curcumin.

1. Introduction

The main ingredient of the Curcuma longa is the rhizome [1], a low-molecular-weight lipophilic molecule that can pass through the cellular membrane easily [2]. By its chemical structure, it belongs to the group of polyphenols [3]. Because of its intensive yellow colour, it is used as a natural food colouring agent [1]. The simple molecular structure and arrangement of functional groups are suitable for examining the relationship between the structure and its activity [4]. The ability of curcumin to interact with different proteins facilitates selective modulation of multiple cellular signalling pathways associated with various chronic diseases [5]. The transcription factors, mediators of inflammation and enzymes such as protein kinase, reductase, and histone acetyltransferase are important molecules for curcumin binding. Curcumin is a powerful epigenetic regulator in many diseases, such as neurological disorders, inflammatory diseases, diabetes, and different types of cancer [6]. Furthermore, it modulates various proteasomal pathways and reduces glycogen metabolism through selective inhibition of phosphorylase kinase enzyme [7]. The studies have shown that curcumin exhibits anti-inflammatory, hypoglycemic, antioxidant, antimicrobial, antiviral, anticancer, neuroprotective, and many other effects [8]. However, the main obstacle to the effective manifestation of the pharmacological activity of curcumin is its poor aqueous solubility and low bioavailability [9,10,11]. The main factors contributing to the low bioavailability of curcumin in the blood plasma and tissues are its poor absorption, fast metabolism, and rapid systemic elimination [12]. The enhancement of the solubility and bioavailability of this promising molecule is crucial for potential clinical application. Different approaches in developing curcumin formulations can improve its physicochemical characteristics and enable safe and efficient use. For that purpose, formulations including nanoparticles, liposomes, micelles, phospholipid complexes, hydrogels, etc., have been described in the reference sources [11,13]. The aim of this paper is to analyse the factors influencing the bioavailability of curcumin, as well as to review pharmacological activities of curcumin and strategies in order to enhance its bioavailability.

2. Curcumin: Background

Turmeric (Curcuma longa) is an aromatic plant from the ginger family (Zingiberaceae). It is grown in the southern and southwestern regions of Asia. It occupies an important place in the cuisine of Iran, Malaysia, India, China, Polynesia, and Thailand. It is used as a spice, and it affects the nature, colour, and taste of food. Curry is the best-known spice that contains turmeric rhizome powder. Curcumin is also used as an ecological dye; it is known as Natural yellow 3 and has been assigned an E number-E100, when used as a food colouring agent [14,15]. Figure 1 shows curcuminoids (curcumin, demethoxycurcumin, and bis-demethoxycurcumin), the main components of turmeric rhizome.
Curcumin, a yellow-orange pigment isolated from turmeric two centuries ago, is a widely studied natural compound that has shown enormous in vitro therapeutic potential. For centuries, it has been used in Ayurvedic medicine and traditional Chinese medicine [11,16]. An overview of the discovery and application of curcumin is shown in Table 1.
Curcumin has been found to possess pleiotropic activities owing to the potential of this polyphenol to modulate multiple signalling molecules. Curcumin exhibits anti-inflammatory, antioxidant, proapoptotic, chemopreventive, chemotherapeutic, antinociceptive, antiproliferative, antiparasitic, and antimalarial effects, and it is used as a wound-healing agent. The research on curcumin and its pharmacological activities has become increasingly important in recent years [9].

3. Isolation of Curcumin from Turmeric Rhizome and Methods of Identification

Turmeric rhizome contains two main classes of pharmacologically active secondary metabolites: curcuminoids and essential oil [28]. Curcuminoids (curcumin, demethoxycurcumin and bis-demethoxycurcumin) are most responsible for the biological activity of turmeric [29]. The isolation of curcuminoids from turmeric rhizomes is achieved by applying conventional and modern extraction methods [30]. Soxhlet extraction [31] and maceration [32] are classic extraction methods. Of the modern methods for extraction of curcuminoids, ultrasound extraction [33], enzyme-assisted extraction [34], microwave extraction [35], supercritical fluid extraction [36], and pressurized fluid extraction [37] are used. The most commonly used solvents for curcuminoid extraction are ethanol, dichloromethane, ethyl acetate, isopropanol, methanol, n-butanol, and acetone [29,38,39]. Sahne et al. used acetone as a solvent in conventional and unconventional extraction processes due to its high solubilization capacity [31]. In the paper by Muthukumar et al., various organic solvents for curcumin extraction were examined. The research findings show that acetone is the most efficient extraction solvent [39]. Thin-layer chromatography (TLC) is a classical analytical technique for separating curcumin from the extraction mixture [29,38]. Curcumin is quantified in the extract using high-performance liquid chromatography (HPLC). After extraction, the organic solvents are removed from the extract by evaporation on a vacuum evaporator. The residue (oleoresin) is then dissolved in methanol and subjected to HPLC analysis [40]. The yield and stability of curcumin depend on the extraction method used. Sahne et al. analysed curcumin extraction from the turmeric rhizome using several advanced methods, and the results were compared with the results obtained by Soxhlet extraction, the most commonly used reference method. The result showed that the yield of curcumin extraction obtained using the Soxhlet method (6.9%) was significantly higher than the one obtained by extraction using microwaves (3.72%), ultrasound (3.92%), and enzymes (4.1%). Although modern extraction methods do not show high extraction yields similar to the Soxhlet method, their advantages (low temperature and short extraction time, use of a very small amount of solvent) make them more favourable methods for curcumin extraction [31].
The kinetic degradation of curcumin from a natural mixture of curcuminoids in different conditions (pH, temperature, and dielectric constant of the solvent), as well as the degradation of pure curcumin in defined conditions, were examined in the paper by Naksuriya et al. An aqueous buffer/methanol 50:50 (v/v) mixture was used as a standard medium to examine the kinetics of curcumin degradation. The results showed that the degradation of pure curcumin present in the curcuminoid mixture underwent a first-order reaction. An increase in pH, temperature, and dielectric constant of the medium lead to an increase in the rate of curcumin degradation. Curcumin showed rapid degradation due to autoxidation in aqueous buffer pH = 8.0 with a constant rate of 0.28 h−1, which corresponds to a half-life (t1/2) of 2.5 h. Curcumin incorporated as a mixture of curcuminoids into ω-methoxypoly(ethylene glycol)-b-(N-(2-benzoyloxypropyl) methacrylamide) polymer micelles was about 300–500 times more stable than pure curcumin in a mixture of phosphate buffer and methanol.
Incorporating curcumin into polymer micelles is a promising approach for stabilising this compound and developing formulations suitable for further pharmaceutical and clinical trials [41]. Liu et al. examined natural deep eutectic solvents formed from organic acids and sugars for the efficiency of curcuminoid extraction. In optimal conditions (the temperature of 50 °C, 0.1/10 g/mL ratio of solid and liquid components, and extraction time of 30 min), higher extraction yields were achieved when a solvent with a ratio of citric acid and glucose 1:1 and 15% water was used, compared to the conventional extraction solvents. The proposed method is an excellent alternative for extracting natural pigments since it is environmentally friendly and sustainable [42]. During the isolation and purification of curcuminoids from oleoresin, the volatile oil of turmeric dissolves curcumin, thus creating a problem in the recrystallization process. Different organic solvents and their combinations for selective recrystallization of curcuminoids were examined. A mixture of isopropyl alcohol and hexane (1:1.5, v/v) proved to be the best solvent for recrystallization in the purification of curcuminoids. The total curcumin content in the raw curcuminoid powder was 76.82% w/w, whereas, in the recrystallized powder, the purity was increased up to 99.45% w/w [29].

4. Physico-Chemical Properties of Curcumin

Curcumin (1,7-bis-(4-hydroxy-3-methoxyphenyl)-1,6-heptadiene-3,5-dione) [43] or diferuloylmethane [44] is an integral component of turmeric (up to ~5%), a well-known traditional spice [45]. It is a lipophilic compound, insoluble in water, acidic, and neutral solutions, and soluble in ethanol, dimethylsulfoxide, and acetone. Curcumin can be extracted from turmeric rhizomes by using organic solvents. The molecular formula of curcumin is C21H20O6, and the molecular weight is 368.38 g/mol. The melting point of curcumin is 183 °C. Curcumin is a tautomeric compound due to the presence of β-diketone in the molecular structure and shows diketo/keto–enol tautomerism (Figure 2) [46,47].
The diketo tautomer can exist in cis and trans forms. Solvent polarity, pH, and temperature significantly affect curcumin’s keto–enol balance [48]. The ratio of keto and enol tautomers of curcumin, on the other hand, strongly influences pharmacological activities [49]. Manolova et al. examined the tautomerism of curcumin in ethanol/water binary mixtures using ultraviolet–visible (UV–VIS) spectroscopy and advanced quantum chemical calculations. The results show that only enol–keto tautomer is present in ethanol. The addition of water leads to the emergence of a new spectral range, which is assigned to the diketo tautomeric form. The diketo form is dominant in the mixture of water and ethanol 90:10 (v/v). The observed equilibrium shift is explained by quantum chemical calculations, which show that water molecules stabilise the diketo tautomer by forming stable complexes [50]. Kawano et al. analysed keto–enol tautomers of curcumin by using liquid chromatography/mass spectrometry. The research findings show that the enol form is the main form in solution (water/acetonitrile) [51]. In nonpolar solvents (carbon tetrachloride) in the solid state and solution, curcumin exists in enol form [50,51].
Curcumin is unstable in the solution form. It has an intense yellow colour, which changes to dark red in the basic solution [52].

5. Structure, Bioavailability, and Safety of the Application of Curcumin

Curcumin, a polyphenol from the diarylheptanoid group, has two aromatic rings symmetrically substituted by methoxy and a phenolic OH group in the ortho position, which are connected to a conjugated seven-membered hydrocarbon chain with an enone part and a 1,3-diketone group (Figure 3). The active functional groups of curcumin are two o-methoxy and two phenolic groups, two double bonds in the hydrocarbon chain and the 1,3-keto–enol part of the structure [53].
Aromatic groups provide hydrophobicity, whereas the α,β-unsaturated β-diketo part of the structure allows flexibility to the molecule. These unique properties of curcumin make it capable of binding to various biomacromolecules. Biologically critical chemical reactions of curcumin are realized through the H-bond of the β-dicarbonyl group and phenolic hydroxyl residues, as well as the ether residue of the methoxy group, and by binding with metals and nonmetals. It has been demonstrated that curcumin binds directly to numerous signalling molecules, such as inflammatory molecules, protein kinase, protein reductase, histone acetyltransferase, histone deacetylase, glyoxalase I, xanthine oxidase, human immunodeficiency virus (HIV1) integrase, HIV1 protease, sarco/endoplasmic reticulum calcium ATPase, deoxyribonucleic acid (DNA) methyltransferase 1, carrier proteins, and metal ions. The diketo group forms chelates with transition metals, reducing metal-induced toxicity, while some of the metal complexes exhibit enhanced antioxidant activity because they mimic enzymes [54]. Curcumin can also bind directly to DNA and ribonucleic acid (RNA). The ability of curcumin to bind to carrier proteins improves its solubility and bioavailability. Curcumin is unstable at physiological pH and degrades rapidly in an autoxidation reaction to the major bicyclopentadione product in which a 7-carbon chain has undergone oxygenation and double cyclization [55]. The alkaline hydrolysis products (ferulic acid, vanillin, ferulaldehyde, and feruloylmethane), as well as oxidation products (such as bicyclopentadione), show biological activity but are significantly less active than curcumin [56].
The clinical trials with curcumin have clearly demonstrated its safety, tolerability, and efficacy against different chronic diseases in humans [8]. The human studies did not indicate any toxic effects when curcumin was administered orally in the dosage of 6 g/day during 4–7 weeks [57]. The study on safety, tolerability, and activity of liposomal curcumin (Lipocurc™) on patients with locally advanced or metastatic cancer was conducted by Greil et al. It demonstrated that 300 mg/m2 of liposomal curcumin was the maximum safe dosage for patients with cancer treatment [58]. Saghatelyan et al. assessed the efficacy and safety of intravenous infusion of curcumin in combination with paclitaxel in patients with metastatic and advanced breast cancer. After a 12-week treatment, curcumin administered intravenously did not cause any significant health issues, nor did it deteriorate the quality of life [59].

6. The Metabolism of Curcumin

Poor bioavailability of curcumin in humans at a dose of 12 g/day is a consequence of poor absorption in the small intestine, fast metabolism in the liver, and rapid systemic elimination [60]. Most of the orally administered curcumin is excreted through faeces, without metabolism, while a smaller, absorbed part undergoes metabolic modification. The metabolism of curcumin takes place in two stages. The first phase involves the reduction in the presence of reductases, which takes place in enterocytes and hepatocytes. The reduction products are dihydrocurcumin, tetrahydrocurcumin, hexahydrocurcumin, and octahydrocurcumin (hexahydrocurcuminol) [61]. The curcumin reduction reaction is catalyzed by enzymes such as nicotinamide adenine dinucleotide phosphate (NADPH)-dependent reductase, alcohol dehydrogenase, and an unidentified microsomal enzyme [62]. In the paper by Hassaninasab et al., the enzyme for curcumin reduction was purified from Escherichia coli and characterized. It was found that the microbial metabolism of curcumin by a purified enzyme involved reduction in two steps, in which, depending on NADPH, curcumin was converted into an intermediate product, dihydrocurcumin, and then into the final product, tetrahydrocurcumin [63]. Curcumin and its reduced metabolites are readily conjugated to glucuronic acid and sulfate in vivo and in vitro. Glucuronidation and sulfation reactions take place in the presence of glucuronyl transferase and sulfotransferase, respectively. Glucuronidation and sulfation of curcumin occur in the liver and intestines of rats and humans [61]. After oral administration in humans, a portion of curcumin is absorbed and can be identified as a water-soluble glucuronide and sulfate conjugate in the plasma. Human phenol sulfotransferase 1A1 (SULT1A1) and human phenol sulfotransferase 1A3 (SULT1A3) are responsible for the sulfation of curcumin in humans and in the intestines of rats, while uridine diphosphate-glucuronosyltransferase (UGT) catalyses glucuronidation of curcumin [54]. The reduction or conjugation of curcumin generates species with a reduced ability to inhibit the expression of cyclooxygenase-2 (COX-2) compared with curcumin. Tetrahydrocurcumin, hexahydrocurcumin, and curcumin sulfate exhibit weaker inhibition of prostaglandin E2, while hexahydrocurcuminol is inactive [64]. The biological activity of curcumin metabolites other than tetrahydrocurcumin is significantly reduced compared with curcumin [65,66]. To enhance the bioavailability of curcumin, piperine which interferes with glucuronidation, curcumin in liposomes, curcumin nanoparticles, curcumin phospholipid complexes, and structural curcumin analogues are used. Figure 4 shows the metabolic and nonmetabolic transformations of curcumin.

7. Characterization of Curcumin

Curcuminoids are widely used in the food processing and pharmaceutical industries due to their properties. The detection and characterization of curcuminoids in different matrices are of great importance. The choice of the analytical method for curcuminoid analysis depends on the sample type, the purpose of the analysis, and the detection and quantification limits [67,68]. The techniques based on chromatography and electrophoresis are among the selected methods for determining curcuminoids. TLC is one of the methods used for fractioning turmeric extracts [38]. The use of the TLC method for turmeric analysis has declined due to prolonged separation time and poor resolution, although it is selective, easy to perform, and inexpensive. New high-performance thin-layer chromatography (HPTLC) methods that overcome these limitations have been developed [69]. The principle of operation is the same as with TLC. Higher resolution, lower detection limit and improved image scanning are advantages of HPTLC methods [68]. HPLC, in combination with UV–VIS detector, is the most commonly used chromatographic method for the qualitative and quantitative analysis of curcumin due to its high precision, accuracy, and low detection limit. Various HPLC methods have been developed to analyse curcuminoids (Table 2).
Liquid chromatography–mass spectrometry (LC/MS) [67] is used to identify and quantify traces of curcumin in biological fluids, food, or other complex matrices. Various LC/MS methods have been developed to detect and quantify curcumin in different matrices [77,78,79]. A rapid and sensitive, selective high-throughput ultrahigh performance liquid chromatography method with tandem mass spectrometry (UPLC/Q-TOF-MS) was developed and validated to quantify curcuminoids to reduce analysis time and improve sensitivity [80]. The UV–VIS spectroscopy can also quantify curcuminoids if the sample matrix or other present components do not absorb within this range. Curcumin shows an absorption maximum at 425 nm [62,68,81]. Fourier transform infrared spectroscopy (FTIR), near-infrared spectroscopy (NIR), Raman’s spectroscopy, nuclear magnetic resonance spectroscopy (NMR), and fluorescence spectroscopy are also used to characterize curcumin [82,83,84,85].
Curcumin exists in three polymorphic forms: monoclinic form and two orthorhombic forms. Pandey et al. examined polymorphs using X-ray diffraction (XRD) and differential scanning calorimetry (DSC) and found that curcumin polymorphs were monotropically linked to each other, with the monoclinic form being the most stable [86].
Electron paramagnetic resonance (EPR) spectroscopy is an efficient and noninvasive spectroscopic method for analysing samples with unpaired electrons. It is used to quantify the types of radicals and analyse the antioxidative effects of substances [87]. EPR spectroscopy was applied for determining the potential and capacity of curcumin against free radicals (DPPH, nitric oxide radical (NO⋅), hydroxyl radical (HO∙) and superoxide anion radical (O2) [88,89]. In the study by Nikolić et al., EPR spectroscopy was used for assessing the antioxidant activity of curcumin-loaded low-energy nanoemulsions according to Tempol stable nitroxide free radical. The research findings show that nanoemulsion with curcumin exhibits swift activity, thus neutralising free radicals within the first five minutes from the beginning of the reaction [90].

8. Formulations

A large number of the curcumin formulations with volatile oil (volatile oil formulation) [91,92], piperine [93], and lecithin [94] have been designed. These formulations increase the absorption of curcumin after oral administration compared with pure curcumin. Liposomes, micelles, phospholipid complexes, cyclodextrins, nanoparticles, emulsions, hydrogels, and phytosomes are new promising curcumin formulations. Such formulations provide more prolonged circulation, better absorption and resistance to metabolic processes, increase absorption from the small intestine, and prolong half-life in the plasma, and thus, increase the efficiency of curcumin [95,96,97] (Figure 5).
Phytosome formulations with curcumin, formulations with volatile oils of turmeric rhizome, and curcumin formulations with a combination of hydrophilic carrier, cellulose derivatives, and natural antioxidants (CHC), compared to a standardized mixture of curcumin (CS), were tested in a study on healthy volunteers. The CHC formulation of curcumin significantly increases the content of curcuminoids in the blood compared with standard curcumin [92]. Cyclodextrins (CDs) can form molecular inclusion complexes with lipophilic compounds, thus enhancing water solubility, dispersion, and absorption of active components. The bioavailability of the curcumin formulation with γ-cyclodextrin was investigated. This formulation was compared with a standardized curcumin extract and two commercially available formulations, the curcumin phytosome formulation (CSL) and the curcumin formulation with rhizome-extracted turmeric essential oils (CEO). The formulation of curcumin with γ-cyclodextrin significantly enhances the absorption of curcuminoids in healthy people [9]. The inclusion complex of curcumin with β-cyclodextrin was prepared using the coprecipitation method. The solubility of curcumin in water increased from 0.00122 to 0.721 mg/mL by forming an inclusion complex. The release of the inclusion complex from nanocomposite and conventional poly (N-isopropylacrylamide/sodium alginate) hydrogels cross-linked with nanoclay and N,N′-methylenebis(acrylamide) (BIS), respectively, was tested under simulated gastrointestinal conditions. At pH = 1.2 and pH = 6.8, hydrogels showed the lowest and the highest release-swelling ratio, respectively. The swelling coefficient and cumulative release decreased with increasing nanoclay content in nanocomposite hydrogels. Conversely, as the BIS ratio in conventional hydrogels increased, the swelling ratio and cumulative release increased [98]. The polyol dilution method was used to formulate liposomes with curcumin in the paper by Kongkaneramit et al. Lipid phase was a mixture of hydrogenated phosphatidylcholine and cholesterol in a molar ratio of 9:1. Propylene glycol, glycerin, and polyethylene glycol 400 were used as polyol solvents. The type and amount of polyol affect both the size of the liposomes and the amount of curcumin incorporated. The preparation temperature is also an important factor in liposome production [99]. Tai et al. studied the stability and release performance of curcumin from liposomes with different contents of hydrogenated phospholipids [100].
Chitosan-coated liposomes may be an alternative carrier for drug delivery in humans. In the work of Cuomo et al., the applicability of chitosan-coated liposomes with curcumin, as well as anionic liposomes with curcumin, was evaluated. The applicability of the formulations was examined in vitro by measuring the bioavailability of ingested curcumin. It has been shown that the presence of a positively charged liposome surface enables better absorption of curcumin in the small intestine, which increases its overall bioavailability [101]. Curcumin nanoemulsion was formulated as a low-energy emulsion and converted to a nanoemulgel using cross-linked polyacrylic acid (Carbopol® 934) as a gelling agent to increase the solubility and absorption of curcumin after topical application to the skin. The nanoemulgel formulation showed faster and earlier wound healing in psoriatic mice compared with curcumin and betamethasone-17-valerate gel. The research findings show that curcumin nanoemugel formulation is a promising candidate for successful long-term treatment of psoriasis [102].
Curcumin nanoemulsions are highly effective in preventing tumour recurrence after surgery and metastasis [103]. A formulation of eye drops (thermosensitive hydrogel containing latanoprost and curcumin nanoparticles) for dual drug delivery has been developed and characterized. The developed dual drug delivery system has shown a prolonged release profile, in vitro and in vivo biocompatibility, reduced levels of inflammation and apoptosis of cells, and protection of trabecular mesh (TM) cells from oxidative damage [104]. PLGA curcumin nanoparticles have shown increased oral and intravenous bioavailability [105]. The oral formulation of nanocurcumin can significantly reduce recovery time in hospitalized patients with COVID-19 [106]. The hybrid curcumin-phospholipid complex was used as a system for oral drug administration to inhibit the metastasis of breast and lung cancer [107]. A high-performance formulation of curcumin phospholipid complex, which can improve the flow, solubility, and oral bioavailability of curcumin, was developed by Wang et al. [108]. Polymer micelles made using block copolymer methoxy-poly(ethylene glycol) (mPEG)-poly(caprolactone) (PCL) enable delayed release of curcumin [109].
In the study by Karavasili et al., the activity of peptide hydrogel with simultaneous delivery of doxorubicin and curcumin in the therapy of head and neck cancer cells was examined. The findings showed the therapeutic utility of a double peptide hydrogel with built-in drugs for the local treatment of head and neck cancer [110]. The amylopectin–chitosan composite hydrogel (LRA–CS) for curcumin delivery was synthesized and tested by Liu et al. The release characteristics of encapsulated curcumin in the simulated gastric and intestinal fluid were observed. The findings showed that LRA–CS hydrogel provided stability of curcumin in the stomach and its release in the small intestine [111]. Chitosan–nanocellulose hydrogel with nonionic surfactant was also used for the delivery of curcumin [112]. Cyclodextrin nanospongoid-based hydrogel (CDNS) was used for transdermal codelivery of curcumin (CUR) and resveratrol (RES). Nanosponges enhanced the in vitro release of curcumin 10 times and the release of resveratrol 2.5 times compared with regular curcumin and resveratrol. The combination of CUR–CDNS and RES–CDNS demonstrated a synergistic cytotoxic effect on MCF-7 cells. A hydrogel base was developed with carbomer and propylene glycol, in which CUR-CDNS and RES-CDNS were incorporated. The photostability of curcumin and resveratrol in the CDNS hydrogel formulation increased almost five and seven times, respectively, compared with the hydrogel formulated without CDNS. Curcumin and resveratrol intake is significantly enhanced when delivered using a CDNS-hydrogel base [113]. In Shef et al., curcumin was incorporated into the oxidized cellulose–polyvinyl alcohol hydrogel system by the freezing process. In vitro studies on rats have shown that this can be an effective method for natural wound healing [114]. In the work of Sahin et al., a new, highly bioavailable formulation of curcumin, advanced ultrasol curcumin (AUC), with improved intestinal stability, was developed. In administered doses, AUC effectively improves the pathophysiology of osteoarthritis in experimentally induced osteoarthritis in rats [115]. An overview of curcumin formulations tested on humans and animals is shown in Table 3 and Table 4, respectively.
A study was conducted to compare the oral bioavailability of the newly developed formulation of curcumin Curene® with a formulation of curcumin containing volatile turmeric oil (CP-01) and standard curcuminoids 95%, on healthy volunteers. It was found that the oral bioavailability of Curene® is significantly higher compared with CP-01 and standard curcuminoids (95%) and that it is safe to be administered to healthy people in trial conditions [136]. The anti-inflammatory activity of Longvida® Optimized Curcumin (LC) was examined on two-month-old wild-type mice and GFAP-IL6. LC can alleviate inflammation and minimize neurodegeneration and motoric defects in GFAP-IL6 mice [137]. There is a range of commercial formulations of curcumin with defined bioavailability and pharmacokinetics such as Meriva®, LongVida®, CurQfenTM, MicroActive curcumin, Micronized curcumin, NovaSol® (micellar curcumin) CurcuWin®, BiocurcumaxTM Curcumin C3 Complex®+Bioperine, Cavacurmin® TheracurminTM. Of the commercial formulations, NovaSol® (185), Curcuwin® (136) and LongVida® (100) stand out as they show bioavailability over 100 times higher than the reference curcumin [11]. The formulations CurcuminRich, Biomor, Liposomal curcumin mango, Liposomal curcumin, and Dr. Mercola Curcumin Advanced are also available on the market for oral administration of curcumin [95].

9. Biological Activities of Curcumin

This section describes in detail the biological activities of curcumin, with a special emphasis on its antimicrobial activity (Figure 6).

9.1. Antimicrobial Activity

9.1.1. Antibacterial Activity

Curcumin shows a wide range of antibacterial effects. It causes membrane damage in the cells of Gram-positive (Staphylococcus aureus and Enterococcus faecalis) and Gram-negative (Escherichia coli and Pseudomonas aeruginosa) bacteria [138]. It blocks the growth of bacteria owing to its structural characteristics and the formation of antioxidant products, inhibits bacterial virulence factors and the formation of bacterial biofilm, and prevents bacterial adhesion to host receptors. As a photosensitizer, curcumin induces phototoxicity and inhibits bacterial growth under blue light [139]. In the study by Adamczak et al., the effectiveness of curcumin was tested in vitro on over 100 strains of pathogens within 19 species. The antimicrobial activity was determined by the broth microdilution method and by calculating the minimum inhibitory concentration (MIC). The results confirmed a much higher susceptibility to Gram-positive than to Gram-negative bacteria. The MIC was also high in Staphylococcus aureus, Staphylococcus haemolyticus, Escherichia coli, and Proteus mirabilis resistant to a large number of drugs (≥2000 µg/mL). However, curcumin was effective against some species and strains: Streptococcus pyogenes (mean MIC = 31.25 µg/mL), Staphylococcus aureus sensitive to methicillin (250 µg/mL), Acinetobacter lwoffii (250 µg/mL) and single strains of Enterococcus faecalis and Pseudomonas aeruginosa (62.5 μg/mL). Curcumin shows poor activity against clinical isolates of Candida species. Curcumin can be considered a promising antibacterial agent but with very selective activity [140]. The antimicrobial activity of curcumin against pathogens in burn wounds is shown in Table 5.

9.1.2. Antiviral Activity

The antiviral effect of curcumin is manifested through interference with virus replication or through suppression of cellular signalling pathways that are essential for virus replication, such as the phosphatidylinositol 3-kinase/protein kinase B (PI3K/Akt) and NF-κB [142]. Curcumin exhibits antiviral activity against DNA and RNA viruses [143]. Jeong et al. established a mechanism by which curcumin pretreatment controlled the early stage of viral haemorrhagic septicaemia virus (VHSV) infection in fathead minnow cells. By rearranging the F-actin/G-actin ratio through reduced regulation of heat shock cognate 71 (HSC71), virus entry into cells is suppressed [144]. Ferreira et al. found that curcumin significantly reduced the replication of HIV-1 and herpes simplex virus type 2 (HSV-2) in chronically infected T cells and human primary genital epithelial cells [145]. Curcumin is an inhibitor of the redox function apurinic/apyrimidinic endonuclease 1 (APE1), affecting many genes, thus accounting for the wide range of curcumin effects on various human diseases. Curcumin effectively blocks the replication of the herpes virus associated with sarcoma and inhibits the pathogenic processes of angiogenesis and cell invasion [146]. Curcumin also exhibits antiviral activity against Zika and chikungunya viruses, dengue virus, hepatitis C virus [147], coxsackievirus, human papilloma virus, severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2), etc., [147,148,149].
SARS-CoV-2 is an infectious virus that causes coronavirus disease-2019 (COVID-19) [150]. The disease with significant mortality worldwide poses a global threat due to the difficulties in treatment because there is currently no approved antiviral drug with proven efficacy and minor adverse effects [151,152]. The severity of the pandemic has prompted scientists to examine existing drugs with the potential for treating COVID-19 [150]. Studies show that curcumin is a good candidate for treating the COVID-19 virus and preventing fatal complications of this disease due to its thoroughly tested and confirmed anti-inflammatory, antiviral, antinociceptive immunomodulatory, antipyretic, antifibrotic, pulmoprotective, and antifatigue effects. Curcumin can interact with spike proteins or angiotensin 2 (ACE2) proteins in the signalling pathway induced by COVID-19. Curcumin also inhibits several important signalling pathways in viral infection, such as transcription factors (NF-κB, signal transducer and activator of transcription 3 (STAT-3), Vnt/b-catenin, nuclear factor E2-related factor (Nrf2), p38/MAPK, and virus-induced inflammation by modulating the manifestation of various factors (IL-10, Interleukin-18 (IL-18), IL-6, tumour necrosis factor (TNF) α/β and COX-2) in COVID-19 [153,154]. Valizadeha et al. investigated the effect of nanocurcumin on the modulation of inflammatory cytokines in patients with COVID-19. Messenger ribonucleic acid (mRNA) expression and cytokine secretion levels of Interleukin-1 β (IL-1β), IL-6, TNF-α, and IL-18 were assessed by polymerase chain reaction (PCR) in real-time and enzyme-linked immunosorbent assay (ELISA), respectively. The results showed that the expression of IL-1β and IL-6 mRNA was dramatically reduced after nanocurcumin administration. This study suggests that by regulating the inflammatory response, nanocurcumin can be used as an innovative therapeutic agent for patients with COVID-19 [155].

9.1.3. Antiparasitic and Antimalarial Activity

Curcumin shows activity against different types of parasites in vitro and in vivo. The antiprotozoal activity of curcumin is shown against Leishmania major, Leishmania donovani [156,157], Trichomonas vaginalis [158], Entamoeba histolytica [159], Giardia lamblia [160], Toxoplasma gondii [161], Neospora caninum [162], etc. The combinations of netilmicin and curcumin and metronidazole and curcumin show a synergistic effect and can be used to treat leishmaniasis and amoebiasis, respectively [156,159]. Curcumin also exhibits anthelmintic activity on the nematode Ascaridia galli and the cestode Raillietina cesticillus [163,164]. The studies show that curcumin is used to treat malaria and that it can increase the effectiveness of antimalarial drugs [165]. Busari et al. examined the in vivo antiplasmodial activity and the assessment of the toxicity of curcumin incorporated into poly(lactic-co-glycolic) nanoparticles. The formulated drug with nanoparticles demonstrated better activity against the malaria parasite than free curcumin. The antimalarial activity of the drug is better at lower concentrations. In vivo toxicity studies have confirmed the safety of the formulated drug at the tested doses [166]. Novaes et al. evaluated the efficacy of curcumin as a complementary strategy in benznidazole-based chemotherapy in mice acutely infected with Trypanosoma cruzi. The results of the research show that the combination of benznidazole with curcumin may be a relevant therapy in the treatment of Chagas’ disease caused by T. cruzi because it reduces the toxic effects of benznidazole and increases its antiparasitic activity [165]. An overview of the recent research of curcumin’s antiviral, antiparasitic, and antimalarial activities is shown in Table 6.

9.2. Anti-Inflammatory Activity

Reactive oxygen species (ROS) play a key role in enhancing inflammation by activating transcription factors NF-κB, the activator protein 1 (AP-1), in acetylation and deacetylation of nuclear histone in a range of inflammatory diseases [183]. The anti-inflammatory effect of curcumin is based on its ability to inhibit COX-2, lipoxygenase (LOX), inducible nitric oxide synthase (iNOS) [184], arachidonic acid metabolism [185], cytokines (interleukins) [186,187], and tumour necrosis factor [188], NF-κB [184] and the release of steroid hormones [189]. COX-2, LOX, and iNOS are important enzymes that mediate inflammatory processes. Improper regulation of COX-2 and iNOS has been associated with the pathophysiology of certain types of cancer in humans, as well as with inflammatory disorders [184]. Curcumin and rutin downregulate COX-2 and reduce tumour-associated inflammation in HPV16-transgenic mice [190] The findings of preclinical studies in animal models with invasive pneumonia showed that curcumin exhibits a protective effect. It regulates the expression of pro and anti-inflammatory factors (interleukin-6 (IL-6), interleukin-8 (IL-8), interleukin-10 (IL-10), and COX-2), induces apoptosis of polymorphonuclear neutrophilic (PMN) cells, and removes ROS, thus improving the inflammatory response. These studies indicate that curcumin can be used as a therapeutic agent against pneumonia and acute lung injury (ALI)/fatal acute respiratory distress syndrome (ARDS) in humans, resulting from coronavirus infection [191].

9.3. Antioxidant Activity

ROS and reactive nitrogen species (RNS) are generated in the human body in various endogenous systems, in pathophysiological conditions or exposure to various physical and chemical factors. Free radicals can change lipids (lipid peroxidation), proteins (loss of enzyme activity), and DNA (mutagenesis and carcinogenesis); they contribute to ageing and many human diseases. Natural protective antioxidant mechanisms include superoxide dismutase (SOD), catalase (CAT), glutathione (GSH), glutathione peroxidase (GPx), and reductase, vitamin E (tocopherols and tocotrienols), vitamin C, etc. [4]. Curcumin also shows strong antioxidant activity. The antioxidant property is attributed to the presence of various functional groups, including methoxy, phenoxy, and carbon–carbon double bonds in its structure. Curcumin is a classic phenolic antioxidant that donates H atoms from phenolic groups [192]. In the work of Samarghandian et al., it was found that curcumin can inhibit oxidative damage caused by stress in the brain, liver, and kidneys of rats [193]. Lipid peroxidation is significantly reduced in rats treated with curcumin before applying γ-radiation [194]. Curcumin increases enzymatic antioxidant activity by increasing the expression of methionine sulfoxide reductase (MSRA) and increasing the levels of the enzymes MSRA, SOD, CAT and GPx [195]. Curcumin may act as an antioxidant against oxidative stress in rats with diabetes mellitus by increased SOD expression in cochlear fibroblasts [196]. The antioxidant activity of curcumin was assessed using the 1,1-diphenyl-2-picryl hydrazyl (DPPH) radical assay compared to ascorbic acid, a known antioxidant. The percentage of free radical removal using curcumin and ascorbic acid was 69 and 62%, respectively, at a concentration of 0.1 mM. No significant difference was observed between curcumin and ascorbic acid in antioxidant potential [197]. Curcumin has shown a large capacity to remove smaller oxidative molecules such as H2O2, HO•, ROO•. Curcumin can be used as an effective antioxidant to protect against ROS in the cytoplasm of cells [198]. Curcumin formulations with different carriers that are stable and protected from various influences are used as antioxidants [199,200].

9.4. Antinociceptive Activity

Preclinical studies have shown that curcumin has an antinociceptive effect on inflammatory and neuropathic pain. The effects of curcumin on postoperative pain in rats were investigated in the work of Zhu et al. The results of the study show that curcumin can alleviate postoperative pain and accelerate recovery from surgery. However, treatment with curcumin before surgery did not affect the threshold of postoperative pain and recovery rate [201]. The antinociceptive effect of poly(d,l-lactide-co-glycolide) nanovesicles with curcumin (PLGA-CUR) administered intravenously or intrathecally in mice in small and high doses was tested using formalin test, zymosan-induced hyperalgesia and sciatic nerve ligation that causes neuropathic allodynia and hyperalgesia. PLGA-CUR administered intravenously managed to reduce the response to nociceptive stimuli in the formalin test and zymosan-induced hyperalgesia, while pure curcumin was inactive. The low doses of intrathecally administered PLGA-CUR significantly reduced allodynia produced by sciatic nerve ligation. Long-lasting antinociceptive effects were observed when high doses of PLGA-CUR were administered intrathecally. At high doses, intrathecally applied pure curcumin had only rapid and transient antinociceptive effects. Measuring cytokine levels and brain-derived neurotrophic factor (BDNF) in the spinal cord of neuropathic mice shows that the antinociceptive effects of PLGA-CUR depend on the decline in the release of cytokines and BDNF in the spinal cord. The study results show the efficacy of PLGA-CUR and suggest that the nanoformulation of PLGA-CUR could be a new potential drug in the treatment of pain [202].

9.5. Wound Healing Agent

Curcumin has strong modulating effects on the wound healing process. The wound healing process consists of four phases: coagulation, inflammation, proliferation, and tissue remodelling. Curcumin induces apoptosis of inflammatory cells during the early phase of wound healing, inhibits the activity of the transcription factor NF-κB, reduces the production of cytokines (TNF-α and IL-1), removes ROS, affects the production of antioxidant enzymes and thus reduces inflammation and shortens the inflammatory phase in the wound healing process. The studies have shown that during the proliferation phase, curcumin enhances fibroblast migration, enhances granulation tissue formation, collagen deposition, and re-epithelialization. In the final phase of wound healing, by increasing the production of the transforming growth factor β, curcumin enhances wound contractions and therefore increases fibroblast proliferation [203]. Various topical curcumin formulations such as films, fibres, emulsions, hydrogels, and various nanoformulations have been developed for targeted delivery of curcumin at the wounds [203,204,205]. Sodium alginate-g-poly(N-isopropylacrylamide), (Alg-pNIPAM), a thermosensitive hydrogel with incorporated curcumin as an in vivo wound dressing was synthesized by Zakerikhoob et al. In vivo studies have shown accelerated collagenesis, re-epithelialization, and wound contraction using the Alg-pNIPAM formulation with curcumin. The formulation showed a more significant anti-inflammatory effect than the free curcumin solution. Given the antioxidant and anti-inflammatory properties of curcumin and the concomitant effect of alginate in keeping wound areas moist, the developed thermosensitive formulation of curcumin could help accelerate wound healing [204]. An overview of the recent research about the biological activities of curcumin is shown in Table 7.

10. Conclusions

Curcumin is a widely studied natural compound that has exhibited enormous in vitro and in vivo therapeutic potential. Curcumin has anti-inflammatory, antioxidant, antiviral, proapoptotic, chemopreventive, chemotherapeutic, antinociceptive, antiproliferative, antiparasitic, and antimalarial effects and is used as a wound-healing agent. Poor absorption of curcumin in the small intestine, rapid metabolism, and rapid systemic elimination cause poor bioavailability of curcumin in humans. Different curcumin formulations are used for more prolonged circulation, better permeability, and resistance to metabolic processes, and thus to increase the efficacy of curcumin. Liposomes, micelles, phospholipid complexes, cyclodextrins, nanoparticles, emulsions, hydrogels, and phytosomes have been described in the reference sources as promising formulations that improve the physicochemical properties of curcumin and enable its safe and efficient use.

Author Contributions

Conceptualization, V.N.; methodology, V.N. and L.N.; investigation, M.U., I.G., A.D. and V.M.; writing—original draft preparation, M.U., I.G., A.D. and V.M.; writing—review and editing, V.N., L.N., M.U., I.G., A.D. and V.M.; visualization, V.N. and L.N.; supervision, V.N. and L.N. All authors have read and agreed to the published version of the manuscript.

Funding

The Republic of Serbia-Ministry of Education, Science, and Technological Development, Program for financing scientific research work, number 451-03-9/2021-14/200133.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data available in a publicly accessible repository.

Acknowledgments

Republic of Serbia-Ministry of Education, Science and Technological Development, Program for financing scientific research work, number 451-03-9/2021-14/200133.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hosseini, A.; Hosseinzadeh, H. Antidotal or protective effects of Curcuma longa (turmeric) and its active ingredient, curcumin, against natural and chemical toxicities: A review. Biomed. Pharmacother. 2018, 99, 411–421. [Google Scholar] [CrossRef] [PubMed]
  2. Toden, S.; Goel, A. The Holy Grail of Curcumin and its Efficacy in Various Diseases: Is Bioavailability Truly a Big Concern? J. Restor. Med. 2017, 6, 27–36. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Den Hartogh, D.J.; Gabriel, A.; Tsiani, E. Antidiabetic Properties of Curcumin II: Evidence from In Vivo Studies. Nutrients 2020, 12, 58. [Google Scholar] [CrossRef] [Green Version]
  4. Arshad, L.; Haque, M.A.; Abbas Bukhari, S.N.; Jantan, I. An overview of structure–activity relationship studies of curcumin analogs as antioxidant and anti-inflammatory agents. Future Med. Chem. 2017, 9, 605–626. [Google Scholar] [CrossRef] [PubMed]
  5. Hassan, F.U.; Rehman, M.S.-u.; Khan, M.S.; Ali, M.A.; Javed, A.; Nawaz, A.; Yang, C. Curcumin as an Alternative Epigenetic Modulator: Mechanism of Action and Potential Effects. Front. Genet. 2019, 10, 514. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Giordano, A.; Tommonaro, G. Curcumin and Cancer. Nutrients 2019, 11, 2376. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Heng, M. Phosphorylase Kinase Inhibition Therapy in Burns and Scalds. BioDiscovery 2017, 20, e11207. [Google Scholar] [CrossRef] [Green Version]
  8. 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]
  9. Purpura, M.; Lowery, R.P.; Wilson, J.M.; Mannan, H.; Münch, 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] [Green Version]
  10. Hewlings, S.J.; Kalman, D.S. Curcumin: A Review of Its Effects on Human Health. Foods 2017, 6, 92. [Google Scholar] [CrossRef]
  11. Jamwal, R. Bioavailable curcumin formulations: A review of pharmacokinetic studies in healthy volunteers. J. Integr. Med. 2018, 16, 367–374. [Google Scholar] [CrossRef]
  12. Ma, Z.; Wang, N.; He, H.; Tang, X. Pharmaceutical strategies of improving oral systemic bioavailability of curcumin for clinical application. J. Control. Release 2019, 316, 359–380. [Google Scholar] [CrossRef]
  13. 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]
  14. Kocaadam, B.; Şanlier, 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]
  15. Almeida, H.H.; Barros, L.; Barreira, J.C.; Calhelha, R.C.; Heleno, S.A.; Sayer, C.; Miranda, C.G.; Leimann, F.V.; Barreiro, M.F.; Ferreira, I.C. Bioactive evaluation and application of different formulations of the natural colorant curcumin (E100) in a hydrophilic matrix (yogurt). Food Chem. 2018, 261, 224–232. [Google Scholar] [CrossRef] [Green Version]
  16. Vogel, H.; Pelletier, J. Curcumin-biological and medicinal properties. J. Pharma 1815, 2, 24–29. [Google Scholar]
  17. Vogel, A. Memoire sur la Curcumine. J. Pharm. Chim. 1842, 3, 20–27. [Google Scholar]
  18. Milobedzka, J.; Kostanecki, S.; Lampe, V. Zur Kenntnis des Curcumins. Ber. Deut. Chem. Ges. 1910, 43, 2163–2170. [Google Scholar] [CrossRef] [Green Version]
  19. Lampe, V.; Milobedzka, J. Studien über Curcumin. Ber. Deut. Chem. Ges. 1913, 46, 2235–2240. [Google Scholar] [CrossRef]
  20. Schraufstätter, E.; Bernt, H. Antibacterial Action of Curcumin and Related Compounds. Nature 1949, 164, 456–457. [Google Scholar] [CrossRef]
  21. Srinivasan, K.R. A Chromatographic Study of the Curcuminoids in Curcuma Longa, L. Pharm. Pharmacol. 1953, 5, 448–457. [Google Scholar] [CrossRef] [PubMed]
  22. Patil, T.N.; Srinivasan, M. Hypocholesteremic effect of curcumin in induced hypercholesteremic rats. Indian J. Exp. Boil. 1971, 9, 167–169. [Google Scholar]
  23. Srinivasan, M. Effect of curcumin on blood sugar as seen in a diabetic subject. Indian J. Med. Sci. 1972, 26, 269–270. [Google Scholar] [PubMed]
  24. Srimal, R.C.; Dhawan, B.N. Pharmacology of diferuloyl methane (curcumin), a non-steroidal anti-inflammatory agent. J. Pharm. Pharmacol. 1973, 25, 447–452. [Google Scholar] [CrossRef] [PubMed]
  25. Sharma, O.P. Antioxidant activity of curcumin and related compounds. Biochem. Pharmacol. 1976, 25, 1811–1812. [Google Scholar] [CrossRef]
  26. Kuttan, R.; Bhanumathy, P.; Nirmala, K.; George, M.C. Potential anticancer activity of turmeric (Curcuma longa). Cancer Lett. 1985, 29, 197–202. [Google Scholar] [CrossRef]
  27. Singh, S.; Aggarwal, B.B. Activation of Transcription Factor NF-κB Is Suppressed by Curcumin (Diferuloylmethane). J. Biol. Chem. 1995, 270, 24995–25000. [Google Scholar] [CrossRef] [Green Version]
  28. Guimarães, A.F.; Vinhas, A.C.A.; Gomes, A.F.; Souza, L.H.; Krepsky, P.B. Essential Oil of Curcuma longa L. Rhizomes Chemical Composition, Yield Variation and Stability. Quím. Nova 2020, 43, 909–913. [Google Scholar] [CrossRef]
  29. Pawar, H.A.; Gavasane, A.J.; Choudhary, P.D. A Novel and Simple Approach for Extraction and Isolation of Curcuminoids from Turmeric Rhizomes. Nat. Prod. Chem. Res. 2018, 6, 1–4. [Google Scholar] [CrossRef]
  30. Tripathy, S.; Verma, D.K.; Thakur, M.; Patel, A.R.; Srivastav, P.P.; Singh, S.; Gupta, A.K.; Chávez-González, M.L.; Aguilar, C.N.; Chakravorty, N.; et al. Curcumin Extraction, Isolation, Quantification and Its Application in Functional Foods: A Review with a Focus on Immune Enhancement Activities and COVID-19. Front. Nutr. 2021, 8, 675. [Google Scholar] [CrossRef]
  31. Sahne, F.; Mohammadi, M.; Najafpour, G.D.; Moghadamnia, A.A. Extraction of bioactive compound curcumin from turmeric (Curcuma longa L.) via different routes: A comparative study. Pak. J. Biotechnol. 2016, 13, 173–180. [Google Scholar]
  32. Monton, C.; Settharaksa, S.; Luprasong, C.; Songsak, T. An optimization approach of dynamic maceration of Centella asiatica to obtain the highest content of four centelloids by response surface methodology. Rev. Bras. Farmacogn. 2019, 29, 254–261. [Google Scholar] [CrossRef]
  33. Patil, S.S.; Rathod, V.K. Synergistic Effect of Ultrasound and Three Phase Partitioning for the Extraction of Curcuminoids from Curcuma longa and its Bioactivity Profile. Process Biochem. 2020, 93, 85–93. [Google Scholar] [CrossRef]
  34. Sahne, F.; Mohammadi, M.; Najafpour, G.D.; Moghadamnia, A.A. Enzyme-assisted ionic liquid extraction of bioactive compound from turmeric (Curcuma longa L.): Isolation, purification and analysis of curcumin. Ind. Crops Prod. 2017, 95, 686–694. [Google Scholar] [CrossRef]
  35. Liang, H.; Wang, W.; Xu, J.; Zhang, Q.; Shen, Z.; Zeng, Z.; Li, Q. Optimization of ionic liquid-based microwave-assisted extraction technique for curcuminoids from Curcuma longa L. Food Bioprod. Process. 2017, 104, 57–65. [Google Scholar] [CrossRef]
  36. Nagavekar, N.; Singhal, R.S. Supercritical fluid extraction of Curcuma longa and Curcuma amada oleoresin: Optimization of extraction conditions, extract profiling, and comparison of bioactivities. Ind. Crops Prod. 2019, 134, 134–145. [Google Scholar] [CrossRef]
  37. Chao, I.-C.; Wang, C.-M.; Marcotullio, M.C.; Lin, L.-G.; Ye, W.-C.; Zhang, Q.-W. Simultaneous Quantification of Three Curcuminoids and Three Volatile Components of Curcuma longa Using Pressurized Liquid Extraction and High-Performance Liquid Chromatography. Molecules 2018, 23, 1568. [Google Scholar] [CrossRef] [Green Version]
  38. Nurjanah, N.; Saepudin, E. Curcumin isolation, synthesis and characterization of curcumin isoxazole derivative compound. AIP Conf. Proc. 2019, 2168, 020065. [Google Scholar] [CrossRef]
  39. Muthukumar, V.P.; Vaishnavi, M.; Theepapriys, S.; Saravanaraj, A. Process Development for the Effective Extraction of Curcumin from Curcuma Longa L (Turmeric). Int. J. Eng. Technol. 2018, 7, 151–155. [Google Scholar] [CrossRef] [Green Version]
  40. Yadav, D.K.; Sharma, K.; Dutta, A.; Kundu, A.; Awasthi, A.; Goon, A.; Banerjee, K.; Saha, S. Purity Evaluation of Curcuminoids in the Turmeric Extract Obtained by Accelerated Solvent Extraction. J. AOAC Int. 2017, 100, 586–591. [Google Scholar] [CrossRef]
  41. Naksuriya, O.; Van Steenbergen, M.J.; Toraño, J.S.; Okonogi, S.; Hennink, W.E. A Kinetic Degradation Study of Curcumin in Its Free Form and Loaded in Polymeric Micelles. AAPS J. 2016, 18, 777–787. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Liu, Y.; Li, J.; Fu, R.; Zhang, L.; Wang, D.; Wang, S. Enhanced extraction of natural pigments from Curcuma longa L. using natural deep eutectic solvents. Ind. Crop. Prod. 2019, 140, 111620. [Google Scholar] [CrossRef]
  43. Nair, D.S.; Krishnakumar, K.; Krishnan, B. Pharmacological profile of curcumin: A review. J. Biol. Innov. 2017, 6, 533–541. [Google Scholar]
  44. Rathore, S.; Mukim, M.; Sharma, P.; Devi, S.; Nagar, J.C.; Khalid, M. Curcumin: A Review for Health Benefits. Int. J. Res. Rev. 2020, 7, 273–290. [Google Scholar]
  45. Nelson, K.M.; Dahlin, J.L.; Bisson, J.; Graham, J.; Pauli, G.F.; Walters, M.A. The Essential Medicinal Chemistry of Curcumin: Miniperspective. J. Med. Chem. 2017, 60, 1620–1637. [Google Scholar] [CrossRef]
  46. Zielińska, A.; Alves, H.; Marques, V.; Durazzo, A.; Lucarini, M.; Alves, T.; Morsink, M.; Willemen, N.; Eder, P.; Chaud, M.; et al. Properties, Extraction Methods, and Delivery Systems for Curcumin as a Natural Source of Beneficial Health Effects. Medicina 2020, 56, 336. [Google Scholar] [CrossRef]
  47. Angelini, G.; Pasc, A.; Gasbarri, C. Curcumin in silver nanoparticles aqueous solution: Kinetics of keto-enol tautomerism and effects on AgNPs. Colloids Surf. A Physicochem. Eng. Asp. 2020, 603, 125235. [Google Scholar] [CrossRef]
  48. Girardon, M.; Parant, S.; Monari, A.; Dehez, F.; Chipot, C.; Rogalska, E.; Canilho, N.; Pasc, A. Triggering Tautomerization of Curcumin by Confinement into Liposomes. ChemPhotoChem 2019, 3, 1034–1041. [Google Scholar] [CrossRef]
  49. Rege, S.A.; Megha, A.; Momin, S.A. Mini review on Keto-Enol ratio of curcuminoids. Ukr. J. Food Sci. 2019, 7, 27–32. [Google Scholar] [CrossRef]
  50. Manolova, Y.; Deneva, V.; Antonov, L.; Drakalska, E.; Momekova, D.; Lambov, N. The effect of the water on the curcumin tautomerism: A quantitative approach. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2014, 132, 815–820. [Google Scholar] [CrossRef]
  51. Kawano, S.-I.; Inohana, Y.; Hashi, Y.; Lin, J.-M. Analysis of keto-enol tautomers of curcumin by liquid chromatography/mass spectrometry. Chin. Chem. Lett. 2013, 24, 685–687. [Google Scholar] [CrossRef]
  52. Liu, J.; Wang, H.; Wang, P.; Guo, M.; Jiang, S.; Li, X.; Jiang, S. Films based on κ-carrageenan incorporated with curcumin for freshness monitoring. Food Hydrocoll. 2018, 83, 134–142. [Google Scholar] [CrossRef]
  53. Yang, H.; Du, Z.; Wang, W.; Song, M.; Sanidad, K.; Sukamtoh, E.; Zheng, J.; Tian, L.; Xiao, H.; Liu, Z.; et al. Structure–Activity Relationship of Curcumin: Role of the Methoxy Group in Anti-inflammatory and Anticolitis Effects of Curcumin. J. Agric. Food Chem. 2017, 65, 4509–4515. [Google Scholar] [CrossRef] [PubMed]
  54. Heger, M.; Van Golen, R.F.; Broekgaarden, M.; Michel, M.C. The Molecular Basis for the Pharmacokinetics and Pharmacodynamics of Curcumin and Its Metabolites in Relation to Cancer. Pharmacol. Rev. 2014, 66, 222–307. [Google Scholar] [CrossRef] [PubMed]
  55. Gordon, O.N.; Luis, P.B.; Sintim, H.O.; Schneider, C. Unraveling Curcumin Degradation. J. Biol. Chem. 2015, 290, 4817–4828. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Zhu, J.; Sanidad, K.Z.; Sukamtoh, E.; Zhang, G. Potential roles of chemical degradation in the biological activities of curcumin. Food Funct. 2017, 8, 907–914. [Google Scholar] [CrossRef]
  57. Soleimani, V.; Sahebkar, A.; Hosseinzadeh, H. Turmeric (Curcuma longa) and its major constituent (curcumin) as nontoxic and safe substances: Review. Phytother. Res. 2018, 32, 985–995. [Google Scholar] [CrossRef]
  58. Greil, R.; Greil-Ressler, S.; Weiss, L.; Schönlieb, C.; Magnes, T.; Radl, B.; Bolger, G.T.; Vcelar, B.; Sordillo, P.P. A phase 1 dose-escalation study on the safety, tolerability and activity of liposomal curcumin (Lipocurc™) in patients with locally advanced or metastatic cancer. Cancer Chemother. Pharmacol. 2018, 82, 695–706. [Google Scholar] [CrossRef] [Green Version]
  59. Saghatelyan, T.; Tananyan, A.; Janoyan, N.; Tadevosyan, A.; Petrosyan, H.; Hovhannisyan, A.; Hayrapetyan, L.; Arustamyan, M.; Arnhold, J.; Rotmann, A.-R.; et al. Efficacy and safety of curcumin in combination with paclitaxel in patients with advanced, metastatic breast cancer: A comparative, randomized, double-blind, placebo-controlled clinical trial. Phytomedicine 2020, 70, 153218. [Google Scholar] [CrossRef]
  60. Prasad, S.; Tyagi, A.K.; Aggarwal, B.B. Recent Developments in Delivery, Bioavailability, Absorption and Metabolism of Curcumin: The Golden Pigment from Golden Spice. Cancer Res. Treat. 2014, 46, 2–18. [Google Scholar] [CrossRef] [Green Version]
  61. Dei Cas, M.; Ghidoni, R. Dietary Curcumin: Correlation between Bioavailability and Health Potential. Nutrients 2019, 11, 2147. [Google Scholar] [CrossRef] [Green Version]
  62. Kotha, R.R.; Luthria, D.L. Curcumin: Biological, Pharmaceutical, Nutraceutical, and Analytical Aspects. Molecules 2019, 24, 2930. [Google Scholar] [CrossRef] [Green Version]
  63. 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]
  64. Ireson, C.; Orr, S.; Jones, D.J.; Verschoyle, R.; Lim, C.K.; Luo, J.L.; Howells, L.; Plummer, S.; Jukes, R.; Williams, M.; et al. Characterization of metabolites of the chemopreventive agent curcumin in human and rat hepatocytes and in the rat in vivo, and evaluation of their ability to inhibit phorbol ester-induced prostaglandin E2 production. Cancer Res. 2001, 61, 1059–1064. [Google Scholar]
  65. Schneider, C.; Gordon, O.N.; Edwards, R.L.; Luis, P.B. Degradation of Curcumin: From Mechanism to Biological Implications. J. Agric. Food Chem. 2015, 63, 7606–7614. [Google Scholar] [CrossRef] [Green Version]
  66. Aggarwal, B.B.; Deb, L.; Prasad, S. Curcumin Differs from Tetrahydrocurcumin for Molecular Targets, Signaling Pathways and Cellular Responses. Molecules 2015, 20, 185–205. [Google Scholar] [CrossRef] [Green Version]
  67. Zaghary, W.; Hanna, E.; Zanoun, M.; Abdallah, N.; Sakr, T. Curcumin: Analysis and Stability. J. Adv. Pharm. Res. 2019, 3, 47–58. [Google Scholar] [CrossRef]
  68. Kotra, V.S.R.; Satyabanta, L.; Goswami, T.K. A critical review of analytical methods for determination of curcuminoids in turmeric. J. Food Sci. Technol. 2019, 56, 5153–5166. [Google Scholar] [CrossRef]
  69. Kushwaha, P.; Shukla, B.; Dwivedi, J.; Saxena, S. Validated high-performance thin-layer chromatographic analysis of curcumin in the methanolic fraction of Curcuma longa L. rhizomes. Futur. J. Pharm. Sci. 2021, 7, 178. [Google Scholar] [CrossRef]
  70. Jadhav, B.-K.; Mahadik, K.-R.; Paradkar, A.-R. Development and Validation of Improved Reversed Phase-HPLC Method for Simultaneous Determination of Curcumin, Demethoxycurcumin and Bis-Demethoxycurcumin. Chromatographia 2007, 65, 483–488. [Google Scholar] [CrossRef]
  71. Wichitnithad, W.; Jongaroonngamsang, N.; Pummangura, S.; Rojsitthisak, P. A simple isocratic HPLC method for the simultaneous determination of curcuminoids in commercial turmeric extracts. Phytochem. Anal. 2009, 20, 314–319. [Google Scholar] [CrossRef]
  72. Jayaprakasha, G.K.; Rao, L.J.M.; Sakariah, K.K. Improved HPLC Method for the Determination of Curcumin, Demethoxycurcumin, and Bisdemethoxycurcumin. J. Agric. Food Chem. 2002, 50, 3668–3672. [Google Scholar] [CrossRef]
  73. Thorat, B.; Jangle, R. Reversed-phase High-performance Liquid Chromatography Method for Analysis of Curcuminoids and Curcuminoid-loaded Liposome Formulation. Indian J. Pharm. Sci. 2013, 75, 60–66. [Google Scholar] [CrossRef] [Green Version]
  74. Radha, A.; Ragavendran, P.; Thomas, A.; Kumar, D.S. A cost effective hplc method for the analysis of curcuminoids. Hygeia J. Drugs Med. 2016, 8, 1–15. [Google Scholar] [CrossRef]
  75. Nugroho, A.; Lukitaning, E.; Rakhmawati, N.; Rohman, A. Analysis of Curcumin in Ethanolic Extract of Curcuma longa Linn. and Curcuma xanthorriza Roxb. Using High Performance Liquid Chromatography with UV-Detection. Res. J. Phytochem. 2015, 9, 188–194. [Google Scholar] [CrossRef] [Green Version]
  76. Hwang, K.-W.; Son, D.; Jo, H.-W.; Kim, C.H.; Seong, K.C.; Moon, J.-K. Levels of curcuminoid and essential oil compositions in turmerics (Curcuma longa L.) grown in Korea. J. Appl. Biol. Chem. 2016, 59, 209–215. [Google Scholar] [CrossRef]
  77. Na Bhuket, P.R.; Niwattisaiwong, N.; Limpikirati, P.; Khemawoot, P.; Towiwat, P.; Ongpipattanakul, B.; Rojsitthisak, P. Simultaneous determination of curcumin diethyl disuccinate and its active metabolite curcumin in rat plasma by LC–MS/MS: Application of esterase inhibitors in the stabilization of an ester-containing prodrug. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2016, 1033–1034, 301–310. [Google Scholar] [CrossRef]
  78. Ma, W.; Wang, J.; Guo, Q.; Tu, P. Simultaneous determination of doxorubicin and curcumin in rat plasma by LC–MS/MS and its application to pharmacokinetic study. J. Pharm. Biomed. Anal. 2015, 111, 215–221. [Google Scholar] [CrossRef]
  79. Ramalingam, P.; Ko, Y.T. A validated LC-MS/MS method for quantitative analysis of curcumin in mouse plasma and brain tissue and its application in pharmacokinetic and brain distribution studies. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2014, 969, 101–108. [Google Scholar] [CrossRef]
  80. Ashraf, K.; Mujeeb, M.; Ahmad, A.; Ahmad, N.; Amir, M. Determination of Curcuminoids in Curcuma longa Linn. by UPLC/Q-TOF–MS: An Application in Turmeric Cultivation. J. Chromatogr. Sci. 2015, 53, 1346–1352. [Google Scholar] [CrossRef] [Green Version]
  81. Van Nong, H.; Hung, L.X.; Thang, P.N.; Chinh, V.D.; Van Vu, L.; Dung, P.T.; Van Trung, T.; Nga, P.T. Fabrication and vibration characterization of curcumin extracted from turmeric (Curcuma longa) rhizomes of the northern Vietnam. SpringerPlus 2016, 5, 1147. [Google Scholar] [CrossRef] [Green Version]
  82. Sathisaran, I.; Dalvi, S.V. Crystal Engineering of Curcumin with Salicylic Acid and Hydroxyquinol as Coformers. Cryst. Growth Des. 2017, 17, 3974–3988. [Google Scholar] [CrossRef]
  83. Thangavel, K.; Dhivya, K. Determination of curcumin, starch and moisture content in turmeric by Fourier transform near infrared spectroscopy (FT-NIR). Eng. Agric. Environ. Food 2019, 12, 264–269. [Google Scholar] [CrossRef]
  84. Pöppler, A.-C.; Lübtow, M.M.; Schlauersbach, J.; Wiest, J.; Meinel, L.; Luxenhofer, R. Loading-Dependent Structural Model of Polymeric Micelles Encapsulating Curcumin by Solid-State NMR Spectroscopy. Angew. Chem. Int. Ed. 2019, 58, 18540–18546. [Google Scholar] [CrossRef] [Green Version]
  85. Ali, Z.; Saleem, M.; Atta, B.M.; Khan, S.S.; Hammad, G. Determination of curcuminoid content in turmeric using fluorescence spectroscopy. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2019, 213, 192–198. [Google Scholar] [CrossRef]
  86. Pandey, K.U.; Dalvi, S.V. Understanding stability relationships among three curcumin polymorphs. Adv. Powder Technol. 2019, 30, 266–276. [Google Scholar] [CrossRef]
  87. Iravani, S.; Soufi, G.J. Electron paramagnetic resonance (EPR) spectroscopy: Food, biomedical and pharmaceutical analysis. Biomed. Spectrosc. Imaging 2020, 9, 165–182. [Google Scholar] [CrossRef]
  88. Dudylina, A.L.; Ivanova, M.V.; Shumaev, K.B.; Ruuge, E.K. Superoxide Formation in Cardiac Mitochondria and Effect of Phenolic Antioxidants. Cell Biochem. Biophys. 2018, 77, 99–107. [Google Scholar] [CrossRef]
  89. Morales, N.P.; Sirijaroonwong, S.; Yamanont, P.; Phisalaphong, C. Electron Paramagnetic Resonance Study of the Free Radical Scavenging Capacity of Curcumin and Its Demethoxy and Hydrogenated Derivatives. Biol. Pharm. Bull. 2015, 38, 1478–1483. [Google Scholar] [CrossRef] [Green Version]
  90. Nikolić, I.; Mitsou, E.; Damjanović, A.; Papadimitriou, V.; Antić-Stanković, J.; Stanojevic, B.; Xenakis, A.; Savić, S. Curcumin-loaded low-energy nanoemulsions: Linking EPR spectroscopy-analysed microstructure and antioxidant potential with in vitro evaluated biological activity. J. Mol. Liq. 2020, 301, 112479. [Google Scholar] [CrossRef]
  91. Gopi, S.; Ac, K.V.; Varma, K.; Jude, S.; Amalraj, A.; Arundhathy, C.; George, R.; Sreeraj, T.; 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]
  92. Jäger, 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] [Green Version]
  93. Baspinar, Y.; Üstündas, M.; Bayraktar, O.; Sezgin, C. Curcumin and piperine loaded zein-chitosan nanoparticles: Development and in-vitro characterisation. Saudi Pharm. J. 2018, 26, 323–334. [Google Scholar] [CrossRef]
  94. Kim, L.; Kim, J.Y. Chondroprotective effect of curcumin and lecithin complex in human chondrocytes stimulated by IL-1β via an anti-inflammatory mechanism. Food Sci. Biotechnol. 2019, 28, 547–553. [Google Scholar] [CrossRef]
  95. Henriques, M.C.; Faustino, M.A.F.; Braga, S.S. Curcumin Innovative Delivery Forms: Paving the “Yellow Brick Road” of Antitumoral Phytotherapy. Appl. Sci. 2020, 10, 8990. [Google Scholar] [CrossRef]
  96. Liu, W.; Zhai, Y.; Heng, X.; Che, F.Y.; Chen, W.; Sun, D.; Zhai, G. Oral bioavailability of curcumin: Problems and advancements. J. Drug Target. 2016, 24, 694–702. [Google Scholar] [CrossRef]
  97. 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] [Green Version]
  98. Kasapoglu-Calik, M.; Ozdemir, M. Synthesis and controlled release of curcumin-β-cyclodextrin inclusion complex from nanocomposite poly(N-isopropylacrylamide/sodium alginate) hydrogels. J. Appl. Polym. Sci. 2019, 136, 47554. [Google Scholar] [CrossRef]
  99. Kongkaneramit, L.; Aiemsumang, P.; Kewsuwan, P. Development of curcumin liposome formulations using polyol dilution method. Songklanakarin J. Sci. Technol. 2016, 38, 605–610. [Google Scholar]
  100. Tai, K.; Rappolt, M.; Mao, L.; Gao, Y.; Yuan, F. Stability and release performance of curcumin-loaded liposomes with varying content of hydrogenated phospholipids. Food Chem. 2020, 326, 126973. [Google Scholar] [CrossRef]
  101. Cuomo, F.; Cofelice, M.; Venditti, F.; Ceglie, A.; Miguel, M.; Lindman, B.; Lopez, F. In-vitro digestion of curcumin loaded chitosan-coated liposomes. Colloids Surf. B Biointerfaces 2018, 168, 29–34. [Google Scholar] [CrossRef]
  102. Algahtani, M.S.; Ahmad, M.Z.; Ahmad, J. Nanoemulsion loaded polymeric hydrogel for topical delivery of curcumin in psoriasis. J. Drug Deliv. Sci. Technol. 2020, 59, 101847. [Google Scholar] [CrossRef]
  103. Guerrero, S.; Inostroza-Riquelme, M.; Contreras-Orellana, P.; Diaz-Garcia, V.; Lara, P.; Vivanco-Palma, A.; Cárdenas, A.; Miranda, V.; Robert, P.; Leyton, L.; et al. Curcumin-loaded nanoemulsion: A new safe and effective formulation to prevent tumor reincidence and metastasis. Nanoscale 2018, 10, 22612–22622. [Google Scholar] [CrossRef]
  104. Cheng, Y.-H.; Ko, Y.-C.; Chang, Y.-F.; Huang, S.-H.; Liu, C.J.-L. Thermosensitive chitosan-gelatin-based hydrogel containing curcumin-loaded nanoparticles and latanoprost as a dual-drug delivery system for glaucoma treatment. Exp. Eye Res. 2019, 179, 179–187. [Google Scholar] [CrossRef]
  105. Gera, M.; Sharma, N.; Ghosh, M.; Huynh, D.L.; Lee, S.J.; Min, T.; Kwon, T.; Jeong, D.K. Nanoformulations of curcumin: An emerging paradigm for improved remedial application. Oncotarget 2017, 8, 66680–66698. [Google Scholar] [CrossRef] [Green Version]
  106. Saber-Moghaddam, N.; Salari, S.; Hejazi, S.; Amini, M.; Taherzadeh, Z.; Eslami, S.; Rezayat, S.M.; Jaafari, M.R.; Elyasi, S. Oral nano-curcumin formulation efficacy in management of mild to moderate 28 hospitalized coronavirus disease -19 patients: An open label nonrandomized clinical trial. Phytother. Res. 2021, 35, 2616–2623. [Google Scholar] [CrossRef]
  107. Liu, Y.; Huang, P.; Hou, X.; Yan, F.; Jiang, Z.; Shi, J.; Xie, X.; Shen, J.; Fan, Q.; Wang, Z.; et al. Hybrid curcumin–phospholipid complex-near-infrared dye oral drug delivery system to inhibit lung metastasis of breast cancer. Int. J. Nanomed. 2019, 14, 3311–3330. [Google Scholar] [CrossRef] [Green Version]
  108. Wang, J.; Wang, L.; Zhang, L.; He, D.; Ju, J.; Li, W. Studies on the curcumin phospholipid complex solidified with Soluplus®. J. Pharm. Pharmacol. 2018, 70, 242–249. [Google Scholar] [CrossRef]
  109. Gupta, A.; Costa, A.P.; Xu, X.; Lee, S.-L.; Cruz, C.N.; Bao, Q.; Burgess, D.J. Formulation and characterization of curcumin loaded polymeric micelles produced via continuous processing. Int. J. Pharm. 2020, 583, 119340. [Google Scholar] [CrossRef]
  110. Karavasili, C.; Andreadis, D.A.; Katsamenis, O.L.; Panteris, E.; Anastasiadou, P.; Kakazanis, Z.; Zoumpourlis, V.; Markopoulou, C.K.; Koutsopoulos, S.; Vizirianakis, I.S.; et al. Synergistic Antitumor Potency of a Self-Assembling Peptide Hydrogel for the Local Co-delivery of Doxorubicin and Curcumin in the Treatment of Head and Neck Cancer. Mol. Pharm. 2019, 16, 2326–2341. [Google Scholar] [CrossRef]
  111. Liu, K.; Huang, R.-L.; Zha, X.-Q.; Li, Q.-M.; Pan, L.-H.; Luo, J.-P. Encapsulation and sustained release of curcumin by a composite hydrogel of lotus root amylopectin and chitosan. Carbohydr. Polym. 2020, 232, 115810. [Google Scholar] [CrossRef]
  112. Gunathilake, T.M.S.U.; Ching, Y.C.; Chuah, C.H.; Illias, H.A.; Ching, K.Y.; Singh, R.; Nai-Shang, L. Influence of a nonionic surfactant on curcumin delivery of nanocellulose reinforced chitosan hydrogel. Int. J. Biol. Macromol. 2018, 118, 1055–1064. [Google Scholar] [CrossRef]
  113. Pushpalatha, R.; Selvamuthukumar, S.; Kilimozhi, D. Cyclodextrin nanosponge based hydrogel for the transdermal co-delivery of curcumin and resveratrol: Development, optimization, in vitro and ex vivo evaluation. J. Drug Deliv. Sci. Technol. 2019, 52, 55–64. [Google Scholar] [CrossRef]
  114. Shefa, A.A.; Sultana, T.; Park, M.K.; Lee, S.Y.; Gwon, J.-G.; Lee, B.-T. Curcumin incorporation into an oxidized cellulose nanofiber-polyvinyl alcohol hydrogel system promotes wound healing. Mater. Des. 2020, 186, 108313. [Google Scholar] [CrossRef]
  115. Sahin, K.; Orhan, C.; Er, B.; Durmus, A.S.; Ozercan, I.H.; Sahin, N.; Padigaru, M.; Morde, A.; Rai, D. Protective Effect of a Novel Highly Bioavailable Formulation of Curcumin in Experimentally Induced Osteoarthritis Rat Model. Curr. Dev. Nutr. 2020, 4 (Suppl. 2), 1765. [Google Scholar] [CrossRef]
  116. Fakhri, S.; Shakeryan, S.; Alizadeh, A.; Shahryari, A. Effect of 6 Weeks of High Intensity Interval Training with Nano curcumin Supplement on Antioxidant Defense and Lipid Peroxidation in Overweight Girls- Clinical Trial. Iran. J. Diabetes Obes. 2020, 11, 173–180. [Google Scholar] [CrossRef]
  117. Bateni, Z.; Rahimi, H.R.; Hedayati, M.; Afsharian, S.; Goudarzi, R.; Sohrab, G. The effects of nano-curcumin supplementation on glycemic control, blood pressure, lipid profile, and insulin resistance in patients with the metabolic syndrome: A randomized, double-blind clinical trial. Phytother. Res. 2021, 35, 3945–3953. [Google Scholar] [CrossRef] [PubMed]
  118. Asadi, S.; Gholami, M.S.; Siassi, F.; Qorbani, M.; Khamoshian, K.; Sotoudeh, G. Nano curcumin supplementation reduced the severity of diabetic sensorimotor polyneuropathy in patients with type 2 diabetes mellitus: A randomized double-blind placebo- controlled clinical trial. Complement. Ther. Med. 2019, 43, 253–260. [Google Scholar] [CrossRef] [PubMed]
  119. Abdolahi, M.; Sarraf, P.; Javanbakht, M.H.; Honarvar, N.M.; Hatami, M.; Soveyd, N.; Tafakhori, A.; Sedighiyan, M.; Djalali, M.; Jafarieh, A.; et al. A Novel Combination of ω-3 Fatty Acids and Nano-Curcumin Modulates Interleukin-6 Gene Expression and High Sensitivity C-reactive Protein Serum Levels in Patients with Migraine: A Randomized Clinical Trial Study. CNS Neurol. Disord. Drug Targets 2018, 17, 430–438. [Google Scholar] [CrossRef] [PubMed]
  120. Jazayeri-Tehrani, S.A.; Rezayat, S.M.; Mansouri, S.; Qorbani, M.; Alavian, S.M.; Daneshi-Maskooni, M.; Hosseinzadeh-Attar, M.J. Nano-curcumin improves glucose indices, lipids, inflammation, and Nesfatin in overweight and obese patients with non-alcoholic fatty liver disease (NAFLD): A double-blind randomized placebo-controlled clinical trial. Nutr. Metab. 2019, 16, 8. [Google Scholar] [CrossRef] [PubMed]
  121. Afshar, G.V.; Rasmi, Y.; Yagmayee, P.; Khadem-Ansari, M.-H.; Makhdomii, K.; Rasooli, J. The Effects of Nano-curcumin Supplementation on Serum Level of hs-CRP, Adhesion Molecules, and Lipid Profiles in Hemodialysis Patients, A Randomized Controlled Clinical Trial. Iran. J. Kidney Dis. 2020, 14, 52–61. [Google Scholar]
  122. Saraf-Bank, S.; Ahmadi, A.; Paknahad, Z.; Maracy, M.; Nourian, M. Effects of curcumin supplementation on markers of inflammation and oxidative stress among healthy overweight and obese girl adolescents: A randomized placebo-controlled clinical trial. Phytother. Res. 2019, 33, 2015–2022. [Google Scholar] [CrossRef] [Green Version]
  123. Pawar, K.S.; Mastud, R.N.; Pawar, S.K.; Pawar, S.S.; Bhoite, R.R.; Bhoite, R.R.; Kulkarni, M.V.; Deshpande, A.R. Oral Curcumin with Piperine as Adjuvant Therapy for the Treatment of COVID-19: A Randomized Clinical Trial. Front. Pharmacol. 2021, 12, 669362. [Google Scholar] [CrossRef]
  124. Kedia, S.; Bhatia, V.; Thareja, S.; Garg, S.; Mouli, V.P.; Bopanna, S.; Tiwari, V.; Makharia, G.; Ahuja, V. Low dose oral curcumin is not effective in induction of remission in mild to moderate ulcerative colitis: Results from a randomized double blind placebo controlled trial. World J. Gastrointest. Pharmacol. Ther. 2017, 8, 147–154. [Google Scholar] [CrossRef]
  125. Sadeghi, N.; Mansoori, A.; Shayesteh, A.; Hashemi, S.J. The effect of curcumin supplementation on clinical outcomes and inflammatory markers in patients with ulcerative colitis. Phytother. Res. 2019, 34, 1123–1133. [Google Scholar] [CrossRef]
  126. Zhang, W.-Y.; Guo, Y.-J.; Han, W.-X.; Yang, M.-Q.; Wen, L.-P.; Wang, K.-Y.; Jiang, P. Curcumin relieves depressive-like behaviors via inhibition of the NLRP3 inflammasome and kynurenine pathway in rats suffering from chronic unpredictable mild stress. Int. Immunopharmacol. 2019, 67, 138–144. [Google Scholar] [CrossRef]
  127. Hamam, F.; Nasr, A. Curcumin-loaded mesoporous silica particles as wound-healing agent: An In vivo study. Saudi J. Med. Med. Sci. 2020, 8, 17–24. [Google Scholar] [CrossRef]
  128. Peng, K.-T.; Chiang, Y.-C.; Huang, T.-Y.; Chen, P.-C.; Chang, P.-J.; Lee, C.-W. Curcumin nanoparticles are a promising anti-bacterial and anti-inflammatory agent for treating periprosthetic joint infections. Int. J. Nanomed. 2019, 14, 469–481. [Google Scholar] [CrossRef] [Green Version]
  129. Doukas, S.G.; Doukas, P.G.; Sasaki, C.T.; Vageli, D. The in vivo preventive and therapeutic properties of curcumin in bile reflux-related oncogenesis of the hypopharynx. J. Cell. Mol. Med. 2020, 24, 10311–10321. [Google Scholar] [CrossRef]
  130. Paolillo, F.R.; Rodrigues, P.G.S.; Bagnato, V.S.; Alves, F.; Pires, L.; Corazza, A.V. The effect of combined curcumin-mediated photodynamic therapy and artificial skin on Staphylococcus aureus–infected wounds in rats. Lasers Med. Sci. 2020, 36, 1219–1226. [Google Scholar] [CrossRef]
  131. Guorgui, J.; Wang, R.; Mattheolabakis, G.; Mackenzie, G.G. Curcumin formulated in solid lipid nanoparticles has enhanced efficacy in Hodgkin’s lymphoma in mice. Arch. Biochem. Biophys. 2018, 648, 12–19. [Google Scholar] [CrossRef] [Green Version]
  132. Sherin, S.; Balachandran, S.; Abraham, A. Curcumin incorporated titanium dioxide nanoparticles as MRI contrasting agent for early diagnosis of atherosclerosis- rat model. Veter. Anim. Sci. 2020, 10, 100090. [Google Scholar] [CrossRef]
  133. Guo, Y.; Wu, R.; Gaspar, J.M.; Sargsyan, D.; Su, Z.-Y.; Zhang, C.; Gao, L.; Cheng, D.; Li, W.; Wang, C.; et al. DNA methylome and transcriptome alterations and cancer prevention by curcumin in colitis-accelerated colon cancer in mice. Carcinogenesis 2018, 39, 669–680. [Google Scholar] [CrossRef]
  134. Pham, L.; Dang, L.H.; Truong, M.D.; Nguyen, T.H.; Le, L.; Le, V.T.; Nam, N.D.; Bach, L.G.; Nguyen, V.T.; Tran, N.Q. A dual synergistic of curcumin and gelatin on thermal-responsive hydrogel based on Chitosan-P123 in wound healing application. Biomed. Pharmacother. 2019, 117, 109183. [Google Scholar] [CrossRef]
  135. Niranjan, R.; Kaushik, M.; Prakash, J.; Venkataprasanna, K.S.; Arpana, C.; Balashanmugam, P.; Venkatasubbu, G.D. Enhanced wound healing by PVA/Chitosan/Curcumin patches: In vitro and in vivo study. Colloids Surf. B Biointerfaces 2019, 182, 110339. [Google Scholar] [CrossRef]
  136. Panda, S.K.; Parachur, V.A.; Mohanty, N.; Swain, T.; Sahu, S. A Comparative Pharmacokinetic Evaluation of a Bioavailable Curcumin Formulation Curene® with Curcumin Formulation Containing Turmeric Volatile Oil and Standard Curcuminoids 95% in Healthy Human Subjects. Funct. Foods Health Dis. 2019, 9, 134–144. [Google Scholar] [CrossRef] [Green Version]
  137. Ullah, F.; Asgarov, R.; Venigalla, M.; Liang, H.; Niedermayer, G.; Münch, G.; Gyengesi, E. Effects of a solid lipid curcumin particle formulation on chronic activation of microglia and astroglia in the GFAP-IL6 mouse model. Sci. Rep. 2020, 10, 100090. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Tyagi, P.; Singh, M.; Kumari, H.; Kumari, A.; Mukhopadhyay, K. Bactericidal Activity of Curcumin I Is Associated with Damaging of Bacterial Membrane. PLoS ONE 2015, 10, e0121313. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Zheng, D.; Huang, C.; Huang, H.; Zhao, Y.; Khan, M.R.U.; Zhao, H.; Huang, L. Antibacterial Mechanism of Curcumin: A Review. Chem. Biodivers. 2020, 17, e2000171. [Google Scholar] [CrossRef] [PubMed]
  140. Adamczak, A.; Ożarowski, M.; Karpiński, T.M. Curcumin, a Natural Antimicrobial Agent with Strain-Specific Activity. Pharmaceuticals 2020, 13, 153. [Google Scholar] [CrossRef] [PubMed]
  141. Sharahi, J.Y.; Ahovan, Z.; Maleki, D.T.; Rad, Z.R.; Rad, Z.R.; Goudarzi, M.; Shariati, A.; Bostanghadiri, N.; Abbasi, E.; Hashemi, A. In vitro antibacterial activity of curcumin-meropenem combination against extensively drug-resistant (XDR) bacteria isolated from burn wound infections. Avicenna J. Phytomedicine 2020, 10, 3–10. [Google Scholar] [CrossRef]
  142. Mathew, D.; Hsu, W.-L. Antiviral potential of curcumin. J. Funct. Foods 2018, 40, 692–699. [Google Scholar] [CrossRef]
  143. Balasubramanian, A.; Pilankatta, R.; Teramoto, T.; Sajith, A.M.; Nwulia, E.; Kulkarni, A.; Padmanabhan, R. Inhibition of dengue virus by curcuminoids. Antivir. Res. 2019, 162, 71–78. [Google Scholar] [CrossRef]
  144. Jeong, E.-H.; Vaidya, B.; Cho, S.-Y.; Park, M.-A.; Kaewintajuk, K.; Kim, S.R.; Oh, M.-J.; Choi, J.-S.; Kwon, J.; Kim, D. Identification of regulators of the early stage of viral hemorrhagic septicemia virus infection during curcumin treatment. Fish Shellfish. Immunol. 2015, 45, 184–193. [Google Scholar] [CrossRef]
  145. Ferreira, V.H.; Nazli, A.; Dizzell, S.E.; Mueller, K.; Kaushic, C. The Anti-Inflammatory Activity of Curcumin Protects the Genital Mucosal Epithelial Barrier from Disruption and Blocks Replication of HIV-1 and HSV-2. PLoS ONE 2015, 10, e0124903. [Google Scholar] [CrossRef] [Green Version]
  146. Li, H.; Zhong, C.; Wang, Q.; Chen, W.; Yuan, Y. Curcumin is an APE1 redox inhibitor and exhibits an antiviral activity against KSHV replication and pathogenesis. Antivir. Res. 2019, 167, 98–103. [Google Scholar] [CrossRef]
  147. Mounce, B.C.; Cesaro, T.; Carrau, L.; Vallet, T.; Vignuzzi, M. Curcumin inhibits Zika and chikungunya virus infection by inhibiting cell binding. Antivir. Res. 2017, 142, 148–157. [Google Scholar] [CrossRef]
  148. Teymouri, M.; Pirro, M.; Johnston, T.P.; Sahebkar, A. Curcumin as a multifaceted compound against human papilloma virus infection and cervical cancers: A review of chemistry, cellular, molecular, and preclinical features. BioFactors 2017, 43, 331–346. [Google Scholar] [CrossRef]
  149. Babaei, F.; Nassiri-Asl, M.; Hosseinzadeh, H. Curcumin (a constituent of turmeric): New treatment option against COVID-19. Food Sci. Nutr. 2020, 8, 5215–5227. [Google Scholar] [CrossRef]
  150. Soni, V.K.; Mehta, A.; Ratre, Y.K.; Tiwari, A.K.; Amit, A.; Singh, R.P.; Sonkar, S.C.; Chaturvedi, N.; Shukla, D.; Vishvakarma, N.K. Curcumin, a traditional spice component, can hold the promise against COVID-19? Eur. J. Pharmacol. 2020, 886, 173551. [Google Scholar] [CrossRef]
  151. Dourado, D.; Freire, D.T.; Pereira, D.T.; Amaral-Machado, L.; Alencar, N.; de Barros, A.L.B.; Egito, E.S.T. Will curcumin nanosystems be the next promising antiviral alternatives in COVID-19 treatment trials? Biomed. Pharmacother. 2021, 139, 111578. [Google Scholar] [CrossRef]
  152. Zahedipour, F.; Hosseini, S.A.; Sathyapalan, T.; Majeed, M.; Jamialahmadi, T.; Al-Rasadi, K.; Banach, M.; Sahebkar, A. Potential effects of curcumin in the treatment of COVID -19 infection. Phytother. Res. 2020, 34, 2911–2920. [Google Scholar] [CrossRef]
  153. Subhan, F.; Khalil, A.A.K.; Zeeshan, M.; Haider, A.; Tauseef, I.; Haleem, S.K.; Ibrahim, A.S. Curcumin: From Ancient Spice to Modern Anti-Viral Drug in COVID-19 Pandemic. Life Sci. 2020, 1, 69–73. [Google Scholar] [CrossRef]
  154. Thimmulappa, R.K.; Mudnakudu-Nagaraju, K.K.; Shivamallu, C.; Subramaniam, K.; Radhakrishnan, A.; Bhojraj, S.; Kuppusamy, G. Antiviral and immunomodulatory activity of curcumin: A case for prophylactic therapy for COVID-19. Heliyon 2021, 7, e06350. [Google Scholar] [CrossRef]
  155. Valizadeh, H.; Abdolmohammadi-Vahid, S.; Danshina, S.; Gencer, M.Z.; Ammari, A.; Sadeghi, A.; Roshangar, L.; Aslani, S.; Esmaeilzadeh, A.; Ghaebi, M.; et al. Nano-curcumin therapy, a promising method in modulating inflammatory cytokines in COVID-19 patients. Int. Immunopharmacol. 2020, 89, 107088. [Google Scholar] [CrossRef]
  156. Khanra, S.; Kumar, Y.P.; Dash, J.; Banerjee, R. In vitro screening of known drugs identified by scaffold hopping techniques shows promising leishmanicidal activity for suramin and netilmicin. BMC Res. Notes 2018, 51, 990–997. [Google Scholar] [CrossRef]
  157. Bafghi, A.F.; Haghirosadat, B.F.; Yazdian, F.; Mirzaei, F.; Pourmadadi, M.; Pournasir, F.; Hemati, M.; Pournasir, S. A novel delivery of curcumin by the efficient nanoliposomal approach against Leishmania major. Prep. Biochem. Biotechnol. 2021, 51, 990–997. [Google Scholar] [CrossRef]
  158. Mallo, N.; Lamas, J.; Sueiro, R.A.; Leiro, J.M. Molecular Targets Implicated in the Antiparasitic and Anti-Inflammatory Activity of the Phytochemical Curcumin in Trichomoniasis. Molecules 2020, 25, 5321. [Google Scholar] [CrossRef] [PubMed]
  159. Rangel-Castañeda, I.A.; Hernández-Hernández, J.M.; Pérez-Rangel, A.; González-Pozos, S.; Carranza-Rosales, P.; Charles-Niño, C.L.; Tapia-Pastrana, G.; Ramírez-Herrera, M.A.; Castillo-Romero, A. Amoebicidal activity of curcumin on Entamoeba histolytica trophozoites. J. Pharm. Pharmacol. 2018, 70, 426–433. [Google Scholar] [CrossRef] [PubMed]
  160. Gutiérrez-Gutiérrez, F.; Palomo-Ligas, L.; Hernández-Hernández, J.M.; Pérez-Rangel, A.; Aguayo-Ortiz, R.; Hernández-Campos, A.; Castillo, R.; González-Pozos, S.; Cortés-Zárate, R.; Ramírez-Herrera, M.A.; et al. Curcumin alters the cytoskeleton and microtubule organization on trophozoites of Giardia lamblia. Acta Trop. 2017, 172, 113–121. [Google Scholar] [CrossRef] [PubMed]
  161. El-Shafey, A.A.M.; Hegab, M.H.A.; Seliem, M.M.E.; Barakat, A.M.A.; Mostafa, N.E.; Abdel-Maksoud, H.A.; Abdelhameed, R.M. Curcumin@metal organic frameworks nano-composite for treatment of chronic toxoplasmosis. J. Mater. Sci. Mater. Med. 2020, 31, 1–13. [Google Scholar] [CrossRef]
  162. Qian, W.; Wang, H.; Shan, D.; Li, B.; Liu, J.; Liu, Q. Activity of several kinds of drugs against Neospora caninum. Parasitol. Int. 2015, 64, 597–602. [Google Scholar] [CrossRef]
  163. Bazh, E.K.A.; El-Bahy, N.M. In vitro and in vivo screening of anthelmintic activity of ginger and curcumin on Ascaridia galli. Parasitol. Res. 2013, 112, 3679–3686. [Google Scholar] [CrossRef]
  164. El-Bahy, N.M.; Bazh, E.K.A. Anthelmintic activity of ginger, curcumin, and praziquentel against Raillietina cesticillus (in vitro and in vivo). Parasitol. Res. 2015, 114, 2427–2434. [Google Scholar] [CrossRef]
  165. Novaes, R.D.; Sartini, M.V.P.; Rodrigues, J.P.F.; Gonçalves, R.V.; Santos, E.C.; Souza, R.L.M.; Caldas, I.S. Curcumin Enhances the Anti-Trypanosoma cruzi Activity of Benznidazole-Based Chemotherapy in Acute Experimental Chagas Disease. Antimicrob. Agents Chemother. 2016, 60, 3355–3364. [Google Scholar] [CrossRef] [Green Version]
  166. Busari, Z.A.; Dauda, K.A.; Morenikeji, O.A.; Afolayan, F.; Oyeyemi, O.T.; Meena, J.; Sahu, D.; Panda, A.K. Antiplasmodial Activity and Toxicological Assessment of Curcumin PLGA-Encapsulated Nanoparticles. Front. Pharmacol. 2017, 8, 622. [Google Scholar] [CrossRef]
  167. Naseri, S.; Darroudi, M.; Aryan, E.; Gholoobi, A.; Rahimi, H.R.; Ketabi, K.; Movaqar, A.; Abdoli, M.; Gouklani, H.; Tei-mourpour, R. The antiviral effects of curcumin nanomicelles on the attachment and entry of hepatitis C virus. Iran. J. Virol. 2017, 11, 29–35. [Google Scholar]
  168. Ahmed, J.; Tan, Y.; Ambegaokar, S. Effects of Curcumin on Vesicular Stomatitis Virus (VSV) Infection and Dicer-1 Expression. FASEB J. 2017, 31, 622.11. [Google Scholar] [CrossRef]
  169. Sharma, R.K.; Cwiklinski, K.; Aalinkeel, R.; Reynolds, J.L.; Sykes, D.E.; Quaye, E.; Oh, J.; Mahajan, S.D.; Schwartz, S.A. Immunomodulatory activities of curcumin-stabilized silver nanoparticles: Efficacy as an antiretroviral therapeutic. Immunol. Investig. 2017, 46, 833–846. [Google Scholar] [CrossRef]
  170. Huang, H.-I.; Chio, C.-C.; Lin, J.-Y. Inhibition of EV71 by curcumin in intestinal epithelial cells. PLoS ONE 2018, 13, e0191617. [Google Scholar] [CrossRef] [Green Version]
  171. Poursina, Z.; Mohammadi, A.; Yazdi, S.Z.; Humpson, I.; Vakili, V.; Boostani, R.; Rafatpanah, H. Curcumin increased the expression of c-FLIP in HTLV-1-associated myelopathy/tropical spastic paraparesis (HAM/TSP) patients. J. Cell. Biochem. 2019, 120, 15740–15745. [Google Scholar] [CrossRef] [PubMed]
  172. Sharma, A.; Yadav, A.; Gupta, N.; Sharma, S.; Kakkar, R.; Cwiklinski, K.; Quaye, E.; Mahajan, S.D.; Schwartz, S.A.; Sharma, R.K. Multifunctional mesoporous curcumin encapsulated iron-phenanthroline nanocluster: A new Anti-HIV agent. Colloids Surf. B Biointerfaces 2019, 180, 289–297. [Google Scholar] [CrossRef] [PubMed]
  173. Nabila, N.; Suada, N.K.; Denis, D.; Yohan, B.; Adi, A.C.; Veterini, A.S.; Anindya, A.L.; Sasmono, R.T.; Rachmawati, H. Antiviral Action of Curcumin Encapsulated in Nanoemulsion against Four Serotypes of Dengue Virus. Pharm. Nanotechnol. 2020, 8, 54–62. [Google Scholar] [CrossRef] [PubMed]
  174. Li, Y.; Wang, J.; Liu, Y.; Luo, X.; Lei, W.; Xie, L. Antiviral and virucidal effects of curcumin on transmissible gastroenteritis virus in vitro. J. Gen. Virol. 2020, 101, 1079–1084. [Google Scholar] [CrossRef]
  175. Zhang, C.; Zhang, K.; Zang, G.; Chen, T.; Lu, N.; Wang, S.; Zhang, G. Curcumin Inhibits Replication of Human Parainfluenza Virus Type 3 by Affecting Viral Inclusion Body Formation. BioMed Res. Int. 2021, 2021, 13. [Google Scholar] [CrossRef]
  176. Thongsri, P.; Pewkliang, Y.; Borwornpinyo, S.; Wongkajornsilp, A.; Hongeng, S.; Sa-Ngiamsuntorn, K. Curcumin inhibited hepatitis B viral entry through NTCP binding. Sci. Rep. 2021, 11, 19125. [Google Scholar] [CrossRef]
  177. Tiwari, B.; Pahuja, R.; Kumar, P.; Rath, S.K.; Gupta, K.C.; Goyal, N. Nanotized Curcumin and Miltefosine, a Potential Combination for Treatment of Experimental Visceral Leishmaniasis. Antimicrob. Agents Chemother. 2017, 61, e01169–16. [Google Scholar] [CrossRef] [Green Version]
  178. Ullah, R.; Rehman, A.; Zafeer, M.F.; Rehman, L.; Khan, Y.A.; Khan, M.A.H.; Khan, S.N.; Khan, A.U.; Abidi, S.M.A. Anthelmintic Potential of Thymoquinone and Curcumin on Fasciola gigantica. PLoS ONE 2017, 12, e0171267. [Google Scholar] [CrossRef]
  179. Asadpour, M.; Namazi, F.; Razavi, S.M.; Nazifi, S. Comparative efficacy of curcumin and paromomycin against Cryptosporidium parvum infection in a BALB/c model. Veter. Parasitol. 2018, 250, 7–14. [Google Scholar] [CrossRef]
  180. Ghosh, A.; Banerjee, T. Nanotized curcumin-benzothiophene conjugate: A potential combination for treatment of cerebral malaria. IUBMB Life 2020, 72, 2637–2650. [Google Scholar] [CrossRef]
  181. Elmi, T.; Ardestani, M.S.; Hajialiani, F.; Motevalian, M.; Mohamadi, M.; Sadeghi, S.; Zamani, Z.; Tabatabaie, F. Novel chloroquine loaded curcumin based anionic linear globular dendrimer G2: A metabolomics study on Plasmodium falciparum in vitro using 1H NMR spectroscopy. Parasitology 2020, 147, 747–759. [Google Scholar] [CrossRef]
  182. Wang, J.; Zhao, L.; Wei, Z.; Zhang, X.; Wang, Y.; Li, F.; Fu, Y.; Liu, B. Inhibition of histone deacetylase reduces lipopolysaccharide-induced-inflammation in primary mammary epithelial cells by regulating ROS-NF-кB signaling pathways. Int. Immunopharmacol. 2018, 56, 230–234. [Google Scholar] [CrossRef]
  183. Shehzad, A.; Qureshi, M.; Anwar, M.N.; Lee, Y.S. Multifunctional Curcumin Mediate Multitherapeutic Effects. J. Food Sci. 2017, 82, 2006–2015. [Google Scholar] [CrossRef] [Green Version]
  184. Banik, U.; Parasuraman, S.; Adhikary, A.K.; Othman, N.H. Curcumin: The spicy modulator of breast carcinogenesis. J. Exp. Clin. Cancer Res. 2017, 36, 98. [Google Scholar] [CrossRef] [Green Version]
  185. Chai, Y.-S.; Chen, Y.-Q.; Lin, S.-H.; Xie, K.; Wang, C.-J.; Yang, Y.-Z.; Xu, F. Curcumin regulates the differentiation of naïve CD4+T cells and activates IL-10 immune modulation against acute lung injury in mice. Biomed. Pharmacother. 2020, 125, 109946. [Google Scholar] [CrossRef]
  186. Hui, S.; Liu, K.; Zhu, X.; Kang, C.; Mi, M.T. Effect of Curcumin on IL-6 and IL-8: A Meta-analysis and Systematic Review. J. Nutr. Food Sci. 2018, 8, 1–7. [Google Scholar] [CrossRef]
  187. Yu, Y.; Shen, Q.; Lai, Y.; Park, S.Y.; Ou, X.; Lin, D.; Jin, M.; Zhang, W. Anti-inflammatory Effects of Curcumin in Microglial Cells. Front. Pharmacol. 2018, 9, 386. [Google Scholar] [CrossRef] [Green Version]
  188. Castaño, P.R.; Parween, S.; Pandey, A.V. Bioactivity of Curcumin on the Cytochrome P450 Enzymes of the Steroidogenic Pathway. Int. J. Mol. Sci. 2019, 20, 4606. [Google Scholar] [CrossRef] [Green Version]
  189. Moutinho, M.S.; Aragão, S.; Carmo, D.; Casaca, F.; Silva, S.; Ribeiro, J.; Sousa, H.; Pires, I.; Queiroga, F.; Colaço, B.; et al. Curcumin and Rutin Down-regulate Cyclooxygenase-2 and Reduce Tumor-associated Inflammation in HPV16-Transgenic Mice. Anticancer Res. 2018, 38, 1461–1466. [Google Scholar] [CrossRef]
  190. Liu, Z.; Ying, Y. The Inhibitory Effect of Curcumin on Virus-Induced Cytokine Storm and Its Potential Use in the Associated Severe Pneumonia. Front. Cell Dev. Biol. 2020, 8, 479. [Google Scholar] [CrossRef]
  191. Pizzino, G.; Irrera, N.; Cucinotta, M.; Pallio, G.; Mannino, F.; Arcoraci, V.; Squadrito, F.; Altavilla, D.; Bitto, A. Oxidative Stress: Harms and Benefits for Human Health. Oxid. Med. Cell. Longev. 2017, 2017, 8416763. [Google Scholar] [CrossRef]
  192. Samarghandian, S.; Azimi-Nezhad, M.; Farkhondeh, T.; Samini, F. Anti-oxidative effects of curcumin on immobilization-induced oxidative stress in rat brain, liver and kidney. Biomed. Pharmacother. 2017, 87, 223–229. [Google Scholar] [CrossRef]
  193. Jagetia, G.C.; Rajanikant, G.K. Curcumin Stimulates the Antioxidant Mechanisms in Mouse Skin Exposed to Fractionated γ-Irradiation. Antioxidants 2015, 4, 25–41. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Meshkibaf, M.H.; Maleknia, M.; Noroozi, S. Effect of curcumin on gene expression and protein level of methionine sulfoxide reductase A (MSRA), SOD, CAT and GPx in Freund’s adjuvant inflammation-induced male rats. J. Inflamm. Res. 2019, 12, 241–249. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Haryuna, T.S.H.; Munir, D.; Maria, A.; Bashiruddin, J. The Antioxidant Effect of Curcumin on Cochlear Fibroblasts in Rat Models of Diabetes Mellitus. Iran. J. Otorhinolaryngol. 2017, 29, 197–202. [Google Scholar] [CrossRef]
  196. Asouri, M.; Ataee, R.; Ahmadi, A.A.; Amini, A.; Moshaei, M.R. Antioxidant and Free Radical Scavenging Activities of Curcumin. Asian J. Chem. 2013, 25, 7593–7595. [Google Scholar] [CrossRef]
  197. Barzegar, A.; Moosavi-Movahedi, A.A. Intracellular ROS Protection Efficiency and Free Radical-Scavenging Activity of Curcumin. PLoS ONE 2011, 6, e26012. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  198. Chen, S.; Wu, J.; Tang, Q.; Xu, C.; Huang, Y.; Huang, D.; Luo, F.; Wu, Y.; Yan, F.; Weng, Z.; et al. Nano-micelles based on hydroxyethyl starch-curcumin conjugates for improved stability, antioxidant and anticancer activity of curcumin. Carbohydr. Polym. 2020, 228, 115398. [Google Scholar] [CrossRef]
  199. Ma, Q.; Ren, Y.; Wang, L. Investigation of antioxidant activity and release kinetics of curcumin from tara gum/polyvinyl alcohol active film. Food Hydrocoll. 2017, 70, 286–292. [Google Scholar] [CrossRef]
  200. Zhu, Q.; Sun, Y.; Yun, X.; Ou, Y.; Zhang, W.; Li, J.-X. Antinociceptive effects of curcumin in a rat model of postoperative pain. Sci. Rep. 2014, 4, 4932. [Google Scholar] [CrossRef] [Green Version]
  201. Pieretti, S.; Ranjan, A.P.; Di Giannuario, A.; Mukerjee, A.; Marzoli, F.; Di Giovannandrea, R.; Vishwanatha, J.K. Curcumin-loaded Poly (d,l-lactide-co-glycolide) nanovesicles induce antinociceptive effects and reduce pronociceptive cytokine and BDNF release in spinal cord after acute administration in mice. Colloids Surf. B Biointerfaces 2017, 158, 379–386. [Google Scholar] [CrossRef]
  202. Barchitta, M.; Maugeri, A.; Favara, G.; Lio, R.M.S.; Evola, G.; Agodi, A.; Basile, G. Nutrition and Wound Healing: An Overview Focusing on the Beneficial Effects of Curcumin. Int. J. Mol. Sci. 2019, 20, 1119. [Google Scholar] [CrossRef] [Green Version]
  203. Mohanty, C.; Sahoo, S.K. Curcumin and its topical formulations for wound healing applications. Drug Discov. Today 2017, 22, 1582–1592. [Google Scholar] [CrossRef]
  204. Zakerikhoob, M.; Abbasi, S.; Yousefi, G.; Mokhtari, M.; Noorbakhsh, M.S. Curcumin-incorporated crosslinked sodium alginate-g-poly (N-isopropyl acrylamide) thermo-responsive hydrogel as an in-situ forming injectable dressing for wound healing: In vitro characterization and in vivo evaluation. Carbohydr. Polym. 2021, 271, 118434. [Google Scholar] [CrossRef]
  205. Krausz, A.E.; Adler, B.L.; Cabral, V.; Navati, M.; Doerner, J.; Charafeddine, R.A.; Chandra, D.; Liang, H.; Gunther, L.; Clendaniel, A.; et al. Curcumin-encapsulated nanoparticles as innovative antimicrobial and wound healing agent. Nanomed. Nanotechnol. Biol. Med. 2015, 11, 195–206. [Google Scholar] [CrossRef] [Green Version]
  206. Dai, C.; Ciccotosto, G.D.; Cappai, R.; Tang, S.; Li, D.; Xie, S.; Xiao, X.; Velkov, T. Curcumin Attenuates Colistin-Induced Neurotoxicity in N2a Cells via Anti-inflammatory Activity, Suppression of Oxidative Stress, and Apoptosis. Mol. Neurobiol. 2018, 55, 421–434. [Google Scholar] [CrossRef]
  207. Yuan, J.; Liu, R.; Ma, Y.; Zhang, Z.; Xie, Z. Curcumin Attenuates Airway Inflammation and Airway Remolding by Inhibiting NF-κB Signaling and COX-2 in Cigarette Smoke-Induced COPD Mice. Inflammation 2018, 41, 1804–1814. [Google Scholar] [CrossRef]
  208. Vasanthkumar, T.; Hanumanthappa, M.; Lakshminarayana, R. Curcumin and capsaicin modulates LPS induced expression of COX-2, IL-6 and TGF-β in human peripheral blood mononuclear cells. Cytotechnology 2019, 71, 963–976. [Google Scholar] [CrossRef]
  209. Wang, H.; Gong, X.; Guo, X.; Liu, C.; Fan, Y.-Y.; Zhang, J.; Niu, B.; Li, W. Characterization, release, and antioxidant activity of curcumin-loaded sodium alginate/ZnO hydrogel beads. Int. J. Biol. Macromol. 2019, 121, 1118–1125. [Google Scholar] [CrossRef]
  210. Özçelik, M.; Erisir, M.; Guler, O.; Baykara, M.; Kirman, E. The effect of curcumin on lipid peroxidation and selected antioxidants in irradiated rats. Acta Veter. Brno 2019, 87, 379–385. [Google Scholar] [CrossRef]
  211. Alizadeh, M.; Kheirouri, S. Curcumin reduces malondialdehyde and improves antioxidants in humans with diseased conditions: A comprehensive meta-analysis of randomized controlled trials. BioMedicine 2019, 9, 23. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  212. Jakubczyk, K.; Drużga, A.; Katarzyna, J.; Skonieczna-Żydecka, K. Antioxidant Potential of Curcumin—A Meta-Analysis of Randomized Clinical Trials. Antioxidants 2020, 9, 1092. [Google Scholar] [CrossRef] [PubMed]
  213. Zhao, G.; Shi, Y.; Gong, C.; Liu, T.; Nan, W.; Ma, L.; Wu, Z.; Da, C.; Zhou, K.; Zhang, H. Curcumin Exerts Antinociceptive Effects in Cancer-Induced Bone Pain via an Endogenous Opioid Mechanism. Front. Neurosci. 2021, 15, 696861. [Google Scholar] [CrossRef] [PubMed]
  214. Ju, J.; Shin, J.Y.; Yoon, J.J.; Yin, M.; Yoon, M.H. Differential expression of spinal γ-aminobutyric acid and opioid receptors modulates the analgesic effects of intrathecal curcumin on postoperative/inflammatory pain in rats. Anesth. Pain Med. 2018, 13, 82–92. [Google Scholar] [CrossRef] [Green Version]
  215. Wu, Y.; Qin, D.; Yang, H.; Fu, H. Evidence for the Participation of Acid-Sensing Ion Channels (ASICs) in the Antinociceptive Effect of Curcumin in a Formalin-Induced Orofacial Inflammatory Model. Cell. Mol. Neurobiol. 2016, 37, 635–642. [Google Scholar] [CrossRef]
Figure 1. Structural formulae of curcuminoids.
Figure 1. Structural formulae of curcuminoids.
Antibiotics 11 00135 g001
Figure 2. Keto-enol tautomerism of curcumin.
Figure 2. Keto-enol tautomerism of curcumin.
Antibiotics 11 00135 g002
Figure 3. Important functional parts of curcumin: 1,3-keto-enol part (A), o-methoxy and phenolic groups (B) and a double bond (C).
Figure 3. Important functional parts of curcumin: 1,3-keto-enol part (A), o-methoxy and phenolic groups (B) and a double bond (C).
Antibiotics 11 00135 g003
Figure 4. Important metabolic and nonmetabolic transformations of curcumin.
Figure 4. Important metabolic and nonmetabolic transformations of curcumin.
Antibiotics 11 00135 g004
Figure 5. Formulations of curcumin.
Figure 5. Formulations of curcumin.
Antibiotics 11 00135 g005
Figure 6. Biological activities of curcumin.
Figure 6. Biological activities of curcumin.
Antibiotics 11 00135 g006
Table 1. The History of curcumin.
Table 1. The History of curcumin.
YearDiscoveryReference
1815Vogel and Pelletier were the first to report the “Orange-yellow Substance” isolated from the rhizome of Curcuma longa and named it curcumin.[16]
1842Vogel Extracted pure preparation of curcumin but did not report its formula.[17]
1910Milobedzka and Lampe identified chemical structure of curcumin as diferuloylmethane, or 1,6-heptadiene-3,5-dione-1,7-bis-(4-hidroxy-3-methoxyphenyl)-(1E, 6E).[18]
1913The synthesis of curcumin was published.[19]
1949Schraufstatter et al. Reported that curcumin is a biologically active compound with antibacterial properties.[20]
1953Srinivasan separated and quantified the components of curcumin using chromatography.[21]
1971It was discovered that curcumin lowers cholesterol [22]
1972It was discovered that curcumin lowers the level of sugar in the blood [23]
1973It was discovered that curcumin has an anti-inflammatory effect [24]
1976It was discovered that curcumin has an antioxidant effect [25]
1980Kuttan et al. demonstrated anticancer activity of curcumin both in vitro and in vivo.[26]
1995Curcumin exhibits anti-inflammatory activity by suppressing the proinflammatory transcription factor, nuclear factor-kappa B (NF-κB)[27]
Table 2. HPLC methods for curcuminoid analysis.
Table 2. HPLC methods for curcuminoid analysis.
Matrix SampleColumnMobile Phaseλ, nmLimit of DetectionReference
Turmeric PowderRP C18Acetonitrile and 0.1% Trifluro-Acetic Acid (50:50, v/v)42027.99 ng/mL[70]
Turmeric ExtractsAlltima C18 columnAcetonitrile and 2% Acetic Acid (40:60, v/v)4250.90 μg/mL[71]
Commercial Samples of TurmericC18Methanol, 2% Acetic acid, and Acetonitrile4250.05 µg/mL[72]
Curcuminoids-Loaded LiposomeZorbax Eclipse XDB C18 (4 × 150mm, 5 µm)Acetonitrile and 0.1% OrthoPhosphoric Acid (50:50, v/v)4250.124 µg/mL[73]
Samples of TurmericZorbax SB-C18 column (4.6 × 250 mm, 5 µm)Acetonitrile and 0.4% Aqueous Acetic Acid4300.31 μg/mL[37]
Extract of TurmericC18 (4.6 × 150mm, 5 µm)Acetonitrile and 2% Acetic Acid (55:45, v/v)4250.0738 ppm[74]
Extract of TurmericWaters Xterra MS C18 column (4.6 × 250 mm, 5 µm)Distilled Water and Acetonitrile (65:35, v/v) Containing 1% Acetic Acid4251.13 μg/mL[75]
Turmeric RhizomeBrownlee SPP C18 column (4.6 × 100 mm, 2.7 µm)Water and Acetonitrile (70:30, v/v)4201.0 μg/mL[76]
Table 3. Clinical applications of curcumin.
Table 3. Clinical applications of curcumin.
DiseaseDoseDurationPatientsResultsReference
Overweight80 mg/Day6 Weeks48 Overweight Girl StudentsPositive antioxidant effect and prevention of lipid peroxidation in overweight individuals.[116]
Metabolic syndrome(MetS)80 mg/Day12 Weeks50 PatientsSupplementation with Nanomicelle curcumin Significantly improved serum triglyceride in MetS patients.[117]
Diabetic sensorimotor polyneuropathy80 mg/Day8 Weeks80 Diabetic patientsNanocurcumin supplementation reduced the severity of diabetic sensorimotor polyneuropathy in patients with type 2 diabetes mellitus.[118]
Migraine80 mg/Day2 Months80 PatientsCombination of omega-3 fatty acids and nanocurcumin modulates interleukin-6 gene Expression and high-sensitivity C-reactive protein serum levels in patients with migraine.[119]
Nonalcoholic fatty liver disease (NAFLD)80 mg/Day3 Months84 PatientsNanocurcumin improves glucose indices, lipids, inflammation, and nesfatin in overweight and obese patients with nonalcoholic fatty liver disease (NAFLD).[120]
Hemodialysis (HD)120 mg/Day12 Weeks60 PatientsNanocurcumin shows beneficial effects in lowering inflammation and Hs-CRP levels, as well as adhesion molecules (ICAM-1, VCAM-1), in hemodialysis patients.[121]
Overweight and obesity500 mg/Day10 Weeks60 AdolescentTen weeks of curcumin supplementation had beneficial effects on inflammation and oxidative stress markers among postpubescent overweight and obese girl adolescents.[122]
Coronavirus disease-20191050 mg/Day14 Days158 PatientsCurcumin is a safe and natural therapeutic option to prevent post-COVID-19 thromboembolic events.[123]
Ulcerative colitis (UC)450 mg/Day8 Weeks41 PatientsLow-dose oral curcumin is not effective in inducing remission in mild-to-moderate ulcerative colitis.[124]
Ulcerative colitis (UC)1500 mg/Day8 Weeks70 PatientsConsumption of the curcumin supplement, along with drug therapy, significant improvement of the clinical outcomes, quality of life, Hs-CRP, and ESR in patients with mild-to-moderate UC.[125]
Table 4. Application of curcumin in vivo-animal models.
Table 4. Application of curcumin in vivo-animal models.
Curcumin FormActivityAnimal ModelReference
Curcumin
Nanocurcumin
Antidepressive effect
Wound-healing Agent
Sprague–Dawley rats
Male Wistar rats
[126]
[127]
Curcumin, nanoparticlesAntibacterial and anti-inflammatory agentMale C57BL/6 mice[128]
CurcuminInhibitors of NF-κBMus musculus, C57BL/6J[129]
CurcuminDecontaminate and accelerate the Wound contractionWistar Rats[130]
Curcumin, NanoparticlesAdjuvant agent for the treatment of
Hodgkin’s Lymphoma
Mice[131]
Curcumin, NanoparticlesContrasting agentSprague–Dawley rats[132]
Curcumin C3 ComplexCancer preventionMale C57BL/6 wild-type mice[133]
Curcumin, HydrogelWound-healing agentMus musculus var. albino mice[134]
PVA/Chitosan/Curcumin PatchesWound-healing agentWistar Rats[135]
Table 5. Minimum inhibitory concentrations (MIC) of curcumin and fractional inhibitory concentration indices (FICIs) for potentially important pathogens of burn wounds [141].
Table 5. Minimum inhibitory concentrations (MIC) of curcumin and fractional inhibitory concentration indices (FICIs) for potentially important pathogens of burn wounds [141].
IsolateGenesCurcumin
MIC µg/mL
FICI
Klebsiella pneumonieDHA1280.5
Pseudomonas aeruginosaVEB1280.5
Acinetobacter baumanniOXA-23, OXA-241280.37
Acinetobacter baumanniOXA-23, OXA-241281
Pseudomonas aeruginosaIMP-11281
Enterococcus faecalis ATCC 29212Type strain1280.26
Pseudomonas aeruginosaGES1280.75
Acinetobacter baumanniOXA-23, OXA-245120.25
Acinetobacter baumanni ATCC19606Type Strain5120.5
Acinetobacter baumanniOXA-23, OXA-245120.25
Pseudomonas aeruginosaIMP-15120.064
Pseudomonas aeruginosaVIM-15120.064
Escherichia coli ATCC 25922Type Strain2560.4
Klebsiella pneumonie ATCC700603Type Strain2560.5
Klebsiella pneumonieNDM-62560.28
Klebsiella pneumonieNDM-12560.56
Klebsiella pneumonieNDM-62560.56
Pseudomonas aeruginosaIMP-22560.56
Table 6. Antiviral, antiparasitic and antimalarial activity of curcumin.
Table 6. Antiviral, antiparasitic and antimalarial activity of curcumin.
ActivitySubstanceType of MicroorganismTherapeutic EffectReference
Antiviral
Curcumin,
nanomicelles
Hepatitis C virus The antiviral effects of curcumin nanomicelles on hepatitis C virus.[167]
CurcuminVesicular stomatitis virus Determination of curcumin effects on vesicular stomatitis virus Dicer-1 expression.[168]
CurcuminChikungunya virus, zika virus Antiviral activity of curcumin against zika and chikungunya virus.[147]
Curcumin,
nanoparticles
Human immunodeficiency virus 1 (HIV-1)Immunomodulatory activities of curcumin-stabilized silver nanoparticles on HIV-1.[169]
CurcuminEnterovirus 71 (EV71)Antiviral effects of curcumin on EV71.[170]
CurcuminHuman T lymphotropic virus 1 (HTLV-1)Determination of curcumin on the expression of c-FLIP in HTLV-1-associated myelopathy/tropical spastic paraparesis (HAM/TSP) patients.[171]
CurcuminKaposi’s sarcoma-associated herpesvirus (KSHV or HHV8)Antiviral activity of curcumin against KSHV replication and pathogenesis.[146]
CurcuminHuman immunodeficiency virus 1 (HIV-1)Multifunctional mesoporous curcumin encapsulated iron phenanthroline Nanocluster on HIV-1.[172]
CurcuminZika virus Inhibitory effects of novel natural products against zika virus.[173]
Curcumin,
nanocurcumin
Dengue virus Antiviral activity of curcumin encapsulated in nanoemulsion against dengue virus serotypes.[174]
CurcuminTransmissible gastroenteritis virus Antiviral effects of curcumin on transmissible gastroenteritis virus.[175]
CurcuminHuman parainfluenza virus type 3 Evaluation of curcumin on replication of human parainfluenza virus type 3.[176]
CurcuminHepatitis B virus Evaluation of curcumin on viral entry of hepatitis B.[156]
Antiparasitic and
Antimalarial
Curcumin and netilmicinLeishmania major, Leishmania donovaniAntileishmanial activity of netilmicin combined with curcumin significantly enhanced compared with when used alone.[177]
Nanoformulation of curcumin and miltefosineLeishmania donovaniCombination therapy of curcumin with miltefosine exhibited a synergistic effect on both promastigotes and amastigotes under in vitro conditions.[166]
Curcumin
Encapsulated to
PLGA
Plasmodium bergheiEncapsulation of curcumin in PLGA led to increased parasite suppression about 56.8% at 5 mg/kg of nanoformulation, which was higher than in free curcumin (40.5%) at 10 mg/kg.[178]
Curcumin aloneGiardia lambliaCurcumin inhibited giardia proliferation disrupted the cytoskeletal structures of trophozoites in the dose-dependent mode.[160]
Curcumin aloneFasciola giganticaA significant decrease was observed in the expression of glutathione-S-transferase and superoxide dismutase.[179]
Curcumin aloneCryptosporidium parvumThe anticryptosporidial and antioxidant activity of curcumin against C. parvum were confirmed.[180]
Nanotized
curcumin-
benzothiophene conjugate
Plasmodium falciparumThe improved oral bioavailability of the nanotized formulation lowered the dosage at which the pharmacological effect was achieved while avoiding any observable adverse side effects.[181]
Curcumin,
nanocomposite
Plasmodium falciparumThe antiparasitic effect of the nanocomposite on the metabolites of plasmodium falciparum[182]
Table 7. Biological activities of curcumin.
Table 7. Biological activities of curcumin.
ActivitySubstanceTargetTherapeutic EffectReference
Anti-inflammatory
CurcuminCOX-2
NF-κB
p-IκB
ROS
Attenuates colistin-induced neurotoxicity in N2a cells via anti-inflammatory activity, suppression of oxidative stress, and a apoptosis.[206]
CurcuminNF-κB
COX-2
Attenuates airway inflammation and airway remoulding in cigarette smoke-induced COPD mice.[207]
Curcumin and rutinCOX-2Reduce tumour-associated inflammation in HPV16-transgenic mice.[189]
Curcumin, curcumin and capsaicinCOX-2
IL-6
TGF-β
Combined curcumin and capsaicin are efficient against the lipopolysaccharide Induced expression of proinflammatory cytokines in peripheral blood mononuclear cells.[208]
Antioxidant
Curcumin-loaded sodium alginate/ZnO hydrogel beadsDPPH AssayComposite hydrogel beads have protected curcumin from light degradation can therefore prolong its antioxidant activity.[209]
CurcuminMDA
SOD
GSH-Px
Curcumin protects the liver, kidneys and brain from the oxidative damage caused by irradiation.[210]
Curcumin,
curcumin and piperine
MDA
SOD
Catalase
Curcumin may be used as an adjunct therapy in individuals with oxidative stress.[211]
CurcuminMDA
total Antioxidant Capacity (TAC)
Pure curcumin reduces MDA concentration and increases total antioxidant capacity.[212]
Antinociceptive
CurcuminDRG Neurons β-Endorphin and EnkephalinThe curcumin attenuates cancer-induced bone pain[213]
Curcuminγ-Aminobutyric Acid (GABA) and Opioid ReceptorsAntinociception of curcumin[214]
Curcumin-loaded PLGA nanovesicles
(PLGA-CUR)
Cytokine and BDNFAntinociceptive effects of PLGA-CUR[201]
Curcumin
The acid-Sensing Ion Channels (ASICs)Antinociceptive Effects of Curcumin[215]
Wound healing agent
PVA/chitosan/curcumin patchesCell Line Studies and MTT AssayAntibacterial activity of PVA/Chi/Cur against four major bacterial strains commonly found in wound sites and water retainability indicates it to be a perfect material for wound treatment.[135]
NanocurcuminFibroblast, Collagen, ReepithelizationCurcumin nanoformulation enhanced wound repair by inhibiting the inflammatory response, stimulating angiogenesis, inducing fibroblast proliferation and enhancing reepithelization and synthesis of collagen.[127]
Curcumin,
Hydrogel
L929 Fibroblast CellsCurcumin incorporation accelerates full-thickness skin wound healing[114]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Urošević, M.; Nikolić, L.; Gajić, I.; Nikolić, V.; Dinić, A.; Miljković, V. Curcumin: Biological Activities and Modern Pharmaceutical Forms. Antibiotics 2022, 11, 135. https://doi.org/10.3390/antibiotics11020135

AMA Style

Urošević M, Nikolić L, Gajić I, Nikolić V, Dinić A, Miljković V. Curcumin: Biological Activities and Modern Pharmaceutical Forms. Antibiotics. 2022; 11(2):135. https://doi.org/10.3390/antibiotics11020135

Chicago/Turabian Style

Urošević, Maja, Ljubiša Nikolić, Ivana Gajić, Vesna Nikolić, Ana Dinić, and Vojkan Miljković. 2022. "Curcumin: Biological Activities and Modern Pharmaceutical Forms" Antibiotics 11, no. 2: 135. https://doi.org/10.3390/antibiotics11020135

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