Next Article in Journal
Influence of the Core Formulation on Features and Drug Delivery Ability of Carbamate-Based Nanogels
Next Article in Special Issue
Setup and Validation of a Reliable Docking Protocol for the Development of Neuroprotective Agents by Targeting the Sigma-1 Receptor (S1R)
Previous Article in Journal
Silencing of HvGSK1.1—A GSK3/SHAGGY-Like Kinase–Enhances Barley (Hordeum vulgare L.) Growth in Normal and in Salt Stress Conditions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Obstacles against the Marketing of Curcumin as a Drug

by
Kambiz Hassanzadeh
1,2,3,
Lucia Buccarello
1,
Jessica Dragotto
1,
Asadollah Mohammadi
3,
Massimo Corbo
4 and
Marco Feligioni
1,4,*
1
European Brain Research Institute (EBRI) Rita Levi Montalcini Foundation, Viale Regina Elena 295, 00161 Rome, Italy
2
Department of Biotechnology and Applied Clinical Sciences, University of L’Aquila, 67100 L’Aquila, Italy
3
Cellular and Molecular Research Center, Research Institute for Health Development, Kurdistan University of Medical Sciences, Sanandaj 66177-15175, Iran
4
Department of Neurorehabilitation Sciences, Casa Cura Policlinico, 20144 Milano, Italy
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2020, 21(18), 6619; https://doi.org/10.3390/ijms21186619
Submission received: 13 August 2020 / Revised: 7 September 2020 / Accepted: 9 September 2020 / Published: 10 September 2020

Abstract

:
Among the extensive public and scientific interest in the use of phytochemicals to prevent or treat human diseases in recent years, natural compounds have been highly investigated to elucidate their therapeutic effect on chronic human diseases including cancer, cardiovascular disease, and neurodegenerative disease. Curcumin, an active principle of the perennial herb Curcuma longa, has attracted an increasing research interest over the last half-century due to its diversity of molecular targets, including transcription factors, enzymes, protein kinases, growth factors, inflammatory cytokines, receptors, and it’s interesting pharmacological activities. Despite that, the clinical effectiveness of the native curcumin is weak, owing to its low bioavailability and rapid metabolism. Preclinical data obtained from animal models and phase I clinical studies done in human volunteers confirmed a small amount of intestinal absorption, hepatic first pass effect, and some degree of intestinal metabolism, might explain its poor systemic availability when it is given via the oral route. During the last decade, researchers have attempted with new pharmaceutical methods such as nanoparticles, liposomes, micelles, solid dispersions, emulsions, and microspheres to improve the bioavailability of curcumin. As a result, a significant number of bioavailable curcumin-based formulations were introduced with a varying range of enhanced bioavailability. This manuscript critically reviews the available scientific evidence on the basic and clinical effects and molecular targets of curcumin. We also discuss its pharmacokinetic and problems for marketing curcumin as a drug.

1. Introduction and History

Natural ingredients have been used during human history for numerous purposes, including alimentary usages but, due to their pharmacological effects, it has grown in the years the scientific interest to use these compounds to prevent or treat human diseases. Therefore, recent years have witnessed a growing understanding of the effect of natural products as sources of new supplements and drugs. Curcumin is one of such compounds with a history that goes back to 5000 years before [1]. Turmeric derived from the rhizome of the plant Curcuma longa (C. longa) has been used by the Indian people for centuries with no known side effects, not only as a food color and flavor but also to treat a wide variety of illnesses [2]. Turmeric was stated in the writings of Marco Polo regarding his journey to India and China in 1280 AC and it was introduced to Europe, for the first time, in the 13th century by Arab traders [1].
The main polyphenol constituents of Turmeric (C. longa) are curcuminoids that have three main chemical components, including curcumin (75–80%), which was found to be the key colored compound, demethoxycurcumin (15–20%), and bisdemethoxycurcumin (3–5%) [3,4]. Curcumin is the yellow pigment of the Indian Spice Turmeric and it was first isolated almost two centuries ago in 1815 by two German scientists, Vogel and Pelletier. Curcumin is a water-insoluble powder obtained in crystalline form in 1870 [5] and eventually identified as 1,6-heptadiene-3,5-dione-1,7-bis(4-hydroxy-3-methoxyphenyl)-(1E,6E)or diferuloylmethane [6].
According to the PubMed database, the first report on antibacterial effect of curcumin was published in 1949 in Nature and the first clinical trial was published in The Lancet in 1937. Although PubMed indicates more than 9200 publications on curcumin, interestingly almost 7500 of those have been published in the last 10 years.
Scientific documents are still published, focusing on the different activity and therapeutic efficacy of curcumin. In 2015, Kumar and colleagues [7] created and developed the Curcumin Resource Database (CRDB) that aims to perform as a gateway to access all relevant data and related information on curcumin and its analogs. This database involves 1186 curcumin analogs, 195 molecular targets, 9075 peer-reviewed publications, 489 patents, and 176 varieties of C. longa obtained by extensive data mining and careful curation from numerous sources [7]. Curcumin has been commonly used in traditional Indian medicine to treat a wide variety of illnesses including hepatic disorders, rheumatism, biliary disorders, anorexia, cough, and sinusitis. Turmeric paste is also available in many Indian homes for treatment of inflammation and wounds.
In recent years, a huge amount of reports and information regarding the different roles of curcumin has been released. Curcumin has been found as a potent antioxidant and reactive oxygen/nitrogen species (ROS, RNS) scavenger [8]. Besides, it has been reported to induce an upregulation of antioxidant enzymes such as superoxide dismutase, glutathione peroxidase and reductase, catalase, etc. [9].
The other main function of curcumin goes back to its protective activity on mitochondrial function and dynamic [10,11] and prevention of neuroinflammation [12], and apoptosis [13,14].
Concerning the above background and history of curcumin, in this paper, we review several aspects including cellular studies to clinical trials related to curcumin, its cellular and molecular targets, its cellular toxicity, and the problems related to its low pharmacokinetic and the current attempts to increase curcumin absorption and bioavailability.
Finally, we discuss the fact that, despite a wide range of cellular, molecular, and clinical studies and interesting results, why curcumin has not yet been licensed as a drug and it couldn’t play a higher-order role in treatment of diseases?

2. Disease Targets of Curcumin: From Cell Lines to Clinical Trials Studies

In this section, we will discuss the several effects of curcumin, including antimicrobial, gastrointestinal, cardiovascular, anti-cancer, ant-inflammatory, neuroprotective in different disorders which have been previously reported. These findings, comprising cell line studies, animal studies, and clinical trials. Then, the possible mechanisms will be described based on disease categories. As shown in Figure 1, there is a wide variety of studies on curcumin ranging from gastrointestinal, central nervous system, and cardiovascular effects, to antimicrobial and anti-cancer characteristics.

2.1. Antimicrobial Effect of Curcumin

As shown in Table 1, there are some in vitro studies with different kinds of curcumin extracts and few number of clinical studies indicating the antimicrobial property of curcumin against organisms like bacteria, fungi, parasite and virus. Antimicrobial studies started with anti-parasitic effect of curcumin and the anti-bacterial characteristic has attracted the most attention.
The stability of FtsZ protofilaments as a vital factor for bacterial cytokinesis has been introduced as a drug target for the antibacterial agents. Inhibition of assembly dynamics of FtsZ in the Z-ring may suppress the bacterial cell proliferation as one of the probable curcumin antibacterial mechanisms of action [15]. Accumulated evidence indicated curcumin plays an inhibitory role against numerous viral infections. The mechanisms involved in antiviral effect of curcumin are either a direct interference of viral replication machinery or suppression of cellular signaling pathways essential for viral replication, such as phosphoinositide 3-kinase (PI3K)/protein kinase B (AKT), nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB) [16].
Studies indicating antimicrobial effects of curcumin in the last decade focused on the pharmacokinetics of curcumin rather than pharmacodynamics, reflecting the concerns of researchers in this area. There is evidence suggesting that mode of antimicrobial activity of curcumin depends on the properties of the delivery system [17]. For instance, in an attempt to compare the antimicrobial activity toward Escherichia coli of two curcumin formulations: methyl-β-cyclodextrin supramolecular inclusion complex and polyelectrolyte-coated monolithic nanoparticles, researchers showed that while curcumin–β-cyclodextrin complexes exhibited a potent bactericidal activity, the curcumin nanoparticles were bacteriostatic [17].
Table 1. Antimicrobial Effect of Curcumin.
Table 1. Antimicrobial Effect of Curcumin.
ProductDose or Concentration UsedEffect and FindingsType of StudyStudied by
CurcuminoidsConcentration 0.1–1 mg/mLAnti-parasitic
Nematocidal activity of mixed Curcuminoids
In vitro Kiuchi F et al. 1993 [18]
Curcumin
Curcumin + boric acid + oxalic acid dihydrate
(boron complex)
Curcumin:
IC50: 100 µM
boron complex:
IC50: 6 µM
Anti-viral
inhibitor of the HIV-l
In vitro Sui Z et al. 1993 [19]
Curcumin
essential oil extracted
Concentration of 0.1% in the mediumAnti-fungal
aflatoxin synthesis by Aspergillus parasiticus
In vitroTantaoui-Elaraki A et al. 1994 [20]
Curcumin2.5 g which was repeated 7 days laterAnti-viral
inhibitor of the HIV-1
Clinical Trial (3 Subjects)Jordan W.C et al. 1996 [21]
Turmeric extractantifungal activity against Candida albicans at 1 µg/mLAnti-fungal
Candida albicans
In vitroRoth GN et al. 1998 [22]
Turmeric oilAnti-bacterial activity in 50–200 ppmAnti-bacterialIn vitroNegi P.S et al.1999 [23]
Curcumincytotoxicity against leishmania in vitro. The LD50 = 37.6 µMAnti-parasitic
against leishmania
In vitro Koide T et al. 2002 [24]
Curcumin extractAnti-fungal at 50–500 mg/LAnti-fungalIn vitroKim MK et al. 2003 [25]
CurcuminIn vitro: IC50: 5 µM
Animal: once daily for 5 days at a dose of 100 mg/kg
Anti-malarialIn vitro & Animal modelReddy RC et al. 2005 [26]
Curcumin at 30 and 100 μMAnti-parasitic
Giardia lamblia trophozoites
In vitro Pérez-Arriaga L et al. 2006 [27]
Curcumin 30 mg every 12 h for 7 daysAnti-bacterialClinical trial (25 Subjects)Di Mario F et al. 2007 [28]
Curcumin extract
quercitin and curcumin (FlogMEV) extracts
In patients with prostatitis
quercitin (100 mg) and curcumin (200 mg) for 14 days
Anti-bacterialClinical trial (284 Subjects)Cai T et al. 2009 [29]
Curcumin nanoparticleConcentration of 260 μMAnti-bacterialIn/vitroTrigo Gutierrez JK et al. 2017 [30]
Curcumin nanoparticle0.1 and 0.2 mg per well concentrationAnti-bacterialIn vitroFakhrullina G et al. 2019 [31]
Curcumin nanoparticle
curcumin-silver nanoparticles
Minimum inhibitory concentration 20 mg/LAnti-bacterialIn vitroJaiswal S et al. 2018 [32]
Iodinated curcuminMinimum inhibitory concentration 150 and 120 µg/mLAnti-bacterialIn vitroManchanda G et al. 2018 [33]
In Table 1, there is an overview of the multipotent antimicrobial character of curcumin that would be useful for investigators to additional studies on curcumin’s antimicrobial role against new microorganisms.

2.2. Gastrointestinal Effect of Curcumin

Similar to the antimicrobial studies, the gastrointestinal activity of curcumin has been tested in animal models and there are also few human clinical studies. Curcumin has been suggested as a remedy for liver and digestive diseases such as irritable bowel syndrome, colitis, Crohn’s disease and bacterial and parasitic diseases. Data in Table 2 shows that most of the reports are related to liver protection and some limited studies indicate the intestine and colon protection properties of curcumin. On the other hand, some reports are about its beneficial effects in animal models recapitulating inflammatory bowel disease. In addition to these effects, in several animal studies, curcumin has been revealed to influence the composition of gut microbiome. For instance, adding the curcumin to a high-fat diet in animals, prevented the fat-induced changes of the gut microbiota [34].
Several lines of evidence support the hepatoprotective effect of curcumin. The findings in this area indicated that curcumin exerts notable protective and therapeutic effects in oxidative liver diseases through numerous molecular mechanisms including the pro-inflammatory cytokines and lipid perodixation products suppression, activation of PI3K/AKT and hepatic stellate cells and ameliorating cellular responses to oxidative stress [35]. As shown in Table 2, the data is mostly obtained from animal studies and further clinical studies are needed to clarify the mechanisms of curcumin in oxidative associated liver diseases.
Research on colitis in animal models indicate a significant decrease in pro-inflammatory biomarkers and modulation of inflammatory cells [36,37,38]. In a study, it has been strongly suggested that curcumin diminishes neutrophil recruitment to the inflammatory sites through interference in chemokine gradient formation in addition to the direct effect on neutrophil polarization, chemotaxis, and chemokinesis [38]. These mechanisms likely significantly contribute to the described protective and potent effect of curcumin in ulcerative colitis. In patients with ulcerative colitis, curcumin has been well-tolerated and led to ameliorate the symptoms and inflammatory markers in clinical trials [39,40].
Curcumin has been reported to be promising in the treatment of intestinal inflammatory diseases (IBD). It provides some beneficial effects on the microbiome, inhibition of toll-like receptors (TLR4)/NF-κB/activator protein1 (AP-1) signal transduction, alterations in cytokine profiles and immune cell maturation and differentiation. These are possible mechanisms contributing to IBD pathology thatare affected by curcumin to strengthen the intestinal barrier [41].
Effects of curcumin on the microbiome have been widely studied and it has been reported that these effects are different based on the disease characteristics. For instance, in a research on mice living in specific-pathogen-free conditions, curcumin reduced the microbial richness and diversity [42]. In another study on rat model of hepatic steatosis, administration of curcumin decreased species richness and diversity induced notable microbiota compositional changes compared to both high-fat diet and control groups. Also, it shifted the structure of the gut microbiota [43]. These microbiome changes inhibit intestinal inflammation. In fact, curcumin maintains short-chain fatty acid-producing bacteria, which acts to provide intestinal mucosal protection [44,45]. Table 2 shows some studies indicating the effect of curcuminoids and curcumin extract on different pathological conditions in the gastrointestinal system ranging from cell line to clinical studies.
Table 2. Gastrointestinal Effect of Curcumin.
Table 2. Gastrointestinal Effect of Curcumin.
ProductDose or Concentration UsedEffect and FindingsType of StudyStudied by
CurcuminConcentrations
5–30μM
Liver protective through inhibiting hepatic stellate cells activationIn vitroTang Y et al. 2010 [46]
Curcumin Dose: 25 μg daily for 10 weeks, intraperitoneal Liver protective: effectively limits the development and progression of fibrosisAnimal modelVizzutti F et al. 2010 [47]
Curcumin 300 mg/kg, by gavage daily for 12 weeksLiver protective: inhibited the development of liver cirrhosis mainly due to its anti-inflammatory activities and not by a direct anti-fibrotic effectAnimal modelBruck R et al. 2007 [48]
Curcumin 1 g after the evening meal for 6 monthsAmeliorate ulcerative colitis
remission in patients with ulcerative colitis
Clinical trial (89 Subjects)Hanai H et al. 2006 [39]
Curcumin 550 mg of curcumin
twice daily for 1 month and then 550 mg three times daily
for another month.
Reductions in concomitant medications Crohn’s diseaseClinical trial (5 Subjects)Holt PRet al. 2005 [49]
Curcumin Concentrations
10–30 μM
Ameliorate Inflammatory bowel disease: dose-dependent suppression of metalloproteinase-3 in colonic myofibroblasts from children and adults with active IBDIn vitroEpstein J et al. 2010 [50]
Curcumin Dose: 75 mg/kg/day orally daily for 6 weeksLiver protective: prevents chronic alcohol-induced liver disease involving decreasing ROS generation and enhancing antioxidative capacityAnimal modelRong S et al. 2012 [51]
Curcumin Dose: 150 mg/kg, orally daily for 6 weeksLiver protective: by inhibition of oxidative stress via mitogen-activated protein kinase/nuclear factor E2-related factor 2 Animal modelXiong ZE et al. 2015 [52]
Curcumin Dose: 150 mg/kg, orally daily for 8 weeksLiver protective: prevention of the oxidative stress induced by chronic alcoholAnimal modelVaratharajalu R et al. 2016 [53]
Curcumin Dose: 70 mg/kg, orally daily for 8 weeksLiver protective: improvement of different features of Non-alcoholic fatty liver disease after a short-term supplementationClinical trial (80 Subjects)Rahmani S et al. 2016 [54]
Curcumin Curcumin (2%) diet from 4 to 18 weeks of ageIntestine protective: beneficial effects of dietary curcumin on intestinal tumorigenesis in rodent models of colon cancer.Animal modelMurphy E.A et al. 2011 [55]
CurcuminDose: 50 mg/kg, orally daily for 10 Inflammatory bowel disease: beneficial effects in experimental colitis and may, therefore, be useful in the treatment of IBD. Animal modelUkil A et al. 2003 [56]
CurcuminCurcumin (2%) diet from 9 days ulcerative colitis: dietary curcumin may be of different value for Crohn’s disease and ulcerative colitis. Animal modelBillerey-Larmonier C et al. 2008 [57]

2.3. Cardiovascular Protective Effect of Curcumin

The cardiovascular protective effects of curcumin have been extensively studied and its use as an adjuvant or therapeutic agent to alleviate cardiovascular disease and other vascular dysfunctions is currently being investigated. As shown in Table 3, the most evaluations on cardiovascular effects of curcumin have been made in the last decade and cardioprotective and anti-hyperlipidemia have more considered for curcumin. The majority of studies are in animal models using curcumin extract, while there are few assessments with nanoparticles.
Curcumin at a dose of 200 mg/kg, nine days (7 days before and 2 days following Adriamycin), significantly prevented adriamycin-induced cardiotoxicity. The mechanism of cardioprotective effects of curcumin is unclear and multiple mechanisms have been proposed in this area including (1) curcumin prevents lipid peroxidation through free radicals scavenging, leading to a blocking of the lipid chain reaction. (2) Curcumin exerts membrane-stabilizing effect, which is supported by the fact that it could prevent the ECG changes induced by adriamycin [58].
In diabetic patients with cardiovascular complications, curcumin has been shown to downregulate nitric oxide synthase (NOS) and reduce NO oxidation, which plays a key role in the pathogenesis of cardiovascular complications in diabetes [59]. In an animal study, researchers demonstrated that the myocardial tissue from diabetic rats exhibited increased levels of NOS enzyme mRNA as compared to control rats which was prevented by curcumin treatment showing a decrease in the oxidative DNA damage [60].
Cardiac hypertrophy is the heart’s response to some variety stimuli such as workload or myocardial infarction that increase biomechanical stress. It is characterized by an increase in the size of the individual cardiac myocytes and enlargement of the myocardium and the whole heart [61]. This disorder threatens affected patients with progression to overt heart failure or sudden death.
Histone acetylation is one of the important control points for the regulation of genes in the hypertrophic myocardium [62]. Curcumin has been reported to be an inhibitor of histone acetyltransferases (HATs), p300-HAT, which perform the acetylation of histone tails [63], therefore, it may have an effect in the prevention of the cardiac hypertrophy and heart failure [64].
There is also limited evidence on anti-hyperchlostrolemia effect of curcumin. In an animal study, administration of curcumin extract had hypolipidemic effects in high cholesterol-induced hypercholesterolemic mice. The authors reported that a mixture of Nelumbo nucifera leaf (NL) and C. longa provided more advanced protection against high cholesterol diet-related lipid accumulation and liver dysfunction and may be a more effective functional food for the management of hypercholesterolemia [65]. In another animal study, curcumin has been found to have hypocholesterolemic effect on both normal and hypercholesterolemic rats and was more effective in hypercholesterolemic animals. The possible proposed mechanism for this effect was, interfering with intestinal cholesterol uptake, augmentation of the conversion of cholesterol into bile acids, and increasing the excretion of bile acids [66].
Despite the extensive research regarding the effect of curcumin on the cardiovascular diseases in animals, additional studies in humans on curcumin’s cardiovascular role and mechanisms using new formulations exerting a high degree of bioavailability are required.
Table 3. Cardiovascular Protective Effect of Curcumin.
Table 3. Cardiovascular Protective Effect of Curcumin.
ProductDose or Concentration UsedEffect and FindingsType of StudyStudied by
Curcumin Dose: 25–50–100–200 mg/kg, orally daily for 10 daysCardioprotective: curcumin (50 mg/kg) with piperine (20 mg/kg) exhibited profound cardioprotection compared to curcumin (200 mg/kg) alone-treated group.Animal modelChakraborty M et al. 2017 [67]
Curcumin Dose: 120 mg/kg, orally daily for 67 daysCardioprotective: through direct antioxidant effects and indirect strategies that could be related to protein kinase C-activated downstream signaling.Animal modelCorrea F et al. 2013 [68]
Curcumin Dose: 200 mg/kg, orally daily for 4 weeksCardioprotective: cardioprotective effect could be attributed to antioxidant.Animal modelSwamy AV et al. 2012 [69]
Curcumin Dose: curcumin (100 mg/kg) plus piperine (5 mg/kg) orally daily for 4 weeksAnti-hypercholesterolemia: co-administration of curcumin plus piperine increasing the activity and gene expression of ApoAI, CYP7A1, LCAT, and LDLR, providing a promising combination for the treatment of hyperlipidemia.Animal modelTu Y et al. 2014 [70]
Curcumin Dose: curcumin 100 mg/kg orally daily for 6 weeksCardioprotective: concomitant co-administration of curcumin and metformin revealed more protection than metformin alone through Inhibition of JAK/STAT pathway and activation of Nrf2/HO-1 pathway Animal modelAbdelsamia E.M et al. 2019 [71]
Curcumin nanoparticle: curcumin and nisin based poly lactic acid nanoparticle (CurNisNp)Dose: 10 and 21 mg/kg injection daily for 7 daysCardioprotective: curcumin nanoparticle confers a significant level of cardioprotection in the guinea pig and is nontoxic.Animal modelNabofa W.E.E et al. 2018 [72]
Curcumin Dose: curcumin 100 mg/kg orally daily for 24 daysCardioprotective: Curcumin improve the heart function and structural changes in doxorubicin-treated ratsAnimal modelJafarinezhad Z et al. 2019 [73]
Curcumin nanoparticleDose: 100–150 mg/kg orally daily for 15 daysCardioprotective: curcumin nanoparticles exert better antioxidative effects on MI compared to conventional curcumin, thus improving myocardial function more effectively and limiting the extension of heart damage.Animal modelBoarescu PM, et al. 2019 [74]
Curcumin Dose: 100 mg/kg orally daily for 7 daysCardioprotective: protects against myocardial infarction-induced cardiac fibrosis via SIRT1 activationIn vitro and in vivoXiao J et al. 2016 [75]
CurcuminoidsDose: 4 g orally daily for 8 daysCardioprotective: significantly decreased MI associated with coronary artery bypass grafting through the antioxidant and anti-inflammatory effects Clinical trial (121 Subjects)Wongcharoen W et al. 2014 [76]
Curcumin Concentration: 5 μmol/LVascular protective: effectively reverses the endothelial dysfunction induced by homocysteineIn vitroRamaswami G et al. 2004 [77]
CurcuminCurcumin (0.05-g/100-g diet) for 10 weeksAnti-hyperlipidemia: curcumin exhibits an obvious hypolipidemic effect by increasing plasma paraoxonase activity, ratios of high-density lipoprotein cholesterol to total cholesterol and of apo A-I to apo B, and hepatic fatty acid oxidation activity with simultaneous inhibition of hepatic fatty acid and cholesterol biosynthesis in high-fat–fed hamsters.Animal modelJang EM et al. 2008 [78]
Curcumin Curcumin (0.02% w/w diet) for 18 weeksAnti-atherogenic: Long-term curcumin treatment lowers plasma and hepatic cholesterol and suppresses early atherosclerotic lesions comparable to the protective effects of lovastatin.Animal modelShin S.K et al. 2011 [79]
Curcumin extract: hydro-alcoholic extract of rhyzome of C. longa containing ∼10 mg of curcuminDose: 20 mg orally daily for 30 daysAnti-hyperlipidemia: decreases significantly the LDL and apo B and increases the HDL and apo A of healthy subjectsClinical trial (30 Subjects)Ramírez-Boscá A et al. 2000 [80]

2.4. Anti-Cancer Effect of Curcumin

Numerous molecular targets have been proposed concerning the chemotherapeutical effect of curcumin against different types of cancer. Extensive studies, mainly in vitro or in animal models, indicate that it can modulate all kinds of cancer hallmarks, including uncontrolled cell proliferation, cancer-associated inflammation, cancer cell death, signaling pathways, cancer angiogenesis, and metastasis [81]. According to a report on curcumin in 2019, 37% of all researches on curcumin focus on its anti-cancer effect or relevant mechanisms. As shown in Table 4, effect of curcumin in different organ cancers have been widely studied, almost all studies have been done in vitro or in animal models. Based on these reports, curcumin, through different signaling pathways, represents a promising candidate as an effective anti-cancer agent to be used alone or in combination with other drugs. It affects molecular targets involved in the development of several cancers.
Several lines of evidence revealed that inflammatory pathways disorder plays a key role in cancer development [82]. Increase in the production of pro-inflammatory cytokines such as tumor necrosis factor alpha (TNF-α) and interleukins, transcription factors including nuclear factor κB (NF-κB), ROS, cyclooxygenase (COX-2), protein kinases B (AKT), activator protein 1 (AP1), signal transducer and activator of transcription 3 (STAT3), causing the initiation and development of cancer [83,84].
Curcumin exerts its immunomodulatory characteristics by interacting with above-mentioned immune mediators, hence its anti-cancer property.
There are few clinical trials for use of curcumin in cancers, either as a monotherapy or in combination with other anti-cancer agents. In a phase I clinical trial, curcumin was administered orally in 15 colorectal cancer patients. The findings reveal that there was not a significant toxicity except a diarrhea symptom reported in two patients, and two patients showed stable disease after two months of curcumin treatment [85]. Another monotherapy phase II clinical trial study of oral formulation of curcumin administered to 25 advanced pancreatic cancer patients showed low levels of curcumin in plasma (22–41 ng/mL) but two patients showed clinical biological activity. In one patient, it was reported disease stability for >18 months. In another patient, a brief but marked tumor regression (73%) was found [86].
There are some combination therapy studies using curcumin in combination with other chemotherapeutic agents used in standard treatments of cancer disease. In a clinical study, a combination therapy of curcumin with imatinib (tyrosine kinase inhibitor) has been evaluated in 50 chronic myeloid leukemia patients in which the synergic action of the two drugs was more efficient than imatinib alone, although additional studies are needed to confirm this efficacy [87]. Furthermore, a combination of curcumin with anti-EGFR (epidermal growth factor receptor) monoclonal antibodies in cutaneous squamous cell carcinoma patients has been shown as a highly effective strategy in disease control in another clinical [88]. Table 4 summarized some of studies indicating anti-cancer effects of curcumin in different organs in the last two decades.
Table 4. Anti-cancer Effect of Curcumin.
Table 4. Anti-cancer Effect of Curcumin.
ProductDose or Concentration UsedEffect and FindingsType of StudyStudied by
Curcumin Concentration
15 µM
Prostate cancer. chronic inflammation can induce a metastasis prone phenotype in prostate cancer cells: Curcumin disrupts this feedback loop by the inhibition of NFκB signalingIn vitroKillian PH et al. 2012 [89]
Curcumin Concentration
50 µM for 1–4 h
Colon cancer: curcumin is an activator of PTPN1 and can reduce cell motility in colon cancer via dephosphorylation of pTyr(421)-CTTN, which could be exploited for novel therapeutic approaches in colon cancerIn vitroRadhakrishnan VM et al. 2014 [90]
Curcumin or tetrahydrocurcumin (THC)Curcumin: 300 mg/kg
THC: 3000 mg/kg for 21 days
Anti-cancer: anti-angiogenic properties of Curcumin and THC represent a common potential mechanism for their anti-cancer actions.Animal modelYoysungnoen P et al. 2008 [91]
Curcumin Concentration
0–20 μM
Breast cancer: curcumin suppresses chemokine-like ECM-associated protein osteopontin-induced VEGF expression and tumor angiogenesisIn vitroChakraborty G et al. 2008 [92]
Curcumin Concentration
3.12–50 µM
ovarian and endometrial cancers: curcumin suppresses JAK-STAT signaling via activation of PIAS-3, thus attenuating STAT-3 phosphorylation and tumor cell growth.In vitroSaydmohammed M et al. 2010 [93]
CurcuminConcentration
20–40 µM
Liver cancer: suppresses migration and proliferation of Hep3B hepatocarcinoma cells through inhibition of the Wnt signaling pathwayIn vitroKim HJ et al. 2013 [94]
Curcumin Concentration
(2, 20, and 50 μM) for 4 h
Burkitt’s lymphoma: curcumin might play an important role in radiotherapy of high-grade non-Hodgkin’s lymphoma through inhibition of the PI3K/AKT-dependent NF-κB pathway.In vitroQiao Q et al. 2013 [95]
Curcumin Concentration
0–20 μg/mL for 24 h
Osteosarcoma: curcumin caused death of HOS cells by blocking cells successively in G(1)/S and G(2)/M phases and activating the caspase-3 pathwayIn vitroLee DS et al. 2009 [96]
Curcumin Concentration
4–10 µM for 24 h
Glioma: curcumin exerts inhibitory action on glioma cell growth and proliferation through induction of cell cycle arrest In vitroLiu E et al. 2007 [97]
Curcumin Concentration
10, 25 µM for 24 h
Breast cancer: Curcumin induces apoptosis in human breast cancer cells through p53-dependent Bax inductionIn vitroChoudhuri T et al. 2002 [98]
Curcumin Concentration
0 to 20 μM for 24 h
Gastric carcinoma: curcumin inhibited the growth of the AGS cells and induced apoptosisIn vitroCao AL et al. 2015 [99]
Curcumin Concentration
0 to 100 μM for 72 h
Adenocarcinoma: curcumin-induced growth inhibition through G2/M arrest in Ras-driven cells and by apoptosis induction in Src-driven cells, In vitroOno M et al. 2013 [100]
Curcumin Concentration
0 to 40 μM for 24–72 h
Colon cancer: Curcumin suppresses proliferation of colon cancer cells by targeting Cyclin-dependent kinase 2In vitroLim TG et al. 2014 [101]
Curcumin micellesConcentration
0 to 100 μg/mL for 24 h
Lung cancer: mixed micelles of PF127 and GL44 significant improvement in curcumin oral bioavailability.In vitroPatil S et al. 2015 [102]
CurcuminoidsDose:
8 caplets daily for 8 weeks. Each caplet contains 1 g curcuminoids (900 mg curcumin, 80 mg desmethoxycurcumin, and 20 mg bisdesmethoxycurcumin)
Pancreatic cancer: Oral curcumin has biological activity in some patients with pancreatic cancer.Clinical trial (25 cases) Dhillon N et al. 2008 [86]
Curcumin Dose: 0.45 and 3.6 g daily for up to 4 months.Colorectal cancer: a daily dose of 3.6 g of curcumin are suitable for its evaluation in the prevention of malignancies at sites other than the gastrointestinal tract.Clinical trial (15 cases) phase ISharma RA et al. 2004 [85]

2.5. Effect of Curcumin on Skin Diseases

Growing evidence suggests that curcumin may show an effective role in the treatment of several skin disorders [103] as shown in the provided in the table below. The studies imply on the anti-inflammatory effect of curcumin except those that show apoptotic effects of curcumin for cancer indication (Table 5). Study on the possible interactions between curcumin and other chemicals, commonly used in topical skin treatments, may provide useful insights for the development of new effective preparations, tailored for different conditions.
In the topical route of administration, curcumin showed a good efficiency, especially when incorporated in new formulations such as polymeric bandages, chitosan-alginate sponges, nano-emulsion, alginate foams, collagen films, hydrogel, and β-cyclodextrin-curcumin nanoparticle complex, making curcumin suitable as a therapeutic agent for the topical treatment of skin diseases [104,105,106,107,108,109,110].
Table 5. Therapeutic Effect of Curcumin in skin diseases.
Table 5. Therapeutic Effect of Curcumin in skin diseases.
ProductDose or Concentration UsedEffect and FindingsType of StudyStudied by
CurcuminDose: 40 mg/kg orally daily for 20 daysPsoriasis: all psoriasis indexes including ear redness, weight, thickness and lymph node weight were significantly improvedAnimal modelKang D et al. 2016 [111]
Turmeric tonic Topical tonic Twice a day for 9 weeksPsoriasis: turmeric tonic significantly reduced the erythema, scaling and induration of lesions (PASI score), and also improved the patients’ quality of life Clinical trial (40 subjects)Bahraini P et al. 2018 [112]
Curcumin nano-fiber
chrysin-curcumin nano-fiber
Topical 5–7.5–10% w/w for 5, 10, 15 daysWound healing: chrysin-curcumin-loaded nanofibers have anti-inflammatory properties in several stages of the wound-healing process by affecting the IL-6, MMP-2, TIMP-1, TIMP-2, and iNOS gene expression.Animal modelMohammadi Z et al. 2019 [113]
Curcumin nanocapsuleDose: 6 mg/kg, intra-peritoneally, twice a week for 21 daysSkin cancer: curcumin caused significant reduction of cell viability in a concentration- and time-dependent manner.Animal modelMazzarino L et al. 2011 [114]
Curcumin Concentration
0 to 20 μM for 6, 12 h
Skin cancer, melanoma: curcumin-induced cell death and apoptosisIn vitroYu T et al. 2010 [115]

2.6. Neuroprotective Effect of Curcumin

Curcumin has several appropriate characteristics for a neuroprotective agent, including potent antioxidant, anti-inflammatory, and anti-protein-aggregate activities that have been previously reviewed [116,117]. Due to its above-mentioned pluripotent benefits, curcumin has countless potential for the prevention of many neurological conditions that existing therapeutics are less than optimal. These disorders include Parkinson’s, Alzheimer’s, Huntingtin’s, brain injury, stroke, and aging [118]. Data in Table 6 show a summary of the main findings on the potential beneficial effects of curcumin against neurodegeneration, and principal attention has been given to Alzheimer’s and Parkinson’s disease (PD).
Alzheimer’s disease (AD) is the most prevalent form of age-related dementia and currently, there are millions of AD patients in the world and this number is expected to increase dramatically with the demographic shift toward a more aged population, unless preventive processes would be achieved [119]. Classically, AD is characterized by the accumulation of amyloid and tau aggregation with the development of neurodegeneration and cognitive defects.
Curcumin has been tested in animal model of Alzheimer’s disease in which not only reduced oxidative stress damage and inflammation markers, but it also mitigates amyloid plaques accumulation and cognitive deficits in rats [120,121]. The possible mechanism for curcumin protection against tau aggregates might be related to its antioxidant characteristics since the initial step for tau dimerization is driven also by oxidative damage, lipid peroxidation, or redox-regulated disulfides [118,122].
Evidence in recent years supports the efficacy of curcumin in PD. In both in vitro and in vivo models of PD curcumin could prevent oxidative stress toxicity by reducing the production of ROS and malondialdehyde and restoring GSH levels [123,124,125] which shields against alpha-synuclein-induced toxicity in the brain [123]. More specifically, antioxidative and anti-apoptotic activity of curcumin has been reported in in vitro studies, indicating the neuroprotective effect of curcumin on dopaminergic neurons [124,125].
More recently, we found that curcumin, at the concentration of 5 µM, protected neuroblastoma (SH-SY5Y) cells against H2O2-induced cell death by modulating of Small Ubiquitin-like Modifier (SUMO)-1-JNK(c-Jun N-terminal kinases)-TAU axis, indicating that curcumin might be a promising therapeutic agent against not only oxidative stress, but also pathologies characterized through SUMO, JNK, and Tau alterations in many neurodegenerative diseases [126].
Other mechanisms of the neuroprotective effect of curcumin are from oxidative damage by restoring mitochondrial membrane potential, upregulation of Cu-Zn superoxide dismutase, and inhibiting the production of intracellular ROS [127].
Although, in all these studies, authors are pleased with their findings, adapted from in vitro and animal models studies, more preclinical and clinical trials are needed to verify the exact effect of curcumin in neurodegenerative disease using different formulations.
Table 6. Neuroprotective Effect of Curcumin in Neurodegenerative Diseases.
Table 6. Neuroprotective Effect of Curcumin in Neurodegenerative Diseases.
ProductDose or Concentration UsedEffect and FindingsType of StudyStudied by
Curcumin
a natural dietary supplement (NDS), containing extracts from Curcuma longa, silymarin, guggul, chlorogenic acid, and inulin
Dose: daily administration of NDS (0.9 mg/mouse) for 16 weeksNeuroprotective: NDS exerts neuroprotective effects in high fat diet-fed mice by reducing brain fat accumulation, oxidative stress and inflammation, and improving brain insulin resistance.Animal modelNuzzo D et al. 2018 [128]
CurcuminoidsConcentration
0.1–30 μM
few minutes before addition to artificial cerebrospinal fluid for the perfusion
Neuroprotective: curcuminoids can restore susceptibility for plastic changes in CA1 excitability that is injured by exposure to Aβ peptide and rescue sinking PS LTP in A β-peptide-exposed hippocampal CA1 neurons.In vitroAhmed Tet al. 2011 [129]
Curcumin Concentration
0–8 μM
Alzheimer’s Disease: curcumin effectively disaggregates Abeta as well as prevents fibril and oligomer formationAnimal modelYang F et al. 2005 [130]
CurcuminoidsConcentration
10 μM
Alzheimer’s Disease: curcumin binds to Aβ oligomers and to Aβ fibrilsIn vitroYanagisawa D et al. 2011 [131]
Curcumin Concentration
0–30 μM
Alzheimer’s Disease: curcumin significantly attenuated β amyloid-induced radical oxygen species production and β-sheet structure formation.In vitroShimmyo Y et al. 2008 [132]
Curcumin Concentration
0–10 μM
Alzheimer’s Disease: curcumin downregulated the expression of amyloid precursor protein and amyloid-β in swAPP695-HEK293 cells, which was through miR-15b-5pIn vitroLiu HY et al. 2019 [133]
CurcuminoidsDose: 3–30 mg/kgAlzheimer’s Disease: increased PSD-95, synaptophysin and camkIV expression levels in the hippocampus in the rat AD modelAnimal modelAhmed T et al. 2010 [134]
Ethanolic extract of turmericDose: 80 mg/kg orally, daily for three weeksAlzheimer’s Disease: effectively prevented cognitive deficitsAnimal modelIshrat T et al. 2009 [135]
Curcumin C3 Complex(®) an extract derived from the rhizomes (roots) of the plant Curcuma longaDose: 2, 4 g/day, orally for 24 weeks.Alzheimer’s Disease: Results were unable to demonstrate clinical or biochemical evidence of efficacy of this formulation.Clinical trial (36 Subjects)Ringman JM et al. 2012 [136]
Tumeric powder capsulesDose: 764 mg/day turmeric (100 mg/day curcumin) orally for 12 weeksAlzheimer’s Disease: a significant improvement of the behavioral symptoms in the AD with the turmeric treatment,Clinical trial (3 Subjects)Hishikawa N et al. 2012 [137]
Curcumin Concentration
0–1 μM
Parkinson’s Disease: Curcumin protected brain mitochondria against peroxynitrite by direct detoxification and inhibition of 3-nitrotyrosine formation and by elevation of total cellular glutathione levels in vivoIn vitroMythri RB et al. 2007 [138]
Curcumin nanoparticle
polymeric nanoparticle encapsulated curcumin
In vitro: (1, 10, 50, 100, 500 nM, 1, 5 μM)
In vivo: 25 mg/kg intraperitoneally twice daily for 4 weeks
Alzheimer’s Disease: NanoCurc™ ameliorated ROS-mediated damage in both cell culture and in animal modelsAnimal model/In vitroRay B et al. 2011 [139]
Curcumin Concentration 0–10 μM for 24 hNeuroprotection: curcumin enhanced neuronal survival against NMDA toxicity In vitroLin MS et al. 2011 [140]
Curcumin diet of 500 ppm curcumin for 4 weeks Traumatic brain injury (TBI): curcumin reduced oxidative damage, normalized levels of BDNF, synapsin I, and CREB and counteracted the cognitive impairment caused by TBI.Animal modelWu A et al. 2006 [141]
Curcumin Dose: 1.25, 2.5, 5, 10 mg/kg, intraperitoneally daily single doseDepression: exerts antidepressant-like effects through the central monoaminergic neurotransmitter systems.Animal modelXu Y et al. 2005 [142]
Curcumin Dose: 200 mg/kg,
intraperitoneally daily for 7 days.
Brain ischemia: curcumin attenuated forebrain ischemia-induced neuronal injury and oxidative stress in hippocampal tissue.Animal modelAl-Omar FA et al. 2006 [143]
Curcumin Dose: 100, 200, 300 mg/kg,
Orally, single dose
Epilepsy: Curcumin (300 mg/kg) significantly increased the latency to myoclonic jerks, clonic seizures as well as generalized tonic–clonic seizures and reduced oxidative stress and cognitive impairmentAnimal modelMehla J et al. 2010 [144]
Curcumin Dose: 50 mg/kg,
Orally, daily for 4 days
Parkinson’s Disease: curcumin protects the tyrosine hydroxylase-positive cells in the substantia nigra and dopamine levels in the striatum through its antioxidant capabilities Animal modelZbarsky V et al. 2005 [145]
Curcumin Concentration 0–25 μM for 24 hParkinson’s Disease: these protective effects are attributed to the antioxidative properties also modulation of nuclear factor kappaB translocation.In vitroWang J et al. 2009 [146]
Curcumin/its metaboliteDose: 80 mg/kg,
intraperitoneally, daily for 7 days
Parkinson’s Disease: curcumin and tetrahydrocurcumin reversed the MPTP induced depletion of dopamine and DOPAC through inhibition of MAO-B activity.Animal modelRajeswari A et al. 2008 [147]
Curcumin Concentration 4 μM for 48 hParkinson’s Disease: curcumin could alleviate α-synuclein-induced toxicity, decreased ROS levels and protected cells against apoptosis.In vitroWang MS et al. 2010 [148]
Curcumin Concentration 0–1 μM for 2 times changing in 6 days treatmentParkinson’s Disease: curcumin protects cells against A53T mutant α-synuclein-induced cell death through prevention of oxidative stress and the mitochondrial rescueIn vitroLiu Zet al. 2011 [123]
Manganese complexes of curcuminIn vitro: 0–5 μg/mL for 3 h
In vivo: 3 times (50 mg/kg × 3) at time points 1, 3, and 7 h post first MPTP sc injection, intraperitoneally
Neuroprotection: treatment with this complex attenuated MPTP-induced striatal dopamine depletion significantlyAnimal model/In vitroVajragupta O et al. 2003 [149]

2.7. Protective Effect of Curcumin in Eye Diseases

Table 7 summarized some studies suggesting the efficiency of curcumin in eye diseases. All the studies were carried out on experimental models using curcumin except one in which researcher used nanoparticles. Curcumin beneficial effects have been proved for major eye diseases [150] such as glaucoma [151,152], age-related macular degeneration [153], diabetic retinopathy [154], cataract [155,156], corneal neovascularization [157], dry eye disease [158] and conjunctivitis [159]. In general, a wide variety of mechanisms have been raised for these effects of which antioxidative stress, anti-angiogenesis, anti-inflammatory and anti-apoptotic are more important [150].

3. Cellular and Molecular Targets of Curcumin

The cellular and molecular targets of curcumin have been summarized in Figure 2 [162,163,164,165]. It has been categorized based on the role of targets and the effect of curcumin on those. Curcumin clearly diminishes mRNA production of pro-inflammatory mediators, including cytokines and relative enzymes such as cyclooxygenase (COX)-2, and inducible nitric oxide synthase (iNOS) [166,167,168]. Apparently, this is due to it inhibitor effect of transcription factors such as activator protein (AP)-1 and nuclear factor (NF)-κB-mediated gene [143,169]; however, the direct molecular targets at low doses are not entirely clear.
There are some controversies regarding the effect of curcumin on these factors. To our knowledge, some of these effects are probably dose-dependent. For example, curcumin increase and decrease the apoptosis proteins and markers depending on its dose or concentration. In fact, in order to clarify the exact mechanism, further investigations using a wide range of doses are needed to explore the dose-dependent effects of curcumin. This hypothesis has been supported by the results of previous studies for instance, curcumin at low doses (<100 nM) prevents AP-1 and NF-κB-mediated transcription, which might rely on the inhibition effect on histone acetylase (HAT) or activation of histone deacetylase (HDAC) activity [170] while at high doses (>3 μM) curcumin can act as an alkylating agent in a study on colon cancer [171]. In fact, some of these curcumin effects at high doses in vitro are obviously toxic and beyond its usage in cancer therapy.

4. Curcumin Metabolism and Degradation

Following oral administration of curcumin, it is metabolized extensively. It has been reported that the phase I reduction reaction and phase II conjugation reaction were the major metabolic pathways of curcumin in animals, and dihydrocurcumin, tetrahydrocurcumin, octahydrocurcumin, and hexahydrocurcumin were the main metabolites in phase I [172].
There is some evidence indicating that metabolites are the major contributors of the pharmacological activity of curcumin. For instance, dihydrocurcumin diminished lipid accumulation, oxidative stress, and insulin resistance in oleic acid-induced L02 and HepG2 cells [173].
Tetrahydrocurcumin, another metabolite of curcumin, has been reported to show a wide range of therapeutic properties [174,175]. The properties of this metabolite are comparable to those of curcumin. However, studies revealed that Tetrahydrocurcumin is more potent than curcumin as an anti-inflammatory, antioxidant, neuroprotective agent and anti-cancer [174]. Several lines of evidence demonstrated that this curcumin metabolite has higher antioxidant activity and upregulates the antioxidant enzymes in different pathological conditions including atherosclerosis [176], diabetes [177], hyperlipidemia [178], and neurotoxicity [179]. Having additional hydrogen molecules, Tetrahydrocurcumin is more hydrophilic than curcumin [180] and pharmacokinetic assessments reveal that it is more stable than curcumin in 0.1 M phosphate buffers at neutral pH and plasma [181]. Moreover, Its half-life in in vitro studies is significantly longer than that of curcumin [182]. Together, these properties suggest that THC may be more potent and efficacious against human diseases than curcumin, owing to its distinct chemical properties and stability. Recent studies showed that curcumin quickly degrades in aqueous buffer and some degradation compounds are produced [183] comprising alkaline hydrolysis products (ferulic acid, vanillin, ferulaldehyde, and feruloyl methane) which are formed through ahydroxyl ion (OH-)-mediated hydrolysis reaction [184] and autoxidation products (bicyclopentadione), are formed through a radical-mediated process [185]. Evidence suggests that curcumin mostly degrades via oxidation pathway rather than through hydrolysis [183].
Although some of these degradation products are biologically active, they are noticeably less-active in comparison to curcumin [183], supporting the idea that chemical degradation of curcumin has a limited contribution to its biological and pharmacological activities.

5. Curcumin Toxicity

In addition to the wide variety therapeutic effects of curcumin discussed above, there are some reports about its toxic potential that could be discussed from 2 aspects; first, even in the therapeutic doses its reactivity against a number of enzymes which act through different cellular mechanisms, provides adverse effects. For instance, its glutathione S-transferase (GST) inhibition can lead to impaired detoxification and potential toxic drug−drug contraindications [186] or the human ether-a-go-go-related gene (hERG) channel inhibition which may lead to cardiotoxicity [187,188]. Second, high doses of curcumin have been reported to be toxic for cells inducing apoptosis [189]. In some studies of therapeutic utility, it has been shown as cytotoxic against a number of important cancer cell lines as mentioned before or even cytotoxic against normal human lymphocytes [190] and noncancerous cell lines [191].
Despite these effects, curcumin is known to be safe by Food and Drug Administration (FDA). Several studies including preclinical and clinical evaluated the safety of this compound [192,193,194]. The maximum proposed dose of curcumin varies, ranging from daily consumption of 3 mg/kg to 10 g [195]. In a clinical study, no serious adverse effect was observed in any of the healthy subjects who used a daily dose of 12 g [196].
According to the literature, curcumin didn’t show mutagenic and genotoxic effects and is safe in pregnant animals. However, more studies in humans are needed. In addition, animal studies didn’t report reproductive toxicity of curcumin following oral administration. Curcumin was safe even at the high doses: in a phase I human trial 25 subjects used up to 8000 mg of curcumin per day for 3 months without showing toxicity. Five other human trials using 1125–2500 mg of curcumin per day have also found it to be safe [197] as well as in another study in which it was used 6 g/day orally for 4–7 weeks.
In addition, a good safety profile has been reported for curcumin in patients with cardiovascular risk factors or pre-malignant lesions of internal organs who took a dose of curcumin ranging from 500 to 8000 mg/day for a 3 months period [192,198]. This safety has been observed in patients with different cancers, including colorectal (taking, ranging from 36 to 180 mg/day for up to 4 months), breast (taking up to 6000 mg/day) and pancreatic cancer (taking 8000 mg/day of curcumin for 2 months) [86,199,200].
Having in mind that the majority of studies reporting curcumin safety has been performed for short periods of time so far and, there is not complete evidence regarding the consequences of chronic administration of this compound. Therefore, trials and more studies are needed specially on novel formulations and in long term use to find out all aspects of toxicity of curcumin.
However, only limited adverse effects such as gastrointestinal upsets have been stated in human. In addition, abdominal pain has also been reported followed by curcumin consumption at a dose of 8000 mg/day in patients with pancreatic cancer [201]. These gastrointestinal side effects might be related to the curcumin-induced COX inhibition and the subsequent inhibition of prostaglandin (PG) synthesis and/or its influence on the composition of the gut microbiome, which needs further study to be established. In addition to gastrointestinal side effects, a mild headache or nausea have been reported in few patients affected by primary sclerosing cholangitis taking curcumin in dose of up to 1400 mg/day [202].
Moreover, recent reports of liver diseases related to curcumin attracted the medical community’s attention to its possible hepatotoxicity [203]. Whether this effect belongs to curcumin molecule or other possible contamination has to be elucidated.
Evaluating the other routes of administration, in a short-term intravenous dosing of liposomal formulation, curcumin has been shown to be safe up to a dose of 120 mg/m2 on healthy subjects in a clinical trial, while in a dose escalation study in patients with metastatic cancer a dose of 300 mg/m2 over 6 h reported to be the maximum tolerated dosage [204,205]. Besides, one case of hemolysis and one death associated with intravenous curcumin preparations were reported, indicating that regarding the safety of intravenous administration of curcumin, further studies and data are required [205,206,207].

6. Pharmacokinetic Deficiency of Curcumin and Current Attempts

As described previously, the major constituent of extracts of C. longa are called curcuminoids, which includes curcumin demethoxycurcumin, and bisdemethoxycurcumin, along with numerous and less abundant secondary metabolites [208]. According to our literature review, many in vitro studies have used synthetic curcumin, while most animal studies and clinical trials used the curcuminoids mixture. In fact, these compounds are completely different in pharmacokinetic parameters.
Gastrointestinal absorption and bioavailability are two most important and critical characteristics in the pharmacokinetics of any compound. From the preclinical and clinical studies, we have found that curcumin is poorly absorbed following oral administration. It has been reported that oral bioavailability of curcumin was only 1% and the highest amount of plasma concentration was 0.051 μg/mL from 12 g curcumin in human, 1.35 μg/mL from 2 g/kg in rat, and 0.22 μg/mL from 1 g/kg in mouse [196]. Pharmacokinetic assessments showed that there were just negligible quantities detected in liver and kidney (<20 μg/tissue) following curcumin oral administration. These studies indicated the nano-molar plasmatic concentration of this compound, with limited biological effects [209,210,211,212,213].
Several studies, including clinical trials, have been performed using curcumin in a wide variety of oral formulations. From low oral doses to high amount of 12 g/day have been given and tested [196]. Data shown in Table 8 summarize some of formulations and their pharmacokinetic properties. In order to increase curcumin bioavailability, different pharmaceutical strategies, including combination with adjuvant substances such as piperine, encapsulation, formulation in nanoparticles, liposome, micelles, nanomicellizing solid dispersion etc., have been proposed, shedding new light to overcome this pivotal limitation [214,215,216,217].
As presented below, there is a remarkable difference between the amount of plasma level of curcumin in different products. Interestingly, new formulations including nanoparticles, liposomes, and micelles exert a higher amount of C-max, the maximum (or peak) serum concentration that a drug achieves, compared to the other formulations (Table 8).
Based on the Nutraceutical Bioavailability Classification Scheme (NuBACS), curcumin exerts poor bio-accessibility, due to its low solubility in water and low stability [218].
It is well documented that water-solubility has a crucial role in oral absorption of compounds. Pharmacokinetics assessments revealed that curcumin is a poorly water-soluble agent (about 11 ng/mL) [219] and considerable efforts have been made to improve this characteristic.
Another important point that has to be taken into account in pharmacokinetic of curcumin, is its susceptibility to degradation, which has been shown to be pH-dependent. Indeed, under the alkaline conditions (pH > 7), curcumin degrades to Trans-6-(40-hydroxy-30-methoxyphenyl)-2, 4-dioxo-5-hexanal, ferulic acid, feruloylmethane, and vanillin within 30 min. while, in acidic medium, curcumin degradation is much slower, with less than 20% of total curcumin at 60 min [220,221].
Following oral administration, the major portion of curcumin is excreted through the feces while a small portion, which is absorbed within the intestine, rapidly metabolized in the livers and plasma [222]. Curcumin is extensively converted to its water-soluble metabolites (glucuronides and sulfates) and excreted through urine [222,223].
Curcumin also undergoes extensive hepatic first-pass metabolism through glucuronidation and sulfation, with the metabolites that exerts a remarkably lower biological function in comparison with parent curcumin [214]. Taking advantage of extensive research on curcumin metabolism, a curcumin-converting enzyme named “Nicotinamide adenine dinucleotide phosphate (NADPH)-dependent curcumin/dihydrocurcumin reductase” has been known which has purified from E. coli, clarifying the role of human intestinal microorganisms in the metabolism of curcumin in vivo [224].
The table provided below shows the studies on curcumin formulations, routes of administration, dose Plasma/tissue level (Cmax) and the time to maximum concentration (Tmax, min) in animals and humans. There is a significant variation in the doses used from the formulations included in this review as well as their Cmax and Tmax were really different.
Having two phenolic hydroxyls and one enolic hydroxyl group, which can form hydrogen bonds with phospholipids polar groups, in one formulation named Meriva, curcumin has complexed with phosphatidylcholine, and this structure protects it from degradation and augmenting the cellular uptake across lipophilic cell membranes by facilitated diffusion mechanism [225,226]. Taking advantage of this formulation, curcumin absorption and bioavailability were significantly increased compared to unformulated curcumin in a randomized, double-blind, study in human [225]. The Tmax has been reported 15 to 30 min, which showed a considerable increase in speed of absorption compared to the other formulations.
In another compound named LongVida®, curcumin is included in a pharmaceutical form of solid lipid particle (SLP)-based formulation. This complex keeps the curcumin from rapid degradation and excretion so improving the systemic curcumin plasma concentration and half-life and eventually improved bioavailability compared to unformulated curcumin. In a clinical trial, oral administration of 650 mg in human exerts 22.43 ng/mL plasma level, which was shown to be higher than curcuma extract and unformulated curcumin [227].
There are also some micronized formulations for curcumin, which were reported to have 9-fold increased bioavailable than unformulated. The micronized curcumin has a smaller diameter of drug particles, which increases the surface area to drug ratio and finally augmentation in the dissolution rate that directly increase the bioavailability of curcumin [228]. Interestingly in a human study, administration of 500 mg of a micronized formulation exerts a considerable 0.60 μg/mL plasma level, which was a promising results for this compound [229].
One of the early and basic strategies for increasing the bioavailability of curcumin was its combination with other compounds that could change its characteristics; among them piperine was known to improve the oral absorption of curcumin in humans. It is a P-glycoprotein inhibitor and enhances the curcumin absorption in the intestine by reducing the efflux phenomenon [230].
P-glycoprotein (P-gp) is a drug transporter that effluxes drugs and other foreign substances out of cells from gastrointestinal tract, brain, liver, and kidney and exerts an important role in drug pharmacokinetics and pharmacoresistance. Piperine has been reported to enhance the bioavailability of curcumin, as a substrate of P-gp by at least 2000% [231].
Piperine also inhibits uridine diphosphate-glucuronosyltransferase (UGT) thus increasing the curcumin availability in the systemic circulation. As shown in Table 8, adding piperine 20 mg/kg to curcumin 2 g/kg in rats exerts a C-max of 1.55 µg/mL in animals following oral administration [231].
In another attempt, micellization technique was employed to improve the solubility of curcumin in a compound named NovaSol®. In this formulation, curcumin incorporated in a nonionic surfactant Tween 80, caused the formation of liquid micelles, which increased dissolution and improved its absorption. This compound has been studied in a single-blind crossover study in healthy subjects and interestingly they showed a 185 folds higher than that of unformulated curcumin [232].
As previously described in this review, nanoparticle-based formulations have been extensively tried to improve the bioavailability and reduction of adverse drug reactions. In this regard, a colloidal nanoparticle dispersion of curcumin has been produced and its pharmacokinetic characteristic has been studied in humans. This formulation which is called Theracurmin™ has been showed to have relative bioavailability almost 16 time as compared to unformulated curcumin. In fact, the colloidal nanoparticle dispersion improves the solubility of curcumin and its oral bioavailability in healthy volunteers. In one study on this formulation, authors stated that two factors including improved solubility by adding a compound named “gum ghatti” and reduced particle size improved the clinical bioavailability of Theracurmin™ [233].
Table 8. Pharmacokinetic characteristics of some curcumin formulations in animal and clinical studies.
Table 8. Pharmacokinetic characteristics of some curcumin formulations in animal and clinical studies.
ProductSpeciesRoute of AdministrationDosePlasma/Tissue Level (Cmax)Time to Maximum Concentration (Tmax) minRef.
CurcuminoidsRatOral500 mg/kg0.06 µg/mL41.7[234]
CurcuminRatOral200 mg/kg1.2 µg/mLno[235]
Curcumin & Curcumin phospholipid complex (Meriva)RatOral340 mg/kg6.5 nM & 33.4 nM30 & 15[236]
CurcuminoidsHumanOral450–3600 mg10 nM/g tissueNo data[237]
CurcuminHumanOral3600 mg12.7 nmol/g tissueNo data[238]
CurcuminoidsRatOral100 mg/kgtrace60[239]
CurcuminRatOral400 mgtraceNo data[209]
CurcuminMouseIntraperitoneal100 mg/kgtraceNo data[240]
CurcuminHumanOral3600 mg10 nMNo data[85]
CurcuminHumanOral1200 mg51 ng/mLNo data[196]
CurcuminIn vitroExposure5–75 µg/mL3% in tissueNo data[241]
CurcuminRatOral10, 80, 400 mg65–66%No data[242]
Curcuma extractHumanOral440 and 2200 mg/day175 to 310 μg/LNo data[199]
Phospholipid formulationHumanOral200−300 mg50 ng/mL240[225]
Solid lipid curcumin particleHumanOral650 mg22.43 ng/mL160[227]
Curcumin-impregnated soluble dietary fiber dispersionsHumanOral600 mg0.37 μg/g tissue60[243]
Micronized formulationHumanOral500 mg0.60 μg/mLNo data[229]
Micronized formulationHumanOral500 mg50.6 nM460[232]
Liquid micelles
formulation
HumanOral500 mg3701 nM66[232]
Curcumin/piperine co-administrationRatOralCurcumin 2 g/kg
piperine 20 mg/kg
1.55 µg/mL120[231]
Lipophilic matrixHumanOral376 mg18 ng/mL60[244]
γ-cyclodextrin complexHumanOral376 mg87 ng/mL60[244]
Colloidal nanoparticleHumanOral30 mg29.5 ng/mL60[233]
CurcuminRatIntravenous40 mg/kgNo dataNo data[245]
BCM-95® CG
A patent formulation
HumanOral2000 mg456.88 ng/g tissue206[246]

7. Conclusions and Future Prospects

Similar to any drug, curcumin still requires extensive preclinical and clinical assessments to clarify its pharmacokinetics, dosing, and potential toxicity, the issues that are of essential concern of drug companies.
According to the therapeutic effects that were summarized in this review, curcumin plays different roles through apposite mechanisms such as cytotoxic and cytoprotective role in different disease. For understanding this controversy and diversity of effects, it is crucial to pay attention to the doses and concentrations that have been used in different situations. It seems that curcumin’s behavior changes in various doses and conditions, and this issue should be highly considered in clinical trials, especially in new formulations with higher bioavailability characteristics. Confirming this idea about the dose-dependent effect of curcumin, recently, we showed that low concentration of curcumin (5 µM) exerts cytoprotective effect in SH-SY5Y neuroblastoma cells while higher concentration (15 µM) induced apoptosis and cell death [126].
Despite having a wide range of pharmacological activity, there is still a key question on curcumin to be answered, that is, why it remains as a supplement and couldn’t play a higher-order role as a drug in treatment of diseases?
In response to this essential concern, we believe that pharmacokinetic deficiency is the main obstacle and studies to address this issue should be more considered before doing further preclinical and clinical studies. Confirming this idea, researchers have discussed that given its unstable, reactive, and poor bioavailability nature, further clinical trials on curcumin are unwarranted [34,191]. Therefore, this decade has witness an escalation in curcumin formulations addressing its pharmacokinetics (Figure 3).
A wide variety of compounds have been formulated and tried to improve the pharmacokinetic problems, especially absorption, which is the first step in oral administration of drugs.
Although the absorption and bioavailability are the crucial issues, curcumin distribution in body tissues should be take into account for its biological activity. In fact, there are just limited number of investigations have addressed this problem so far. In this regard, researchers reported that oral administration of curcumin (400 mg) in rats exerts only traces of unchanged drug in other organs such as liver and kidney [209].
Novel delivery strategies including nanoparticles, liposome, phospholipid complex, micro-emulsions, or polymer micelles suggest significant promise and are worthy of further investigation to enhance the oral bioavailability of curcumin. Here we reviewed some new formulations on the market, and some of their pharmacokinetic characteristics have been compared in Table 8. Most clinical studies on these new formulations were conducted randomized and blinded (neither researchers nor volunteers were aware of the treatment). However, the most important limitation of these investigations was the low number of subjects were recruited in the studies. In addition, there are remarkable differences in the sampling time following oral administration of curcumin that may affect the pharmacokinetic parameters including Tmax and Cmax. In fact, a wide-ranging of variability in this research, including route of administration, study design, sample size, and even the type of reports, makes it very difficult to directly compare and conclude which formulation is the best in pharmacokinetic characteristics overall. However, an overall view to these results shows that novel formulations involving nanoparticles and micronized products showed improved solubility and a relatively more amount of blood concentration following oral administration compared to the others.
With regards to the fact that most of the investigations reported in this review have been designed and performed in vitro, extensive harmonized large clinical study including different formulations in humans with the same design will make great sense and is needed to compare the bioavailability and efficacy of these curcumin formulations.
Considering the pharmacokinetic properties of curcumin metabolites and their beneficial effects, such as regulation of oxidative stress, prevention of neurodegenerative disorders and their potential mechanism to prevent or treat human diseases, developing new formulations and performing clinical trials using these compounds is also suggested.
Finally, bypassing the limitations of curcumin mentioned in this review, such as low oral bioavailability, distribution, and metabolism, will pave the way to do more applicative research and clinical studies on curcumin as a drug in the near future.

Author Contributions

K.H.: study design, performing bibliographic search, interpreting the relevant literature and writing the original draft; L.B.: Revising the manuscript critically for important intellectual content; J.D.: Revising the manuscript critically for important intellectual content. A.M.: Revising the manuscript critically for important intellectual content; M.C.: Revising the manuscript critically for important intellectual content; M.F.: Conceiving the idea, design of the study, support the research with his funding grants and supervision of project. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AH RAryl hydrocarbon receptor
ADAlzheimer’s disease
AP-1Activator protein 1
BaxBcl-2-associated X protein
BDNFBrain-derived neurotrophic factor
CDPKCalcium-dependent protein kinases
CRDBCurcumin Resource Database
CREBcAMP response element-binding protein
COX-2Cyclooxygenase-2
CTClinical Trial
CXCR 4C-X-C Motif Chemokine Receptor 4
EGFEpidermal growth factor
ER-alfaEstrogen receptor alfa
ERKExtracellular signal-regulated kinases
FADDFas Associated via death domain
FAKFocal adhesion kinase
FASFas cell surface death receptor
FGFFibroblast growth factors
GSTGlutathione-S-transferase
HATHistone acetylase
H2 RHistamine H2 receptor
HDACHistone deacetylase
HGFHepatocyte growth factor
HMG-CoA-R3-hydroxy-3-methyl-glutaryl-CoA reductase
HSP-70Heat shock protein 70
IBDIntestinal inflammatory diseases
ICAMsIntercellular cell adhesion molecules
ILInterleukin
iNOSInducible nitric oxide synthase
JAKJanus kinase
JNKc-Jun N-terminal kinases
LDL RLow-Density Lipoprotein Receptor
MCP-1Monocyte chemoattractant protein-1
MIP-1αMacrophage inflammatory proteins
MMPMatrix metallopeptidases
MRPMultidrug resistance-associated protein
NADPHNicotinamide adenine dinucleotide phosphate
NFκBNuclear Factor kappa-light-chain-enhancer of ctivated B cells
NGFNerve growth factor
Nrf2Nuclear factor erythroid 2–related factor 2
NuBACSNutraceutical Bioavailability Classification cheme
P38-MAPKP38 mitogen-activated protein kinases
PDParkinson’s disease
PDGFPlatelet-derived growth factor
P-gpP-glycoprotein
PhKPhosphorylase kinase
PKAProtein kinase A
PLA2Phospholipase A2
PPAR-gammaPeroxisome proliferator-activated receptor amma
ROSReactive oxygen species
RNSReactive nitrogen species
SLPSolid lipid particle
STATSignal transducer and activator of transcription
SyKSpleen tyrosine kinase
TFTissue factor
TGF-αTransforming growth factor alpha
TGF-βTransforming growth factor beta
TLRToll-like receptors
TNF-αTumor necrosis factor alpha
UGTUridine diphosphate-glucuronosyltransferase
VCAMVascular cell adhesion molecule
XOXanthine oxidase
5-LOX5-Lipoxygenase

References

  1. Aggarwal, B.B.; Sundaram, C.; Malani, N.; Ichikawa, H. Curcumin: The Indian solid gold. Adv. Exp. Med. Biol. 2007, 595, 1–75. [Google Scholar] [CrossRef] [PubMed]
  2. Bigford, G.E.; Del Rossi, G. Supplemental substances derived from foods as adjunctive therapeutic agents for treatment of neurodegenerative diseases and disorders. Adv. Nutr. 2014, 5, 394–403. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Lampe, V.; Milobedzka, J. Studien über Curcumin. Berichte Dtsch. Chem. Ges. 1913, 46, 2235–2240. [Google Scholar] [CrossRef]
  4. 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]
  5. Daybe, F.V. Uber Curcumin. Den Farbstoff der Curcumawurzzel Ber 1870, 3, 609. [Google Scholar]
  6. Govindarajan, V.S. Turmeric—Chemistry, technology, and quality. Crit. Rev. Food Sci. Nutr. 1980, 12, 199–301. [Google Scholar] [CrossRef]
  7. Kumar, A.; Chetia, H.; Sharma, S.; Kabiraj, D.; Talukdar, N.C.; Bora, U. Curcumin Resource Database. Database J. Biol. Databases Curation 2015, 2015, bav070. [Google Scholar] [CrossRef] [Green Version]
  8. Mishra, B.; Priyadarsini, K.I.; Bhide, M.K.; Kadam, R.M.; Mohan, H. Reactions of superoxide radicals with curcumin: Probable mechanisms by optical spectroscopy and EPR. Free Radic. Res. 2004, 38, 355–362. [Google Scholar] [CrossRef]
  9. Trujillo, J.; Chirino, Y.I.; Molina-Jijón, E.; Andérica-Romero, A.C.; Tapia, E.; Pedraza-Chaverrí, J. Renoprotective effect of the antioxidant curcumin: Recent findings. Redox Biol. 2013, 1, 448–456. [Google Scholar] [CrossRef] [Green Version]
  10. Liu, Z.-J.; Liu, H.-Q.; Xiao, C.; Fan, H.-Z.; Huang, Q.; Liu, Y.-H.; Wang, Y. Curcumin protects neurons against oxygen-glucose deprivation/reoxygenation-induced injury through activation of peroxisome proliferator-activated receptor-γ function. J. Neurosci. Res. 2014, 92, 1549–1559. [Google Scholar] [CrossRef]
  11. Reddy, P.H.; Manczak, M.; Yin, X.; Grady, M.C.; Mitchell, A.; Kandimalla, R.; Kuruva, C.S. Protective effects of a natural product, curcumin, against amyloid β induced mitochondrial and synaptic toxicities in Alzheimer’s disease. J. Investig. Med. 2016, 64, 1220–1234. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Motaghinejad, M.; Motevalian, M.; Fatima, S.; Hashemi, H.; Gholami, M. Curcumin confers neuroprotection against alcohol-induced hippocampal neurodegeneration via CREB-BDNF pathway in rats. Biomed. Pharmacother. Biomed. Pharmacother. 2017, 87, 721–740. [Google Scholar] [CrossRef] [PubMed]
  13. Wang, Q.; Sun, A.Y.; Simonyi, A.; Jensen, M.D.; Shelat, P.B.; Rottinghaus, G.E.; MacDonald, R.S.; Miller, D.K.; Lubahn, D.E.; Weisman, G.A.; et al. Neuroprotective mechanisms of curcumin against cerebral ischemia-induced neuronal apoptosis and behavioral deficits. J. Neurosci. Res. 2005, 82, 138–148. [Google Scholar] [CrossRef] [PubMed]
  14. Chen, J.; Tang, X.Q.; Zhi, J.L.; Cui, Y.; Yu, H.M.; Tang, E.H.; Sun, S.N.; Feng, J.Q.; Chen, P.X. Curcumin protects PC12 cells against 1-methyl-4-phenylpyridinium ion-induced apoptosis by bcl-2-mitochondria-ROS-iNOS pathway. Apoptosis Int. J. Program. Cell Death 2006, 11, 943–953. [Google Scholar] [CrossRef]
  15. Zorofchian Moghadamtousi, S.; Abdul Kadir, H.; Hassandarvish, P.; Tajik, H.; Abubakar, S.; Zandi, K. A Review on Antibacterial, Antiviral, and Antifungal Activity of Curcumin. Available online: https://www.hindawi.com/journals/bmri/2014/186864/ (accessed on 13 April 2020).
  16. Mathew, D.; Hsu, W.-L. Antiviral potential of curcumin. J. Funct. Foods 2018, 40, 692–699. [Google Scholar] [CrossRef]
  17. Shlar, I.; Droby, S.; Choudhary, R.; Rodov, V. The mode of antimicrobial action of curcumin depends on the delivery system: Monolithic nanoparticles vs. supramolecular inclusion complex. RSC Adv. 2017, 7, 42559–42569. [Google Scholar] [CrossRef] [Green Version]
  18. Kiuchi, F.; Goto, Y.; Sugimoto, N.; Akao, N.; Kondo, K.; Tsuda, Y. Nematocidal activity of turmeric: Synergistic action of curcuminoids. Chem. Pharm. Bull. 1993, 41, 1640–1643. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Sui, Z.; Salto, R.; Li, J.; Craik, C.; Ortiz de Montellano, P.R. Inhibition of the HIV-1 and HIV-2 proteases by curcumin and curcumin boron complexes. Bioorg. Med. Chem. 1993, 1, 415–422. [Google Scholar] [CrossRef]
  20. Tantaoui-Elaraki, A.; Beraoud, L. Inhibition of growth and aflatoxin production in Aspergillus parasiticus by essential oils of selected plant materials. J. Environ. Pathol. Toxicol. Oncol. Off. Organ Int. Soc. Environ. Toxicol. Cancer 1994, 13, 67–72. [Google Scholar]
  21. Jordan, W.C.; Drew, C.R. Curcumin—A natural herb with anti-HIV activity. J. Natl. Med. Assoc. 1996, 88, 333. [Google Scholar]
  22. Roth, G.N.; Chandra, A.; Nair, M.G. Novel bioactivities of Curcuma longa constituents. J. Nat. Prod. 1998, 61, 542–545. [Google Scholar] [CrossRef]
  23. Negi, P.S.; Jayaprakasha, G.K.; Jagan Mohan Rao, L.; Sakariah, K.K. Antibacterial activity of turmeric oil: A byproduct from curcumin manufacture. J. Agric. Food Chem. 1999, 47, 4297–4300. [Google Scholar] [CrossRef] [PubMed]
  24. Koide, T.; Nose, M.; Ogihara, Y.; Yabu, Y.; Ohta, N. Leishmanicidal effect of curcumin in vitro. Biol. Pharm. Bull. 2002, 25, 131–133. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Kim, M.-K.; Choi, G.-J.; Lee, H.-S. Fungicidal property of Curcuma longa L. rhizome-derived curcumin against phytopathogenic fungi in a greenhouse. J. Agric. Food Chem. 2003, 51, 1578–1581. [Google Scholar] [CrossRef] [PubMed]
  26. Reddy, R.C.; Vatsala, P.G.; Keshamouni, V.G.; Padmanaban, G.; Rangarajan, P.N. Curcumin for malaria therapy. Biochem. Biophys. Res. Commun. 2005, 326, 472–474. [Google Scholar] [CrossRef] [PubMed]
  27. Pérez-Arriaga, L.; Mendoza-Magaña, M.L.; Cortés-Zárate, R.; Corona-Rivera, A.; Bobadilla-Morales, L.; Troyo-Sanromán, R.; Ramírez-Herrera, M.A. Cytotoxic effect of curcumin on Giardia lamblia trophozoites. Acta Trop. 2006, 98, 152–161. [Google Scholar] [CrossRef]
  28. Di Mario, F.; Cavallaro, L.G.; Nouvenne, A.; Stefani, N.; Cavestro, G.M.; Iori, V.; Maino, M.; Comparato, G.; Fanigliulo, L.; Morana, E.; et al. A curcumin-based 1-week triple therapy for eradication of Helicobacter pylori infection: Something to learn from failure? Helicobacter 2007, 12, 238–243. [Google Scholar] [CrossRef]
  29. Cai, T.; Mazzoli, S.; Bechi, A.; Addonisio, P.; Mondaini, N.; Pagliai, R.C.; Bartoletti, R. Serenoa repens associated with Urtica dioica (ProstaMEV) and curcumin and quercitin (FlogMEV) extracts are able to improve the efficacy of prulifloxacin in bacterial prostatitis patients: Results from a prospective randomised study. Int. J. Antimicrob. Agents 2009, 33, 549–553. [Google Scholar] [CrossRef]
  30. Trigo Gutierrez, J.K.; Zanatta, G.C.; Ortega, A.L.M.; Balastegui, M.I.C.; Sanitá, P.V.; Pavarina, A.C.; Barbugli, P.A.; de Mima, E.G.O. Encapsulation of curcumin in polymeric nanoparticles for antimicrobial Photodynamic Therapy. PLoS ONE 2017, 12, e0187418. [Google Scholar] [CrossRef] [Green Version]
  31. Fakhrullina, G.; Khakimova, E.; Akhatova, F.; Lazzara, G.; Parisi, F.; Fakhrullin, R. Selective Antimicrobial Effects of Curcumin@Halloysite Nanoformulation: A Caenorhabditis elegans Study. ACS Appl. Mater. Interfaces 2019, 11, 23050–23064. [Google Scholar] [CrossRef]
  32. Jaiswal, S.; Mishra, P. Antimicrobial and antibiofilm activity of curcumin-silver nanoparticles with improved stability and selective toxicity to bacteria over mammalian cells. Med. Microbiol. Immunol. 2018, 207, 39–53. [Google Scholar] [CrossRef] [PubMed]
  33. Manchanda, G.; Sodhi, R.K.; Jain, U.K.; Chandra, R.; Madan, J. Iodinated curcumin bearing dermal cream augmented drug delivery, antimicrobial and antioxidant activities. J. Microencapsul. 2018, 35, 49–61. [Google Scholar] [CrossRef] [PubMed]
  34. Lopresti, A.L. The Problem of Curcumin and Its Bioavailability: Could Its Gastrointestinal Influence Contribute to Its Overall Health-Enhancing Effects? Adv. Nutr. 2018, 9, 41–50. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Farzaei, M.H.; Zobeiri, M.; Parvizi, F.; El-Senduny, F.F.; Marmouzi, I.; Coy-Barrera, E.; Naseri, R.; Nabavi, S.M.; Rahimi, R.; Abdollahi, M. Curcumin in Liver Diseases: A Systematic Review of the Cellular Mechanisms of Oxidative Stress and Clinical Perspective. Nutrients 2018, 10, 855. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Villegas, I.; Sánchez-Fidalgo, S.; de la Lastra, C.A. Chemopreventive effect of dietary curcumin on inflammation-induced colorectal carcinogenesis in mice. Mol. Nutr. Food Res. 2011, 55, 259–267. [Google Scholar] [CrossRef]
  37. Larmonier, C.B.; Uno, J.K.; Lee, K.-M.; Karrasch, T.; Laubitz, D.; Thurston, R.; Midura-Kiela, M.T.; Ghishan, F.K.; Sartor, R.B.; Jobin, C.; et al. Limited effects of dietary curcumin on Th-1 driven colitis in IL-10 deficient mice suggest an IL-10-dependent mechanism of protection. Am. J. Physiol. Gastrointest. Liver Physiol. 2008, 295, G1079–G1091. [Google Scholar] [CrossRef]
  38. Larmonier, C.B.; Midura-Kiela, M.T.; Ramalingam, R.; Laubitz, D.; Janikashvili, N.; Larmonier, N.; Ghishan, F.K.; Kiela, P.R. Modulation of neutrophil motility by curcuminImplications for inflammatory bowel disease. Inflamm. Bowel Dis. 2011, 17, 503–515. [Google Scholar] [CrossRef] [Green Version]
  39. Hanai, H.; Iida, T.; Takeuchi, K.; Watanabe, F.; Maruyama, Y.; Andoh, A.; Tsujikawa, T.; Fujiyama, Y.; Mitsuyama, K.; Sata, M.; et al. Curcumin maintenance therapy for ulcerative colitis: Randomized, multicenter, double-blind, placebo-controlled trial. Clin. Gastroenterol. Hepatol. Off. Clin. Pract. J. Am. Gastroenterol. Assoc. 2006, 4, 1502–1506. [Google Scholar] [CrossRef]
  40. Durgaprasad, S.; Pai, C.G.; Vasanthkumar, N.; Alvres, J.F.; Namitha, S. A pilot study of the antioxidant effect of curcumin in tropical pancreatitis. Indian J. Med. Res. 2005, 122, 315–318. [Google Scholar]
  41. Burge, K.; Gunasekaran, A.; Eckert, J.; Chaaban, H. Curcumin and Intestinal Inflammatory Diseases: Molecular Mechanisms of Protection. Int. J. Mol. Sci. 2019, 20, 1912. [Google Scholar] [CrossRef] [Green Version]
  42. Shen, L.; Liu, L.; Ji, H.-F. Regulative effects of curcumin spice administration on gut microbiota and its pharmacological implications. Food Nutr. Res. 2017, 61, 1361780. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Feng, W.; Wang, H.; Zhang, P.; Gao, C.; Tao, J.; Ge, Z.; Zhu, D.; Bi, Y. Modulation of gut microbiota contributes to curcumin-mediated attenuation of hepatic steatosis in rats. Biochim. Biophys. Acta Gen. Subj. 2017, 1861, 1801–1812. [Google Scholar] [CrossRef] [PubMed]
  44. Maslowski, K.M.; Vieira, A.T.; Ng, A.; Kranich, J.; Sierro, F.; Yu, D.; Schilter, H.C.; Rolph, M.S.; Mackay, F.; Artis, D.; et al. Regulation of inflammatory responses by gut microbiota and chemoattractant receptor GPR43. Nature 2009, 461, 1282–1286. [Google Scholar] [CrossRef] [PubMed]
  45. De Filippo, C.; Cavalieri, D.; Di Paola, M.; Ramazzotti, M.; Poullet, J.B.; Massart, S.; Collini, S.; Pieraccini, G.; Lionetti, P. Impact of diet in shaping gut microbiota revealed by a comparative study in children from Europe and rural Africa. Proc. Natl. Acad. Sci. USA 2010, 107, 14691–14696. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Tang, Y.; Chen, A. Curcumin protects hepatic stellate cells against leptin-induced activation in vitro by accumulating intracellular lipids. Endocrinology 2010, 151, 4168–4177. [Google Scholar] [CrossRef] [Green Version]
  47. Vizzutti, F.; Provenzano, A.; Galastri, S.; Milani, S.; Delogu, W.; Novo, E.; Caligiuri, A.; Zamara, E.; Arena, U.; Laffi, G.; et al. Curcumin limits the fibrogenic evolution of experimental steatohepatitis. Lab. Investig. J. Tech. Methods Pathol. 2010, 90, 104–115. [Google Scholar] [CrossRef] [Green Version]
  48. Bruck, R.; Ashkenazi, M.; Weiss, S.; Goldiner, I.; Shapiro, H.; Aeed, H.; Genina, O.; Helpern, Z.; Pines, M. Prevention of liver cirrhosis in rats by curcumin. Liver Int. Off. J. Int. Assoc. Study Liver 2007, 27, 373–383. [Google Scholar] [CrossRef]
  49. Holt, P.R.; Katz, S.; Kirshoff, R. Curcumin therapy in inflammatory bowel disease: A pilot study. Dig. Dis. Sci. 2005, 50, 2191–2193. [Google Scholar] [CrossRef] [Green Version]
  50. Epstein, J.; Docena, G.; MacDonald, T.T.; Sanderson, I.R. Curcumin suppresses p38 mitogen-activated protein kinase activation, reduces IL-1beta and matrix metalloproteinase-3 and enhances IL-10 in the mucosa of children and adults with inflammatory bowel disease. Br. J. Nutr. 2010, 103, 824–832. [Google Scholar] [CrossRef] [Green Version]
  51. Rong, S.; Zhao, Y.; Bao, W.; Xiao, X.; Wang, D.; Nussler, A.K.; Yan, H.; Yao, P.; Liu, L. Curcumin prevents chronic alcohol-induced liver disease involving decreasing ROS generation and enhancing antioxidative capacity. Phytomed. Int. J. Phytother. Phytopharm. 2012, 19, 545–550. [Google Scholar] [CrossRef]
  52. Xiong, Z.E.; Dong, W.G.; Wang, B.Y.; Tong, Q.Y.; Li, Z.Y. Curcumin attenuates chronic ethanol-induced liver injury by inhibition of oxidative stress via mitogen-activated protein kinase/nuclear factor E2-related factor 2 pathway in mice. Pharmacogn. Mag. 2015, 11, 707–715. [Google Scholar] [CrossRef] [Green Version]
  53. Varatharajalu, R.; Garige, M.; Leckey, L.C.; Reyes-Gordillo, K.; Shah, R.; Lakshman, M.R. Protective Role of Dietary Curcumin in the Prevention of the Oxidative Stress Induced by Chronic Alcohol with respect to Hepatic Injury and Antiatherogenic Markers. Oxid. Med. Cell. Longev. 2016, 2016, 5017460. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Rahmani, S.; Asgary, S.; Askari, G.; Keshvari, M.; Hatamipour, M.; Feizi, A.; Sahebkar, A. Treatment of Non-alcoholic Fatty Liver Disease with Curcumin: A Randomized Placebo-controlled Trial. Phytother. Res. PTR 2016, 30, 1540–1548. [Google Scholar] [CrossRef] [PubMed]
  55. Murphy, E.A.; Davis, J.M.; McClellan, J.L.; Gordon, B.T.; Carmichael, M.D. Curcumin’s Effect on Intestinal Inflammation and Tumorigenesis in the ApcMin/+ Mouse. J. Interferon Cytokine Res. 2011, 31, 219–226. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Ukil, A.; Maity, S.; Karmakar, S.; Datta, N.; Vedasiromoni, J.R.; Das, P.K. Curcumin, the major component of food flavour turmeric, reduces mucosal injury in trinitrobenzene sulphonic acid-induced colitis. Br. J. Pharmacol. 2003, 139, 209–218. [Google Scholar] [CrossRef] [Green Version]
  57. Billerey-Larmonier, C.; Uno, J.K.; Larmonier, N.; Midura, A.J.; Timmermann, B.; Ghishan, F.K.; Kiela, P.R. Protective effects of dietary curcumin in mouse model of chemically induced colitis are strain dependent. Inflamm. Bowel Dis. 2008, 14, 780–793. [Google Scholar] [CrossRef] [Green Version]
  58. Wongcharoen, W.; Phrommintikul, A. The protective role of curcumin in cardiovascular diseases. Int. J. Cardiol. 2009, 133, 145–151. [Google Scholar] [CrossRef]
  59. Brouet, I.; Ohshima, H. Curcumin, an Anti-tumor Promoter and Anti-inflammatory Agent, Inhibits Induction of Nitric Oxide Synthase in Activated Macrophages. Biochem. Biophys. Res. Commun. 1995, 206, 533–540. [Google Scholar] [CrossRef]
  60. Farhangkhoee, H.; Khan, Z.A.; Chen, S.; Chakrabarti, S. Differential effects of curcumin on vasoactive factors in the diabetic rat heart. Nutr. Metab. 2006, 3, 27. [Google Scholar] [CrossRef] [Green Version]
  61. Frey, N.; Olson, E.N. Cardiac Hypertrophy: The Good, the Bad, and the Ugly. Annu. Rev. Physiol. 2003, 65, 45–79. [Google Scholar] [CrossRef]
  62. Backs, J.; Olson, E.N. Control of Cardiac Growth by Histone Acetylation/Deacetylation. Circ. Res. 2006, 98, 15–24. [Google Scholar] [CrossRef] [PubMed]
  63. Marcu, M.G.; Jung, Y.-J.; Lee, S.; Chung, E.-J.; Lee, M.-J.; Trepel, J.; Neckers, L. Curcumin is an Inhibitor of p300 Histone Acetylatransferase. Available online: https://www.ingentaconnect.com/content/ben/mc/2006/00000002/00000002/art00006 (accessed on 14 April 2020).
  64. Morimoto, T.; Sunagawa, Y.; Kawamura, T.; Takaya, T.; Wada, H.; Nagasawa, A.; Komeda, M.; Fujita, M.; Shimatsu, A.; Kita, T.; et al. The dietary compound curcumin inhibits p300 histone acetyltransferase activity and prevents heart failure in rats. J. Clin. Investig. 2008, 118, 868–878. [Google Scholar] [CrossRef] [PubMed]
  65. Jung, M.-A.; Lee, S.Y.; Han, S.H.; Hong, J.; Na, J.-R.; Lee, J.Y.; Kim, Y.; Kim, S. Hypocholesterolemic effects of Curcuma longa L. with Nelumbo nucifera leaf in an in vitro model and a high cholesterol diet-induced hypercholesterolemic mouse model. Anim. Cells Syst. 2015, 19, 133–143. [Google Scholar] [CrossRef]
  66. Soliman, G. Effect of Curcumin, Mixture of Curcumin and Piperine and Curcum (Turmeric) on Lipid Profile of Normal and Hyperlipidemic Rats. Egypt. J. Hosp. Med. 2005, 21, 145–161. [Google Scholar] [CrossRef]
  67. Chakraborty, M.; Bhattacharjee, A.; Kamath, J.V. Cardioprotective effect of curcumin and piperine combination against cyclophosphamide-induced cardiotoxicity. Indian J. Pharmacol. 2017, 49, 65–70. [Google Scholar] [CrossRef] [PubMed]
  68. Correa, F.; Buelna-Chontal, M.; Hernández-Reséndiz, S.; García-Niño, W.R.; Roldán, F.J.; Soto, V.; Silva-Palacios, A.; Amador, A.; Pedraza-Chaverrí, J.; Tapia, E.; et al. Curcumin maintains cardiac and mitochondrial function in chronic kidney disease. Free Radic. Biol. Med. 2013, 61, 119–129. [Google Scholar] [CrossRef] [PubMed]
  69. Swamy, A.V.; Gulliaya, S.; Thippeswamy, A.; Koti, B.C.; Manjula, D.V. Cardioprotective effect of curcumin against doxorubicin-induced myocardial toxicity in albino rats. Indian J. Pharmacol. 2012, 44, 73–77. [Google Scholar] [CrossRef] [Green Version]
  70. Tu, Y.; Sun, D.; Zeng, X.; Yao, N.; Huang, X.; Huang, D.; Chen, Y. Piperine potentiates the hypocholesterolemic effect of curcumin in rats fed on a high fat diet. Exp. Ther. Med. 2014, 8, 260–266. [Google Scholar] [CrossRef] [Green Version]
  71. Abdelsamia, E.M.; Khaleel, S.A.; Balah, A.; Abdel Baky, N.A. Curcumin augments the cardioprotective effect of metformin in an experimental model of type I diabetes mellitus; Impact of Nrf2/HO-1 and JAK/STAT pathways. Biomed. Pharmacother. 2019, 109, 2136–2144. [Google Scholar] [CrossRef]
  72. Nabofa, W.E.E.; Alashe, O.O.; Oyeyemi, O.T.; Attah, A.F.; Oyagbemi, A.A.; Omobowale, T.O.; Adedapo, A.A.; Alada, A.R.A. Cardioprotective Effects of Curcumin-Nisin Based Poly Lactic Acid Nanoparticle on Myocardial Infarction in Guinea Pigs. Sci. Rep. 2018, 8, 1–11. [Google Scholar] [CrossRef] [PubMed]
  73. Jafarinezhad, Z.; Rafati, A.; Ketabchi, F.; Noorafshan, A.; Karbalay-Doust, S. Cardioprotective effects of curcumin and carvacrol in doxorubicin-treated rats: Stereological study. Food Sci. Nutr. 2019, 7, 3581–3588. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Boarescu, P.-M.; Chirilă, I.; Bulboacă, A.E.; Bocșan, I.C.; Pop, R.M.; Gheban, D.; Bolboacă, S.D. Effects of Curcumin Nanoparticles in Isoproterenol-Induced Myocardial Infarction. Available online: https://www.hindawi.com/journals/omcl/2019/7847142/ (accessed on 20 December 2019).
  75. Xiao, J.; Sheng, X.; Zhang, X.; Guo, M.; Ji, X. Curcumin protects against myocardial infarction-induced cardiac fibrosis via SIRT1 activation in vivo and in vitro. Drug Des. Devel. Ther. 2016, 10, 1267–1277. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Wongcharoen, W.; Jai-aue, S.; Phrommintikul, A.; Nawarawong, W.; Woragidpoonpol, S.; Tepsuwan, T.; Sukonthasarn, A.; Apaijai, N.; Chattipakorn, N. Effects of Curcuminoids on Frequency of Acute Myocardial Infarction After Coronary Artery Bypass Grafting. Am. J. Cardiol. 2012, 110, 40–44. [Google Scholar] [CrossRef] [PubMed]
  77. Ramaswami, G.; Chai, H.; Yao, Q.; Lin, P.H.; Lumsden, A.B.; Chen, C. Curcumin blocks homocysteine-induced endothelial dysfunction in porcine coronary arteries. J. Vasc. Surg. 2004, 40, 1216–1222. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Jang, E.-M.; Choi, M.-S.; Jung, U.J.; Kim, M.-J.; Kim, H.-J.; Jeon, S.-M.; Shin, S.-K.; Seong, C.-N.; Lee, M.-K. Beneficial effects of curcumin on hyperlipidemia and insulin resistance in high-fat–fed hamsters. Metabolism 2008, 57, 1576–1583. [Google Scholar] [CrossRef] [PubMed]
  79. Shin, S.-K.; Ha, T.-Y.; McGregor, R.A.; Choi, M.-S. Long-term curcumin administration protects against atherosclerosis via hepatic regulation of lipoprotein cholesterol metabolism. Mol. Nutr. Food Res. 2011, 55, 1829–1840. [Google Scholar] [CrossRef]
  80. Ramírez-Boscá, A.; Soler, A.; Carrión, M.A.; Díaz-Alperi, J.; Bernd, A.; Quintanilla, C.; Quintanilla, A.E.; Miquel, J. An hydroalcoholic extract of Curcuma longa lowers the apo B/apo A ratio: Implications for atherogenesis prevention. Mech. Ageing Dev. 2000, 119, 41–47. [Google Scholar] [CrossRef]
  81. Block, K.I.; Gyllenhaal, C.; Lowe, L.; Amedei, A.; Amin, A.R.; Amin, A.; Aquilano, K.; Arbiser, J.; Arreola, A.; Arzumanyan, A.; et al. A Broad-Spectrum Integrative Design for Cancer Prevention and Therapy. Semin. Cancer Biol. 2015, 35, S276–S304. [Google Scholar] [CrossRef]
  82. Mantovani, A. Molecular pathways linking inflammation and cancer. Curr. Mol. Med. 2010, 10, 369–373. [Google Scholar] [CrossRef]
  83. Catanzaro, M.; Corsini, E.; Rosini, M.; Racchi, M.; Lanni, C. Immunomodulators Inspired by Nature: A Review on Curcumin and Echinacea. Molecules 2018, 23, 2778. [Google Scholar] [CrossRef] [Green Version]
  84. Mohamed, S.I.A.; Jantan, I.; Haque, M.A. Naturally occurring immunomodulators with antitumor activity: An insight on their mechanisms of action. Int. Immunopharmacol. 2017, 50, 291–304. [Google Scholar] [CrossRef] [PubMed]
  85. Sharma, R.A.; Euden, S.A.; Platton, S.L.; Cooke, D.N.; Shafayat, A.; Hewitt, H.R.; Marczylo, T.H.; Morgan, B.; Hemingway, D.; Plummer, S.M.; et al. Phase I clinical trial of oral curcumin: Biomarkers of systemic activity and compliance. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2004, 10, 6847–6854. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Dhillon, N.; Aggarwal, B.B.; Newman, R.A.; Wolff, R.A.; Kunnumakkara, A.B.; Abbruzzese, J.L.; Ng, C.S.; Badmaev, V.; Kurzrock, R. Phase II trial of curcumin in patients with advanced pancreatic cancer. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2008, 14, 4491–4499. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Ghalaut, V.S.; Sangwan, L.; Dahiya, K.; Ghalaut, P.S.; Dhankhar, R.; Saharan, R. Effect of imatinib therapy with and without turmeric powder on nitric oxide levels in chronic myeloid leukemia. J. Oncol. Pharm. Pract. Off. Publ. Int. Soc. Oncol. Pharm. Pract. 2012, 18, 186–190. [Google Scholar] [CrossRef] [PubMed]
  88. Capalbo, C.; Belardinilli, F.; Filetti, M.; Parisi, C.; Petroni, M.; Colicchia, V.; Tessitore, A.; Santoni, M.; Coppa, A.; Giannini, G.; et al. Effective treatment of a platinum-resistant cutaneous squamous cell carcinoma case by EGFR pathway inhibition. Mol. Clin. Oncol. 2018, 9, 30–34. [Google Scholar] [CrossRef]
  89. Killian, P.H.; Kronski, E.; Michalik, K.M.; Barbieri, O.; Astigiano, S.; Sommerhoff, C.P.; Pfeffer, U.; Nerlich, A.G.; Bachmeier, B.E. Curcumin inhibits prostate cancer metastasis in vivo by targeting the inflammatory cytokines CXCL1 and -2. Carcinogenesis 2012, 33, 2507–2519. [Google Scholar] [CrossRef] [Green Version]
  90. Radhakrishnan, V.M.; Kojs, P.; Young, G.; Ramalingam, R.; Jagadish, B.; Mash, E.A.; Martinez, J.D.; Ghishan, F.K.; Kiela, P.R. pTyr421 cortactin is overexpressed in colon cancer and is dephosphorylated by curcumin: Involvement of non-receptor type 1 protein tyrosine phosphatase (PTPN1). PLoS ONE 2014, 9, e85796. [Google Scholar] [CrossRef]
  91. Yoysungnoen, P.; Wirachwong, P.; Changtam, C.; Suksamrarn, A.; Patumraj, S. Anti-cancer and anti-angiogenic effects of curcumin and tetrahydrocurcumin on implanted hepatocellular carcinoma in nude mice. World J. Gastroenterol. 2008, 14, 2003–2009. [Google Scholar] [CrossRef]
  92. Chakraborty, G.; Jain, S.; Kale, S.; Raja, R.; Kumar, S.; Mishra, R.; Kundu, G.C. Curcumin suppresses breast tumor angiogenesis by abrogating osteopontin-induced VEGF expression. Mol. Med. Rep. 2008, 1, 641–646. [Google Scholar] [CrossRef] [Green Version]
  93. Saydmohammed, M.; Joseph, D.; Syed, V. Curcumin suppresses constitutive activation of STAT-3 by up-regulating protein inhibitor of activated STAT-3 (PIAS-3) in ovarian and endometrial cancer cells. J. Cell. Biochem. 2010, 110, 447–456. [Google Scholar] [CrossRef]
  94. Kim, H.J.; Park, S.Y.; Park, O.J.; Kim, Y.-M. Curcumin suppresses migration and proliferation of Hep3B hepatocarcinoma cells through inhibition of the Wnt signaling pathway. Mol. Med. Rep. 2013, 8, 282–286. [Google Scholar] [CrossRef]
  95. Qiao, Q.; Jiang, Y.; Li, G. Inhibition of the PI3K/AKT-NF-κB pathway with curcumin enhanced radiation-induced apoptosis in human Burkitt’s lymphoma. J. Pharmacol. Sci. 2013, 121, 247–256. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Lee, D.S.; Lee, M.K.; Kim, J.H. Curcumin induces cell cycle arrest and apoptosis in human osteosarcoma (HOS) cells. Anticancer Res. 2009, 29, 5039–5044. [Google Scholar] [PubMed]
  97. Liu, E.; Wu, J.; Cao, W.; Zhang, J.; Liu, W.; Jiang, X.; Zhang, X. Curcumin induces G2/M cell cycle arrest in a p53-dependent manner and upregulates ING4 expression in human glioma. J. Neurooncol. 2007, 85, 263–270. [Google Scholar] [CrossRef] [PubMed]
  98. Choudhuri, T.; Pal, S.; Agwarwal, M.L.; Das, T.; Sa, G. Curcumin induces apoptosis in human breast cancer cells through p53-dependent Bax induction. FEBS Lett. 2002, 512, 334–340. [Google Scholar] [CrossRef] [Green Version]
  99. Cao, A.-L.; Tang, Q.-F.; Zhou, W.-C.; Qiu, Y.-Y.; Hu, S.-J.; Yin, P.-H. Ras/ERK signaling pathway is involved in curcumin-induced cell cycle arrest and apoptosis in human gastric carcinoma AGS cells. J. Asian Nat. Prod. Res. 2015, 17, 56–63. [Google Scholar] [CrossRef]
  100. Ono, M.; Higuchi, T.; Takeshima, M.; Chen, C.; Nakano, S. Differential anti-tumor activities of curcumin against Ras- and Src-activated human adenocarcinoma cells. Biochem. Biophys. Res. Commun. 2013, 436, 186–191. [Google Scholar] [CrossRef]
  101. Lim, T.-G.; Lee, S.-Y.; Huang, Z.; Lim, D.Y.; Chen, H.; Jung, S.K.; Bode, A.M.; Lee, K.W.; Dong, Z. Curcumin suppresses proliferation of colon cancer cells by targeting CDK2. Cancer Prev. Res. Phila. PA 2014, 7, 466–474. [Google Scholar] [CrossRef] [Green Version]
  102. Patil, S.; Choudhary, B.; Rathore, A.; Roy, K.; Mahadik, K. Enhanced oral bioavailability and anticancer activity of novel curcumin loaded mixed micelles in human lung cancer cells. Phytomed. Int. J. Phytother. Phytopharm. 2015, 22, 1103–1111. [Google Scholar] [CrossRef]
  103. Vollono, L.; Falconi, M.; Gaziano, R.; Iacovelli, F.; Dika, E.; Terracciano, C.; Bianchi, L.; Campione, E. Potential of Curcumin in Skin Disorders. Nutrients 2019, 11, 2169. [Google Scholar] [CrossRef] [Green Version]
  104. Dai, M.; Zheng, X.; Xu, X.; Kong, X.; Li, X.; Guo, G.; Luo, F.; Zhao, X.; Wei, Y.Q.; Qian, Z. Chitosan-alginate sponge: Preparation and application in curcumin delivery for dermal wound healing in rat. J. Biomed. Biotechnol. 2009, 2009, 595126. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Mohanty, C.; Das, M.; Sahoo, S.K. Sustained wound healing activity of curcumin loaded oleic acid based polymeric bandage in a rat model. Mol. Pharm. 2012, 9, 2801–2811. [Google Scholar] [CrossRef] [PubMed]
  106. Hegge, A.B.; Andersen, T.; Melvik, J.E.; Bruzell, E.; Kristensen, S.; Tønnesen, H.H. Formulation and bacterial phototoxicity of curcumin loaded alginate foams for wound treatment applications: Studies on curcumin and curcuminoides XLII. J. Pharm. Sci. 2011, 100, 174–185. [Google Scholar] [CrossRef] [PubMed]
  107. Gopinath, D.; Ahmed, M.R.; Gomathi, K.; Chitra, K.; Sehgal, P.K.; Jayakumar, R. Dermal wound healing processes with curcumin incorporated collagen films. Biomaterials 2004, 25, 1911–1917. [Google Scholar] [CrossRef]
  108. Gong, C.; Wu, Q.; Wang, Y.; Zhang, D.; Luo, F.; Zhao, X.; Wei, Y.; Qian, Z. A biodegradable hydrogel system containing curcumin encapsulated in micelles for cutaneous wound healing. Biomaterials 2013, 34, 6377–6387. [Google Scholar] [CrossRef]
  109. Mohanty, C.; Sahoo, S.K. Curcumin and its topical formulations for wound healing applications. Drug Discov. Today 2017, 22, 1582–1592. [Google Scholar] [CrossRef]
  110. Rachmawati, H.; Edityaningrum, C.A.; Mauludin, R. Molecular inclusion complex of curcumin-β-cyclodextrin nanoparticle to enhance curcumin skin permeability from hydrophilic matrix gel. AAPS PharmSciTech 2013, 14, 1303–1312. [Google Scholar] [CrossRef] [Green Version]
  111. Kang, D.; Li, B.; Luo, L.; Jiang, W.; Lu, Q.; Rong, M.; Lai, R. Curcumin shows excellent therapeutic effect on psoriasis in mouse model. Biochimie 2016, 123, 73–80. [Google Scholar] [CrossRef]
  112. Bahraini, P.; Rajabi, M.; Mansouri, P.; Sarafian, G.; Chalangari, R.; Azizian, Z. Turmeric tonic as a treatment in scalp psoriasis: A randomized placebo-control clinical trial. J. Cosmet. Dermatol. 2018, 17, 461–466. [Google Scholar] [CrossRef]
  113. Mohammadi, Z.; Sharif Zak, M.; Majdi, H.; Mostafavi, E.; Barati, M.; Lotfimehr, H.; Ghaseminasab, K.; Pazoki-Toroudi, H.; Webster, T.J.; Akbarzadeh, A. The effect of chrysin-curcumin-loaded nanofibres on the wound-healing process in male rats. Artif. Cells Nanomed. Biotechnol. 2019, 47, 1642–1652. [Google Scholar] [CrossRef]
  114. Mazzarino, L.; Silva, L.F.C.; Curta, J.C.; Licínio, M.A.; Costa, A.; Pacheco, L.K.; Siqueira, J.M.; Montanari, J.; Romero, E.; Assreuy, J.; et al. Curcumin-loaded lipid and polymeric nanocapsules stabilized by nonionic surfactants: An in vitro and In vivo antitumor activity on B16-F10 melanoma and macrophage uptake comparative study. J. Biomed. Nanotechnol. 2011, 7, 406–414. [Google Scholar] [CrossRef] [PubMed]
  115. Yu, T.; Li, J.; Sun, H. C6 ceramide potentiates curcumin-induced cell death and apoptosis in melanoma cell lines in vitro. Cancer Chemother. Pharmacol. 2010, 66, 999–1003. [Google Scholar] [CrossRef] [PubMed]
  116. Cole, G.M.; Yang, F.; Lim, G.P.; Cummings, J.L.; Frautschy, D.L.M. A Rationale for Curcuminoids for the Prevention or Treatment of Alzheimers Disease. Available online: http://www.eurekaselect.com/91646/article (accessed on 21 April 2020).
  117. Ringman, J.M.; Frautschy, S.A.; Cole, G.M.; Masterman, D.L.; Cummings, J.L. A potential role of the curry spice curcumin in Alzheimer’s disease. Curr. Alzheimer Res. 2005, 2, 131–136. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Cole, G.M.; Teter, B.; Frautschy, S.A. Neuroprotective effects of curcumin. Adv. Exp. Med. Biol. 2007, 595, 197–212. [Google Scholar] [CrossRef] [Green Version]
  119. Brookmeyer, R.; Gray, S.; Kawas, C. Projections of Alzheimer’s disease in the United States and the public health impact of delaying disease onset. Am. J. Public Health 1998, 88, 1337–1342. [Google Scholar] [CrossRef] [Green Version]
  120. Lim, G.P.; Chu, T.; Yang, F.; Beech, W.; Frautschy, S.A.; Cole, G.M. The curry spice curcumin reduces oxidative damage and amyloid pathology in an Alzheimer transgenic mouse. J. Neurosci. Off. J. Soc. Neurosci. 2001, 21, 8370–8377. [Google Scholar] [CrossRef]
  121. Frautschy, S.A.; Hu, W.; Kim, P.; Miller, S.A.; Chu, T.; Harris-White, M.E.; Cole, G.M. Phenolic anti-inflammatory antioxidant reversal of Abeta-induced cognitive deficits and neuropathology. Neurobiol. Aging 2001, 22, 993–1005. [Google Scholar] [CrossRef]
  122. Santacruz, K.; Lewis, J.; Spires, T.; Paulson, J.; Kotilinek, L.; Ingelsson, M.; Guimaraes, A.; DeTure, M.; Ramsden, M.; McGowan, E.; et al. Tau suppression in a neurodegenerative mouse model improves memory function. Science 2005, 309, 476–481. [Google Scholar] [CrossRef] [Green Version]
  123. Liu, Z.; Yu, Y.; Li, X.; Ross, C.A.; Smith, W.W. Curcumin protects against A53T alpha-synuclein-induced toxicity in a PC12 inducible cell model for Parkinsonism. Pharmacol. Res. 2011, 63, 439–444. [Google Scholar] [CrossRef]
  124. Cui, Q.; Sun, S. Curcumin Antagonizes Rotenone-induced Injury of PC12 Cells by Antioxidant Activity. Acta Med. Univ. Sci. Technol. Huazhong 2010, 39, 37–41, 46. [Google Scholar]
  125. Buratta, S.; Chiaradia, E.; Tognoloni, A.; Gambelunghe, A.; Meschini, C.; Palmieri, L.; Muzi, G.; Urbanelli, L.; Emiliani, C.; Tancini, B. Effect of Curcumin on Protein Damage Induced by Rotenone in Dopaminergic PC12 Cells. Int. J. Mol. Sci. 2020, 21, 2761. [Google Scholar] [CrossRef]
  126. Buccarello, L.; Dragotto, J.; Iorio, F.; Hassanzadeh, K.; Corbo, M.; Feligioni, M. The pivotal role of SUMO-1-JNK-Tau axis in an in vitro model of oxidative stress counteracted by the protective effect of curcumin. Biochem. Pharmacol. 2020, 114066. [Google Scholar] [CrossRef] [PubMed]
  127. Jagatha, B.; Mythri, R.B.; Vali, S.; Bharath, M.M.S. Curcumin treatment alleviates the effects of glutathione depletion in vitro and in vivo: Therapeutic implications for Parkinson’s disease explained via in silico studies. Free Radic. Biol. Med. 2008, 44, 907–917. [Google Scholar] [CrossRef] [PubMed]
  128. Nuzzo, D.; Amato, A.; Picone, P.; Terzo, S.; Galizzi, G.; Bonina, F.P.; Mulè, F.; Di Carlo, M. A Natural Dietary Supplement with a Combination of Nutrients Prevents Neurodegeneration Induced by a High Fat Diet in Mice. Nutrients 2018, 10, 1130. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Ahmed, T.; Gilani, A.-H.; Hosseinmardi, N.; Semnanian, S.; Enam, S.A.; Fathollahi, Y. Curcuminoids rescue long-term potentiation impaired by amyloid peptide in rat hippocampal slices. Synap 2011, 65, 572–582. [Google Scholar] [CrossRef]
  130. Yang, F.; Lim, G.P.; Begum, A.N.; Ubeda, O.J.; Simmons, M.R.; Ambegaokar, S.S.; Chen, P.P.; Kayed, R.; Glabe, C.G.; Frautschy, S.A.; et al. Curcumin inhibits formation of amyloid beta oligomers and fibrils, binds plaques, and reduces amyloid in vivo. J. Biol. Chem. 2005, 280, 5892–5901. [Google Scholar] [CrossRef] [Green Version]
  131. Yanagisawa, D.; Taguchi, H.; Yamamoto, A.; Shirai, N.; Hirao, K.; Tooyama, I. Curcuminoid binds to amyloid-β1-42 oligomer and fibril. J. Alzheimers Dis. JAD 2011, 24 (Suppl. 2), 33–42. [Google Scholar] [CrossRef]
  132. Shimmyo, Y.; Kihara, T.; Akaike, A.; Niidome, T.; Sugimoto, H. Epigallocatechin-3-gallate and curcumin suppress amyloid beta-induced beta-site APP cleaving enzyme-1 upregulation. Neuroreport 2008, 19, 1329–1333. [Google Scholar] [CrossRef]
  133. Liu, H.-Y.; Fu, X.; Li, Y.-F.; Li, X.-L.; Ma, Z.-Y.; Zhang, Y.; Gao, Q.-C. miR-15b-5p targeting amyloid precursor protein is involved in the anti-amyloid eflect of curcumin in swAPP695-HEK293 cells. Neural Regen. Res. 2019, 14, 1603–1609. [Google Scholar] [CrossRef]
  134. Ahmed, T.; Enam, S.A.; Gilani, A.H. Curcuminoids enhance memory in an amyloid-infused rat model of Alzheimer’s disease. Neuroscience 2010, 169, 1296–1306. [Google Scholar] [CrossRef]
  135. Ishrat, T.; Hoda, M.N.; Khan, M.B.; Yousuf, S.; Ahmad, M.; Khan, M.M.; Ahmad, A.; Islam, F. Amelioration of cognitive deficits and neurodegeneration by curcumin in rat model of sporadic dementia of Alzheimer’s type (SDAT). Eur. Neuropsychopharmacol. J. Eur. Coll. Neuropsychopharmacol. 2009, 19, 636–647. [Google Scholar] [CrossRef] [PubMed]
  136. Ringman, J.M.; Frautschy, S.A.; Teng, E.; Begum, A.N.; Bardens, J.; Beigi, M.; Gylys, K.H.; Badmaev, V.; Heath, D.D.; Apostolova, L.G.; et al. Oral curcumin for Alzheimer’s disease: Tolerability and efficacy in a 24-week randomized, double blind, placebo-controlled study. Alzheimers Res. Ther. 2012, 4, 43. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Hishikawa, N.; Takahashi, Y.; Amakusa, Y.; Tanno, Y.; Tuji, Y.; Niwa, H.; Murakami, N.; Krishna, U.K. Effects of turmeric on Alzheimer’s disease with behavioral and psychological symptoms of dementia. Ayu 2012, 33, 499–504. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Mythri, R.B.; Jagatha, B.; Pradhan, N.; Andersen, J.; Bharath, M.M.S. Mitochondrial complex I inhibition in Parkinson’s disease: How can curcumin protect mitochondria? Antioxid. Redox Signal. 2007, 9, 399–408. [Google Scholar] [CrossRef] [PubMed]
  139. Ray, B.; Bisht, S.; Maitra, A.; Maitra, A.; Lahiri, D.K. Neuroprotective and neurorescue effects of a novel polymeric nanoparticle formulation of curcumin (NanoCurc™) in the neuronal cell culture and animal model: Implications for Alzheimer’s disease. J. Alzheimers Dis. JAD 2011, 23, 61–77. [Google Scholar] [CrossRef] [PubMed]
  140. Lin, M.-S.; Hung, K.-S.; Chiu, W.-T.; Sun, Y.-Y.; Tsai, S.-H.; Lin, J.-W.; Lee, Y.-H. Curcumin enhances neuronal survival in N-methyl-d-aspartic acid toxicity by inducing RANTES expression in astrocytes via PI-3K and MAPK signaling pathways. Prog. Neuropsychopharmacol. Biol. Psychiatry 2011, 35, 931–938. [Google Scholar] [CrossRef]
  141. Wu, A.; Ying, Z.; Gomez-Pinilla, F. Dietary curcumin counteracts the outcome of traumatic brain injury on oxidative stress, synaptic plasticity, and cognition. Exp. Neurol. 2006, 197, 309–317. [Google Scholar] [CrossRef]
  142. Xu, Y.; Ku, B.-S.; Yao, H.-Y.; Lin, Y.-H.; Ma, X.; Zhang, Y.-H.; Li, X.-J. The effects of curcumin on depressive-like behaviors in mice. Eur. J. Pharmacol. 2005, 518, 40–46. [Google Scholar] [CrossRef]
  143. Al-Omar, F.A.; Nagi, M.N.; Abdulgadir, M.M.; Al Joni, K.S.; Al-Majed, A.A. Immediate and delayed treatments with curcumin prevents forebrain ischemia-induced neuronal damage and oxidative insult in the rat hippocampus. Neurochem. Res. 2006, 31, 611–618. [Google Scholar] [CrossRef]
  144. Mehla, J.; Reeta, K.H.; Gupta, P.; Gupta, Y.K. Protective effect of curcumin against seizures and cognitive impairment in a pentylenetetrazole-kindled epileptic rat model. Life Sci. 2010, 87, 596–603. [Google Scholar] [CrossRef]
  145. Zbarsky, V.; Datla, K.P.; Parkar, S.; Rai, D.K.; Aruoma, O.I.; Dexter, D.T. Neuroprotective properties of the natural phenolic antioxidants curcumin and naringenin but not quercetin and fisetin in a 6-OHDA model of Parkinson’s disease. Free Radic. Res. 2005, 39, 1119–1125. [Google Scholar] [CrossRef]
  146. Wang, J.; Du, X.-X.; Jiang, H.; Xie, J.-X. Curcumin attenuates 6-hydroxydopamine-induced cytotoxicity by anti-oxidation and nuclear factor-kappa B modulation in MES23.5 cells. Biochem. Pharmacol. 2009, 78, 178–183. [Google Scholar] [CrossRef]
  147. Rajeswari, A.; Sabesan, M. Inhibition of monoamine oxidase-B by the polyphenolic compound, curcumin and its metabolite tetrahydrocurcumin, in a model of Parkinson’s disease induced by MPTP neurodegeneration in mice. Inflammopharmacology 2008, 16, 96–99. [Google Scholar] [CrossRef] [PubMed]
  148. Wang, M.S.; Boddapati, S.; Emadi, S.; Sierks, M.R. Curcumin reduces alpha-synuclein induced cytotoxicity in Parkinson’s disease cell model. BMC Neurosci. 2010, 11, 57. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Vajragupta, O.; Boonchoong, P.; Watanabe, H.; Tohda, M.; Kummasud, N.; Sumanont, Y. Manganese complexes of curcumin and its derivatives: Evaluation for the radical scavenging ability and neuroprotective activity. Free Radic. Biol. Med. 2003, 35, 1632–1644. [Google Scholar] [CrossRef] [PubMed]
  150. Radomska-Leśniewska, D.M.; Osiecka-Iwan, A.; Hyc, A.; Góźdź, A.; Dąbrowska, A.M.; Skopiński, P. Therapeutic potential of curcumin in eye diseases. Cent. Eur. J. Immunol. 2019, 44, 181–189. [Google Scholar] [CrossRef] [PubMed]
  151. Yue, Y.-K.; Mo, B.; Zhao, J.; Yu, Y.-J.; Liu, L.; Yue, C.-L.; Liu, W. Neuroprotective effect of curcumin against oxidative damage in BV-2 microglia and high intraocular pressure animal model. J. Ocul. Pharmacol. Ther. Off. J. Assoc. Ocul. Pharmacol. Ther. 2014, 30, 657–664. [Google Scholar] [CrossRef] [PubMed]
  152. Wang, L.; Li, C.; Guo, H.; Kern, T.S.; Huang, K.; Zheng, L. Curcumin Inhibits Neuronal and Vascular Degeneration in Retina after Ischemia and Reperfusion Injury. PLoS ONE 2011, 6, e23194. [Google Scholar] [CrossRef] [Green Version]
  153. Zhu, W.; Wu, Y.; Meng, Y.-F.; Wang, J.-Y.; Xu, M.; Tao, J.-J.; Lu, J. Effect of curcumin on aging retinal pigment epithelial cells. Drug Des. Devel. Ther. 2015, 9, 5337–5344. [Google Scholar] [CrossRef] [Green Version]
  154. Gupta, S.K.; Kumar, B.; Nag, T.C.; Agrawal, S.S.; Agrawal, R.; Agrawal, P.; Saxena, R.; Srivastava, S. Curcumin prevents experimental diabetic retinopathy in rats through its hypoglycemic, antioxidant, and anti-inflammatory mechanisms. J. Ocul. Pharmacol. Ther. Off. J. Assoc. Ocul. Pharmacol. Ther. 2011, 27, 123–130. [Google Scholar] [CrossRef]
  155. Manikandan, R.; Thiagarajan, R.; Beulaja, S.; Chindhu, S.; Mariammal, K.; Sudhandiran, G.; Arumugam, M. Anti-cataractogenic effect of curcumin and aminoguanidine against selenium-induced oxidative stress in the eye lens of Wistar rat pups: An in vitro study using isolated lens. Chem. Biol. Interact. 2009, 181, 202–209. [Google Scholar] [CrossRef] [PubMed]
  156. Suryanarayana, P.; Saraswat, M.; Mrudula, T.; Krishna, T.P.; Krishnaswamy, K.; Reddy, G.B. Curcumin and turmeric delay streptozotocin-induced diabetic cataract in rats. Investig. Ophthalmol. Vis. Sci. 2005, 46, 2092–2099. [Google Scholar] [CrossRef] [PubMed]
  157. Pradhan, N.; Guha, R.; Chowdhury, S.; Nandi, S.; Konar, A.; Hazra, S. Curcumin nanoparticles inhibit corneal neovascularization. J. Mol. Med. Berl. Ger. 2015, 93, 1095–1106. [Google Scholar] [CrossRef] [PubMed]
  158. Chen, M.; Hu, D.-N.; Pan, Z.; Lu, C.-W.; Xue, C.-Y.; Aass, I. Curcumin protects against hyperosmoticity-induced IL-1beta elevation in human corneal epithelial cell via MAPK pathways. Exp. Eye Res. 2010, 90, 437–443. [Google Scholar] [CrossRef]
  159. Chung, S.-H.; Choi, S.H.; Choi, J.A.; Chuck, R.S.; Joo, C.-K. Curcumin suppresses ovalbumin-induced allergic conjunctivitis. Mol. Vis. 2012, 18, 1966–1972. [Google Scholar] [PubMed]
  160. Burugula, B.; Ganesh, B.S.; Chintala, S.K. Curcumin Attenuates Staurosporine-Mediated Death of Retinal Ganglion Cells. Investig. Ophthalmol. Vis. Sci. 2011, 52, 4263–4273. [Google Scholar] [CrossRef] [Green Version]
  161. Davis, B.M.; Pahlitzsch, M.; Guo, L.; Balendra, S.; Shah, P.; Ravindran, N.; Malaguarnera, G.; Sisa, C.; Shamsher, E.; Hamze, H.; et al. Topical Curcumin Nanocarriers are Neuroprotective in Eye Disease. Sci. Rep. 2018, 8, 11066. [Google Scholar] [CrossRef] [Green Version]
  162. Kasi, P.D.; Tamilselvam, R.; Skalicka-Woźniak, K.; Nabavi, S.F.; Daglia, M.; Bishayee, A.; Pazoki-Toroudi, H.; Nabavi, S.M. Molecular targets of curcumin for cancer therapy: An updated review. Tumour Biol. J. Int. Soc. Oncodev. Biol. Med. 2016, 37, 13017–13028. [Google Scholar] [CrossRef]
  163. Zhou, H.; Beevers, C.S.; Huang, S. The targets of curcumin. Curr. Drug Targets 2011, 12, 332–347. [Google Scholar] [CrossRef]
  164. Aggarwal, B.B.; Surh, Y.-J.; Shishodia, S. (Eds.) The Molecular Targets and Therapeutic Uses of Curcumin in Health and Disease; Advances in Experimental Medicine and Biology; Springer: New York, NY, USA, 2007; ISBN 978-0-387-46400-8. [Google Scholar]
  165. Song, X.; Zhang, M.; Dai, E.; Luo, Y. Molecular targets of curcumin in breast cancer (Review). Mol. Med. Rep. 2019, 19, 23–29. [Google Scholar] [CrossRef] [Green Version]
  166. Hsu, H.-Y.; Wen, M.-H. Lipopolysaccharide-mediated reactive oxygen species and signal transduction in the regulation of interleukin-1 gene expression. J. Biol. Chem. 2002, 277, 22131–22139. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Nanji, A.A.; Jokelainen, K.; Tipoe, G.L.; Rahemtulla, A.; Thomas, P.; Dannenberg, A.J. Curcumin prevents alcohol-induced liver disease in rats by inhibiting the expression of NF-kappa B-dependent genes. Am. J. Physiol. Gastrointest. Liver Physiol. 2003, 284, G321–G327. [Google Scholar] [CrossRef] [Green Version]
  168. Rahimi, K.; Ahmadi, A.; Hassanzadeh, K.; Soleimani, Z.; Sathyapalan, T.; Mohammadi, A.; Sahebkar, A. Targeting the balance of T helper cell responses by curcumin in inflammatory and autoimmune states. Autoimmun. Rev. 2019, 18, 738–748. [Google Scholar] [CrossRef] [PubMed]
  169. Kang, G.; Kong, P.-J.; Yuh, Y.-J.; Lim, S.-Y.; Yim, S.-V.; Chun, W.; Kim, S.-S. Curcumin suppresses lipopolysaccharide-induced cyclooxygenase-2 expression by inhibiting activator protein 1 and nuclear factor kappab bindings in BV2 microglial cells. J. Pharmacol. Sci. 2004, 94, 325–328. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  170. Rahman, I.; Marwick, J.; Kirkham, P. Redox modulation of chromatin remodeling: Impact on histone acetylation and deacetylation, NF-kappaB and pro-inflammatory gene expression. Biochem. Pharmacol. 2004, 68, 1255–1267. [Google Scholar] [CrossRef] [PubMed]
  171. Fang, J.; Lu, J.; Holmgren, A. Thioredoxin reductase is irreversibly modified by curcumin: A novel molecular mechanism for its anticancer activity. J. Biol. Chem. 2005, 280, 25284–25290. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Wang, K.; Qiu, F. Curcuminoid Metabolism and its Contribution to the Pharmacological Effects. Curr. Drug Metab. 2013, 14, 791–806. [Google Scholar] [CrossRef]
  173. Yu, Q.; Liu, Y.; Wu, Y.; Chen, Y. Dihydrocurcumin ameliorates the lipid accumulation, oxidative stress and insulin resistance in oleic acid-induced L02 and HepG2 cells. Biomed. Pharmacother. 2018, 103, 1327–1336. [Google Scholar] [CrossRef]
  174. 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]
  175. Wu, J.-C.; Tsai, M.-L.; Lai, C.-S.; Wang, Y.-J.; Ho, C.-T.; Pan, M.-H. Chemopreventative effects of tetrahydrocurcumin on human diseases. Food Funct. 2014, 5, 12–17. [Google Scholar] [CrossRef]
  176. Naito, M.; Wu, X.; Nomura, H.; Kodama, M.; Kato, Y.; Kato, Y.; Osawa, T. The Protective Effects of Tetrahydrocurcumin on Oxidative Stress in Cholesterol-fed Rabbits. J. Atheroscler. Thromb. 2002, 9, 243–250. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Murugan, P.; Pari, L. Effect of Tetrahydrocurcumin on Lipid Peroxidation and Lipids in Streptozotocin-Nicotinamide-Induced Diabetic Rats. Basic Clin. Pharmacol. Toxicol. 2006, 99, 122–127. [Google Scholar] [CrossRef] [PubMed]
  178. Pari, L.; Murugan, P. Antihyperlipidemic Effect of Curcumin and Tetrahydrocurcumin in Experimental Type 2 Diabetic Rats. Ren. Fail. 2007, 29, 881–889. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Gao, Y.; Li, J.; Wu, L.; Zhou, C.; Wang, Q.; Li, X.; Zhou, M.; Wang, H. Tetrahydrocurcumin provides neuroprotection in rats after traumatic brain injury: Autophagy and the PI3K/AKT pathways as a potential mechanism. J. Surg. Res. 2016, 206, 67–76. [Google Scholar] [CrossRef] [PubMed]
  180. Anand, P.; Thomas, S.G.; Kunnumakkara, A.B.; Sundaram, C.; Harikumar, K.B.; Sung, B.; Tharakan, S.T.; Misra, K.; Priyadarsini, I.K.; Rajasekharan, K.N.; et al. Biological activities of curcumin and its analogues (Congeners) made by man and Mother Nature. Biochem. Pharmacol. 2008, 76, 1590–1611. [Google Scholar] [CrossRef] [PubMed]
  181. Pfeiffer, E.; Hoehle, S.I.; Walch, S.G.; Riess, A.; Sólyom, A.M.; Metzler, M. Curcuminoids Form Reactive Glucuronides In Vitro. J. Agric. Food Chem. 2007, 55, 538–544. [Google Scholar] [CrossRef]
  182. Vijaya Saradhi, U.V.R.; Ling, Y.; Wang, J.; Chiu, M.; Schwartz, E.B.; Fuchs, J.R.; Chan, K.K.; Liu, Z. A liquid chromatography–tandem mass spectrometric method for quantification of curcuminoids in cell medium and mouse plasma. J. Chromatogr. B 2010, 878, 3045–3051. [Google Scholar] [CrossRef] [Green Version]
  183. 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]
  184. Wang, Y.-J.; Pan, M.-H.; Cheng, A.-L.; Lin, L.-I.; Ho, Y.-S.; Hsieh, C.-Y.; Lin, J.-K. Stability of curcumin in buffer solutions and characterization of its degradation products. J. Pharm. Biomed. Anal. 1997, 15, 1867–1876. [Google Scholar] [CrossRef]
  185. Gordon, O.N.; Luis, P.B.; Sintim, H.O.; Schneider, C. Unraveling curcumin degradation: Autoxidation proceeds through spiroepoxide and vinylether intermediates en route to the main bicyclopentadione. J. Biol. Chem. 2015, 290, 4817–4828. [Google Scholar] [CrossRef] [Green Version]
  186. Zhang, G.; Nitteranon, V.; Chan, L.Y.; Parkin, K.L. Glutathione conjugation attenuates biological activities of 6-dehydroshogaol from ginger. Food Chem. 2013, 140, 1–8. [Google Scholar] [CrossRef] [PubMed]
  187. Schramm, A.; Jähne, E.A.; Baburin, I.; Hering, S.; Hamburger, M. Natural products as potential human ether-a-go-go-related gene channel inhibitors—Outcomes from a screening of widely used herbal medicines and edible plants. Planta Med. 2014, 80, 1045–1050. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Hu, C.-W.; Sheng, Y.; Zhang, Q.; Liu, H.-B.; Xie, X.; Ma, W.-C.; Huo, R.; Dong, D.-L. Curcumin inhibits hERG potassium channels in vitro. Toxicol. Lett. 2012, 208, 192–196. [Google Scholar] [CrossRef] [PubMed]
  189. Li, W.; Zhou, Y.; Yang, J.; Zhang, H.H.; Zhao, S.L.; Zhang, T.; Huo, J.; Zheng, P. Curcumin induces apoptosis and protective autophagy in human gastric cancer cells with different degree of differentiation. Zhonghua Zhong Liu Za Zhi 2017, 39, 490–496. [Google Scholar] [CrossRef]
  190. 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]
  191. Nelson, K.M.; Dahlin, J.L.; Bisson, J.; Graham, J.; Pauli, G.F.; Walters, M.A. The Essential Medicinal Chemistry of Curcumin. J. Med. Chem. 2017, 60, 1620–1637. [Google Scholar] [CrossRef]
  192. Cheng, A.L.; Hsu, C.H.; Lin, J.K.; Hsu, M.M.; Ho, Y.F.; Shen, T.S.; Ko, J.Y.; Lin, J.T.; Lin, B.R.; Ming-Shiang, W.; et al. Phase I clinical trial of curcumin, a chemopreventive agent, in patients with high-risk or pre-malignant lesions. Anticancer Res. 2001, 21, 2895–2900. [Google Scholar]
  193. Prasad, S.; Gupta, S.C.; Tyagi, A.K.; Aggarwal, B.B. Curcumin, a component of golden spice: From bedside to bench and back. Biotechnol. Adv. 2014, 32, 1053–1064. [Google Scholar] [CrossRef]
  194. Rahimnia, A.-R.; Panahi, Y.; Alishiri, G.; Sharafi, M.; Sahebkar, A. Impact of Supplementation with Curcuminoids on Systemic Inflammation in Patients with Knee Osteoarthritis: Findings from a Randomized Double-Blind Placebo-Controlled Trial. Drug Res. 2015, 65, 521–525. [Google Scholar] [CrossRef]
  195. 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]
  196. Lao, C.D.; Ruffin, M.T.; Normolle, D.; Heath, D.D.; Murray, S.I.; Bailey, J.M.; Boggs, M.E.; Crowell, J.; Rock, C.L.; Brenner, D.E. Dose escalation of a curcuminoid formulation. BMC Complement. Altern. Med. 2006, 6, 10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Chainani, W.N. Safety and anti-inflammatory activity of curcumin: A component of tumeric (Curcuma longa). J. Altern. Complement. Med. 2003, 9, 161–168. [Google Scholar] [CrossRef] [Green Version]
  198. Qin, S.; Huang, L.; Gong, J.; Shen, S.; Huang, J.; Ren, H.; Hu, H. Efficacy and safety of turmeric and curcumin in lowering blood lipid levels in patients with cardiovascular risk factors: A meta-analysis of randomized controlled trials. Nutr. J. 2017, 16, 68. [Google Scholar] [CrossRef] [PubMed]
  199. Sharma, R.A.; McLelland, H.R.; Hill, K.A.; Ireson, C.R.; Euden, S.A.; Manson, M.M.; Pirmohamed, M.; Marnett, L.J.; Gescher, A.J.; Steward, W.P. Pharmacodynamic and pharmacokinetic study of oral Curcuma extract in patients with colorectal cancer. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2001, 7, 1894–1900. [Google Scholar]
  200. Ryan, J.L.; Heckler, C.E.; Ling, M.; Katz, A.; Williams, J.P.; Pentland, A.P.; Morrow, G.R. Curcumin for radiation dermatitis: A randomized, double-blind, placebo-controlled clinical trial of thirty breast cancer patients. Radiat. Res. 2013, 180, 34–43. [Google Scholar] [CrossRef] [Green Version]
  201. Epelbaum, R.; Schaffer, M.; Vizel, B.; Badmaev, V.; Bar-Sela, G. Curcumin and gemcitabine in patients with advanced pancreatic cancer. Nutr. Cancer 2010, 62, 1137–1141. [Google Scholar] [CrossRef]
  202. Eaton, J.E.; Nelson, K.M.; Gossard, A.A.; Carey, E.J.; Tabibian, J.H.; Lindor, K.D.; LaRusso, N.F. Efficacy and safety of curcumin in primary sclerosing cholangitis: An open label pilot study. Scand. J. Gastroenterol. 2019, 54, 633–639. [Google Scholar] [CrossRef] [PubMed]
  203. Lukefahr, A.L.; McEvoy, S.; Alfafara, C.; Funk, J.L. Drug-induced autoimmune hepatitis associated with turmeric dietary supplement use. BMJ Case Rep. 2018, 2018. [Google Scholar] [CrossRef]
  204. Storka, A.; Vcelar, B.; Klickovic, U.; Gouya, G.; Weisshaar, S.; Aschauer, S.; Bolger, G.; Helson, L.; Wolzt, M. Safety, tolerability and pharmacokinetics of liposomal curcumin in healthy humans. Int. J. Clin. Pharmacol. Ther. 2015, 53, 54–65. [Google Scholar] [CrossRef]
  205. 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]
  206. Sun, J.; Bi, C.; Chan, H.M.; Sun, S.; Zhang, Q.; Zheng, Y. Curcumin-loaded solid lipid nanoparticles have prolonged in vitro antitumour activity, cellular uptake and improved in vivo bioavailability. Colloids Surf. B Biointerfaces 2013, 111, 367–375. [Google Scholar] [CrossRef] [PubMed]
  207. Lasoff, D.R.; Cantrell, F.L.; Ly, B.T. Death associated with intravenous turmeric (Curcumin) preparation. Clin. Toxicol. 2018, 56, 384–385. [Google Scholar] [CrossRef] [PubMed]
  208. Priyadarsini, K.I. The Chemistry of Curcumin: From Extraction to Therapeutic Agent. Molecules 2014, 19, 20091–20112. [Google Scholar] [CrossRef] [Green Version]
  209. Ravindranath, V.; Chandrasekhara, N. Absorption and tissue distribution of curcumin in rats. Toxicology 1980, 16, 259–265. [Google Scholar] [CrossRef]
  210. Nabavi, S.M.; Russo, G.L.; Tedesco, I.; Daglia, M.; Orhan, I.E.; Nabavi, S.F.; Bishayee, A.; Nagulapalli Venkata, K.C.; Abdollahi, M.; Hajheydari, Z. Curcumin and Melanoma: From Chemistry to Medicine. Nutr. Cancer 2018, 70, 164–175. [Google Scholar] [CrossRef]
  211. Aggarwal, B.B.; Sung, B. Pharmacological basis for the role of curcumin in chronic diseases: An age-old spice with modern targets. Trends Pharmacol. Sci. 2009, 30, 85–94. [Google Scholar] [CrossRef]
  212. Xie, X.; Tao, Q.; Zou, Y.; Zhang, F.; Guo, M.; Wang, Y.; Wang, H.; Zhou, Q.; Yu, S. PLGA Nanoparticles Improve the Oral Bioavailability of Curcumin in Rats: Characterizations and Mechanisms. J. Agric. Food Chem. 2011, 59, 9280–9289. [Google Scholar] [CrossRef]
  213. Yang, C.; Su, X.; Liu, A.; Zhang, L.; Yu, A.; Xi, Y.; Zhai, G. Advances in clinical study of curcumin. Curr. Pharm. Des. 2013, 19, 1966–1973. [Google Scholar]
  214. 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. Off. J. Korean Cancer Assoc. 2014, 46, 2–18. [Google Scholar] [CrossRef] [Green Version]
  215. Hou, Y.; Wang, H.; Zhang, F.; Sun, F.; Xin, M.; Li, M.; Li, J.; Wu, X. Novel self-nanomicellizing solid dispersion based on rebaudioside A: A potential nanoplatform for oral delivery of curcumin. Int. J. Nanomed. 2019, 14, 557–571. [Google Scholar] [CrossRef] [Green Version]
  216. Sunagawa, Y.; Hirano, S.; Katanasaka, Y.; Miyazaki, Y.; Funamoto, M.; Okamura, N.; Hojo, Y.; Suzuki, H.; Doi, O.; Yokoji, T.; et al. Colloidal submicron-particle curcumin exhibits high absorption efficiency-a double-blind, 3-way crossover study. J. Nutr. Sci. Vitaminol. 2015, 61, 37–44. [Google Scholar] [CrossRef] [Green Version]
  217. Patel, S.S.; Acharya, A.; Ray, R.S.; Agrawal, R.; Raghuwanshi, R.; Jain, P. Cellular and molecular mechanisms of curcumin in prevention and treatment of disease. Crit. Rev. Food Sci. Nutr. 2020, 60, 887–939. [Google Scholar] [CrossRef] [PubMed]
  218. McClements, D.J.; Li, F.; Xiao, H. The Nutraceutical Bioavailability Classification Scheme: Classifying Nutraceuticals According to Factors Limiting their Oral Bioavailability. Annu. Rev. Food Sci. Technol. 2015, 6, 299–327. [Google Scholar] [CrossRef] [PubMed]
  219. Zhang, L.; Zhu, W.; Yang, C.; Guo, H.; Yu, A.; Ji, J.; Gao, Y.; Sun, M.; Zhai, G. A novel folate-modified self-microemulsifying drug delivery system of curcumin for colon targeting. Int. J. Nanomed. 2012, 7, 151–162. [Google Scholar] [CrossRef] [Green Version]
  220. 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]
  221. Tønnesen, H.H.; Karlsen, J. Studies on curcumin and curcuminoids. VI. Kinetics of curcumin degradation in aqueous solution. Z. Lebensm. Unters. Forsch. 1985, 180, 402–404. [Google Scholar] [CrossRef]
  222. Hoehle, S.I.; Pfeiffer, E.; Sólyom, A.M.; Metzler, M. Metabolism of curcuminoids in tissue slices and subcellular fractions from rat liver. J. Agric. Food Chem. 2006, 54, 756–764. [Google Scholar] [CrossRef]
  223. Ireson, C.R.; Jones, D.J.L.; Orr, S.; Coughtrie, M.W.H.; Boocock, D.J.; Williams, M.L.; Farmer, P.B.; Steward, W.P.; Gescher, A.J. Metabolism of the cancer chemopreventive agent curcumin in human and rat intestine. Cancer Epidemiol. Biomark. Prev. Publ. Am. Assoc. Cancer Res. Cosponsored Am. Soc. Prev. Oncol. 2002, 11, 105–111. [Google Scholar]
  224. 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]
  225. Cuomo, J.; Appendino, G.; Dern, A.S.; Schneider, E.; McKinnon, T.P.; Brown, M.J.; Togni, S.; Dixon, B.M. Comparative absorption of a standardized curcuminoid mixture and its lecithin formulation. J. Nat. Prod. 2011, 74, 664–669. [Google Scholar] [CrossRef]
  226. Semalty, A.; Semalty, M.; Rawat, M.S.M.; Franceschi, F. Supramolecular phospholipids-polyphenolics interactions: The PHYTOSOME strategy to improve the bioavailability of phytochemicals. Fitoterapia 2010, 81, 306–314. [Google Scholar] [CrossRef]
  227. Gota, V.S.; Maru, G.B.; Soni, T.G.; Gandhi, T.R.; Kochar, N.; Agarwal, M.G. Safety and pharmacokinetics of a solid lipid curcumin particle formulation in osteosarcoma patients and healthy volunteers. J. Agric. Food Chem. 2010, 58, 2095–2099. [Google Scholar] [CrossRef]
  228. Rasenack, N.; Müller, B.W. Dissolution rate enhancement by in situ micronization of poorly water-soluble drugs. Pharm. Res. 2002, 19, 1894–1900. [Google Scholar] [CrossRef]
  229. Madhavi, D.; Kagan, D. Bioavailability of a Sustained Release Formulation of Curcumin. Integr. Med. Encinitas Calif. 2014, 13, 24–30. [Google Scholar]
  230. Bhardwaj, R.K.; Glaeser, H.; Becquemont, L.; Klotz, U.; Gupta, S.K.; Fromm, M.F. Piperine, a major constituent of black pepper, inhibits human P-glycoprotein and CYP3A4. J. Pharmacol. Exp. Ther. 2002, 302, 645–650. [Google Scholar] [CrossRef] [Green Version]
  231. Shoba, G.; Joy, D.; Joseph, T.; Majeed, M.; Rajendran, R.; Srinivas, P.S. Influence of piperine on the pharmacokinetics of curcumin in animals and human volunteers. Planta Med. 1998, 64, 353–356. [Google Scholar] [CrossRef] [Green Version]
  232. Schiborr, C.; Kocher, A.; Behnam, D.; Jandasek, J.; Toelstede, S.; Frank, J. The oral bioavailability of curcumin from micronized powder and liquid micelles is significantly increased in healthy humans and differs between sexes. Mol. Nutr. Food Res. 2014, 58, 516–527. [Google Scholar] [CrossRef]
  233. Sasaki, H.; Sunagawa, Y.; Takahashi, K.; Imaizumi, A.; Fukuda, H.; Hashimoto, T.; Wada, H.; Katanasaka, Y.; Kakeya, H.; Fujita, M.; et al. Innovative preparation of curcumin for improved oral bioavailability. Biol. Pharm. Bull. 2011, 34, 660–665. [Google Scholar] [CrossRef] [Green Version]
  234. Yang, K.-Y.; Lin, L.-C.; Tseng, T.-Y.; Wang, S.-C.; Tsai, T.-H. Oral bioavailability of curcumin in rat and the herbal analysis from Curcuma longa by LC-MS/MS. J. Chromatogr. B Analyt. Technol. Biomed. Life. Sci. 2007, 853, 183–189. [Google Scholar] [CrossRef]
  235. Maiti, K.; Mukherjee, K.; Gantait, A.; Saha, B.P.; Mukherjee, P.K. Curcumin–phospholipid complex: Preparation, therapeutic evaluation and pharmacokinetic study in rats. Int. J. Pharm. 2007, 330, 155–163. [Google Scholar] [CrossRef]
  236. Marczylo, T.H.; Verschoyle, R.D.; Cooke, D.N.; Morazzoni, P.; Steward, W.P.; Gescher, A.J. Comparison of systemic availability of curcumin with that of curcumin formulated with phosphatidylcholine. Cancer Chemother. Pharmacol. 2007, 60, 171–177. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  237. Garcea, G.; Jones, D.J.L.; Singh, R.; Dennison, A.R.; Farmer, P.B.; Sharma, R.A.; Steward, W.P.; Gescher, A.J.; Berry, D.P. Detection of curcumin and its metabolites in hepatic tissue and portal blood of patients following oral administration. Br. J. Cancer 2004, 90, 1011–1015. [Google Scholar] [CrossRef] [PubMed]
  238. Garcea, G.; Berry, D.P.; Jones, D.J.L.; Singh, R.; Dennison, A.R.; Farmer, P.B.; Sharma, R.A.; Steward, W.P.; Gescher, A.J. Consumption of the putative chemopreventive agent curcumin by cancer patients: Assessment of curcumin levels in the colorectum and their pharmacodynamic consequences. Cancer Epidemiol. Biomark. Prev. Publ. Am. Assoc. Cancer Res. Cosponsored Am. Soc. Prev. Oncol. 2005, 14, 120–125. [Google Scholar]
  239. Asai, A.; Miyazawa, T. Occurrence of orally administered curcuminoid as glucuronide and glucuronide/sulfate conjugates in rat plasma. Life Sci. 2000, 67, 2785–2793. [Google Scholar] [CrossRef]
  240. Perkins, S.; Verschoyle, R.D.; Hill, K.; Parveen, I.; Threadgill, M.D.; Sharma, R.A.; Williams, M.L.; Steward, W.P.; Gescher, A.J. Chemopreventive efficacy and pharmacokinetics of curcumin in the min/+ mouse, a model of familial adenomatous polyposis. Cancer Epidemiol. Biomark. Prev. Publ. Am. Assoc. Cancer Res. Cosponsored Am. Soc. Prev. Oncol. 2002, 11, 535–540. [Google Scholar]
  241. Ravindranath, V.; Chandrasekhara, N. In vitro studies on the intestinal absorption of curcumin in rats. Toxicology 1981, 20, 251–257. [Google Scholar] [CrossRef]
  242. Ravindranath, V.; Chandrasekhara, N. Metabolism of curcumin—Studies with [3H]curcumin. Toxicology 1981, 22, 337–344. [Google Scholar] [CrossRef]
  243. Im, K.; Ravi, A.; Kumar, D.; Kuttan, R.; Maliakel, B. An enhanced bioavailable formulation of curcumin using fenugreek-derived soluble dietary fibre. J. Funct. Foods 2012, 4, 348–357. [Google Scholar] [CrossRef]
  244. 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]
  245. Boyanapalli, S.S.S.; Huang, Y.; Su, Z.; Cheng, D.; Zhang, C.; Guo, Y.; Rao, R.; Androulakis, I.P.; Kong, A.-N. Pharmacokinetics and Pharmacodynamics of Curcumin in regulating anti-inflammatory and epigenetic gene expression. Biopharm. Drug Dispos. 2018, 39, 289–297. [Google Scholar] [CrossRef]
  246. Antony, B.; Merina, B.; Iyer, V.S.; Judy, N.; Lennertz, K.; Joyal, S. A Pilot Cross-Over Study to Evaluate Human Oral Bioavailability of BCM-95CG (Biocurcumax), A Novel Bioenhanced Preparation of Curcumin. Indian J. Pharm. Sci. 2008, 70, 445–449. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Effect of curcumin in different diseases. The therapeutic benefits obtained from in vitro cell cultures to small and large animal studies as well as clinical trials. CNS: Central Nervous System.
Figure 1. Effect of curcumin in different diseases. The therapeutic benefits obtained from in vitro cell cultures to small and large animal studies as well as clinical trials. CNS: Central Nervous System.
Ijms 21 06619 g001
Figure 2. Cellular and molecular targets of curcumin. Curcumin directly or indirectly interacts with numerous molecular targets and modulates their activity and function. AH R: Aryl hydrocarbon receptor, AP-1: Activator protein 1, Bax: Bcl-2-associated X protein, BDNF: Brain-derived neurotrophic factor, CDPK: Calcium-dependent protein kinases CRDB: Curcumin Resource Database, CREB: cAMP response element-binding protein, COX-2: Cyclooxygenase-2, CXCR 4: C-X-C Motif Chemokine Receptor 4, EGF: Epidermal growth factor, ER-alfa: Estrogen receptor alfa, ERK: Extracellular signal-regulated kinases, FADD: Fas Associated via death domain, FAK: Focal adhesion kinase, FAS: Fas cell surface death receptor, FGF: Fibroblast growth factors, GST: Glutathione-S-transferase, HAT: Histone acetylase, H2 R: Histamine H2 receptor, HDAC: Histone deacetylase, HGF: Hepatocyte growth factor, HMG-CoA-R: 3-hydroxy-3-methyl-glutaryl-CoA reductase, HSP-70: Heat shock protein 70, IBD: Intestinal inflammatory diseases ICAMs: Intercellular cell adhesion molecules, IL: Interleukin, iNOS: Inducible nitric oxide synthase, JAK: Janus kinase, JNK: c-Jun N-terminal kinases, LDL R: Low-Density Lipoprotein Receptor, MCP-1: Monocyte chemoattractant protein-1, MIP-1α: Macrophage inflammatory proteins, MMP: Matrix metallopeptidases, MRP: Multidrug resistance-associated protein, NFκB: Nuclear Factor kappa-light-chain-enhancer of activated B cells, NGF: Nerve growth factor, Nrf2: Nuclear factor erythroid 2–related factor 2, P38-MAPK: P38 mitogen-activated protein kinases PDGF: Platelet-derived growth factor, PhK: Phosphorylase kinase, PKA: Protein kinase A, PLA2: Phospholipase A2, PPAR-gamma: Peroxisome proliferator-activated receptor gamma, ROS: Reactive oxygen species, RNS: Reactive nitrogen species, SLP: Solid lipid particle, STAT: Signal transducer and activator of transcription, SyK: Spleen tyrosine kinase, TF: Tissue factor, TGF-α: Transforming growth factor alpha, TGF-β: Transforming growth factor beta, TLR: Toll-like receptors, TNF-α: Tumor necrosis factor alpha, UGT: Uridine diphosphate-glucuronosyltransferase, VCAM: Vascular cell adhesion molecule, XO: Xanthine oxidase, 5-LOX: 5-Lipoxygenase.
Figure 2. Cellular and molecular targets of curcumin. Curcumin directly or indirectly interacts with numerous molecular targets and modulates their activity and function. AH R: Aryl hydrocarbon receptor, AP-1: Activator protein 1, Bax: Bcl-2-associated X protein, BDNF: Brain-derived neurotrophic factor, CDPK: Calcium-dependent protein kinases CRDB: Curcumin Resource Database, CREB: cAMP response element-binding protein, COX-2: Cyclooxygenase-2, CXCR 4: C-X-C Motif Chemokine Receptor 4, EGF: Epidermal growth factor, ER-alfa: Estrogen receptor alfa, ERK: Extracellular signal-regulated kinases, FADD: Fas Associated via death domain, FAK: Focal adhesion kinase, FAS: Fas cell surface death receptor, FGF: Fibroblast growth factors, GST: Glutathione-S-transferase, HAT: Histone acetylase, H2 R: Histamine H2 receptor, HDAC: Histone deacetylase, HGF: Hepatocyte growth factor, HMG-CoA-R: 3-hydroxy-3-methyl-glutaryl-CoA reductase, HSP-70: Heat shock protein 70, IBD: Intestinal inflammatory diseases ICAMs: Intercellular cell adhesion molecules, IL: Interleukin, iNOS: Inducible nitric oxide synthase, JAK: Janus kinase, JNK: c-Jun N-terminal kinases, LDL R: Low-Density Lipoprotein Receptor, MCP-1: Monocyte chemoattractant protein-1, MIP-1α: Macrophage inflammatory proteins, MMP: Matrix metallopeptidases, MRP: Multidrug resistance-associated protein, NFκB: Nuclear Factor kappa-light-chain-enhancer of activated B cells, NGF: Nerve growth factor, Nrf2: Nuclear factor erythroid 2–related factor 2, P38-MAPK: P38 mitogen-activated protein kinases PDGF: Platelet-derived growth factor, PhK: Phosphorylase kinase, PKA: Protein kinase A, PLA2: Phospholipase A2, PPAR-gamma: Peroxisome proliferator-activated receptor gamma, ROS: Reactive oxygen species, RNS: Reactive nitrogen species, SLP: Solid lipid particle, STAT: Signal transducer and activator of transcription, SyK: Spleen tyrosine kinase, TF: Tissue factor, TGF-α: Transforming growth factor alpha, TGF-β: Transforming growth factor beta, TLR: Toll-like receptors, TNF-α: Tumor necrosis factor alpha, UGT: Uridine diphosphate-glucuronosyltransferase, VCAM: Vascular cell adhesion molecule, XO: Xanthine oxidase, 5-LOX: 5-Lipoxygenase.
Ijms 21 06619 g002
Figure 3. Obstacles against the marketing of curcumin as a drug. Challenges in curcumin oral bioavailability, distribution and metabolism, the main pharmacokinetic parameters, that emerged as major obstacles limiting the therapeutic efficacy and marketing of curcumin as a drug.
Figure 3. Obstacles against the marketing of curcumin as a drug. Challenges in curcumin oral bioavailability, distribution and metabolism, the main pharmacokinetic parameters, that emerged as major obstacles limiting the therapeutic efficacy and marketing of curcumin as a drug.
Ijms 21 06619 g003
Table 7. Effect of Curcumin in eye diseases.
Table 7. Effect of Curcumin in eye diseases.
ProductDose or Concentration UsedEffect and FindingsType of StudyStudied by
CurcuminIn vitro: 0.1, 1, 10 μM for 2 for 1 h
In vivo: 10 mg/kg orally daily for 6 weeks
Glaucoma: curcumin,
increased the cell viability and decreased intracellular ROS and apoptosis significantly.
In vivo study, curcumin protected rat BV-2 microglia from
death significantly
In vitro/Animal modelYue YK et al. 2014 [151]
CurcuminCurcumin (0.01%, 0.05% and 0.25%, which are equivalent to 100, 500 and 2500 ppm in diets) for 2 days before the injury Glaucoma: Curcumin protected retinal neurons and microvessels against Ischemia/Reperfusion injury through inhibition of injury-induced activation of NF-κB and STAT3, and on over-expression of MCP-1. Animal modelWang L et al. 2011 [152]
Curcumin Concentration 0–100 μM for 24 h Age-related macular degeneration: Curcumin improved cell viability and reduced apoptosis and oxidative stress and had a significant influence on expression of apoptosis-associated proteins and oxidative stress biomarkers.In vitroZhu W et al. 2015 [153]
CurcuminDose: 1 g/kg orally, daily for 16 weeksDiabetic retinopathy: curcumin positively controlled the antioxidant system, pro-inflammatory cytokines, tumor necrosis factor-α and vascular endothelial growth factor in the diabetic retinaeAnimal modelGupta SK et al. 2011 [154]
Curcumin + sodium seleniteConcentrations
Curcumin 100, 200 μM
sodium selenite 100 μM
Cataract: Curcumin suppressed selenium-induced oxidative stress and cataract formation through preventing depletion of antioxidants, and inhibiting generation of free radicals, and by inhibiting iNOS expressionIn vitroManikandan R et al. 2009 [155]
Curcumin and Turmeric extract 0.002%–0.01% curcumin and 0.5% turmeric in dietCataract: turmeric and curcumin were effective against the diabetic cataract development in rats.Animal modelSuryanarayana P et al. 2005 [156]
Curcumin nanoparticles (NP)In vitro: curcumin 5–20 μM for 24 h
In vivo: 20-μL solution containing 80 μg curcumin for 14 days
Corneal neovascularization: NP increased the retention of curcumin in the cornea and suppressed the expression of VEGF, inflammatory cytokines, and MMP so prevented corneal neovascularization through suppressing the NFκB pathway.In vitro/Animal modelPradhan N et al. 2015 [157]
Curcumin Concentrations 1–30 μM for 24 hDry eye disease: Curcumin has the potential for dry eye disease. It prevented the hyperosmoticity-induced increase of NF-κB and IL-1β productionIn vitroChen M et al. 2010 [158]
Curcumin Dose:10, 20 mg/kg intraperitoneally twice on days 14 and 17, beginning 1 h before the challenge in the conjunctival sacConjunctivitis: curcumin suppressed the allergic conjunctival inflammation in an experimental model.Animal modelChung SH et al. 2012 [159]
Curcumin Dose: 2.5 and 10 μM) injected into the vitreous of C57BL/6 mice.Retinal degeneration: curcumin attenuated retinal ganglion cell and amacrine cell death by restoring NF-κB expression.Animal modelBurugula B et al. 2011 [160]
Curcumin NanoparticleIn vitro: curcumin 0–20 μM for 24 h
In vivo: topical eye drop daily for 21 days
Neuroprotective in eye disease: Curcumin-loaded nanocarriers protected a retinal cell line against glutamate and hypoxia-induced injuryIn vitro/Animal modelDavis BM et al. 2018 [161]

Share and Cite

MDPI and ACS Style

Hassanzadeh, K.; Buccarello, L.; Dragotto, J.; Mohammadi, A.; Corbo, M.; Feligioni, M. Obstacles against the Marketing of Curcumin as a Drug. Int. J. Mol. Sci. 2020, 21, 6619. https://doi.org/10.3390/ijms21186619

AMA Style

Hassanzadeh K, Buccarello L, Dragotto J, Mohammadi A, Corbo M, Feligioni M. Obstacles against the Marketing of Curcumin as a Drug. International Journal of Molecular Sciences. 2020; 21(18):6619. https://doi.org/10.3390/ijms21186619

Chicago/Turabian Style

Hassanzadeh, Kambiz, Lucia Buccarello, Jessica Dragotto, Asadollah Mohammadi, Massimo Corbo, and Marco Feligioni. 2020. "Obstacles against the Marketing of Curcumin as a Drug" International Journal of Molecular Sciences 21, no. 18: 6619. https://doi.org/10.3390/ijms21186619

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