Next Article in Journal
Impact of Cattle Feeding Strategy on the Beef Metabolome
Next Article in Special Issue
Glucose Metabolism Disorders: Challenges and Opportunities for Diagnosis and Treatment
Previous Article in Journal
The New Era of Salivaomics in Dentistry: Frontiers and Facts in the Early Diagnosis and Prevention of Oral Diseases and Cancer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Reconnoitering the Therapeutic Role of Curcumin in Disease Prevention and Treatment: Lessons Learnt and Future Directions

1
Banerjee Research Group, College of Medicine and Health Sciences, Mohammed Bin Rashid University of Medicine and Health Sciences (MBRU), Dubai 505055, United Arab Emirates
2
Department of Diabetes, Nutrition and Metabolic Diseases, Carol Davila University of Medicine and Pharmacy, 020022 Bucharest, Romania
3
Department of Health Promotion, Mother and Child Care, Internal Medicine and Medical Specialties (Promise), University of Palermo, 90128 Palermo, Italy
4
Centre for Medical Education, University of Dundee, Dundee DD1 4HN, UK
*
Author to whom correspondence should be addressed.
Metabolites 2022, 12(7), 639; https://doi.org/10.3390/metabo12070639
Submission received: 30 April 2022 / Revised: 30 June 2022 / Accepted: 8 July 2022 / Published: 12 July 2022

Abstract

:
Turmeric is a plant with a very long history of medicinal use across different cultures. Curcumin is the active part of turmeric, which has exhibited various beneficial physiological and pharmacological effects. This review aims to critically appraise the corpus of literature associated with the above pharmacological properties of curcumin, with a specific focus on antioxidant, anti-inflammatory, anticancer and antimicrobial properties. We have also reviewed the different extraction strategies currently in practice, highlighting the strengths and drawbacks of each technique. Further, our review also summarizes the clinical trials that have been conducted with curcumin, which will allow the reader to get a quick insight into the disease/patient population of interest with the outcome that was investigated. Lastly, we have also highlighted the research areas that need to be further scrutinized to better grasp curcumin’s beneficial physiological and medicinal properties, which can then be translated to facilitate the design of better bioactive therapeutic leads.

Graphical Abstract

1. Introduction

In recent times, there has been an increased impetus to reconnoitre the medicinal properties of food. Case in point, Rao et al., in a recent exploratory study, assessed the safety and prospective efficacy of Nigella sativa and fenugreek seed-supplemented chapattis (unleavened flatbread originating from the Indian subcontinent) in obese and type–2 diabetic subjects to demonstrate that the consumption of chapattis combined with N. sativa/fenugreek triggered a significant clinical improvement in obesity and diabetes. However, the other key highlight of this study was the long-term compliance of 100% [1]. Although the compliance regarding a clinical trial may vary from the compliance during the regular use of a food product in the community, the adherence to the dietary intervention in this study is a reason for optimism, whereby treatment for chronic diseases can be effectively delivered through food. Hence, this school of thought has coined the term “Functional Foods”.
The term “Functional foods (FFs)” was first created and defined in the 1980s by the Ministry of Health and Welfare of Japan when they established a regulatory system for foods that possess possible health benefits [2]. FFs were defined as foods that keep constructive effects on target functions in the physiological milieu of humans beyond nutritional effects, aiming for health promotion and wellbeing and/or the reduction of chronic diseases. With time, FFs have attained popularity, with a market value of USD 1.7 billion as of today. Further to the above, FFs have given way to the term “nutraceutical”.
The term “nutraceutical” was coined from “nutrition” and “pharmaceutical” in 1989 by the founder and chairman of the Foundation for Innovation in Medicine (FIM) situated in Cranford, NJ, Stephen DeFelice, MD [3]. According to DeFelice, a nutraceutical is “a food (or part of a food) that provides medical or health benefits, including the prevention and/or treatment of a disease” [3]. In other words, the term “nutraceutical” implies the pharmaceutical formulation of the bioactive compound, whose concentration often shifts from its natural concentration in food. No regulatory evaluations and no toxicological assessments are required. Therefore, one can say that when FF aids in the prevention and/or treatment of disease(s) and/or disorder(s) other than anaemia, it is called a nutraceutical [4]. However, an important point to remember is that the term nutraceutical, as commonly used in marketing, has no regulatory definition. Hence, broadly, nutraceuticals are foods or parts of food playing an important role in modifying and maintaining the customary physiological function that maintains healthy human beings. Case in point, allyl sulphur compounds in garlic, quercetin in berries, EPA [5] (Eicosapentaenoic acid) and DHA (Docosahexaenoic acid) in fish oils, curcumin in turmeric, ginsenosides from ginseng roots and polyphenolic catechins in green tea are some of the nutraceuticals that have been studied extensively [6].
There is an increased interest in the niche of nutraceutical research because nutraceutical(s) are generally associated with lesser side effects. For example, oleocanthal (OC), present in significantly high concentrations in extra-virgin olive oil (EVOO), is a structural analogue of ibuprofen and, like ibuprofen, mediates anti-inflammatory properties by the inhibition of cyclooxygenase (COX) enzymes in the prostaglandin biosynthesis pathway [7,8]. Similarly, lycopene, which is a tetraterpene compound abundant in tomato and tomato-based products, is essentially recognized as a potent antioxidant and a non-pro-vitamin A carotenoid. Lycopene has been shown to ameliorate cancer insurgences, diabetes mellitus, cardiac complications, oxidative stress-mediated malfunctions, inflammatory events, skin and bone diseases and hepatic, neural and reproductive disorders [9]. Likewise, reserveratol, which is an activator of SIRT1, one of the mammalian forms of the sirtuin family of proteins, mediates its beneficial effects on metabolism, stress resistance, cell survival, cellular senescence, inflammation–immune function, endothelial functions and circadian rhythms [10]. Curcumin isolated from the turmeric plant is one of those single molecules that have been appraised extensively in numerous in vitro and in vivo studies. However, few of the effects observed in such studies have been replicated efficiently in the numerous clinical trials that have been conducted with curcumin (see below for details). This raises the question, “Should we continue to explore the beneficial properties of curcumin, which some have christened as “The Golden Spice”?” Furthermore, “is there a definite mechanism by which curcumin mediates its effect?” In this review, we tackle these aspects of curcumin structure-function and address some of the surrounding controversies.
Turmeric is a plant has a very long history of medicinal use, especially in Indian culture, dating back nearly 4000 years. It is used as a culinary spice and has immense religious significance. In India, turmeric is colloquially referred to as “Haldi”, which, when literally translated, denotes the color yellow and is responsible for the precise yellow color of the traditional Indian curry. Over time, the use of turmeric also reached other parts of the globe. It possibly reached China by 700 AD, East Africa by 800 AD, West Africa by 1200 AD and Jamaica in the eighteenth century. In 1280, Marco Polo alluded to this spice, marvelling at a vegetable that displayed virtues like saffron. According to Sanskrit medical discourses and the alternative medicinal disciplines of the Ayurveda and Unani systems, turmeric has a long history of therapeutic use in South Asia. In fact, Susruta’s Ayurvedic Compendium, dating back to 250 BC, endorses the use of an ointment containing turmeric to relieve the effects of poisoned food. The medicinal property of turmeric is attributed to the bioactive main natural polyphenol compound curcumin, (1,7-bis(4-hydroxy-3-methoxyphenyl)-1,6-heptadiene-3,5-dione), also called diferuloylmethane. Curcumin mediates a plethora of beneficial physiological effects. Due to its ability to interact with various molecular targets, curcumin is one of the most interesting pleiotropic nutraceuticals [11]. The antioxidant, anti-inflammatory, anticancer and antimicrobial properties of curcumin have been extensively researched and appraised in different in vitro and in vivo experimental models. Figure 1 gives an overview of the pharmacological activities of curcumin, which will be further discussed in detail in the manuscript.
This review aims to critically appraise the corpus of literature associated with the above pharmacological properties of curcumin, with a specific focus (Figure 1) on identifying the gaps that need to be addressed to obtain a better insight into the molecular mechanism through which curcumin mediates these advantageous effects. Lastly, we have also highlighted the research areas that need to be further scrutinized to better grasp the beneficial physiological and medicinal properties of curcumin, which can then be translated to facilitate the design of better bioactive therapeutic leads.

1.1. Source of Curcumin

Curcumin is the principal curcuminoid of the turmeric plant Curcuma longa. Curcumin was discovered around two centuries ago when Vogel and Pelletier reported the isolation of “yellow colouring-matter” from the rhizomes of Curcuma longa (turmeric) and named it curcumin [12]. The turmeric plant, Curcuma longa, is a rhizomatous herbaceous perennial plant belonging to the ginger family Zingiberaceae, which is native to tropical South Asia (Figure 2a–c) [13]. As many as 133 species of Curcuma have been identified worldwide [14]. Turmeric is derived from the mature tuberous rhizome of C. longa. Once the rhizomes mature underground (beneath the foliage), they become yellowish-brown with a dull orange interior (Figure 2b) [15]. This yellowish color is because of the presence of curcuminoids, which are natural polyphenol compounds classified into three diarylheptanoids (diferuloylmethane derivatives): curcumin (77%), demethoxycurcumin (17%), bisdemethoxycurcumin (3–6%) (Figure 2d-f) and other less abundant secondary metabolites.

1.2. Chemistry and Bioavailability of Curcumin

Curcumin has been identified as 1,6-heptadiene-3,5-dione-1,7-bis(4-hydroxy-3-methoxyphenyl)-(1E,6E) or diferuloylmethane (Figure 2d) [16]. It is an orange-yellow crystalline powder that is relatively insoluble in water (limiting its medicinal use in humans when taken orally or injected) and ether but soluble in acetone, ethanol and acetic acid dimethylsulfoxide [17]. It has a melting point of 183 °C, a molecular formula of C21H20O6 and a molecular weight of 368.37 g/mol [18]. Spectrophotometrically, the maximum absorption (lmax) of curcumin in methanol occurs at 430 nm, and, in acetone, it occurs at 415–420 nm [19]. The powder gives a brownish-red color with alkali and a light-yellow color with acids [20]. Curcumin exists in enolic and β-diketonic forms. The fact that curcumin in solutions exists primarily in its enolic form has been the key to the radical scavenging ability of curcumin [21]. It is stable in acidic pH, but in neutral and basic pH, it gets degraded to ferulic acid and feruloylmethane. Curcumin rapidly degrades when placed in a phosphate buffer solution at a pH of 7.2, whereas in ascorbic acid, N-acetylcysteine and glutathione, it does not degrade, which explains the oxidative mechanism of these anti-oxidative agents [22]. Curcumin is poorly absorbed in the gastrointestinal tract (one of the key hurdles to increasing the bioavailability of curcumin). Case in point, a poor absorption from the gut was observed in rats after the oral administration of curcumin at a dose of 1 g/kg bw, which led to 75% fecal excretion with traces in the urine, and the concentration of curcumin was below 5 ug/mL in the plasma [23,24]. The oral administration of radio-labelled curcumin at a 0.6 mg/kg dose in rats resulted in 89% excretion in feces and 6% excretion in the bile after 72 h. At the same dose, when administered peritoneally, 73% fecal excretion was observed, and 11% excretion was observed in the bile [25]. A slightly better absorption rate of 60% was seen upon 400 mg of curcumin being administered orally, and 40% faecal excretion was observed over a period of 5 days [26].

1.3. Extraction of Curcumin

Several extraction strategies have been availed for the isolation of curcumin. Conventional methods such as solvent extraction, Soxhlet extraction and hydro/steam distillation are time-consuming, are not eco-friendly and have low efficiencies. In this review, we have not touched upon these techniques. The readers are referred to the excellent review of Zhang et al. for details, if interested [27]. Novel extraction techniques (summarized in Table 1, with the associated advantages and drawbacks of these techniques) have been effectively strategized to maximize the extraction efficiency, decrease the use of toxic solvents and concomitantly be cost-effective.

2. Methods

Relevant publications were searched for in PubMed (https://pubmed.ncbi.nlm.nih.gov (accessed on 29 April 2022) and Google Scholar (https://scholar.google.com (accessed on 29 April 2022), using the various names of curcumin and its related functions such as antimicrobial, antiinflammation, anti-fungal, antibacterial, rheumatoid arthritis, cancer, diabetes, inflammatory bowel disease and gut microbiota as keywords. The lead author (BMS) and the corresponding author (YB) finalized the keywords, which were vetted and agreed upon by all the authors. The search was conducted independently by two authors (BMS and MA). Overlaps were eliminated, and the final list of selected articles was agreed upon and vetted by all participating authors. The search was restricted to articles published only in English. The images are shown in Figure 2a The turmeric plant, (b) the turmeric rhizome with a yellow-orange interior, (c) the powdered form of turmeric and (d) the chemical structure of curcumin were adopted with minor modifications from the Wiki commons [33,34].

3. Anti-Inflammatory Properties of Curcumin

Inflammation is a response to tissue damage caused by oxidative stress, pathogens, chemicals or radiation and triggers repair. Chronic inflammation lasts from several months to years due to tissue invasion by inflammatory cytokines and growth factors. Curcumin shows an anti-inflammatory property by interacting with Toll-like receptors (TLRs), which play a key role in innate immunity [35]. Upon binding, it regulates the production of inflammatory mediators such as Mitogen-activated protein kinases (MAPK), Activator Protein 1 (AP-1) and Nuclear Factor Kappa-B (NF-κB) [36]. The Janus kinase/Signal transducer and activator of the transcription (JAK/STAT) signaling pathway has been one of the main targets to treat inflammatory diseases such as rheumatoid arthritis and inflammatory bowel diseases. Curcumin has also been proven to regulate JAK/STAT signaling. Another way to decrease inflammation is by regulating inflammatory mediators. Case in point, curcumin has decreased the level of mediators such as Interleukin-1 (IL-1), IL-17, IL-27, IL-6, IL-8, IL-1β [37], Tumor necrosis factor-α, Monocyte chemotactic protein-1 (MCP-1) and Inducible nitric oxide synthase (iNOS). Nuclear factor erythroid 2 p45-related factor (Nrf2) overactivation is seen in neoplasms [38] and has also been linked with insulin resistance in diabetes. Curcumin has suppressed proteins such as Keap1, which interacts with Nrf2, thereby regulating its overexpression. In Figure 3, we have summarized curcumin’s anti-inflammatory property via the inhibition of signalling pathways.
One of the most important complexes that participate in inflammation is the inflammasome. Among the various inflammasomes, the NOD-like receptor pyrin domain containing 3 (NLRP3) detects the products of damaged cells and triggers an immune response. It mainly involves two plausible mechanisms:
(1)
Inflammatory bacterial products such as lipopolysaccharide(LPS) activate the NF-κB pathway to activate NLRP3, leading to Pro-Interleukin-1β (pro-IL-1β) synthesis.
(2)
Stimuli such as nigericin, aluminium crystal and monosodium urate crystal lead to NLRP3 activation, subsequently leading to the activation of caspase-1 along with the promotion of proinflammatory cytokines such as IL-1B and IL-18 [39].
NK-κB plays a vital role in aggregating NLRP3 components to form an active NLRP3 inflammasome. Curcumin has been shown to suppress the activation of the NLRP3 inflammasome and IL-1B secretion by regulating the NK-κB pathway [40]. Additionally, curcumin also inhibits the NLRP3 inflammasome by preventing Ca2+ influx and attenuating K+ efflux, thereby disrupting the formation of NLRP3 components [41]. Therefore, NLRP3 is one of the best targets by which curcumin can treat various inflammatory diseases. The key inflammatory diseases for which the beneficial effects of curcumin have been extensively investigated/appraised are depicted below.

3.1. Rheumatoid Arthritis (RA)

RA is a chronic inflammatory disease affecting the joints and causing irreversible bone, synovium and cartilage degradation, reduced mobility and discomfort. Curcumin has been found to suppress pro-inflammatory pathways crucial in the development of RA. A study by Wang et al. and Murakami et al. demonstrated that curcumin increased macrophage apoptosis and decreased the level of IKBα, thereby reducing the expression of COX-2 and inhibiting the activation of NF-κB [42,43]. Curcumin has also been shown to inhibit lymphocyte proliferation and decrease IL-4 and IL-5 levels and the granulocyte-macrophage colony-stimulating factor in lymphocytes [44]. Moreover, curcumin augments the activity of anti-inflammatory IL-10, inhibits BAFF (B cell-activating factor) expression and suppresses STAT1 signaling [45]. An in vivo study with curcumin by da Silva et al. revealed decreased infiltration and neutrophil activation, which prevents the migration of neutrophils from the blood to inflamed joints, acting as a proapoptotic agent in RA treatment. Curcumin also increases the surface expression of the cluster of differentiation (CD) 16+ and CD 56dim in natural killer cells, proving its immunostimulatory activity [46]. Experiments conducted by Moon et al. [47] and Dai et al. [48] explained the anti-inflammatory property of curcumin in synovial fibroblasts, where it suppresses COX-2 (this blocks the synthesis of prostaglandin E2), reducing synovial cell hyperplasia via the Mtor pathway and downregulating various NK-κB complexes, IL-1β and TNF-α [47,49].
Chondrocyte apoptosis seen in RA has been responsible for joint cartilage damage. In a study, curcumin inhibited IL-1β-induced IKBα phosphorylation and the activation of caspase-3 and COX-2 in chondrocytes isolated from cartilage (this might support cartilage regeneration in RA) and suppressed apoptosis in these chondrocytes [50]. In in vivo RA models, curcumin has been shown to decrease IL-1β, IL-18RA, IL-6, IL-18, TNF-α, IFN-gam, MMP3 [51] and IL-17 [42]. Bone degradation in RA by osteoclasts has been investigated by Shang et al., employing peripheral blood mononuclear cells (PBMCs) obtained from patients with RA with different concentrations of curcumin (2.5–10 µM) for 48 h. The results from this study demonstrated that curcumin inhibited M-CSF and RANKL-stimulated osteoclast differentiation via the suppression of ERK1/2, p38 and JNK activation. Another study evaluated curcumin’s capacity for inhibiting human osteoclastogenesis. Curcumin concentrations in the range of 1–10 µM inhibited osteoclast differentiation and bone-resorption, indicating that curcumin could be a potential therapeutic leading to managing bone deterioration in RA. It has also been reported that curcumin supplementation (500 mg for 8 weeks) [52] or curcumin nanomicelle administration (40 mg, 3 times a day over a period of 12 weeks) in RA patients tend to decrease the tenderness and swelling of the joints [53].

3.2. Osteoarthritis (OA)

Osteoarthritis (OA) is one of the leading causes of morbidity and disability worldwide. The prevalence of OA is projected to increase in the future [54,55]. The disease’s exact pathophysiology is not yet completely understood. Nonetheless, biomechanical (wear and tear), inflammatory and metabolic factors have been implicated in inducing the sterile inflammation and catabolism of the cartilage of the joint [56,57,58]. To date, there is no effective treatment to prevent or halt the progression of the disease that has been discovered [59]. The available pharmacological interventions, including non-steroidal anti-inflammatory drugs (NSAIDs) and acetaminophen, target the symptomatic treatment of pain [60]. Nonetheless, the prolonged use of NSAIDs is associated with significant cardiovascular, renal and gastrointestinal adverse events [61,62,63]. Curcumin emerged as a safe alternative for pain symptom relief and has been studied in preclinical and clinical trials.
In several preclinical studies, curcumin has shown positive effects on the reduction of inflammatory and catabolic markers in OA rat models [64,65]. Yan et al. have examined the effects of intra-articular curcumin injections in OA-induced rat knee models. Inflammatory markers in OA, including the Toll-like receptor (TLR)- 4 and its downstream pathway including NF-κB, IL-1β and TNF-α, were reduced significantly [64]. Additionally, curcumin preserved cartilage thickness and reduced the number of apoptotic chondrocytes in microscopic studies [64]. Zhang et al. showed similar findings with an intraperitoneal injection of curcumin [66,67]. The oral curcumin effects on rat OA models showed similar findings of decreased serum levels of cyclooxygenase-2 (COX-2) and 5-lipoxygenase, which are responsible for pain and inflammation. Matrix metalloproteinase-3 (MMP-3) proteins, which are highly expressed in osteoarthritic tissues and are responsible for breaking down cartilage by degrading the extracellular matrix in osteoarthritic joints, were also reduced [68]. Additional in vitro studies revealed the decreased activation of proapoptotic protein caspase-1 and the decreased expression of MMP-3 and displayed a dose-dependent inverse relationship between curcumin and MMP-3 levels [65]
Reduced autophagy and increased apoptosis have been indicated in the pathophysiology of OA [69,70]. In vivo experiments revealed that curcumin admiration decreased caspase-3 and Bax/Bcl2 levels, reducing apoptosis, while autophagic activity was high through the increased expression levels of light chain-3 (LC-3) [66]. Additionally, in vitro, mechanistic studies revealed the inhibition of the AKT/mTOR pathway by curcumin, which resulted in reduced apoptosis and enhanced autophagy [66,71].
Clinical studies of oral curcumin showed promising results in alleviating OA symptoms [72]. Several studies have shown benefits on the pain and functional scores of OA after administering oral curcumin alone or as an adjunct [73,74,75,76,77]. Previous lab studies revealed the synergism of COX-2 inhibitors and curcumin by reducing the expression of the enzyme and reducing prostaglandin E2 levels, which has translated in clinical trials into reducing pain symptoms and improving functional outcomes [78,79]. Additionally, Shep et al. showed that patients using NSAIDs and curcumin reported reduced GI pain as compared to patients receiving NSAIDs alone [77]. Patients on curcumin were able to decrease their daily dosage of NSAIDs owing to the analgesic effect of curcumin. Kuptniratsaikul et al. compared ibuprofen against ibuprofen and showed similar outcomes on the Western Ontario and McMaster Universities Osteoarthritis Index (WOMAC) [76]. An exploratory trial has shown decreased Coll2-1, a novel OA marker in patients’ serum, after administering curcumin [75]. A recent systematic review and metanalysis by Paultre et al. concluded that heterogenous curcumin is safe and beneficial in terms of the pain and function of patients with knee OA [72].

3.3. Cancer

Inflammation causes an increase in the production of pro-inflammatory molecules such as cytokines, reactive oxygen species (ROS), cyclooxygenase (COX)-2, transcription factors such as NF-κB, protein kinases B, activator protein 1(AP-1) and the signal transducer and activator of transcription 3(STAT3), leading to the initiation and development of cancer [80]. Curcumin shows a similar activity as that seen in RA, where it suppresses NF-κB activity by inhibiting IκB. It downregulates the expression of inflammatory genes such as TNF-α[81] and downregulates cyclin D1, Bcl-2, Bcl-xL, IL-6, COX-2 and MMP9 through NF-κB inhibition [82]. Curcumin has also been shown to downregulate AP-1 (known to be related to anti-apoptotic genes). In addition, Curcumin is directly or indirectly related to the regulation of STAT3 (a protein that promotes oncogenesis) by inhibiting IL-6 [83]. The anticancer effects of curcumin observed in different cancer models are summarized in Table 2.

3.4. Diabetes

Inflammation plays a pivotal role in diabetes. In fact, a review of the possible mechanisms that drive the metabolic pattern in Type 1 Diabetes and Type 2 Diabetes (T1D and T2D) and the involved inflammatory pathways indicates that the effective management of diabetes requires the modulation of the inflammatory pathways. In line with this, in this review, we will critically appraise the different cell and animal models that have been employed in investigating the anti-diabetic effects of curcumin, identifying the key results obtained in these studies. First, stress-causing factors such as obesity stimulate NK-κB activity and cause insulin resistance in adipose tissue, the liver and leukocytes. Second, curcumin supplementation has significantly reduced the NLRP3 inflammasome by inhibiting its activation, downregulating the NK-κB pathway and thereby reducing the caspase-1 activation and IL-1B secretion. Another anti-inflammatory activity of curcumin is the inhibition of ER stress in adipocytes by preventing the phosphorylation of phospho-inositol-requiring kinase 1(p-IRE1) and phospho-eukaryotic Initiation Factor 2 (p-eIF2) [163]. It also reduced the glycerol level and FFA released from adipose tissues [164]. The third mechanism is the inhibition of the pro-inflammatory NF-κB signaling pathway activation. Therefore, curcumin has shown beneficial anti-inflammatory effects by suppressing the expression of IL-6, TNFα, IL-1β [165] and MCP-1 from adipocytes [166] by inhibiting the recruitment of macrophages in adipose tissues and inhibiting NLRP3 inflammasome activity [40]. The effect of curcumin on diabetes has been summarized in Table 3.
In experimental models, chemicals such as streptozotocin (STZ) and alloxan have been used to induce diabetes. In mice, low doses of STZ (i.e., 40 mg/kg intraperitoneally injected for 5 consecutive days) have closely resembled human T1DM, with chronic pancreatic islet inflammation, insulitis and insulin deficiency. In rats, a single dose of STZ (i.e., 65 mg/kg) is required to generate T1DM, and high doses of STZ cause the toxin-induced necrosis of B cells, hypoglycemia and cell death. For T2DM, the exposure to a high-fat diet (60% fat by caloric content) followed by a moderate dose of STZ has resulted in hyperglycemia and insulin resistance [167]. None of the above models mimic human T1DM and T2DM. Therefore, the choice of model depends on the aim of the study. Challenges such as the regulation of STZ specificity and toxicity, the careful monitoring of diets and other factors and the ethics involved in the use of animal models should be kept in mind for the appropriate induction of diabetes using STZ [168].
Alloxan has been effectively administered at 170–200 mg/kg BW intraperitoneally to induce diabetes in animal models. However, alloxan-induced hyperglycemia is not sufficiently stable for the proper evaluation of antidiabetic compounds. It induces diabetes by a mechanism characterized by reactive oxygen species toxicity, ketosis and a high mortality rate. Instability, poor diabetogenicity, easy auto-reversal and the route and speed of administration are the factors to be considered to improve the use of alloxan as a diabetogenic drug [169].
Table 3. Summary of the anti-diabetic role of curcumin and the mechanism of action.
Table 3. Summary of the anti-diabetic role of curcumin and the mechanism of action.
ModelConc. of CurcuminIncreaseDecreaseNo. of Mice/Rats UsedRoute of
Administration
Reference
Albino Wistar rats with Streptozotocin-induced diabetes 0.5% of diet; 8 weeksATPase activity, PUFA/SFA ratioPhospholipid, triglyceride, kidney weight, renal lesion progression, renal damage, urine ALT and AST, kidney alkaline and acid phosphatase, glucose-6- phosphatase48Intraperitoneal[170]
Albino Wistar rats with Streptozotocin-induced diabetes300 mg/kg b.w./day for 8 weeks Creatinine, kidney SOD activity, kidney catalase activityGlucose, total cholesterol, triglyceride, urea, body weight, kidney lipid peroxidation10Intraperitoneal[171]
Wistar Rats with Streptozotocin-induced diabetes80 mg/kg b.w./day; 45 daysInsulin, SOD, catalase, GPx activity, glutathione-S-transferaseGlucose, lipid peroxidation, TBARS, H2O224Intraperitoneal[172]
Sprague–Dawley rats with Streptozotocin-induced diabetes15 and 30 mg/kg b.w./day; 2 weeksCreatinine clearance, SOD activity, catalase activityGlucose, creatinine, renal changes, oxidative stress, urine albumin, proteinuria, lipid peroxidation, MDAN/AIntraperitoneal[173]
Wistar-NIN rats with Streptozotocin-induced diabetes0.01% curcumin; 8 weeksSOD activity, pancreas catalase activityGlucose, insulin, TBARS, pancreas SOD activity, glutathione-S-transferase activity32Intraperitoneal[174]
Sprague–Dawley rats with Streptozotocin induced type 1 diabetes50 mg/kg b.w./day; 6 weeksAlbumin, acetyl-histone H3, phospho-histone H3Urea, creatinine, HSP-27 protein, p38 protein12Intraperitoneal[175]
C57/BL6J mice with Streptozotocin-induced diabetes7.5 mg/kg b.w./day; 10 h prior to STZInsulin, glucose clearance, GLUT2 mRNAGlucose, IL-16, TNF-α, pancreatic IL-6 N/AIntraperitoneal[176]
Wistar rats with Streptozotocin-induced diabetes80 mg/kg b.w./day; 45 dayInsulin, SOD activity, CAT activity, GPx activity, glutathione activityKidney and liver: morphological changes, oxidative stress, TBARS, HP30Intraperitoneal[177]
Swiss albino mice with Streptozotocin-induced diabetes10 mM; 10 µL/mouse i.p.; 28 days and 106 BMCs, a single injectionInsulin, islet regeneration, SOD activity, catalase activity, GPx activityGlucose, MDA levels40 Intraperitoneal[178]
Wistar rats with alloxan-induced diabetes0.08 mg/kg b.w./day; 21 daysHemoglobin, glutathione, GPx activityGlucose, HbA1c, TBARS, SDH activity36Oral[179]
Wistar rats with alloxan-induced diabetes0.1 mg/kg b.w.; 2 h GlucoseN/AOral[180]

3.5. Kidney Diseases

Acute kidney disease (AKD) and chronic kidney disease (CKD) have led to several cases of mortality worldwide. An increase in inflammation and decreased antioxidant activity are mostly seen in kidney diseases and in hemodialysis patients. The supplementation of curcumin has shown favorable effects on renal diseases, mainly due to its anti-inflammatory and anti-oxidant properties. Curcumin has decreased renal damage and inflammation by reducing the expression of inflammatory cytokines such as IL-1β, IL-6, TNF-α, adiponectin (which is associated with arterial stiffness, leading to death) and cystatin in rats with adenine-induced CKD [181].
Nuclear factor-erythroid-2-related factor 2 (Nrf2) is a crucial transcription factor, and in the case of oxidative stress, Nrf2 translocates into the nucleus and induces the production of detoxifying enzymes. The upregulation of transcription factor Nrf2 was seen in the kidney upon the administration of curcumin; this upregulation led to an increase in glutathione reductase and thereby exhibited the antioxidant property by decreasing glutathione levels [182]. Diabetic nephropathy (DN) is the cause of end-stage renal failure, and inflammation plays an important part in the development and progression of DN. Curcumin prevents inflammation by renal macrophage infiltration and modulates transcription factors such as AP-1 and chemokines such as IL-1,IL-6. The oral supplementation of curcumin at 100 mg/kg/day for 8 weeks in STZ-induced diabetic rats prevented macrophage infiltration by inhibiting the activity of NF-κB, IκBα and regulated MCP-1 at the nuclear level, thereby preventing glomerular injury and damage [183]. Curcumin analogues in diabetic rats were administered at 5 mg/kg/day for 6 weeks, causing a similar reduction in kidney inflammation via the inhibition of the JNK pathway and diabetes-related histone acetylation [184]. Chemotherapeutic agents such as cisplatin cause acute kidney injury. Curcumin prevents the mitochondrial bioenergetics alterations and redox balance by preventing the increase in the mitochondrial fission protein and decreasing NAD± dependent deacetylase sirtuin-3 in acute kidney injuries [185]. Heavy metals cause nephrotoxicity due to ROS overproduction, decrease the endogenous antioxidant property and suppress the autophagy flux (leading to cell damage). Curcumin modulates autophagy via the modulation of Akt/mTOR and by increasing the adenosine monophosphate-activated protein kinase (AMPK) and extracellular signal-dependent kinase (ERK) pathways [186]. Curcumin administered orally in wistar rats at 400 mg/kg/day (with AKI via a dose of potassium dichromate) could preserve mitochondrial bioenergetics by increasing the expression of mitochondrial transcription factor A and bring peroxisome proliferator-activated receptor gamma coactivator 1-alpha (PGC-1α) back to a normal level [186]. Therefore, curcumin can be potentially used to treat renal diseases.

3.6. Antioxidant

Oxidative stress results from an imbalance between oxidants and antioxidative measures. It is hypothesized that damage from reactive oxygen species (ROS) and reactive nitrogen species (RNS) results in many chronic diseases (atherosclerosis, Alzheimer’s disease, liver disease) and the senescence of cells [187,188,189,190]. Curcumin has potent antioxidant properties due the fact that it has multiple functional groups including the β-diketo group, carbon–carbon double bonds and phenyl rings containing varying amounts of hydroxyl and methoxy substituents. These properties allow curcumin to protect lipid membranes from peroxidation induced by oxidation agents [191]. In fact, one study has shown that curcumin was more effective as an antioxidant than α-tocopherol [192].
Curcumin has multiple pathways to act as a direct antioxidant. Firstly, curcumin acts as an ROS (specifically H2O2) scavenger, as shown in vitro by Ak et al. [193]. Secondly, curcumin, through its phenolic or central methylenic groups, is associated with its hydrogen donor capacity [194]. Whatever et al. proved that the enol form of curcumin is more stable than the diketo form and that the bond-dissociation energy (BDE) of the phenolic O:H bond is lower than the BDE of the central O:H bond. Therefore, the hydrogen ion abstraction takes place in the phenolic form [195,196]. Thirdly, curcumin degradation products (ferulic acid and vanillin) under basic pH can act as potent antioxidants [191,197]. Lastly, curcumin can chelate heavy metal ions such as ferrous ions through its functional carbonyl group [193].
Additionally, curcumin exhibits indirect effects that combat oxidative stress on the cells. High-dose curcumin administration in albino rats by Faten et al. has shown the increased activity of antioxidant enzymes such as superoxide dismutase, catalase, glutathione peroxidase and glutathione-S-transferase (GST) in different tissues [198]. Furthermore, curcumin increased the mRNA expression (by 2–12 times) and protein levels (by 2–6 times) of antioxidant enzymes including glutamyl-cysteine ligase, quinone oxidoreductase and heme oxygenase 1 (OH-1) in human islet cells [199,200]. The expression of HO-1 was induced by curcumin through the activation of the Nrf2/antioxidant-responsive element (ARE) pathway in rat kidney epithelial cells [201]. Curcumin also increased the expression of the heat shock protein HSP70 [202]. Several studies have shown that curcumin inhibits phase 1 enzymes and activates phase 2 enzymes, leading to reduced toxic metabolites and increased antioxidants effects [203,204,205,206]. Curcumin acts indirectly to reduce oxidative stress through the inhibition of inflammatory pathways through the inhibition of NF-κB, which will be discussed later in the article.
Paradoxically, curcumin can selectively induce oxidative stress in cancer cells, leading to apoptosis and autophagy [207,208,209]. This was further proven when N-acetyl cysteine or glutathione was added and the curcumin effect was nulled [209,210]. The etiology of the paradoxical action of curcumin is unclear, but one study points to the significantly higher intake of curcumin in cancer cells [211]. Multiple studies are leveraging curcumin in the treatment of different types of cancers.
Among the various benefits of curcumin, the regulation of ER (Endoplasmic Reticulum) stress by using curcumin is an important strategy in treating several diseases such as cancer [212], diabetes [213], osteroporosis [214] and neurodegenerative diseases [215]. ER stress is caused due to the accumulation of unfolded or misfolded proteins, leading to a stress response called unfolded protein response (UPR). Curcumin can regulate ER stress by causing cell apoptosis or cell survival based on the type of cell being examined. In normal cells, curcumin scavenges ROS and decreases UPR, thereby suppressing ER stress and inhibiting apoptosis. In the case of inflammatory diseases, curcumin activates the MAPK pathway and increases the proteins involved in apoptosis such as transcription factor 6, the glucose-regulating protein and the C/EBP homologous protein CHOP. In diabetes, ER stress has been shown to trigger beta cells dysfunction or cell death [216]. Curcumin suppresses NF-κB activity and reduces caspase-12 and caspase-3 levels (usually increased due to ER stress) [217]. In murine myelomonocytic leukemia cells, curcumin induced apoptosis by the generation of ROS, the cytosolic release of Ca2+ and the inducing of DNA damage [218,219]. In human lung carcinoma A-549 cells, curcumin prevented cell proliferation by inducing G2/M-phase arrest and increased p53 and p21 levels, which are hallmarks of ER stress [220]. Curcumin caused apoptosis via the activation of CHOP in human leukemia HL-60 cells [221]. The exact cellular mechanism underlying the effect of ER stress on cell death or cell survival still needs more evidence due to its dualistic response.

3.7. Gut Microbiota

The exceptionally complicated and abundant microbial community inhabits the GI tract, with 100 trillion bacteria which are, remarkably, 10–100 times greater than the number of eukaryotic cells [222]. Furthermore, the gut environment differs markedly between different anatomical regions regarding physiology, substrate availability, digesta flow rates, host secretions, oxygen tension and pH [223,224]. Aside from the poor systemic bioavailability of curcumin, it is expected to be found at high concentrations in the gastrointestinal tract after oral administration. Thus, it is suspected that curcumin could exert direct regulative effects on the gut microbiota, which could explain the paradox between curcumin’s poor systemic bioavailability and its widely reported pharmacological effects (Table 4).

3.8. Inflammatory Bowel Disease (IBD)

TNF blockers, immunosuppressants and anti-inflammatory medications are commonly used to treat IBD, but due to the insufficient results and high cost involved in the treatment, there has been a need for alternatives. Bioactives have antioxidant and anti-inflammatory activity that could be used to effectively treat or prevent IBD. The use of curcumin in preclinical studies has suggested that it can target various molecular and cellular pathways involved in IBD pathogenesis. Recent studies have shown that the various molecular signaling pathways that participate in IBD development are targeted by curcumin, including PPAR-gamma, P13K, TLR-4, Akt, mTOR, ERK5, AP1, TGF-β, PAK1, Wnt, β-catenin, Shh, Rac1, p38MAPK, EBPα, NLRP3 inflammasome, Nrf2, Notch-1, AMPK, STAT3 and MyD-88 [230,231] Autophagy suppression has been linked with an excessive inflammatory response in IBD. In this regard, curcumin has shown an autophagy-regulating property by reducing the expression of genes such as Beclin-1, autophagy-related gene 5 and LC3II. In addition, curcumin has shown anti-apoptotic activity by inhibiting apoptotic cell death, thereby preventing damage to the intestinal epithelial barrier [232]. Studies have shown that curcumin suppresses NF-κB in chondrocytes by reducing the expression of cyclooxygenase, prostaglandin E-2 and inflammatory cytokines [233]. Curcumin can also interact with transient potential vanilloid receptor 1 in inflamed tissues to prevent intestinal inflammation [234]. Curcumin analogues such as non-electrophilic curcumin are known to suppress colitis in mice by inhibiting pro-inflammatory signals. Numerous clinical studies have linked inflammatory diseases to NLRP3. We agree with Karthikeyan et al. (2018) that curcumin can be a potential NLRP3 inflammasome suppressant by in vivo studies, and this could be a promising treatment for IBD [235].
However, due to its iron reduction property, curcumin should be carefully used for treatment since poor iron absorption is already seen in IBD patients. Therefore, monitoring the erythroid parameters is essential. Although curcumin alone or in combination with other drugs could be used for treatment by optimising the dosage, rigorous randomized controlled and long-term clinical trials should be conducted to establish the role of curcumin in the treatment of IBD.

4. Anti-Microbial

The antimicrobial activity of curcumin dates back to the old days when it was used as an insect repellent in the house [236]. Later, it was introduced as a potential suppressor of microbial activity in the cotton and wool industries [237]. Curcumin and other antimicrobial compounds have been key ingredients in ointments for skin protection and wound-dressing properties [238]. Several studies have reported the broad-spectrum antimicrobial activity for curcumin, including antibacterial, antiviral, antifungal and antimalarial activities.

4.1. Antiviral

Antiviral drugs are in high demand due to increasing viral infections globally and the lack of preventive and therapeutic options [238]. The first known antiviral activity of curcumin dates back to the 1990s, with the discovery that curcumin inhibits HIV viral protease in vitro. Since then, several studies have been conducted to understand its mechanism of action on different types of viruses. Each stage of the viral replication cycle, such as attachment/penetration, genome replication, gene expression, assembly and release, has been an attractive target for the effective inhibitory activity of curcumin. During the attachment step, the uptake of viral particles by binding to the receptors on the host cell membrane surface and entry into the host cell takes place by receptor-mediated endocytosis [239]. As a result, Curcumin has shown effective activity:
  • Against the viral envelope proteins by (a) modulating the membrane lipid bilayer of the host [240], (b) inhibiting its entry by interacting with viral surface proteins and reducing viral particle production [241], (c) disrupting the integrity of the viral membranes [242].
  • By targeting replication in two ways: (a) targeting the viral replication machinery and (b) modulating cellular factors to interrupt the replication process [243,244].

4.1.1. Human Immunodeficiency Virus (HIV)

Curcumin has been shown to impact the HIV function at several stages of the virus lifecycle. Ferreira et al. conducted a study to understand the anti-inflammatory activity of curcumin in the female genital tract, which leads to the downregulation of tight junction (TJ) proteins, resulting in barrier loss and thereby allowing HIV-1 to traverse the genital epithelium and infect the host. The treatment of genital epithelial cells with 5 µM curcumin reduced the expression of virus replication marker p24 and protected the epithelial barrier by preventing TJ protein downregulation, thus reducing the HIV infection rate [245]. Curcumin can inhibit HIV replication by interacting with the viral integrase, protease and trans-activator of the transcription (Tat) protein. Docking studies have suggested that curcumin could bind effectively to the active site of HIV-1 protease [246]. Pretreatment with curcumin has inhibited the induction of proinflammatory cytokines such as Il-6, TNF and chemokines Il-8, IP-10, RANTES, eotaxin, MIP-1α (Macrophage Inflammatory Protein-1 Alpha) and MCP-1. In one study, curcumin degraded the Tat protein through a proteasomal pathway [247] and reduced its transactivation in HIV-1-infected cells. Even curcumin analogues, such as curcumin A (which lacks the β-diketone moiety of curcumin), have been tested against HIV-1 [248]. This study showed that curcumin A lowered late viral genome copy levels and could inhibit the early reverse transcription of the virus [248]. The therapeutic activity of curcumin is due to its ability to activate heme oxygenase-1, thereby inhibiting HIV-1 [249]. Curcumin-stabilized silver nanoparticles have shown promising activity by lowering HIV-LTR (Long Terminal Repeat) expression and lowering the expression of TNF-α, IL-6, IL-1β and NF-κB [250]. Collectively, these studies show curcumin’s potential against HIV-1.

4.1.2. Severe Acute Respiratory Syndrome Coronavirus 2 (SARS-CoV-2)

Wen et al. (2007) have shown that curcumin can inhibit SARS-CoV-1 replication in the cultures of Vero E6 cells (EC50 of > 10 µM) [251]. Docking studies have concluded that curcumin could bind to target receptors such as protease, spike glycoprotein-receptor binding domain and PD-ACE2 (Angiotension Converting Enzyme-2). The ability of curcumin to modulate a wide range of molecular targets that are responsible for the attachment and internalization of SARS-CoV-2 could be used to effectively manage the coronavirus infection. Furthermore, Curcumin could block the entry of viruses into the cell by altering surface protein structures in the virus. Adding to this, a molecular docking study indicated that curcumin could bind to ACE2 to inhibit COVID19 entry into the cell. Curcumin could also interact with the viral protease, such as the main protease, which could be a potential therapeutic target [252]. Due to growing evidence on the effect of curcumin on interferons in different viral diseases, curcumin could trigger innate immunity by stimulating the production of interferon-stimulating genes and cytokines, as seen in the study on a porcine epidemic diarrhea virus model. Curcumin has played an important role in reducing the expression of crucial chemokines and cytokines such as IFN-γ, MCP-1, IL-6 and IL-10 in lung infection [253] and against the human RSV, preventing viral replication, which could be used to treat pulmonary inflammation due to COVID-19 infection. Reduction in the ACE2 expression could decrease the risk of renal damage. In this regard, curcumin could upregulate ACE2, leading to improved renal blood flow [254]. Curcumin can be used as an effective anti-fibrotic agent in kidneys [143]. To sum up, the antiviral and anti-inflammatory activity of curcumin can be helpful in both preventing and treating COVID-19. Further in vitro studies could help us better understand the mechanism of action, if any exists.

4.1.3. Influenza A Virus (IAV)

Curcumin has been shown to inhibit NF-κB signaling, which is required for IAV replication. Curcumin or its analogues have been shown to inhibit IAV by preventing entry, inhibiting replication or preventing viral exit. A study by Dai et al. showed that curcumin interferes with early-stage virus gene expression and replication and inhibits several IAV-induced toll-like receptor signaling pathways including TLR2/4/7, MyD88, TRIF and TRAF6 [255,256]. Additionally, curcumin reduced IAV replication and lung injury in an in vivo animal model, which explains its role in combating infection and viral-induced disease [257]. Another study by Han et al. made a similar observation on mice infected with the IAV strain PR8 and fed 30 or 100 mg/kg of curcumin. Curcumin-treated mice had lower levels of MCP-1, IL-6 and TNF-α in bronchoalveolar lavage fluid and lung tissues as compared to untreated mice [257]. Curcumin analogues such as monoacetylcurcumin have inhibited plaque formation (IC50-0.2 µM). Although curcumin and MAC mildly reduce neuraminidase activity, they act via different mechanisms to inhibit IAV. The authors have suggested a combined use of the two for better activity [258]. A study by Lai et al. on MDCK cells treated with curcumin showed reduced mRNA levels of the IAV M gene in infected cells. Additionally, curcumin reduced lung pathology in in vivo treated mice [259].

4.1.4. Herpes Simplex Virus (HSV)

Curcumin inhibited plaque formation by 88% and blocked viral adsorption by 92% in HSV1- and HSV2-infected Vero cell lines at a concentration of 30 μM. Curcumin treatment at a 5 μM concentration in primary human GECs reduced HSV-2 replication 1000-fold compared to the control group, and 50 μM of curcumin showed 100% inhibition [245]. To enhance the bioavailability of curcumin, it was encapsulated by Poly-(Lactic-Co-Glycolic Acid) and delivered via an intravaginal route against HSV-2 infection in mice. The results showed that curcumin-PLGA had no effect on the mice’s survival following the low or lethal dose of HSV-2 [260].

4.1.5. Dengue Virus (DENV)

Curcumin reduced the plaque formation of all four strains (DENV-1–4) examined in LLC-MK2 cells, with limited toxicity (CC50 of 59.42 μM). Another study showed that curcumin inhibits DENV-2 by the indirect interaction with cellular systems rather than directly on viral function. A study by Balasubramanian et al. evaluated the anti-DENV (Dengue virus) properties of curcumin and other synthesized analogues. Curcumin and the analogues showed inhibitory activity on viral protease activity (IC50 36–66 μM) [261]. The MOA of curcumin was through cellular lipid metabolism, as it downregulated acetyl-CoA carboxylase and fatty acid synthase and lowered the lipid droplet formation, which is usually seen in a DENV infection. Mainly, actin filament disorganization and defects in polymerization were seen after the curcumin treatment. Therefore, curcumin shows anti-DENV activity by actin filament organization, cell lipogenesis and viral enzymes [261,262].

4.1.6. Enterovirus 71 (EV71)

Huang et al. evaluated the activity of curcumin against EV71 in HT29 human intestinal epithelial cells. A 10 μM concentration of curcumin reduced the protein expression during the early stage of infection and the genome replication of the virus and prevented EV-71-induced cell death. Usually, cells infected with EV71 show the phosphorylated residue Tyr311 of protein kinase C-delta [263], but it was reduced in curcumin-treated cells. Lin et al. evaluated curcumin-derived carbon quantum dot formulations (Cur-CQD) against EV71 [264]. This formulation increased the water solubility of curcumin due to a better antiviral activity. In addition, cur-CQD lowered the expression of viral proteins such as structural protein VP1 and non-structural proteins such as 3CDpro and 3Dpol in a dose-dependent manner. The treatment also reduced the amount of viral mRNA and proteins that were detected in the brain and limb muscle tissue [252].

4.2. Antifungal Activity

Curcumin has the potential to be used as an antifungal against a wide range of fungi in in vitro and in vivo studies including cryptococcus, candida, trichophyton and Paracoccidioides [265,266]. With the emerging antifungal resistance, candida and other fungi species, there is a need for novel antifungal agents [267,268,269]. In addition, traditional anti-fungal medications such as azoles and polyenes possess serious side effects, most commonly resulting in kidney damage. On the other hand, curcumin has displayed minimal toxicity in a few reports, but no long-term trials have been conducted to assess its safety [270,271,272]
The exact mechanism of curcumin is unknown, but evidence by Sharma et al. showed that curcumin affects candida by increasing the production of reactive oxygen species (ROS) through altering membrane ATPase activity, interfering with ergosterol synthesis and inducing apoptosis as a result of reactive oxygen species accumulation [273,274,275]. This was proved further by including an antioxidant that attenuated curcumin’s effects on the fungus. The effect of curcumin on fungal cells also extends to the inactivation of specific genes that affect growth and drug metabolism. Curcumin targets global suppressor thymidine uptake 1 (TUP1) in candida, leading to its transcription and inhibiting hyphae development. Curcumin restored the sensitivity to fluconazole, which might be due to its effect on the active transporters (ABC and MDR) of the drug [273,276]. Curcumin has phytochemical properties when combined with photodynamic therapy and can be genotoxic to many fungi (candida, aspergillus and dermatophyte) since it can prevent the repair process of DNA damage [277,278,279].
A study by Martinez et al. measured the minimal inhibitory concentration (MIC) of Curcumin against 23 human pathogenic strains of fungi in vitro. Although Curcumin was more potent in many strains of Paracoccidioides brasiliensis than fluconazole, the strain MG05 growth was inhibited at an MIC of 0.5 mg/L of curcumin compared to 16 mg/L of fluconazole. Curcumin exhibits the potential to be administered through multiple routes, including intravenous, topical and oral routes depending on the offending agent site of infection. One study isolated the samples of candida from HIV patients with oropharyngeal candidiasis and exposed them to curcumin, which inhibited 90% of the yeast [280]. A study conducted on a vulvovaginal yeast infection model in rats benefited from 1.0% curcumin local cream application [281].
Owning to the phytochemical properties of curcumin. Many studies have examined the effect of curcumin with light on candida biofilm growth and dermatophytes infection. For example, Brasch et al. found that curcumin plus visible light inhibited the growth of different dermatophytes [278]. In addition, an experiment conducted by Dovigo et al. showed that candida growth and biofilm formation were inactivated using curcumin with photodynamic therapy [277].
Curcumin possesses the potential to be used as a monotherapy or in combination with azoles or polyenes. Sharma et al. proved that, when used in combination with Amphotericin b, curcumin showed a synergistic effect and a reduced side effects profile [273]. This can be leveraged in the future to reduce the dosage and, in turn, the side effects of current anti-fungal medications.

4.3. Antibacterial

Several antibiotics are available against specific bacteria. However, due to the extensive use of drugs, it is challenging to eliminate pathogens from the human body due to developed resistance. So, it is important to naturally get rid of bacterial infections. Curcumin, a known spice, shows antibacterial activities against most gram-positive and gram-negative bacteria [282]. Curcumin is known to be a relatively unstable molecule, with a particle size of 500–800 nm, impairing cellular uptake and resulting in low bioavailability [283,284]. A study found that curcumin kills several pathogenic gram-positive bacteria such as Staphylococcus aureus, Staphylococcus epidermidis and Enterococcus, which are the main causative agent of skin diseases, pneumonia, meningitis and urinary tract infections in human beings [282]. In addition, curcumin suppresses the adherence of Streptococcus mutants to human tooth surfaces and the extracellular matrix protein [285]. Curcumin possesses a synergistic effect with important antibiotics such as cefixime, vancomycin and tetracycline against Staphylococcus aureus (S. aureus) [286,287,288]. However, very few studies have demonstrated the mechanism of the antibacterial activity of curcumin, which seems to differ depending on the strain being studied. For instance, studies have shown that the antibacterial activity of curcumin against Bacillus subtilis occurs through the inhibition of bacterial cell proliferation by blocking the assembly dynamics of FtsZ in the Z ring [289]. In the case of Pseudomonas aeruginosa (P. aeruginosa) infection, curcumin was shown to have anti-infective activity by affecting virulence, quorum sensing and biofilm initiation [236].
Moreover, these mechanisms have not been confirmed in the case of other bacterial genera and, hence, could not be generalized for all bacteria. Therefore, a detailed study on the antibacterial mechanism of curcumin, including a large number of bacteria from different genera, is required. Furthermore, due to the increase of resistance in Gram-positive and Gram-negative bacteria, there is an urgent need to identify and assess alternative antimicrobials, including those from plant materials with low human cytotoxicity. Curcumin I showed no toxic effect on human health, even when taken at doses as high as 8 g per day [290].

5. Clinical Trials with Curcumin

Numerous clinical trials have been conducted with curcumin, appraising its therapeutic and pharmacological benefits across different patient populations. A summary of the concluded trials is depicted in Table 5, which will allow the reader to get a quick insight into the disease/patient population of interest. In this study, we have only considered registered and completed trials in the Clinical Trials registry. Of all the clinical trials, trial number NCT00927485 has studied the role of curcumin on intestinal adenomas for a significant duration of over 5 years, whereas others have done so for only a limited period of time. Therefore, the obtained results should be taken with a grain of salt. Case in point, in trial number NCT04012424, the trial was conducted for only a period of two days. This indicates that further trials extending over longer periods are required. Another shortcoming that is observed in most of the trials is the low number of participants. For example, in trial number NCT03568513, the effect of curcumin on IBS was only studied in 4 people out of the 50 that were expected to enroll.
Additionally, the trial has been conducted at a single center and does not provide enough information about the physiological and genetic makeup of the participants, which has been shown to affect the intestinal microbial milieu. Thus, the trial from this and similar trials require further investigation and validation. In addition to the above, most of the trials available in the database have not posted the obtained results; this makes it difficult to conclude if their primary outcome was achieved. The open-label study of the curcumin CS complex in schizophrenia (trial number NCT01875822) (Sl No.18) concluded in 2012; no results are available in the study page or in the literature. Although curcumin exhibits beneficial pharmacological effects in cell and animal models, the results are not very well replicated in human subjects. This has drawn considerable skepticism, and curcumin has been labelled as a pan-assay interference (PAINS) compound in the case of different screening tests such as fluorescence interference, the covalent labelling of proteins, redox reactivity, etc. It is to be noted that these tests are limited to in vitro studies; the real results obtained from human trials and case reports are more valid than any theoretical warning to prove its activity. Additionally, there has not been any experiment to prove that the biological activity of curcumin is due to its unique structure. A paper [291] suggested that curcumin is a “bimolecular sensitive fluorescent probe.” This does not necessarily have to be related to a fluorescence interfering property. We also know that any molecule with an ability to interact with various targets could bring numerous side effects. When it comes to several in vivo studies, curcumin proved to be safe even at a very high dose. A clinical trial in healthy volunteers consuming 500 to 12,000 mg of curcumin showed no toxicity, but a very low serum availability was detected in 2 of the 6 patients who received the highest dose (10,000 mg and 12,000 mg) [292]. This could possibly be due to the genetic modifiers of curcumin metabolism or even to the preparation method of commercially obtained curcumin. The activity of curcumin has also been related to its metabolites, which are more of an advantage, as this could possibly be used to treat diseases with multiple causes such as cancer and diabetes. A letter to the editor by Burgos-Moron et al. entitled “The dark side of curcumin” [293] suggests that the cytotoxicity of curcumin and its ability to intercalate into DNA have been nullified by an experiment conducted by Kurien et al., stating that the cytotoxicity was not due to curcumin but due to the solvent used for the dissolution of curcumin (i.e., ethanol) [294]. The best possible explanation for the ineffectiveness of curcumin in certain studies could be the very low bioavailability or moderate biological activity of curcumin. The non-replication of activity is related to curcumin’s low bioavailability due to the high hydrophobicity caused by the cyclic rings in its structure. Studies have shown that the combination with piperine enhances the serum concentration, the extent of absorption and the bioavailability of curcumin. However, few clinical trials have employed this strategy. In this regard, further investigation is required.
Table 5. Effect of curcumin on the completed clinical trials.
Table 5. Effect of curcumin on the completed clinical trials.
Sl No.Clincal Trial IdentifierTrial TitleNo. of ParticipantsInclusion CriteriaYear of CompletionPrimary OutcomeClinical Trial No.Follow-Up
Period
1.NCT03085680Curcumin and Function in Older Adults21Aged above 65 years with a CRP level greater than 1.0 mg/dL2020To examine the effects of dietary supplementation with curcumin on changes in physical function, walking speed (400 m walk test) and grip strength290 days
2.NCT03211104Comparison of Duration of Treatment Interruption with or without Curcumin During the Off Treatment Periods in Patients with Prostate Cancer Undergoing Intermittent Androgen Deprivation Therapy107Patients with localized prostate cancer or metastatic prostate cancer at the time of diagnosis who received intermittent androgen deprivation therapy (IAD)2015To determine whether the period from the first interruption of the androgen deprivation therapy to the time when androgen deprivation therapy needs to be retreated differs between the curcumin group and placebo groupNA180 days
3.NCT04012424The Effect of Premedication with Curcumin on Post-endodontic Pain44Patients in the age range of 20–55 years with acute pulpitis2020Change in postoperative pain after a single endodontic visitN/A2 days
4.NCT04870060Ability of Curcumin to Decrease Cytokines Involved in Mucositis in the Autologous Transplant40Patients aged 18 years and above with a creatinine clearance greater than 50 mL/min and a serum bilirubin level greater than 2 mg/dl2015To calculate TNFa, IL-1, IL-6, IL-8, IL-17, TGF-B, IFN-gamma and E2 levels228 days
5.NCT01543386Effects of Curcumin on Vascular Reactivity2150- to 70-year-old smokers2012Changes in brachial flow-mediated dilatation25 days
6.NCT03568513Effect of Curcumin on Gut Microbiota in IBS4Patients aged 10 to 18 years with diarrhoea-predominant IBS2020Alterations in gut microbiotaN/A56 days
7NCT03864783The Effect of Curcumin on Liver Fat Content in Obese Subjects39BMI and haemoglobin greater than 30.0 kg/m2
and 7.5 mmol/L, respectively
2020Curcumin’s effect on steatosisN/A42 days
8.NCT04044417Curcumin-Simvastatin-EDTA in the Treatment of Periodontitis30Patients aged 25 to 50 years suffering from at least a single posterior 2–3 wall periodontal pocket of depth 2018Reduction in probing depth4180 days
9.NCT04032132Curcumin Paste as an Adjunctive Therapy in Periodontitis24Patients aged 25 to 45 years with at least a single posterior 2–3 wall periodontal defect of pocket depth2018Evaluate the influence of curcumin paste on the clinical outcomes of the surgical treatment4180 days
10.NCT03746158Interindividual Variation in Excretion of Curcumin818–30-year-old healthy adults.2019Determine the concentration of curcumin and its metabolites in human fecal samplesN/A28 days
11.NCT01179256Effect of Supplemental Oral Curcumin in Patients with Atopic Asthma16Patients aged 18–60 years on low- or medium-dose inhaled corticosteroids2010Improvement in post-bronchodilator FEV1N/AN/A
12.NCT01246973Oral Curcumin for Radiation Dermatitis in Breast Cancer Patients686Females aged 21–120 years2015To measure the Mean Radiation Dermatitis Severity Score242 days
13.NCT04119752Effect of Curcumin on Microvascular Response and Tissue Oxygenation in Older People28Aged 60– 85 years with two or more risk factors for cardiovascular disease2020Changes in microvascular reactivity and tissue oxygen saturation.N/A120 min
14.NCT02255370Curcumin Associated with Thiopurin in the Prevention of Post-op Recurrence in Crohn Disease (POPCUR)61Patients aged 18 years and older with Crohn’s disease 2018Rutgeerts endoscopic score3180 days
15.NCT02298985Curcumin Addition to Antipsychotic Treatment in Chronic Schizophrenia Patients38Patients aged 18–60 years with schizophrenia and a SANS greater than 30 points2017Positive and Negative Symptoms Scale (PANSS)4180 days
16.NCT0138316118-Month Study of Memory Effects of Curcumin46Aged 50–90 years with a modified Ischemic score of less than 42017Change from the baseline to 18 months on the Brief Visual Memory Test-Revised2540 days
17.NCT01333917Curcumin Biomarkers40Healthy volunteers aged 40–80 years2013To understand the changes in gene expression, the ribonucleic acid (RNA) level and apoptosis130 days
18.NCT01875822Open-label Study of Curcumin C-3 Complex in Schizophrenia17Patients aged 18–65 years with DSMIV schizophrenia and a SANS greater than 302012To understand the change from the baseline negative symptoms: alogia, anhedonia, social withdrawal and lack of motivation2112 days
19.NCT02978339A Study Evaluating the Safety and Efficacy of Curcumin in Patients with Primary Sclerosing Cholangitis (PSC)15Diagnosed with primary sclerosing cholangitis with alkaline phosphatase >1.5×2019Change in Serum Alkaline Phosphatase (SAP)284 days
20.NCT04208334The Effect of Curcumin for Treatment of Cancer Anorexia-Cachexia Syndrome in Patients with Stage III-IV of Head and Neck Cancer (CurChexia)20Patients with stage 3–4 head and neck cancer 2021To measure muscle mass260 days
21.NCT01925287Oral Bioavailability of Curcumin from Micronized Powder and Liquid Micelles in Healthy Young Women and Men23Healthy volunteers with a normal range blood chemistry value2013To determine total curcumin, demethoxycurcumin and bisdemethoxycurcumin after deconjugation with beta-glucuronidase124 h
22.NCT02104752Curcumin as a Novel Treatment to Improve Cognitive Dysfunction in Schizophrenia39Volunteers diagnosed with DSM-5 schizophrenia with a corrected vision of at least 20/302017Measurement and treatment research to improve cognition in schizophrenia156 days
23.NCT02369549Micro-Particle Curcumin for the Treatment of Chronic Kidney Disease518Patients with an eGFR between 15 and 60 mL/min/1.73 m2 with a minimum of 300 mg of protein in urine or with a albumin/creatinine ratio of at least 300 mg2020Change in albuminuria and the Estimated Glomerular Filtration Rate (eGFR)3180 days
24.NCT02439385Avastin/FOLFIRI in Combination with Curcumin in Colorectal Cancer Patients with Unresectable Metastasis50Colon or rectal cancer patients aged above 19 years with an ASA score of less than 3 2019To evaluate progression-free survival in colorectal cancer patients2730 days
25.NCT02474953A Study to Compare the Pharmacokinetic Profile of a Proprietary Curcumin Formulation to a Comparator Curcumin Product (15PCHB)12Volunteers aged 18–45 years with a BMI that is 18–29.9 kg/m2(±1 kg/m2)2015To measure the maximum concentration of curcumin and time until the max concentration of curcumin148 h
26.NCT04421716Testing the Bioavailability of Phytonutrients, Curcumin and Ursolic Acid18Men aged 18 years or older2021To evaluate the number, frequency, duration and relation of toxicity events to CURC and UA, the peak serum concentration, the half-life and the time taken to reach the maximum concentration114 days
27.NCT04258501Exploratory Study of Efficacy on Selected Natural Extracts Reducing Post Prandial Blood Glucose Response7220–50-year-old healthy individuals with a normal BMI2012Change in post-prandial blood glucoseNA2 h
28.NCT01035580Trial on Safety and Pharmacokinetics of Intravaginal Curcumin13Volunteers aged 18–45 years currently using a birth control method2012To reach the maximum selected dose or maximum tolerated dose of intravaginal curcumin without a dose-limiting toxicity114 days
29.NCT01403545Evaluation of Liposomal Curcumin in Healthy Volunteers50Volunteers in the age group of 18–45 years with a BMI between 18–27 kg/m22012Safety and tolerability of increasing doses of intravenous liposomal curcumin17 days
30.NCT01225094Curcumin to Prevent Complications After Elective Abdominal Aortic Aneurysm (AAA) Repair606Volunteers aged 18 years or above who have undergone the repair of AAA 2016To measure urine IL-18, NT-ProBNP, hsCRP and serum creatinine 2N/A
31.NCT01160302Curcumin Biomarker Trial in Head and Neck Cancer33Volunteers aged between 18–90 years willing to undergo tumor biopsies 2016Change in tissue biomarkers and pharmacokinetics of microgranular curcumin128 days
32.NCT01917890Radiosensitizing and Radioprotective Effects of Curcumin in Prostate Cancer40Aged between 50–80 years with relapsed or treated basal skin cancer and no severe hypertension 2013Biochemical or clinical progression-free survivalN/A365 days
33.NCT00895167The Effects of Oral Curcumin on Heme Oxygenase-1 (HO-1) in Healthy Male Subjects (CUMAHS)12Aged between 18–45 years with a BMI between 18 and 28 kg/m22009The maximal HO-1 mRNA expression and HO-1 protein level in PBMCs148 h
34.NCT03542240Effects of Curcumin Supplementation on Gut Barrier Function in Patients with Metabolic Syndrome15Waist Circumference—Female: ≥ 88 cm, Male: ≥ 102 cm B. Blood Pressure: ≥ 130/85 mm/Hg. Impaired fasting glucose or HbA1c fasting glucose ≥ 100 mg/dL or HgA1c ≥ 5.7 D. HDL-C—Females: < 50 mg/dL, Males: < 40 mg/dL E. Triglycerides ≥ 150 mg/dL 2020Change in intestinal permeability and intestinal barrier functionN/A365 days
35.NCT00927485Use of Curcumin for Treatment of Intestinal Adenomas in Familial Adenomatous Polyposis (FAP)4421–85 years with FAP (with an intact colon or who have had surgery)2016To determine the number of polyps and the size of polyps 5 years
36.NCT01042938Curcumin for the Prevention of Radiation-induced Dermatitis in Breast Cancer Patients35Females aged 21 years or above with a diagnosis of non-inflammatory breast adenocarcinoma2011Severity of dermatitis in the radiation treatment site in breast cancer patients249 days
37.NCT01490996Combining Curcumin with FOLFOX Chemotherapy in Patients with Inoperable Colorectal Cancer (CUFOX)4118 years or above, diagnosed with metastatic colorectal cancer and with an ECOG status of 0 or 12017Completion of dose escalation over two cycles of therapy2365 days
38.NCT01975363Pilot Study of Curcumin for Women with Obesity and High Risk for Breast Cancer29Females with an increased risk of breast cancer and a BMI between 25–402016Determine the adherence, tolerability and safety of two doses of nanoemulsion curcuminN/A90 days
39.NCT01859858Effect of Curcumin on Dose Limiting Toxicity and Pharmacokinetics of Irinotecan in Patients with Solid Tumors23Aged above 19 years with adequate bone marrow, renal and hepatic function and an ECOG status of 0 or 12016Maximum tolerated dose, pharmacokinetics of irinotecan and SN-38128 days
40.NCT04103788Evaluation of Increased Absorption of a Curcumin Emulsion (CurQ+) in Healthy Volunteers10Aged between 21 and 75 years2018Comparative effect of differing serum sample preparation methodologies on curcumin absorption levelsN/A6 h
41.NCT01925547Micellar Curcumin and Metabolic Syndrome Biomarkers42Total cholesterol > 5.2 mmol/L, LDL cholesterol > 3.4 mmol/L, Triglyceride > 2.26 mmol/L, CRP > 2 mg/L2014To measure the serum CRP level242 days
42.NCT01330810Curcumin Pharmacokinetics12Aged between 16 and 65 years with a BMI in the range of 18–30 kg/m22012To measure the AUC, Cmax, Tmax, Ke, T1/2, Vd and bioequivalence of tissue curcumin concentration148 h
43.NCT02908152Curcumin Supplement in Nonalcoholic Fatty Liver Patients50Patients diagnosed with type 2 diabetes with a CAP score greater than 2632017To measure hepatic steatosis272 days
44.NCT01201694Phase I Study of Surface-Controlled Water Soluble Curcumin (THERACURMIN CR-011L)28Patients aged 13 or older with an ECOG status of 3 or better and normal organ and marrow function2014To measure the Maximum Tolerated Dose (MTD) of surface-controlled water-dispersible curcumin128 days
45.NCT04028739Theracurmin vs. Curcumin Bioavailability Study24Healthy adults aged 19–60 years with a BMI of 18–30 kg/m22019To compare the bioavailability of curcumin in healthy adultsNA12 h
46.NCT03795792Oral Curcumin Administration to Remit Metabolic Syndrome105Men and women aged 20–55 years old with metabolic syndrome according to the ATP III criteria2019Remission of metabolic syndrome (≤2 components according to the ATP III criteria)NA3 months
47.NCT00528151A Randomized, Double-blind, Placebo-controlled Trial of Curcumin in Leber’s Hereditary Optic Neuropathy (LHON)70Aged 8 years or older with Leber’s hereditary optic neuropathy2007Visual outcome31 year
48NCT00889161Curcumin in Pediatric Inflammatory Bowel Disease118–18-year-old patients with IBD who have been on IBD medication for 3 months2010To determine the tolerability of curcumin in pediatric patients with inflammatory bowel disease19 weeks
49NCT01514266Effect of Curcumin on Lung Inflammation57≥45-year-old patients with COPD and a stable clinical course2010Change in sputum dysplasiaNA3 months
50NCT00779493Curcumin (Tumeric) in the Treatment of Irritable Bowel Syndrome: A Randomized-Controlled Trial (CuTIBS)17≥18-year-old patients who conform to the Rome III criteria2009The primary outcome will be defined as at least a 50% reduction in the irritable bowel severity score (IBSS)46 months
51NCT03329781Modulation of Endotoxaemia Via Curcumin Intake in Healthy Overweight Adults (ENDOCUR)1618–45-year-old healthy individuals with a BMI ≥ 25 kg/m22018Level of endotoxin in plasmaNA21 days
52NCT00094445Trial of Curcumin in Advanced Pancreatic Cancer50≥45-year--old patients with unresectable adenocarcinoma of the pancreas20146-month participant survival26 months
53NCT01750359Efficacy and Safety Curcumin in Depression4020–60-year-old patients with a major depressive disorder2011Change in Hamilton Depression Rating Scale and Montgomery–Asberg Depression Rating Scale 46 weeks
54NCT00181662Pharmacokinetics of Curcumin in Healthy Volunteers6≥45year-old healthy female individuals2007Curcumin pharmacologyNANA
55NCT03598205Curcumin and Intravitreal Dexamethasone in Diabetic Macular Edema (DIABEC)7218–90-year-old patients with significant diabetic macular edema and a central retinal thickness of >300 microns2019Mean difference in central retinal thickness from baseline to 6 monthsNA6 months
56NCT00641147Curcumin in Treating Patients with Familial Adenomatous Polyposis4418–85-year-old patients with familial adenomatous polyposis2016The average number of polyps in the placebo arm at the end of the study is compared to the average in the curcumin arm212 months
57NCT04385979Curcumin and Nanocurcumin in Oral Aphthous Ulcer48Patients with minor and recurrent aphthous ulcers with 48 h2020Wound size and pain scoreNA1 week
58NCT01320436Curcumin + aminosalicylic Acid (5ASA) Versus 5ASA Alone in the Treatment of Mild to Moderate Ulcerative Colitis5018–70-year-old patients with confirmed diagnosis of ulcerative colitis on a stable dose of ulcerative colitis medication2014The percentage of patients who achieve clinical remission compared between the two study arms34 weeks
59NCT03072992“Curcumin” in Combination with Chemotherapy in Advanced Breast Cancer15018–75-year-old female patients diagnosed with breast carcinoma and adequate organ function2019Objective response rate, assessed with the Modified Response Evaluation Criteria in Solid Tumours (RECIST)224 weeks
60NCT00113841Curcumin (Diferuloylmethane Derivative) With or Without Bioperine in Patients with Multiple Myeloma42Patients with multiple myeloma and adequate organ function2009Percent change of NF-kB protein expression in peripheral blood mononuclear cellsNA4 weeks
61NCT01909037Exploratory non comparative Study to Evaluate the Efficacy of Highly Bioavailable Curcumin (Flexofytol) in Patients with Knee Osteoarthritis2245–80-year-old patients with osteoarthritis and a symptomatic knee for more than 6 months who can avoid using analgesics during the study2012Change in the serum levels of biomarkers of cartilage metabolism and inflammation184 days
62NCT00365209Phase II A Trial of Curcumin Among Patients with Prevalent Subclinical Neoplastic Lesions (Aberrant Crypt Foci)44≥40-year-old patients with a >3 pack-year smoking history2011Change in prostaglandin E2 (PGE2) values found in rectal aberrant crypt foci (ACF) tissue230 days
63NCT02494141Curcumin Therapy to Treat Vascular Dysfunction in Children and Young Adults With ADPKD686–25-year-old patients with an ADPKD diagnosis and normal renal function2021Change in brachial artery flow-mediated dilation (FMD-BA) and aortic pulse-wave velocity (aPWV)412 months
64NCT04378972Anti-inflammatory Effect of Curcumin, Homotaurine, Vitamin D3 on Human Vitreous in Patients with Diabetic Retinopathy25≥18-year-old patients with diabetic retinopathy requiring vitrectomy2019Analyze human vitreous samples’ pro-inflammatory cytokinesNA7 days
65NCT04972045Bioavailability of Curcumin Capsules in Healthy Adult Subjects1218–55-year-old healthy subjects with a BMI of 18–28 kg/m22021Measure Peak Plasma Concentration, area under the curve, Tmax and bioavailability13 days
66NCT01489592Effect of Curcumin on Iron Metabolism in Healthy Volunteer (CURHEP)1818–35-year-old healthy adults with a BMI of 18–25 and no HFE mutation2012Maximal variation of the serum hepcidin level after the oral administration of curcumin248 h
67NCT01964846Effect of Antioxidant Intake on Cardiovascular Risk2245–70-year-old healthy patients with a stable weight2015Change in the blood levels of anti- and pro-inflammatory markersNA2 weeks
68NCT02100423Curcumin and Cholecalciferol in Treating Patients with Previously Untreated Stage 0-II Chronic Lymphocytic Leukemia or Small Lymphocytic Lymphoma35≥18-year-old patients with a CLL or SLL diagnosis and adequate organ function2018Overall response rate (biologic response rate + complete response [CR] + partial response [PR]) based on NCI-WG (for CLL) and the Cheson criteria (for SLL22 years
69NCT03530436Comparison of Curcumin Bioavailability1218–35-year-old healthy individuals2018Pharmacokinetics of curcuminoids (curcumin, demethoxycurcumin, bisdemethoxycurcumin) at different time framesNA24 h
70NCT02529982Curcumin Supplementation and Patients with Type 2 Diabetes4444–65-year-old patients with type 2 Diabetes Mellitus with a BMI of 18.5–30 kg/m22016Fasting blood sugar, insulin, HbA1c, homeostatic model assessment of insulin resistance and change in pancreatic B-cell functionNA10 weeks
71NCT03066791Turmeric and Curcumin on Sebum Production3018–50-year-old healthy individuals2017Sebum productionNA8 weeks
72NCT01514370Dietary Supplement of Curcumin in Subjects with Active Relapsing Multiple Sclerosis Treated With Subcutaneous Interferon Beta 1a (CONTAIN)8018–60-year-old patients with multiple sclerosis under the treatment of IFN beta-1a for 6–12 months2016Number of subjects with active (new or enlarging) T2 lesions, as assessed by magnetic resonance imaging (MRI) at Month 12224 months
73NCT00475683Curcumin for Prevention of Oral Mucositis in Children Chemotherapy85–30-year-old patients diagnosed with cancer who received doxorubicin containing chemotherapy2010Measured change of an objective measurement of oral mucositis36 weeks
74NCT00164749A Pilot Study of Curcumin and Ginkgo for Treating Alzheimer’s Disease36≥50-year-old patients of Chinese ethnicity with a progressive decline in memory ≥6 months 2006Measured change in the isoprostane level in plasma and the A-beta level in serum26 months
75NCT02152475Photodynamic Therapy (PDT) for Oral Disinfection3020–35-year-old healthy adults who do not perform any oral hygiene2013Microbiological analysis by the total number of colony-forming units12 h
76NCT01831193Effect of Oral Supplementation with Curcumin (Turmeric) in Patients with Proteinuric Chronic Kidney Disease12018–70-year-old patients diagnosed with proteinuric chronic kidney disease taking ARB or ACEi2014Change in proteinuria38 weeks
77NCT02556632Prophylactic Topical Agents in Reducing Radiation-Induced Dermatitis in Patients with Non-inflammatory Breast Cancer (Curcumin-II)191≥21-year-old patients diagnosed with non-inflammatory breast cancer or carcinoma in situ who are undergoing radiation therapy2016Measured mean Radiation Dermatitis Severity (RDS) score, incidence of moist sesquamation and change in the severity of skin reactions using RDS21 week post-radiation chemotherapy
78NCT04465851Effect of Ferrous iROn and cUrcumin sTatus on Inflammatory and Neurotrophic markErs (Fe-ROUTINE)15518–40-year-old healthy individuals2020To assess the influence of curcumin administration on ferrous iron supplementation-associated inflammationNA42 days
79NCT00192842Gemcitabine With Curcumin for Pancreatic Cancer17≥18-year-old patients suffering from advanced or metastatic pancreatic adenocarcinoma with no prior therapy2010time to tumor progression2NA
80NCT00099710Curcumin in Patients with Mild to Moderate Alzheimer’s Disease33≥50-year-olds with a diagnosis of Alzheimer’s disease2007Measured safety and tolerability of curcumin212 months
81NCT01712542Curcumin Bioavailability in Glioblastoma Patients15≥18-year-old patients with glioblastoma2013Measured concentration of curcumin in glioblastomaNAAt time of tumor resection
82NCT01022632Effect of Curcumin as Nutraceutical in Patients of Depression6018–65-year-old patients with a diagnosis of depression2010Measured response and mean change in the Hamilton Depression Rating Scale (HAM-D17) NA6 weeks
83NCT03144882Evaluation of Curcumin’s Effect on Inflammation in Hemodialysis Patients71≥18-year-old clinically stable patients receiving hemodialysis2017Measured mean Interleukin-6 levelsNA1 year
84NCT03141918Effect of Supplementation of Bioactive Compounds on the Energy Metabolism of People Living With HIV/AIDS2018–70-year-old patients with HIV receiving antiretroviral therapy ≥6 months2017Measuring the oxidation of energetic substrates; evaluation at restNA10 days
85NCT01740323Phase II Study of Curcumin vs. Placebo for Chemotherapy-Treated Breast Cancer Patients Undergoing Radiotherapy30≥18-year-old female patients undergoing breast radiotherapy2018Measured change in NF-kB DNA binding, Plasma TNF-alpha, sTNFR2, IL-1ra, IL-6 and CRP26 weeks after the completion of radiotherapy
86NCT04107987Berberine, Curcumin, Inositol, Banaba and Chromium Picolinate in Patients with Fasting Dysglycemia14818–75-year-old patients with impaired fasting glucose who are not on treatment2019Measured progression of dysglycemia33 months
87NCT00027495Curcumin for the Prevention of Colon CancerNA≥18-year-old healthy individuals2007To determine the pharmacokinetics and measure the Maximum Tolerated Dose (MTD) 172 h
88NCT04723849Efficacy Evaluation of a Mixed Compound of Antioxidants in Terms of Endothelium Damage/Function in Pediatric Subjects with Obesity. (OBELIX)486–17-year-old patients with a BMI > 95% for their age based on the CDC standard2020To test the effects of a mixed compound including curcumin on endothelium in a cohort of pediatric subjects with obesityNA6 months
89NCT00768118A Nutritional Supplement Capsule Containing Curcumin, Green Tea Extract, Polygonum Cuspidatum Extract, and Soybean Extract in Healthy Participants11≥18-year-old healthy individuals2008Measure the magnitude of change in the blood lymphocyte NF-kB levelNA15 days
90NCT02017353Effect of Curcumin Addition to Standard Treatment on Tumour-induced Inflammation in Endometrial Carcinoma7≥18-year-old female patients with endometrial carcinoma and no life-threatening metastases2016Measured change in the inflammatory markers in peripheral blood from the baseline221 days
91NCT00792818The Efficacy and Safety of Curcuma Domestica Extracts and Ibuprofen in Knee Osteoarthritis36750–75-year-old patients diagnosed with primary osteoarthritis2012Measured change in mean Western Ontario and McMaster Universities Osteoarthritis (WOMAC) pain subscale312 months
92NCT03290417Correlative Analysis of the Genomics of Vitamin D and Omega-3 Fatty Acid Intake in Prostate Cancer37Patients diagnosed with prostate cancer who are on active surveillance2019Measured gene expression of very low and low-risk prostate cancer patients on active surveillanceNA12 months
93NCT00525421A Clinical Study of Curcuminoids in the Treatment of Oral Lichen Planus20≥21-year-old patients diagnosed with lichen planus2009Measured percent change from the baseline to two weeks in the symptoms and signs of oral lichen planus22 weeks
94NCT02337192Antimicrobial Photodynamic Therapy Applied in Orthodontic Patients.2418–50-year-old healthy individuals with fixed orthodontic treatment 2014Microbiological analysis by the total number of colony-forming units (CFU)11 h
95NCT01288859Physiological Effects of New Polyphenol-enriched Foods in Humans1018–45-year-old healthy individuals2011Measured serum polyphenol concentrations, urinary excretion of total polyphenols and the number of total fecal polyphenolsNA24 h
96NCT01029327Effects of Curcumin on Postprandial Blood Glucose, and Insulin in Healthy Subjects15≥18-year-old healthy individuals2009To study the effect of curcumin on the postprandial blood glucose and plasma concentrations of insulinNANA
97NCT02815475Turmeric Anti-Inflammatory and Cell-Damage Trial (TACT)9018–80-year-old healthy individuals2016Measured change from baseline DNA methylation analyses and baseline oxidative stress determinationNA6 weeks
98NCT03769857NEM® + BIOCURC® Versus Placebo in Exercise-induced Joint Pain, Stiffness, & Cartilage Turnover in Healthy Men & Women8440–75-year-old healthy adults with no diagnosis of joint arthritis2019Measured exercise-induced cartilage turnoverNA2 weeks
99NCT03621865A Comparative Pharmacokinetic Study to Evaluate the Ability of a New Formulation to Enhance Curcuminoids Bioavailability (TURBIO)3018–45-year-old healthy individuals with a normal BMI and a stable weight2018Measured dose-normalized AUC of total curcuminoids plasmatic concentrationNA24 h
100NCT03289832Effect of Orally Delivered Phytochemicals on Aging and Inflammation in the Skin2518–70-year-old healthy individuals willing to avoid sun exposure and follow a diet2019Measured change in erythema 1, 2 and 3 Days after UV exposureNA10 days
101NCT03140657The Effects of Nanocurcumin on Treg Cells and Th17 Cells Responses in Ankylosing Spondylitis Patients2423–46-year-old patients with a diagnosis of ankylosing spondylitis2018Assessments of ankylosing spondylitis signs and symptoms (BASDI)24 months
102NCT03192059Study of Pembrolizumab, Radiation and Immune Modulatory Cocktail in Cervical/Uterine Cancer (PRIMMO)43≥18-year-old female patients with endometrial, cervical or uterine malignancy refractory to treatment2021Measured efficacy (objective response rate) at week 26 according to the immune-related response criteria (irRC)2156 weeks
103NCT03530787Cosmetic Effects of Topical Acetyl Zingerone3130–60-year-old healthy individuals2018Measured change in wrinkle appearance and skin pigmentationNA8 weeks
104NCT03493997Multicentre International STudy for the Prevention with Ialuril® of Radio-induced Cystitis (MISTIC)100≥18-year-old male patients who planned to receive primary therapy for prostate cancer2018Measured rate of patients who stopped treatment with intravesical or oral ialuril due to intolerance or adverse events212 months
105NCT04849182Vertistop® D and Vertistop® L in Preventing Recurrence of High-recurrence BPPV12818–85-year-old patients with benign paroxysmal positional vertigo (BPPV)2020Measured number of BPPV recurrences in patients supplemented with Vertistop DNA6 months
106NCT02099890The Effect of Diet on Chronic Inflammation and Related Disorders Following Spinal Cord Injury20≥18-year-old patients with a spinal cord injury2015Measured change from the baseline in the nerve conduction velocity of somatic nerves at 3 and 6 months36 months
107NCT03483376aPDT for the Remediation of Dental Black Stain30≥12-year-old patients with a dental black stain in at least two teeth2020Area and depth of color of the black stainNA6 months
108NCT00235625Curcuminoids for the Treatment of Chronic Psoriasis Vulgaris1218–75-year-old patients with chronic plaque-type psoriasis2007Physicians Global Assessment (PGA) of change216 weeks
109NCT04382040A Phase II, Controlled Clinical Study Designed to Evaluate the Effect of ArtemiC in Patients Diagnosed With COVID-1950≥18-year-old patients with a diagnosis of SARS-CoV-2 infection who are hospitalized and are in stable condition2020Time to clinical improvement, defined as a national Early Warning Score 2 (NEWS2) of ≤ 2, maintained for 24 h, and measurement of adverse events22 weeks
110NCT03150966The Immunomodulatory Effects of Oral Nanocurcumin in Multiple Sclerosis Patients4118–65-year-old patients who are diagnosed with multiple sclerosis2017Measurement of the Expanded Disability Status Scale (EDSS)26 months
111NCT02442453Effect of Scaling and Root Planing Along with Topical Application of Commercially Available Curcuma Longa Gel on Superoxide Dismutase and Malondialdehyde Levels in Saliva of Chronic Periodontitis Patients10030–55-year-old healthy individuals with chronic generalized periodontitis2014Measurement of the superoxide dismutase antioxidant enzyme levels in the saliva of chronic periodontitis subjects41 month
112NCT02909621Evaluation of FLEXOFYTOL® Versus PLACEBO (COPRA)15045–80-year-old patients with knee osteoarthritis 2017Measuring the variation in the serum levels of the sColl2-1 biomarker between T0 and T3 by specific immunoassays and the variation in the global assessment of disease activity by the patient using a visual analogue scale (VAS)NA6 months
113NCT04439981Curcuma Extract Beneficial for Muscle Damage2014–18-year-old healthy male individuals2019Change in lactic acid, Hb, IL-6 and creatinine kinaseNA21 days
114NCT02251678Evaluate the Effect of Elimune Capsules21≥18-year-old patients with plaque psoriasis with or without arthritis2015Individual subject serum levels of biomarkers (CRP, TNFa, IL-6, IL-12)128 days
115NCT04633551Vascular Inflammation and Anti-inflammatory Supplements After Adverse Pregnancy Outcomes (VIA)818–45-year-old female patients who had a singleton pregnancy of < 3 years complicated by an adverse pregnancy outcome (APO)2021Measurement of blood pressure, arterial stiffness, augmentation index and endothelial functionNA1 month
116NCT02834078Effect of BGG on Glucose Metabolism and Other Markers of Metabolic Syndrome (Glucogold)12620–60-year-old patients with a BMI ≥ 25 suffering from pre-diabetes or early diagnosed diabetes2016Measured change in the oral disposition index and HbA1cNA84 days
117NCT04149639A Study Investigating the Effectiveness of a LifeSeasons NeuroQ Supplement with Lifestyle Changes to Improve Cognitive Function in Healthy Adults Who Have One or More Risk Factors for Cognitive Decline40≥45-year-old patients with risk factors for cognitive decline2020Measured change in cognition as assessed by the change in the Neurocognitive Index (NCI) score from the CNS-Vital Signs (CNS-VS) panelNA135 days
118NCT01716637Short Term Efficacy and Safety of Perispinal Administration of Etanercept in Mild to Moderate Alzheimer’s Disease1260–85-year-old patients with a diagnosis of Alzheimer’s disease2016Difference in the effects of the treatment for 6 weeks with etanercept + nutritional supplements versus nutritional supplements alone on the Mini-Mental Status Examination (MMSE) score116 weeks
119NCT01752868Can Fish Oil and Phytochemical Supplements Mimic Anti-Aging Effects of Calorie Restriction?5640–60-year-old patients with a BMI of 21–30 kg/m2 who are sedentary to moderately active2012Carotid-femoral pulse wave velocityNA6 months
120NCT00799630Effects of Nutraperf Consumption in Runners1418–46-year-old healthy male distance runners 2008Measurement of different metabolic parameters (heart rate, oxygen consumption, respiratory quotient, ventilation, glycemia, lactatemia) on central and peripheral fatigue and on cognitive parametersNANA
121NCT04765527Turmeric and Exercise-Induced Muscle Damage and Oxinflammation5318–50-year-old healthy individuals who are willing to exercise2021Measuring a change in the serum concentration of creatine kinaseNA4 days
122NCT02413099The Efficacy and Safety of New Herbal Formula (KBMSI-2) in the Treatment of Erectile Dysfunction4418–40-year-old male patients with a history of erectile dysfunction2013Measuring a change in the EF domain scores of the IIEF questionnaire from the baseline48 weeks
123NCT01906840Role of Turmeric on Oxidative Modulation in ESRD Patients48≥18-year-old patients who undergo regular dialysis2012Measuring the effects of turmerics on oxidative stress markers28 weeks
124NCT01646047Diabetes Visual Function Supplement Study (DiVFuSS)70≥18-year-old patients with a ≥5-year history of diabetes mellitus2014Measuring changes in visual functionNA6 months
125NCT02369536Efficacy of a Natural Components Mixture in the Treatment of non-Alcoholic Fatty Liver Disease (NAFLD) (NUTRAFAST)12618–80-year-old patients with non-alcoholic fatty liver disease (NAFLD)2016Hematic levels of hepatic enzymes AST, ALT and GGTNA3 months
126NCT02088307Study of the Cardiovascular Vitamin, CardioLife2118–90-year-old patients with cardiovascular disease2016Change in blood pressureNA6 months
127NCT05089318Evaluation of Flexofytol® PLUS in Hand Osteoarthritis.239≥45-year-old patients with hand arthritis and a regular use of analgesia2021Pain using a Visual Analog Scale (VAS)NA84 days
128NCT03482401Disposition of Dietary Polyphenols and Methylxanthines in Mammary Tissues from Breast Cancer Patients (POLYSEN)40≥18-year-old patients diagnosed with breast cancer2019Quantification of dietary polyphenols and methylxanthines in breast tissuesNA24 months
129NCT04890704Curcuminoids and Contrast-induced Acute Kidney Injury9618–80-year-old patients undergoing elective CAG with a stable eGFR of 15–60 mL/min/1.72 m22019The incidence of CI-AKI development between the addition of curcuminoids to the standard protocol and the standard protocol alone in patients who underwent CAG148 h
130NCT00219882Safety Study of Orally Administered Curcuminoids in Adult Subjects with Cystic Fibrosis (SEER)1118–40-year-old patients who suffer from cystic fibrosis (homozygous for the ΔF508 CFTR genotype)2006Safety and tolerability of 14 days of treatment with orally administered curcuminoids, as assessed by adverse events, laboratory parameters and spirometry114 days
131NCT04844658COVID-19, Hospitalized, PatIents, Nasafytol51≥18-year-old patients with a recent hospitalization due to SARS-CoV-22021Improvement of the patient’s clinical condition based on the WHO ordinal outcomes score, the duration of hospitalization, mortality, fever, oxygen therapy, adverse events and several blood parametersNA14 days
132NCT03065504Turmeric and Turmeric-containing Tablets and Sebum Production3018–50-year-old healthy individuals2017Change in facial sebum productionNA4 weeks
133NCT04281758Comparison of Plasma Caffeine Concentration After Oral Consumption of Caffeinated Beverages with Varied Bioactive Compounds in Healthy Volunteers1618–55-year-old healthy individuals willing to avoid caffeine and alcohol for a period of time2020Incremental area-under-the-concentration-curve (iAUC)1210 min
134NCT04258501Exploratory Study of Efficacy on Selected Natural Extracts Reducing Post Prandial Blood Glucose Response7220–50-year-old healthy individuals with a normal BMI2012Change in post-prandial blood glucoseNA2 h

6. Limitations

In this paper, we have only focused on the key therapeutic activity of curcumin. Additionally, our focus has been on those activities of curcumin that are well characterized. Other aspects of curcumin activity, such as those associated with beneficial effects in neurological disorders, were not reviewed in this study. This alludes to the fact that such results have not been investigated in detail or explicated in the clinical trials. For example, for the phase 2 trial—curcumin in patients with mild to moderate Alzheimer’s disease—no results are posted in the trial database. This aspect may be attributed to the fact that the trial did not exhibit any beneficial outcome, more so because the delivery of curcumin across the blood-brain barrier has always been challenging. On the same note, we have not touched upon the aspect of curcumin delivery, as this is not only outside the scope of the review but also requires a detailed discussion which will make the present manuscript inadvertently lengthy. Readers are directed to some excellent reviews published in recent times for further details, if interested [295,296,297,298].

7. Conclusions and Future Directions

Curcumin is a pleiotropic molecule with a flexible structure with diverse biological functions. It is a potent proteasome inhibitor that increases the p53 level and induces apoptosis by mitochondrial caspase activation. Curcumin also disrupts 26S proteasome activity by inhibiting DYRK2 in different cancerous cells, resulting in the inhibition of cell proliferation [299]. However, further research is required to establish curcumin’s precise epigenetic regulatory effect for preventing and curing lethal diseases such as cancers. Curcumin may also act as an epigenetic regulator in neurological disorders, inflammation and diabetes. It can be effectively used as a histones modifier (acetylation/deacetylation), which is among the most important epigenetic changes responsible for gene expression alterations, leading to the modulation of the risks of rheumatoid arthritis and cancer.
Curcumin has shown therapeutic potential against several human diseases. The underlying mechanism for curcumin’s clinical efficacy seems to be the modulation of numerous signaling molecules. However, because of the complex nature of some diseases, the underlying mechanism in many cases remains unclear. Pharmacokinetic data indicate an almost 40-fold increase in blood levels in cases where curcumin was administered via formulation compared to pure form [300]. The poor bioavailability and limited adverse effects reported by some investigators are a major limitation to the therapeutic utility of curcumin. Nanocurcumin has shown a higher solubility and bioavailability in comparison to curcumin in recent studies [301]. Curcumin linked to phosphatidylcholine (which forms the fytosome–curcumin complex) has shown better bioavailability upon oral administration in rats [302]. We hope that the results from ongoing clinical trials will provide a deeper understanding of curcumin’s therapeutic potential and help to place this interesting molecule at the forefront of novel therapeutics.
The use of nanotechnology and a targeted drug delivery system has been shown to improve the cellular uptake, tissue specificity and effectiveness of curcumin. Although several nanosystems have been explored for the delivery of curcumin, due to its ability to inhibit the ABC efflux transporter [303], the combination nanoparticles of curcumin must be tested in cancerous cells once the proper dosage is determined. Most of the experiments using curcumin formulations have only been tested in pre-clinical models. The issue of cellular toxicity needs to be addressed by studying its activity in humans. Cost-effective techniques for curcumin nanoencapsulation are an emerging industrial requirement. The clinical trials to date have been conducted on a limited group of patients. Moreover, the tissue specificity of nanoparticles needs to be evaluated.
The combination of curcumin with other therapeutic reagents can further be explored. Case in point, a recent study showed that fecal microbial transplantation (FMT) leads to favorable outcomes in metabolic syndrome [304]. It would be interesting to appraise what happens in a group where a combined approach of FMT, fiber supplementation and curcumin is employed. For example, Liraglutide is used in the treatment of obesity because it induces weight loss; curcumin can be supplemented in a combined formulation with liraglutide to add to its benefits [305].
Currently, in our group, we are assessing the therapeutic potential of curcumin in combination with vitamin D and lipids of minor physiological abundance in attenuating inflammation in chondrocytes treated with LPS. The rationale behind using such a combination is that lipids [198] of minor physiological abundance, such as lysosulfatide (which is present in HDL particles) and vitamin D (which has a steroid nucleus), will augment the solubility and thus the bioavailability of curcumin [199].

Author Contributions

B.M.S., M.A. and Y.B. collected the literature and prepared the outlines and the first draft of manuscript. M.R., A.P.S. and R.P. gave suggestions and improved the manuscript. 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

ACEAcetyl coenzyme
ADPCAndrogen-dependant prostate cancer
AIPCAndrogen-independent prostate cancer
AKTProtein kinase B (also called PKB)
ALTAlanine transaminase
AMPKAMP-activated protein kinase
APActivator protein
ASTAspartate aminotransferase
BAFFB cell activating factor
BCLB cell lymphocyte
BclB cell lymphoma
Bcl-xLB cell lymphoma-extra large
CDCluster differentiation
CHOPCytoxan hydroxydaunorubicin oncovin prednisone
Coll2Collagen
COXCyclooxygenase
Cur-CQDCurcumin carbon quantom dots
CXCLChemokine (C-X-C motif) ligand
DHADocosahexaenoic acid
DYRKDual specificity tyrosine phosphorylation-regulated kinase
EAEEnzyme-assisted extraction
EBPEnhancer binding protein
EPAEicosapentaenoic acid
ERKExtracellular-regulated kinase
ERKExtracellular-regulated kinase
FFAFree fatty acid
FOXForkhead box protein
FtsZFilamenting temperature sensitive mutant Z
GSTGlutathione S- transferase
HpHaptoglobin
HSPHeat shock protein
IFNInterferon
IKBαInhibitor of kappa light chain gene enhancer in B cells
ILInterlekin
ILEIonic liquid-based extraction
ILInterleukin
iNOSInducible nitric oxide syntase
JAK/STATJanus kinase/signal transducers and activators of transcription
JNKJun N-terminal kinase
LPSLipopolysaccharides
MAEMicrowave-assisted extraction
MAPKMitogen-activated protein kinase
MCPMethyl-accepting chemotaxis protein
M-CSFMacrophage colony stimulating factor
MDAMalondialdehyde
MICMinimum inhibitory concentration
MIPMacrophage inflammatory protein
MMPMatrix metalloproteinase
MPAMedroxyprogesterone acetate
MTORMammalian target of rapamycin
MyDMyeloid differentiation
NF-κkBNuclear factor kappa-light-chain-enhancer of activated B cells
NLRPNod-like receptor protein
NODNucleotide oligomerization domain
NrfNuclear respiratory factor
NSAIDNon-steroidal anti-inflammatory drugs
OCOleocanthal
ODCOrnithine decarboxylase
PECAMPlatelet endothelial cell adhesion molecule
PLGAPoly(D,L-Lactic-co-glycolic acid)
PPARPeroxisome proliferator-activated receptors
Rac 1Rass-related C 3 botulinum toxin substrate 1
RANKLReceptor activator of nuclear factor kappa B ligand
RANTESRegulated on activation, normal T cell expressed and secreted
RNSReactive nitrogen species
ROSReactive oxygen species
SDHSuccinate dehydrogenase
SFESupercritical fluid extraction
ShhSonic hedgehog protein
SODSuperoxide dismutase
STZStreptozotocin
TBARsThiobarbituric acid reactive substances
TLRToll-like receptor
TLRToll-like Receptor
TNFTumor Necrosis Factor
TRAILTumor necrosis factor (TNF)-related apoptosis-inducing ligand
TRAILTumor necrosis factor-related apoptosis-inducing ligand
UAEUltrasound-assisted extraction
UPRUnfolded protein response
VEGFVascular endothelial growth factor
WntWingless related integration site
XIAPX-chromosome-linked inhibitor of apoptosis protein
ADPCAndrogen-dependant prostate cancer
AIPCAndrogen-independent prostate cancer
APActivator protein
BclB cell lymphoma
Bcl-xLB cell lymphoma-extra large
COXCyclooxygenase
CXCLChemokine (C-X-C motif) ligand
FOXForkhead box protein
ILInterleukin
JNKJun N-terminal kinase
MMPMatrix metalloproteinase
MPAMedroxyprogesterone acetate
NF-kBNuclear factor kappa-light-chain-enhancer of activated B cells
ODCOrnithine decarboxylase
PECAMPlatelet endothelial cell adhesion molecule
PLGAPoly(D,L-Lactic-co-glycolic acid)
TLRToll-like receptor
TNFTumor necrosis factor
TRAILTumor necrosis factor-related apoptosis-inducing ligand
TRAILTumor necrosis factor (TNF)-related apoptosis-inducing ligand
VEGFVascular endothelial growth factor
XIAPX-chromosome-linked inhibitor of apoptosis protein

References

  1. Rao, A.S.; Hegde, S.; Pacioretty, L.M.; DeBenedetto, J.; Babish, J.G. Nigella sativa and trigonella foenum-graecum supplemented chapatis safely improve HbA1c, body weight, waist circumference, blood lipids, and fatty liver in overweight and diabetic subjects: A twelve-week safety and efficacy study. J. Med. Food 2020, 23, 905–919. [Google Scholar] [CrossRef] [PubMed]
  2. Konstantinidi, M.; Koutelidakis, A.E. Functional foods and bioactive compounds: A review of its possible role on weight management and obesity’s metabolic consequences. Medicines 2019, 6, 94. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Brower, V. Nutraceuticals: Poised for a healthy slice of the healthcare market? Nat. Biotechnol. 1998, 16, 728–731. [Google Scholar] [CrossRef] [PubMed]
  4. Trottier, G.; Boström, P.J.; Lawrentschuk, N.; Fleshner, N.E. Nutraceuticals and prostate cancer prevention: A current review. Nat. Rev. Urol. 2010, 7, 21–30. [Google Scholar] [CrossRef]
  5. Zeisel, S.H. Regulation of “nutraceuticals”. Science 1999, 285, 1853–1855. [Google Scholar] [CrossRef]
  6. Keservani, R.K.; Kesharwani, R.K.; Vyas, N.; Jain, S.; Raghuvanshi, R.; Sharma, A.K. Nutraceutical and functional food as future food: A review. Der Pharm. Lett. 2010, 2, 106–116. [Google Scholar]
  7. Murphy, R. Extra-virgin olive oil has similar activity to ibuprofen. Nat. Clin. Pract. Rheumatol. 2005, 1, 66. [Google Scholar] [CrossRef]
  8. Patti, A.M.; Carruba, G.; Cicero, A.F.G.; Banach, M.; Nikolic, D.; Giglio, R.V.; Terranova, A.; Soresi, M.; Giannitrapani, L.; Montalto, G. Daily use of extra virgin olive oil with high oleocanthal concentration reduced body weight, waist circumference, alanine transaminase, inflammatory cytokines and hepatic steatosis in subjects with the metabolic syndrome: A 2-month intervention study. Metabolites 2020, 10, 392. [Google Scholar] [CrossRef]
  9. Imran, M.; Ghorat, F.; Ul-Haq, I.; Ur-Rehman, H.; Aslam, F.; Heydari, M.; Shariati, M.A.; Okuskhanova, E.; Yessimbekov, Z.; Thiruvengadam, M. Lycopene as a natural antioxidant used to prevent human health disorders. Antioxidants 2020, 9, 706. [Google Scholar] [CrossRef]
  10. Berman, A.Y.; Motechin, R.A.; Wiesenfeld, M.Y.; Holz, M.K. The therapeutic potential of resveratrol: A review of clinical trials. NPJ Precis. Oncol. 2017, 1, 35. [Google Scholar] [CrossRef] [Green Version]
  11. Canistro, D.; Chiavaroli, A.; Cicia, D.; Cimino, F.; Curro, D.; Dell’Agli, M.; Ferrante, C.; Giovannelli, L.; Leone, S.; Martinelli, G. The pharmacological basis of the curcumin nutraceutical uses: An update. Pharmadvances 2021, 3, 421–466. [Google Scholar] [CrossRef]
  12. Vogel, A.; Pelletier, J. Examen chimique de la racine de Curcuma. J. Pharm. 1815, 1, 289–300. [Google Scholar]
  13. Prasad, S.; Aggarwal, B.B. Turmeric, the golden spice. In Herbal Medicine: Biomolecular and Clinical Aspects, 2nd ed.; CRC Press/Taylor & Francis: Boca Raton, FL, USA, 2011. [Google Scholar]
  14. Hosseini, A.; Hosseinzadeh, H. Antidotal or protective effects of Curcuma longa (turmeric) and its active ingredient, curcumin, against natural and chemical toxicities: A review. Biomed. Pharmacother. 2018, 99, 411–421. [Google Scholar] [CrossRef]
  15. Lal, J. Turmeric, curcumin and our life: A review. Bull. Environ. Pharmacol. Life Sci. 2012, 1, 11–17. [Google Scholar]
  16. Aggarwal, B.B.; Surh, Y.-J.; Shishodia, S. The Molecular Targets and Therapeutic Uses of Curcumin in Health and Disease; Springer Science & Business Media: Cham, Switzerland, 2007. [Google Scholar]
  17. Aggarwal, B.B.; Kumar, A.; Bharti, A.C. Anticancer potential of curcumin: Preclinical and clinical studies. Anticancer Res. 2003, 23, 363–398. [Google Scholar]
  18. Lin, C.-L.; Lin, J.-K. Curcumin: A potential cancer chemopreventive agent through suppressing NF-κB signaling. J. Cancer Mol. 2008, 4, 11–16. [Google Scholar]
  19. Shen, L.; Ji, H.-F. Theoretical study on physicochemical properties of curcumin. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2007, 67, 619–623. [Google Scholar] [CrossRef] [PubMed]
  20. Sharma, S. Synthesis and Anti Diabetic Activity of Curcumin. Int. J. Adv. Res. Ideal Innov. Technol. 2017, 3, 1512–1520. [Google Scholar]
  21. Stanić, Z. Curcumin, a compound from natural sources, a true scientific challenge—A review. Plant Foods Hum. Nutr. 2017, 72, 1–12. [Google Scholar] [CrossRef]
  22. 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]
  23. Volak, L.P.; Hanley, M.J.; Masse, G.; Hazarika, S.; Harmatz, J.S.; Badmaev, V.; Majeed, M.; Greenblatt, D.J.; Court, M.H. Effect of a herbal extract containing curcumin and piperine on midazolam, flurbiprofen and paracetamol (acetaminophen) pharmacokinetics in healthy volunteers. Br. J. Clin. Pharmacol. 2013, 75, 450–462. [Google Scholar] [CrossRef] [Green Version]
  24. Lukita-Atmadja, W.; Ito, Y.; Baker, G.L.; McCuskey, R.S. Effect of curcuminoids as anti-inflammatory agents on the hepatic microvascular response to endotoxin. Shock 2002, 17, 399–403. [Google Scholar] [CrossRef] [PubMed]
  25. Holder, G.M.; Plummer, J.L.; Ryan, A.J. The metabolism and excretion of curcumin (1, 7-bis-(4-hydroxy-3-methoxyphenyl)-1, 6-heptadiene-3, 5-dione) in the rat. Xenobiotica 1978, 8, 761–768. [Google Scholar] [CrossRef] [PubMed]
  26. Ravindranath, V.; Chandrasekhara, N. Absorption and tissue distribution of curcumin in rats. Toxicology 1980, 16, 259–265. [Google Scholar] [CrossRef]
  27. Zhang, Q.-W.; Lin, L.-G.; Ye, W.-C. Techniques for extraction and isolation of natural products: A comprehensive review. Chin. Med. 2018, 13, 20. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Mandal, V.; Mohan, Y.; Hemalatha, S. Microwave assisted extraction of curcumin by sample–solvent dual heating mechanism using Taguchi L9 orthogonal design. J. Pharm. Biomed. Anal. 2008, 46, 322–327. [Google Scholar] [CrossRef]
  29. Sahne, F.; Mohammadi, M.; Najafpour, G.D.; Moghadamnia, A.A. Enzyme-assisted ionic liquid extraction of bioactive compound from turmeric (Curcuma longa L.): Isolation, purification and analysis of curcumin. Ind. Crops Prod. 2017, 95, 686–694. [Google Scholar] [CrossRef]
  30. Osorio-Tobon, J.F.; Carvalho, P.I.N.; Rostagno, M.A.; Petenate, A.J.; Meireles, M.A.A. Extraction of curcuminoids from deflavored turmeric (Curcuma longa L.) using pressurized liquids: Process integration and economic evaluation. J. Supercrit. Fluids 2014, 95, 167–174. [Google Scholar] [CrossRef]
  31. Li, M.; Ngadi, M.O.; Ma, Y. Optimisation of pulsed ultrasonic and microwave-assisted extraction for curcuminoids by response surface methodology and kinetic study. Food Chem. 2014, 165, 29–34. [Google Scholar] [CrossRef]
  32. Xu, J.; Wang, W.; Liang, H.; Zhang, Q.; Li, Q. Optimization of ionic liquid based ultrasonic assisted extraction of antioxidant compounds from Curcuma longa L. using response surface methodology. Ind. Crops Prod. 2015, 76, 487–493. [Google Scholar] [CrossRef]
  33. Search Media-Wikimedia Commons. (n.d.). Available online: https://commons.wikimedia.org/w/index.php?search=Curcumin+structure&title=Special:MediaSearch&go=Go&type=image (accessed on 24 April 2022).
  34. Search Media-Wikimedia Commons. (n.d.). Available online: https://commons.wikimedia.org/w/index.php?search=turmeric&title=Special:MediaSearch&go=Go&type=image (accessed on 24 April 2022).
  35. Gao, Y.; Zhuang, Z.; Lu, Y.; Tao, T.; Zhou, Y.; Liu, G.; Wang, H.; Zhang, D.; Wu, L.; Dai, H. Curcumin mitigates neuro-inflammation by modulating microglia polarization through inhibiting TLR4 axis signaling pathway following experimental subarachnoid hemorrhage. Front. Neurosci. 2019, 13, 1223. [Google Scholar] [CrossRef] [PubMed]
  36. Zhang, J.; Zheng, Y.; Luo, Y.; Du, Y.; Zhang, X.; Fu, J. Curcumin inhibits LPS-induced neuroinflammation by promoting microglial M2 polarization via TREM2/TLR4/NF-B pathways in BV2 cells. Mol. Immunol. 2019, 116, 29–37. [Google Scholar] [CrossRef] [PubMed]
  37. Sadeghi, A.; Rostamirad, A.; Seyyedebrahimi, S.; Meshkani, R. Curcumin ameliorates palmitate-induced inflammation in skeletal muscle cells by regulating JNK/NF-kB pathway and ROS production. Inflammopharmacology 2018, 26, 1265–1272. [Google Scholar] [CrossRef] [PubMed]
  38. Garufi, A.; Giorno, E.; Gilardini Montani, M.S.; Pistritto, G.; Crispini, A.; Cirone, M.; D’Orazi, G. p62/SQSTM1/Keap1/NRF2 axis reduces cancer cells death-sensitivity in response to Zn (II)–curcumin complex. Biomolecules 2021, 11, 348. [Google Scholar] [CrossRef]
  39. Liu, T.; Zhang, L.; Joo, D.; Sun, S.-C. NF-κB signaling in inflammation. Signal Transduct. Target. Ther. 2017, 2, 17023. [Google Scholar] [CrossRef] [Green Version]
  40. Kong, F.; Ye, B.; Cao, J.; Cai, X.; Lin, L.; Huang, S.; Huang, W.; Huang, Z. Curcumin represses NLRP3 inflammasome activation via TLR4/MyD88/NF- B and P2X7R signaling in PMA-induced macrophages. Front. Pharmacol. 2016, 7, 369. [Google Scholar] [CrossRef] [Green Version]
  41. Gong, Z.; Zhou, J.; Li, H.; Gao, Y.; Xu, C.; Zhao, S.; Chen, Y.; Cai, W.; Wu, J. Curcumin suppresses NLRP3 inflammasome activation and protects against LPS induced septic shock. Mol. Nutr. Food Res. 2015, 59, 2132–2142. [Google Scholar] [CrossRef]
  42. Wang, Q.; Ye, C.; Sun, S.; Li, R.; Shi, X.; Wang, S.; Zeng, X.; Kuang, N.; Liu, Y.; Shi, Q. Curcumin attenuates collagen-induced rat arthritis via anti-inflammatory and apoptotic effects. Int. Immunopharmacol. 2019, 72, 292–300. [Google Scholar] [CrossRef]
  43. Murakami, Y.; Kawata, A.; Fujisawa, S. Expression of cyclooxygenase-2, nitric oxide synthase 2 and heme oxygenase-1 mRNA induced by bis-eugenol in RAW264. 7 cells and their antioxidant activity determined using the induction period method. In Vivo 2017, 31, 819–831. [Google Scholar]
  44. Mollazadeh, H.; Cicero, A.F.G.; Blesso, C.N.; Pirro, M.; Majeed, M.; Sahebkar, A. Immune modulation by curcumin: The role of interleukin-10. Crit. Rev. Food Sci. Nutr. 2019, 59, 89–101. [Google Scholar] [CrossRef]
  45. Huang, G.; Xu, Z.; Huang, Y.; Duan, X.; Gong, W.; Zhang, Y.; Fan, J.; He, F. Curcumin protects against collagen-induced arthritis via suppression of BAFF production. J. Clin. Immunol. 2013, 33, 550–557. [Google Scholar] [CrossRef] [PubMed]
  46. Bugatti, S.; Cassione, E.B.; De Stefano, L.; Manzo, A. Established rheumatoid arthritis. The pathogenic aspects. Best Pract. Res. Clin. Rheumatol. 2019, 33, 101478. [Google Scholar] [CrossRef] [PubMed]
  47. Moon, D.-O.; Kim, M.-O.; Choi, Y.H.; Park, Y.-M.; Kim, G.-Y. Curcumin attenuates inflammatory response in IL-1β-induced human synovial fibroblasts and collagen-induced arthritis in mouse model. Int. Immunopharmacol. 2010, 10, 605–610. [Google Scholar] [CrossRef]
  48. Dai, Q.; Di Zhou, L.X.; Song, X. Curcumin alleviates rheumatoid arthritis-induced inflammation and synovial hyperplasia by targeting mTOR pathway in rats. Drug Des. Devel. Ther. 2018, 12, 4095–4105. [Google Scholar] [CrossRef] [Green Version]
  49. Liu, H.; Zhang, Y.; Zheng, S.; Weng, Z.; Ma, J.; Li, Y.; Xie, X.; Zheng, W. Detention of copper by sulfur nanoparticles inhibits the proliferation of A375 malignant melanoma and MCF-7 breast cancer cells. Biochem. Biophys. Res. Commun. 2016, 477, 1031–1037. [Google Scholar] [CrossRef]
  50. Buhrmann, C.; Mobasheri, A.; Matis, U.; Shakibaei, M. Curcumin mediated suppression of nuclear factor-κB promotes chondrogenic differentiation of mesenchymal stem cells in a high-density co-culture microenvironment. Arthritis Res. Ther. 2010, 12, R127. [Google Scholar] [CrossRef] [Green Version]
  51. Nozaki, Y.; Ri, J.; Sakai, K.; Niki, K.; Kinoshita, K.; Funauchi, M.; Matsumura, I. Inhibition of the IL-18 receptor signaling pathway ameliorates disease in a murine model of rheumatoid arthritis. Cells 2019, 9, 11. [Google Scholar] [CrossRef] [Green Version]
  52. Chandran, B.; Goel, A. A randomized, pilot study to assess the efficacy and safety of curcumin in patients with active rheumatoid arthritis. Phyther. Res. 2012, 26, 1719–1725. [Google Scholar] [CrossRef] [PubMed]
  53. Javadi, M.; Khadem Haghighian, H.; Goodarzy, S.; Abbasi, M.; Nassiri-Asl, M. Effect of curcumin nanomicelle on the clinical symptoms of patients with rheumatoid arthritis: A randomized, double-blind, controlled trial. Int. J. Rheum. Dis. 2019, 22, 1857–1862. [Google Scholar] [CrossRef]
  54. Fu, M.; Zhou, H.; Li, Y.; Jin, H.; Liu, X. Global, regional, and national burdens of hip osteoarthritis from 1990 to 2019: Estimates from the 2019 Global Burden of Disease Study. Arthritis Res. Ther. 2022, 24, 8. [Google Scholar] [CrossRef]
  55. Cross, M.; Smith, E.; Hoy, D.; Nolte, S.; Ackerman, I.; Fransen, M.; Bridgett, L.; Williams, S.; Guillemin, F.; Hill, C.L.; et al. The global burden of hip and knee osteoarthritis: Estimates from the Global Burden of Disease 2010 study. Ann. Rheum. Dis. 2014, 73, 1323–1330. [Google Scholar] [CrossRef] [PubMed]
  56. Kalaitzoglou, E.; Griffin, T.M.; Humphrey, M.B. Innate Immune Responses and Osteoarthritis. Curr. Rheumatol. Rep. 2017, 19, 17–22. [Google Scholar] [CrossRef] [PubMed]
  57. Radin, E.; Paul, I.; Rose, R. Role of mechanical factors in pathogenesis of primary osteoarthritis. Lancet 1972, 299, 519–522. [Google Scholar] [CrossRef]
  58. Andriacchi, T.P.; Favre, J.; Erhart-Hledik, J.C.; Chu, C.R. A systems view of risk factors for knee osteoarthritis reveals insights into the pathogenesis of the disease. Ann. Biomed. Eng. 2015, 43, 376–387. [Google Scholar] [CrossRef]
  59. Zhang, W.; Nuki, G.; Moskowitz, R.W.; Abramson, S.; Altman, R.D.; Arden, N.K.; Bierma-Zeinstra, S.; Brandt, K.D.; Croft, P.; Doherty, M.; et al. OARSI recommendations for the management of hip and knee osteoarthritis. Part III: Changes in evidence following systematic cumulative update of research published through January 2009. Osteoarthr. Cartil. 2010, 18, 476–499. [Google Scholar] [CrossRef] [Green Version]
  60. Felson, D.T.; Lawrence, R.C.; Hochberg, M.C.; McAlindon, T.; Dieppe, P.A.; Minor, M.A.; Blair, S.N.; Berman, B.M.; Fries, J.F.; Weinberger, M. Osteoarthritis: New insights. Part 2: Treatment approaches. Ann. Intern. Med. 2000, 133, 726–737. [Google Scholar] [CrossRef]
  61. Smalley, W.E.; Ray, W.A.; Daugherty, J.R.; Griffin, M.R. Nonsteroidal anti-inflammatory drugs and the incidence of hospitalizations for peptic ulcer disease in elderly persons. Am. J. Epidemiol. 1995, 141, 539–545. [Google Scholar] [CrossRef]
  62. Pitt, B.; Pepine, C.; Willerson, J.T. Cyclooxygenase-2 inhibition and cardiovascular events. Circulation 2002, 106, 167–169. [Google Scholar] [CrossRef] [Green Version]
  63. Hörl, W.H. Nonsteroidal anti-inflammatory drugs and the kidney. Pharmaceuticals 2010, 3, 2291–2321. [Google Scholar] [CrossRef]
  64. Yan, D.; He, B.; Guo, J.; Li, S.; Wang, J. Involvement of TLR4 in the protective effect of intra-articular administration of curcumin on rat experimental osteoarthritis. Acta Cir. Bras. 2019, 34, e201900604. [Google Scholar] [CrossRef] [Green Version]
  65. Sun, Y.; Liu, W.; Zhang, H.; Li, H.; Liu, J.; Zhang, F.; Jiang, T.; Jiang, S. Curcumin Prevents Osteoarthritis by Inhibiting the Activation of Inflammasome NLRP3. J. Interf. Cytokine Res. 2017, 37, 449–455. [Google Scholar] [CrossRef] [PubMed]
  66. Zhang, G.; Cao, J.; Yang, E.; Liang, B.; Ding, J.; Liang, J.; Xu, J. Curcumin improves age-related and surgically induced osteoarthritis by promoting autophagy in mice. Biosci. Rep. 2018, 38, 1–11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Csaki, C.; Mobasheri, A.; Shakibaei, M. Synergistic chondroprotective effects of curcumin and resveratrol in human articular chondrocytes: Inhibition of IL-1β-induced NF-κB-mediated inflammation and apoptosis. Arthritis Res. Ther. 2009, 11, R165. [Google Scholar] [CrossRef] [Green Version]
  68. Yabas, M.; Orhan, C.; Er, B.; Tuzcu, M.; Durmus, A.S.; Ozercan, I.H.; Sahin, N.; Bhanuse, P.; Morde, A.A.; Padigaru, M.; et al. A Next Generation Formulation of Curcumin Ameliorates Experimentally Induced Osteoarthritis in Rats via Regulation of Inflammatory Mediators. Front. Immunol. 2021, 12, 1–13. [Google Scholar] [CrossRef]
  69. Benderdour, M.; Martel-Pelletier, J.; Pelletier, J.-P.; Kapoor, M.; Zunzunegui, M.-V.; Fahmi, H. Cellular aging, senescence and autophagy processes in osteoarthritis. Curr. Aging Sci. 2015, 8, 147–157. [Google Scholar] [CrossRef]
  70. Li, X.; Feng, K.; Li, J.; Yu, D.; Fan, Q.; Tang, T.; Yao, X.; Wang, X. Curcumin inhibits apoptosis of chondrocytes through activation ERK1/2 signaling pathways induced autophagy. Nutrients 2017, 9, 414. [Google Scholar] [CrossRef] [Green Version]
  71. Sun, K.; Luo, J.; Guo, J.; Yao, X.; Jing, X.; Guo, F. The PI3K/AKT/mTOR signaling pathway in osteoarthritis: A narrative review. Osteoarthr. Cartil. 2020, 28, 400–409. [Google Scholar] [CrossRef]
  72. Paultre, K.; Cade, W.; Hernandez, D.; Reynolds, J.; Greif, D.; Best, T.M. Therapeutic effects of turmeric or curcumin extract on pain and function for individuals with knee osteoarthritis: A systematic review. BMJ Open Sport Exerc. Med. 2021, 7, e000935. [Google Scholar] [CrossRef] [PubMed]
  73. Panda, S.K.; Nirvanashetty, S.; Parachur, V.A.; Mohanty, N.; Swain, T. A Randomized, Double Blind, Placebo Controlled, Parallel-Group Study to Evaluate the Safety and Efficacy of Curene® versus Placebo in Reducing Symptoms of Knee OA. Biomed. Res. Int. 2018, 2018, 1–8. [Google Scholar] [CrossRef] [Green Version]
  74. Nakagawa, Y.; Mukai, S.; Yamada, S.; Matsuoka, M.; Tarumi, E.; Hashimoto, T.; Tamura, C.; Imaizumi, A.; Nishihira, J.; Nakamura, T. Short-term effects of highly-bioavailable curcumin for treating knee osteoarthritis: A randomized, double-blind, placebo-controlled prospective study. J. Orthop. Sci. 2014, 19, 933–939. [Google Scholar] [CrossRef] [Green Version]
  75. Henrotin, Y.; Gharbi, M.; Dierckxsens, Y.; Priem, F.; Marty, M.; Seidel, L.; Albert, A.; Heuse, E.; Bonnet, V.; Castermans, C. Decrease of a specific biomarker of collagen degradation in osteoarthritis, Coll2-1, by treatment with highly bioavailable curcumin during an exploratory clinical trial. BMC Complement. Altern. Med. 2014, 14, 159. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Kuptniratsaikul, V.; Dajpratham, P.; Taechaarpornkul, W.; Buntragulpoontawee, M.; Lukkanapichonchut, P.; Chootip, C.; Saengsuwan, J.; Tantayakom, K.; Laongpech, S. Efficacy and safety of Curcuma domestica extracts compared with ibuprofen in patients with knee osteoarthritis: A multicenter study. Clin. Interv. Aging 2014, 9, 451–458. [Google Scholar] [CrossRef] [Green Version]
  77. Shep, D.; Khanwelkar, C.; Gade, P.; Karad, S. Efficacy and safety of combination of curcuminoid complex and diclofenac versus diclofenac in knee osteoarthritis: A randomized trial. Medicine 2020, 99, e19723. [Google Scholar] [CrossRef]
  78. Lev-Ari, S.; Strier, L.; Kazanov, D.; Elkayam, O.; Lichtenberg, D.; Caspi, D.; Arber, N. Curcumin synergistically potentiates the growth-inhibitory and pro-apoptotic effects of celecoxib in osteoarthritis synovial adherent cells. Rheumatology 2006, 45, 171–177. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Belcaro, G.; Cesarone, M.R.; Dugall, M.; Pellegrini, L.; Ledda, A.; Grossi, M.G.; Togni, S.; Appendino, G. Efficacy and safety of Meriva®, a curcumin-phosphatidylcholine complex, during extended administration in osteoarthritis patients. Altern. Med. Rev. 2010, 15, 337–344. [Google Scholar]
  80. 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] [PubMed] [Green Version]
  81. Shanmugam, M.K.; Rane, G.; Kanchi, M.M.; Arfuso, F.; Chinnathambi, A.; Zayed, M.E.; Alharbi, S.A.; Tan, B.K.H.; Kumar, A.P.; Sethi, G. The multifaceted role of curcumin in cancer prevention and treatment. Molecules 2015, 20, 2728–2769. [Google Scholar] [CrossRef]
  82. Bimonte, S.; Barbieri, A.; Palma, G.; Rea, D.; Luciano, A.; D’Aiuto, M.; Arra, C.; Izzo, F. Dissecting the role of curcumin in tumour growth and angiogenesis in mouse model of human breast cancer. Biomed Res. Int. 2015, 2015, 878134. [Google Scholar] [CrossRef]
  83. Fetoni, A.R.; Paciello, F.; Mezzogori, D.; Rolesi, R.; Eramo, S.L.M.; Paludetti, G.; Troiani, D. Molecular targets for anticancer redox chemotherapy and cisplatin-induced ototoxicity: The role of curcumin on pSTAT3 and Nrf-2 signalling. Br. J. Cancer 2015, 113, 1434–1444. [Google Scholar] [CrossRef] [Green Version]
  84. Deeb, D.; Jiang, H.; Gao, X.; Divine, G.; Dulchavsky, S.A.; Gautam, S.C. Chemosensitization of hormone-refractory prostate cancer cells by curcumin to TRAIL-induced apoptosis. J. Exp. Ther. Oncol. 2005, 5, 81–91. [Google Scholar]
  85. Deeb, D.; Jiang, H.; Gao, X.; Al-Holou, S.; Danyluk, A.L.; Dulchavsky, S.A.; Gautam, S.C. Curcumin [1, 7-bis (4-hydroxy-3-methoxyphenyl)-1–6-heptadine-3, 5-dione; C21H20O6] sensitizes human prostate cancer cells to tumor necrosis factor-related apoptosis-inducing ligand/Apo2L-induced apoptosis by suppressing nuclear factor-κB via inhibition of the Prosurvival Akt Signaling Pathway. J. Pharmacol. Exp. Ther. 2007, 321, 616–625. [Google Scholar] [PubMed]
  86. De Velasco, M.A.; Lu, Y.; Kura, Y.; China, T.; Inoue, Y.; Nakayama, A.; Okada, H.; Horie, S.; Uemura, H.; Ide, H. Chemopreventive effects of nanoparticle curcumin in a mouse model of Pten-deficient prostate cancer. Hum. Cell 2020, 33, 730–736. [Google Scholar] [CrossRef] [PubMed]
  87. Ma, Q.; Qian, W.; Tao, W.; Zhou, Y.; Xue, B. Delivery of curcumin nanoliposomes using surface modified with CD133 aptamers for prostate cancer. Drug Des. Devel. Ther. 2019, 13, 4021–4033. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Deeb, D.; Jiang, H.; Gao, X.; Hafner, M.S.; Wong, H.; Divine, G.; Chapman, R.A.; Dulchavsky, S.A.; Gautam, S.C. Curcumin sensitizes prostate cancer cells to tumor necrosis factor–related apoptosis-inducing ligand/Apo2L by inhibiting nuclear factor-κB through suppression of IκBα phosphorylation. Mol. Cancer Ther. 2004, 3, 803–812. [Google Scholar] [CrossRef]
  89. Mukhopadhyay, A.; Bueso-Ramos, C.; Chatterjee, D.; Pantazis, P.; Aggarwal, B.B. Curcumin downregulates cell survival mechanisms in human prostate cancer cell lines. Oncogene 2001, 20, 7597–7609. [Google Scholar] [CrossRef] [Green Version]
  90. Liu, S.; Wang, Z.; Hu, Z.; Zeng, X.; Li, Y.; Su, Y.; Zhang, C.; Ye, Z. Anti-tumor activity of curcumin against androgen-independent prostate cancer cells via inhibition of NF-κB and AP-1 pathway in vitro. J. Huazhong Univ. Sci. Technol. Med. Sci. 2011, 31, 530–534. [Google Scholar] [CrossRef]
  91. Hour, T.; Chen, J.; Huang, C.; Guan, J.; Lu, S.; Pu, Y. Curcumin enhances cytotoxicity of chemotherapeutic agents in prostate cancer cells by inducing p21WAF1/CIP1 and C/EBPβ expressions and suppressing NF-κB activation. Prostate 2002, 51, 211–218. [Google Scholar] [CrossRef]
  92. 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]
  93. Deeb, D.; Xu, Y.X.; Jiang, H.; Gao, X.; Janakiraman, N.; Chapman, R.A.; Gautam, S.C. Curcumin (diferuloyl-methane) enhances tumor necrosis factor-related apoptosis-inducing ligand-induced apoptosis in lncap prostate cancer cells1. Mol. Cancer Ther. 2003, 2, 95–103. [Google Scholar]
  94. Andrzejewski, T.; Deeb, D.; Gao, X.; Danyluk, A.; Arbab, A.S.; Dulchavsky, S.A.; Gautam, S.C. Therapeutic efficacy of curcumin/TRAIL combination regimen for hormone-refractory prostate cancer. Oncol. Res. Featur. Preclin. Clin. Cancer Ther. 2008, 17, 257–267. [Google Scholar] [CrossRef]
  95. Gracia, E.; Mancini, A.; Colapietro, A.; Mateo, C.; Gracia, I.; Festuccia, C.; Carmona, M. Impregnation of curcumin into a biodegradable (poly-lactic-co-glycolic acid, PLGA) support, to transfer its well known in vitro effect to an in vivo prostate cancer model. Nutrients 2019, 11, 2312. [Google Scholar] [CrossRef] [Green Version]
  96. Poma, P.; Labbozzetta, M.; D’Alessandro, N.; Notarbartolo, M. NF-κB is a potential molecular drug target in triple-negative breast cancers. Omi. A J. Integr. Biol. 2017, 21, 225–231. [Google Scholar] [CrossRef]
  97. RS, P.; Bomb, K.; Srivastava, R.; Bandyopadhyaya, R. Dual drug delivery of curcumin and niclosamide using PLGA nanoparticles for improved therapeutic effect on breast cancer cells. J. Polym. Res. 2020, 27, 133. [Google Scholar]
  98. Abdel-Hafez, S.M.; Hathout, R.M.; Sammour, O.A. Curcumin-loaded ultradeformable nanovesicles as a potential delivery system for breast cancer therapy. Colloids Surf. B Biointerfaces 2018, 167, 63–72. [Google Scholar] [CrossRef]
  99. Prabhuraj, R.S.; Bomb, K.; Srivastava, R.; Bandyopadhyaya, R. Selection of superior targeting ligands using PEGylated PLGA nanoparticles for delivery of curcumin in the treatment of triple-negative breast cancer cells. J. Drug Deliv. Sci. Technol. 2020, 57, 101722. [Google Scholar] [CrossRef]
  100. Pires, B.R.B.; Mencalha, A.L.; Ferreira, G.M.; de Souza, W.F.; Morgado-Díaz, J.A.; Maia, A.M.; Corrêa, S.; Abdelhay, E.S.F.W. NF-kappaB is involved in the regulation of EMT genes in breast cancer cells. PLoS ONE 2017, 12, e0169622. [Google Scholar]
  101. Park, Y.H. The nuclear factor-kappa B pathway and response to treatment in breast cancer. Pharmacogenomics 2017, 18, 1697–1709. [Google Scholar] [CrossRef] [PubMed]
  102. Leu, T.-H.; Maa, M.-C. The molecular mechanisms for the antitumorigenic effect of curcumin. Curr. Med. Chem. Agents 2002, 2, 357–370. [Google Scholar]
  103. Carroll, C.E.; Ellersieck, M.R.; Hyder, S.M. Curcumin inhibits MPA-induced secretion of VEGF from T47-D human breast cancer cells. Menopause 2008, 15, 570–574. [Google Scholar] [CrossRef]
  104. Song, X.; Zhang, M.; Dai, E.; Luo, Y. Molecular targets of curcumin in breast cancer. Mol. Med. Rep. 2019, 19, 23–29. [Google Scholar] [CrossRef] [Green Version]
  105. Ghosh, S.; Dutta, S.; Sarkar, A.; Kundu, M.; Sil, P.C. Targeted delivery of curcumin in breast cancer cells via hyaluronic acid modified mesoporous silica nanoparticle to enhance anticancer efficiency. Colloids Surfaces B Biointerfaces 2021, 197, 111404. [Google Scholar] [CrossRef] [PubMed]
  106. Song, W.; Su, X.; Gregory, D.A.; Li, W.; Cai, Z.; Zhao, X. Magnetic alginate/chitosan nanoparticles for targeted delivery of curcumin into human breast cancer cells. Nanomaterials 2018, 8, 907. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Kundu, M.; Sadhukhan, P.; Ghosh, N.; Chatterjee, S.; Manna, P.; Das, J.; Sil, P.C. pH-responsive and targeted delivery of curcumin via phenylboronic acid-functionalized ZnO nanoparticles for breast cancer therapy. J. Adv. Res. 2019, 18, 161–172. [Google Scholar] [CrossRef] [PubMed]
  108. Singh, S.P.; Sharma, M.; Gupta, P.K. Cytotoxicity of curcumin silica nanoparticle complexes conjugated with hyaluronic acid on colon cancer cells. Int. J. Biol. Macromol. 2015, 74, 162–170. [Google Scholar] [CrossRef]
  109. Alizadeh, A.M.; Khaniki, M.; Azizian, S.; Mohaghgheghi, M.A.; Sadeghizadeh, M.; Najafi, F. Chemoprevention of azoxymethane-initiated colon cancer in rat by using a novel polymeric nanocarrier–curcumin. Eur. J. Pharmacol. 2012, 689, 226–232. [Google Scholar] [CrossRef] [PubMed]
  110. Goel, A.; Boland, C.R.; Chauhan, D.P. Specific inhibition of cyclooxygenase-2 (COX-2) expression by dietary curcumin in HT-29 human colon cancer cells. Cancer Lett. 2001, 172, 111–118. [Google Scholar] [CrossRef] [Green Version]
  111. Xu, B.; Yu, L.; Zhao, L.-Z. Curcumin up regulates T helper 1 cells in patients with colon cancer. Am. J. Transl. Res. 2017, 9, 1866–1875. [Google Scholar] [PubMed]
  112. Anitha, A.; Deepa, N.; Chennazhi, K.P.; Lakshmanan, V.-K.; Jayakumar, R. Combinatorial anticancer effects of curcumin and 5-fluorouracil loaded thiolated chitosan nanoparticles towards colon cancer treatment. Biochim. Biophys. Acta (BBA) General Subj. 2014, 1840, 2730–2743. [Google Scholar] [CrossRef]
  113. Anitha, A.; Sreeranganathan, M.; Chennazhi, K.P.; Lakshmanan, V.-K.; Jayakumar, R. In vitro combinatorial anticancer effects of 5-fluorouracil and curcumin loaded N, O-carboxymethyl chitosan nanoparticles toward colon cancer and in vivo pharmacokinetic studies. Eur. J. Pharm. Biopharm. 2014, 88, 238–251. [Google Scholar] [CrossRef]
  114. Esmatabadi, M.J.D.; Farhangi, B.; Safari, Z.; Kazerooni, H.; Shirzad, H.; Zolghadr, F.; Sadeghizadeh, M. Dendrosomal curcumin inhibits metastatic potential of human SW480 colon cancer cells through Down-regulation of Claudin1, Zeb1 and Hef1-1 gene expression. Asian Pac. J. Cancer Prev. 2015, 16, 2473–2481. [Google Scholar] [CrossRef] [Green Version]
  115. Chang, T.; Trench, D.; Putnam, J.; Stenzel, M.H.; Lord, M.S. Curcumin-loading-dependent stability of PEGMEMA-based micelles affects endocytosis and exocytosis in colon carcinoma cells. Mol. Pharm. 2016, 13, 924–932. [Google Scholar] [CrossRef]
  116. Waghela, B.N.; Sharma, A.; Dhumale, S.; Pandey, S.M.; Pathak, C. Curcumin conjugated with PLGA potentiates sustainability, anti-proliferative activity and apoptosis in human colon carcinoma cells. PLoS ONE 2015, 10, e0117526. [Google Scholar] [CrossRef] [PubMed]
  117. Weng, W.; Goel, A. Curcumin and colorectal cancer: An update and current perspective on this natural medicine. Semin. Cancer Biol. 2020, 80, 73–86. [Google Scholar] [CrossRef] [PubMed]
  118. Moon, D.-O.; Jin, C.-Y.; Lee, J.-D.; Choi, Y.H.; Ahn, S.-C.; Lee, C.-M.; Jeong, S.-C.; Park, Y.-M.; Kim, G.-Y. Curcumin decreases binding of Shiga-like toxin-1B on human intestinal epithelial cell line HT29 stimulated with TNF-α and IL-1β: Suppression of p38, JNK and NF-κB p65 as potential targets. Biol. Pharm. Bull. 2006, 29, 1470–1475. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Mudduluru, G.; George-William, J.N.; Muppala, S.; Asangani, I.A.; Kumarswamy, R.; Nelson, L.D.; Allgayer, H. Curcumin regulates miR-21 expression and inhibits invasion and metastasis in colorectal cancer. Biosci. Rep. 2011, 31, 185–197. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Wei, X.; Senanayake, T.H.; Bohling, A.; Vinogradov, S.V. Targeted nanogel conjugate for improved stability and cellular permeability of curcumin: Synthesis, pharmacokinetics, and tumor growth inhibition. Mol. Pharm. 2014, 11, 3112–3122. [Google Scholar] [CrossRef] [Green Version]
  121. Yallapu, M.M.; Ebeling, M.C.; Khan, S.; Sundram, V.; Chauhan, N.; Gupta, B.K.; Puumala, S.E.; Jaggi, M.; Chauhan, S.C. Novel curcumin-loaded magnetic nanoparticles for pancreatic cancer treatment. Mol. Cancer Ther. 2013, 12, 1471–1480. [Google Scholar] [CrossRef] [Green Version]
  122. Ali, S.; Ahmad, A.; Banerjee, S.; Padhye, S.; Dominiak, K.; Schaffert, J.M.; Wang, Z.; Philip, P.A.; Sarkar, F.H. Gemcitabine sensitivity can be induced in pancreatic cancer cells through modulation of miR-200 and miR-21 expression by curcumin or its analogue CDF. Cancer Res. 2010, 70, 3606–3617. [Google Scholar] [CrossRef] [Green Version]
  123. Bimonte, S.; Barbieri, A.; Palma, G.; Luciano, A.; Rea, D.; Arra, C. Curcumin inhibits tumor growth and angiogenesis in an orthotopic mouse model of human pancreatic cancer. Biomed Res. Int. 2013, 2013, 810423. [Google Scholar] [CrossRef]
  124. Li, L.; Braiteh, F.S.; Kurzrock, R. Liposome encapsulated curcumin: In vitro and in vivo effects on proliferation, apoptosis, signaling, and angiogenesis. Cancer Interdiscip. Int. J. Am. Cancer Soc. 2005, 104, 1322–1331. [Google Scholar] [CrossRef]
  125. Osterman, C.J.D.; Wall, N.R. Curcumin and pancreatic cancer: A research and clinical update. J. Nat. Sci. 2015, 1, e124. [Google Scholar]
  126. Li, J.; Wang, Y.; Yang, C.; Wang, P.; Oelschlager, D.K.; Zheng, Y.; Tian, D.-A.; Grizzle, W.E.; Buchsbaum, D.J.; Wan, M. Polyethylene glycosylated curcumin conjugate inhibits pancreatic cancer cell growth through inactivation of Jab1. Mol. Pharmacol. 2009, 76, 81–90. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Aggarwal, S.; Ndinguri, M.W.; Solipuram, R.; Wakamatsu, N.; Hammer, R.P.; Ingram, D.; Hansel, W. [DLys6] luteinizing hormone releasing hormone-curcumin conjugate inhibits pancreatic cancer cell growth in vitro and in vivo. Int. J. Cancer 2011, 129, 1611–1623. [Google Scholar] [CrossRef] [Green Version]
  128. Sahu, R.P.; Batra, S.; Srivastava, S.K. Activation of ATM/Chk1 by curcumin causes cell cycle arrest and apoptosis in human pancreatic cancer cells. Br. J. Cancer 2009, 100, 1425–1433. [Google Scholar] [CrossRef] [Green Version]
  129. Zhao, Z.; Li, C.; Xi, H.; Gao, Y.; Xu, D. Curcumin induces apoptosis in pancreatic cancer cells through the induction of forkhead box O1 and inhibition of the PI3K/Akt pathway. Mol. Med. Rep. 2015, 12, 5415–5422. [Google Scholar] [CrossRef] [Green Version]
  130. Lev-Ari, S.; Starr, A.; Vexler, A.; Karaush, V.; Loew, V.; Greif, J.; Fenig, E.; Aderka, D.; Ben-Yosef, R. Inhibition of pancreatic and lung adenocarcinoma cell survival by curcumin is associated with increased apoptosis, down-regulation of COX-2 and EGFR and inhibition of Erk1/2 activity. Anticancer Res. 2006, 26, 4423–4430. [Google Scholar]
  131. Basha, R.; Connelly, S.F.; Sankpal, U.T.; Nagaraju, G.P.; Patel, H.; Vishwanatha, J.K.; Shelake, S.; Tabor-Simecka, L.; Shoji, M.; Simecka, J.W. Small molecule tolfenamic acid and dietary spice curcumin treatment enhances antiproliferative effect in pancreatic cancer cells via suppressing Sp1, disrupting NF-kB translocation to nucleus and cell cycle phase distribution. J. Nutr. Biochem. 2016, 31, 77–87. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Subramaniam, D.; May, R.; Sureban, S.M.; Lee, K.B.; George, R.; Kuppusamy, P.; Ramanujam, R.P.; Hideg, K.; Dieckgraefe, B.K.; Houchen, C.W. Diphenyl difluoroketone: A curcumin derivative with potent in vivo anticancer activity. Cancer Res. 2008, 68, 1962–1969. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Moragoda, L.; Jaszewski, R.; Majumdar, A.P. Curcumin induced modulation of cell cycle and apoptosis in gastric and colon cancer cells. Anticancer Res. 2001, 21, 873–878. [Google Scholar]
  134. Geetha, P.; Sivaram, A.J.; Jayakumar, R.; Mohan, C.G. Integration of in silico modeling, prediction by binding energy and experimental approach to study the amorphous chitin nanocarriers for cancer drug delivery. Carbohydr. Polym. 2016, 142, 240–249. [Google Scholar] [CrossRef]
  135. Qin, H.B.; Wei, L.; Zhang, J.W.; Tang, J.M. Study on functions and mechanism of curcumin in inducing gastric carcinoma BGC apoptosis. Chinese J. Cell. Mol. Immunol. 2011, 27, 1227–1230. [Google Scholar]
  136. Cai, X.Z.; Huang, W.Y.; Qiao, Y.; Du, S.Y.; Chen, Y.; Chen, D.; Yu, S.; Che, R.C.; Liu, N.; Jiang, Y. Inhibitory effects of curcumin on gastric cancer cells: A proteomic study of molecular targets. Phytomedicine 2013, 20, 495–505. [Google Scholar] [CrossRef]
  137. Gupta, N.; Aggarwal, N. Bioavailability enhancement and targeting of stomach tumors using gastro-retentive floating drug delivery system of curcumin—“A technical note”. AAPS PharmSciTech 2008, 9, 810–813. [Google Scholar]
  138. Arya, P.; Pathak, K. Assessing the viability of microsponges as gastro retentive drug delivery system of curcumin: Optimization and pharmacokinetics. Int. J. Pharm. 2014, 460, 1–12. [Google Scholar] [CrossRef]
  139. Dhivya, R.; Ranjani, J.; Bowen, P.K.; Rajendhran, J.; Mayandi, J.; Annaraj, J. Biocompatible curcumin loaded PMMA-PEG/ZnO nanocomposite induce apoptosis and cytotoxicity in human gastric cancer cells. Mater. Sci. Eng. C 2017, 80, 59–68. [Google Scholar] [CrossRef]
  140. Hejazi, R.; Amiji, M. Chitosan-based gastrointestinal delivery systems. J. Control Release 2003, 89, 151–165. [Google Scholar] [CrossRef]
  141. Li, N.; Wang, N.; Wu, T.; Qiu, C.; Wang, X.; Jiang, S.; Zhang, Z.; Liu, T.; Wei, C.; Wang, T. Preparation of curcumin-hydroxypropyl-β-cyclodextrin inclusion complex by cosolvency-lyophilization procedure to enhance oral bioavailability of the drug. Drug Dev. Ind. Pharm. 2018, 44, 1966–1974. [Google Scholar] [CrossRef]
  142. Guerrero, S.; Inostroza-Riquelme, M.; Contreras-Orellana, P.; Diaz-Garcia, V.; Lara, P.; Vivanco-Palma, A.; Cárdenas, A.; Miranda, V.; Robert, P.; Leyton, L. Curcumin-loaded nanoemulsion: A new safe and effective formulation to prevent tumor reincidence and metastasis. Nanoscale 2018, 10, 22612–22622. [Google Scholar] [CrossRef]
  143. Ibrahim, S.; Tagami, T.; Kishi, T.; Ozeki, T. Curcumin marinosomes as promising nano-drug delivery system for lung cancer. Int. J. Pharm. 2018, 540, 40–49. [Google Scholar] [CrossRef]
  144. Muddineti, O.S.; Kumari, P.; Ray, E.; Ghosh, B.; Biswas, S. Curcumin-loaded chitosan—Cholesterol micelles: Evaluation in monolayers and 3D cancer spheroid model. Nanomedicine 2017, 12, 1435–1453. [Google Scholar] [CrossRef]
  145. Kim, K.-C.; Baek, S.-H.; Lee, C. Curcumin-induced downregulation of Axl receptor tyrosine kinase inhibits cell proliferation and circumvents chemoresistance in non-small lung cancer cells. Int. J. Oncol. 2015, 47, 2296–2303. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Chiang, I.; Wang, W.-S.; Liu, H.-C.; Yang, S.-T.; Tang, N.-Y.; Chung, J.-G. Curcumin alters gene expression-associated DNA damage, cell cycle, cell survival and cell migration and invasion in NCI-H460 human lung cancer cells in vitro. Oncol. Rep. 2015, 34, 1853–1874. [Google Scholar] [CrossRef] [PubMed]
  147. Sathuvan, M.; Thangam, R.; Gajendiran, M.; Vivek, R.; Balasubramanian, S.; Nagaraj, S.; Gunasekaran, P.; Madhan, B.; Rengasamy, R. κ-Carrageenan: An effective drug carrier to deliver curcumin in cancer cells and to induce apoptosis. Carbohydr. Polym. 2017, 160, 184–193. [Google Scholar] [CrossRef]
  148. Tunc, D.; Dere, E.; Karakas, D.; Cevatemre, B.; Yilmaz, V.T.; Ulukaya, E. Cytotoxic and apoptotic effects of the combination of palladium (II) 5, 5-diethylbarbiturate complex with bis (2-pyridylmethyl) amine and curcumin on non small lung cancer cell lines. Bioorg. Med. Chem. 2017, 25, 1717–1723. [Google Scholar] [CrossRef] [PubMed]
  149. Jyoti, K.; Pandey, R.S.; Kush, P.; Kaushik, D.; Jain, U.K.; Madan, J. Inhalable bioresponsive chitosan microspheres of doxorubicin and soluble curcumin augmented drug delivery in lung cancer cells. Int. J. Biol. Macromol. 2017, 98, 50–58. [Google Scholar] [CrossRef] [PubMed]
  150. Li, S.; Fang, C.; Zhang, J.; Liu, B.; Wei, Z.; Fan, X.; Sui, Z.; Tan, Q. Catanionic lipid nanosystems improve pharmacokinetics and anti-lung cancer activity of curcumin. Nanomed. Nanotechnol. Biol. Med. 2016, 12, 1567–1579. [Google Scholar] [CrossRef] [PubMed]
  151. Luo, C.-Q.; Xing, L.; Cui, P.-F.; Qiao, J.-B.; He, Y.-J.; Chen, B.-A.; Jin, L.; Jiang, H.-L. Curcumin-coordinated nanoparticles with improved stability for reactive oxygen species-responsive drug delivery in lung cancer therapy. Int. J. Nanomed. 2017, 12, 855–869. [Google Scholar] [CrossRef] [Green Version]
  152. Paunovic, V.; Ristic, B.; Markovic, Z.; Todorovic-Markovic, B.; Kosic, M.; Prekodravac, J.; Kravic-Stevovic, T.; Martinovic, T.; Micusik, M.; Spitalsky, Z. c-Jun N-terminal kinase-dependent apoptotic photocytotoxicity of solvent exchange-prepared curcumin nanoparticles. Biomed. Microdevices 2016, 18, 37. [Google Scholar] [CrossRef]
  153. Chen, H.; Chen, L.; Wang, L.; Zhou, X.; Chan, J.Y.-W.; Li, J.; Cui, G.; Lee, S.M.-Y. Synergistic effect of fenretinide and curcumin for treatment of non-small cell lung cancer. Cancer Biol. Ther. 2016, 17, 1022–1029. [Google Scholar] [CrossRef] [Green Version]
  154. Hosseinzadehdehkordi, M.; Adelinik, A.; Tashakor, A. Dual effect of curcumin targets reactive oxygen species, adenosine triphosphate contents and intermediate steps of mitochondria-mediated apoptosis in lung cancer cell lines. Eur. J. Pharmacol. 2015, 769, 203–210. [Google Scholar] [CrossRef]
  155. Srivastava, S.; Mohammad, S.; Pant, A.B.; Mishra, P.R.; Pandey, G.; Gupta, S.; Farooqui, S. Co-delivery of 5-fluorouracil and curcumin nanohybrid formulations for improved chemotherapy against oral squamous cell carcinoma. J. Maxillofac. Oral Surg. 2018, 17, 597–610. [Google Scholar] [CrossRef] [PubMed]
  156. Srivastava, R.; Kundu, A.; Pradhan, D.; Jyoti, B.; Chokotiya, H.; Parashar, P. A Comparative Study to Evaluate the Efficacy of Curcumin Lozenges (TurmNova®) and Intralesional Corticosteroids with Hyaluronidase in Management of Oral Submucous Fibrosis. J. Contemp. Dent. Pract. 2021, 22, 751–755. [Google Scholar] [CrossRef] [PubMed]
  157. Singh, S.P.; Sharma, M.; Gupta, P.K. Enhancement of phototoxicity of curcumin in human oral cancer cells using silica nanoparticles as delivery vehicle. Lasers Med. Sci. 2014, 29, 645–652. [Google Scholar] [CrossRef]
  158. De Souza Ferreira, S.B.; Slowik, K.M.; de Castro Hoshino, L.V.; Baesso, M.L.; Murdoch, C.; Colley, H.E.; Bruschi, M.L. Mucoadhesive emulgel systems containing curcumin for oral squamous cell carcinoma treatment: From pre-formulation to cytotoxicity in tissue-engineering oral mucosa. Eur. J. Pharm. Sci. 2020, 151, 105372. [Google Scholar]
  159. Ardito, F.; Perrone, D.; Giuliani, M.; Testa, N.F.; Muzio, L.L. Effects of curcumin on squamous cell carcinoma of tongue: An in vitro study. Curr. Top. Med. Chem. 2018, 18, 233–243. [Google Scholar] [CrossRef]
  160. Ba, P.; Xu, M.; Yu, M.; Li, L.; Duan, X.; Lv, S.; Fu, G.; Yang, J.; Yang, P.; Yang, C. Curcumin suppresses the proliferation and tumorigenicity of Cal27 by modulating cancer associated fibroblasts of TSCC. Oral Dis. 2020, 26, 1375–1383. [Google Scholar] [CrossRef]
  161. Jose, A.; Labala, S.; Ninave, K.M.; Gade, S.K.; Venuganti, V.V.K. Effective skin cancer treatment by topical co-delivery of curcumin and STAT3 siRNA using cationic liposomes. AAPS PharmSciTech 2018, 19, 166–175. [Google Scholar] [CrossRef] [PubMed]
  162. Peram, M.R.; Jalalpure, S.; Kumbar, V.; Patil, S.; Joshi, S.; Bhat, K.; Diwan, P. Factorial design based curcumin ethosomal nanocarriers for the skin cancer delivery: In vitro evaluation. J. Liposome Res. 2019, 29, 291–311. [Google Scholar] [CrossRef]
  163. Wang, L.; Zhang, B.; Huang, F.; Liu, B.; Xie, Y. Curcumin inhibits lipolysis via suppression of ER stress in adipose tissue and prevents hepatic insulin resistance. J. Lipid Res. 2016, 57, 1243–1255. [Google Scholar] [CrossRef] [Green Version]
  164. Bradford, P.G. Curcumin and obesity. Biofactors 2013, 39, 78–87. [Google Scholar] [CrossRef]
  165. Gonzales, A.M.; Orlando, R.A. Curcumin and resveratrol inhibit nuclear factor-kappaB-mediated cytokine expression in adipocytes. Nutr. Metab. 2008, 5, 17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Woo, H.-M.; Kang, J.-H.; Kawada, T.; Yoo, H.; Sung, M.-K.; Yu, R. Active spice-derived components can inhibit inflammatory responses of adipose tissue in obesity by suppressing inflammatory actions of macrophages and release of monocyte chemoattractant protein-1 from adipocytes. Life Sci. 2007, 80, 926–931. [Google Scholar] [CrossRef] [PubMed]
  167. Goyal, S.N.; Reddy, N.M.; Patil, K.R.; Nakhate, K.T.; Ojha, S.; Patil, C.R.; Agrawal, Y.O. Challenges and issues with streptozotocin-induced diabetes—A clinically relevant animal model to understand the diabetes pathogenesis and evaluate therapeutics. Chem. Biol. Interact. 2016, 244, 49–63. [Google Scholar] [CrossRef] [PubMed]
  168. Den Hartogh, D.J.; Gabriel, A.; Tsiani, E. Antidiabetic properties of curcumin II: Evidence from in vivo studies. Nutrients 2019, 12, 58. [Google Scholar] [CrossRef] [Green Version]
  169. Kottaisamy, C.P.D.; Raj, D.S.; Prasanth Kumar, V.; Sankaran, U. Experimental animal models for diabetes and its related complications—A review. Lab. Anim. Res. 2021, 37, 23. [Google Scholar] [CrossRef]
  170. Suresh Babu, P.; Srinivasan, K. Amelioration of renal lesions associated with diabetes by dietary curcumin in streptozotocin diabetic rats. Mol. Cell. Biochem. 1998, 181, 87–96. [Google Scholar] [CrossRef]
  171. Hussain, A. Hypoglycemic, hypolipidemic and antioxidant properties of combination ofCurcumin from Curcuma longa, Linn, and partially purified product fromAbroma augusta, Linn. in streptozotocin induced diabetes. Indian J. Clin. Biochem. 2002, 17, 33–43. [Google Scholar] [CrossRef] [Green Version]
  172. Murugan, P.; Pari, L. Antioxidant effect of tetrahydrocurcumin in streptozotocin-nicotinamideinduced diabetic rats. Life Sci. 2006, 79, 1720–1728. [Google Scholar] [CrossRef]
  173. Sharma, S.; Kulkarni, S.K.; Chopra, K. Curcumin, the active principle of turmeric (Curcuma longa), ameliorates diabetic nephropathy in rats. Clin. Exp. Pharmacol. Physiol. 2006, 33, 940–945. [Google Scholar] [CrossRef]
  174. Suryanarayana, P.; Satyanarayana, A.; Balakrishna, N.; Kumar, P.U.; Reddy, G.B. Effect of turmeric and curcumin on oxidative stress and antioxidant enzymes in streptozotocin-induced diabetic rat. Med. Sci. Monit. 2007, 13, BR286–BR292. [Google Scholar]
  175. Tikoo, K.; Meena, R.L.; Kabra, D.G.; Gaikwad, A.B. Change in post translational modifications of histone H3, heat shock protein 27 and MAP kinase p38 expression by curcumin in streptozotocin induced type I diabetic nephropathy. Br. J. Pharmacol. 2008, 153, 1225–1231. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Kanitkar, M.; Gokhale, K.; Galande, S.; Bhonde, R.R. Novel role of curcumin in the prevention of cytokine induced islet death in vitro and diabetogenesis in vivo. Br. J. Pharmacol. 2008, 155, 702–713. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Pari, L.; Karthikesan, K.; Menon, V.P. Comparative and combined effect of chlorogenic acid and tetrahydrocurcumin on antioxidant disparities in chemical induced experimental diabetes. Mol. Cell. Biochem. 2010, 341, 109–117. [Google Scholar] [CrossRef]
  178. El-Azab, M.F.; Attia, F.M.; El-Mowafy, A.M. Novel role of curcumin combined with bone marrow transplantation in reversing experimental diabetes: Effects on pancreatic islet regeneration, oxidative stress, and inflammatory cytokines. Eur. J. Pharmacol. 2011, 658, 41–48. [Google Scholar] [CrossRef]
  179. Arun, N.; Nalini, N. Efficacy of turmeric on blood sugar and polyol pathway in diabetic albino rats. Plant Foods Hum. Nutr. 2002, 57, 41–52. [Google Scholar] [CrossRef] [PubMed]
  180. Das, K.K.; Razzaghi-Asl, N.; Tikare, S.N.; Di Santo, R.; Costi, R.; Messore, A.; Pescatori, L.; Crucitti, G.C.; Jargar, J.G.; Dhundasi, S.A. Hypoglycemic activity of curcumin synthetic analogues in alloxan-induced diabetic rats. J. Enzyme Inhib. Med. Chem. 2016, 31, 99–105. [Google Scholar] [CrossRef] [Green Version]
  181. Ghosh, S.S.; Gehr, T.W.B.; Ghosh, S. Curcumin and chronic kidney disease (CKD): Major mode of action through stimulating endogenous intestinal alkaline phosphatase. Molecules 2014, 19, 20139–20156. [Google Scholar] [CrossRef] [Green Version]
  182. Ali, B.H.; Al Salam, S.; Al Suleimani, Y.; Al Kalbani, J.; Al Bahlani, S.; Ashique, M.; Manoj, P.; Al Dhahli, B.; Al Abri, N.; Naser, H.T. Curcumin ameliorates kidney function and oxidative stress in experimental chronic kidney disease. Basic Clin. Pharmacol. Toxicol. 2018, 122, 65–73. [Google Scholar] [CrossRef]
  183. Soetikno, V.; Sari, F.R.; Veeraveedu, P.T.; Thandavarayan, R.A.; Harima, M.; Sukumaran, V.; Lakshmanan, A.P.; Suzuki, K.; Kawachi, H.; Watanabe, K. Curcumin ameliorates macrophage infiltration by inhibiting NF- B activation and proinflammatory cytokines in streptozotocin induced-diabetic nephropathy. Nutr. Metab. 2011, 8, 35. [Google Scholar] [CrossRef] [Green Version]
  184. Wang, Y.; Wang, Y.; Luo, M.; Wu, H.; Kong, L.; Xin, Y.; Cui, W.; Zhao, Y.; Wang, J.; Liang, G. Novel curcumin analog C66 prevents diabetic nephropathy via JNK pathway with the involvement of p300/CBP-mediated histone acetylation. Biochim. Biophys. Acta (BBA) Molecular Basis Dis. 2015, 1852, 34–46. [Google Scholar] [CrossRef] [Green Version]
  185. Ortega-Domínguez, B.; Aparicio-Trejo, O.E.; García-Arroyo, F.E.; León-Contreras, J.C.; Tapia, E.; Molina-Jijón, E.; Hernández-Pando, R.; Sánchez-Lozada, L.G.; Barrera-Oviedo, D.; Pedraza-Chaverri, J. Curcumin prevents cisplatin-induced renal alterations in mitochondrial bioenergetics and dynamic. Food Chem. Toxicol. 2017, 107, 373–385. [Google Scholar] [CrossRef] [PubMed]
  186. Avila-Rojas, S.H.; Lira-León, A.; Aparicio-Trejo, O.E.; Reyes-Fermín, L.M.; Pedraza-Chaverri, J. Role of autophagy on heavy metal-induced renal damage and the protective effects of curcumin in autophagy and kidney preservation. Medicina 2019, 55, 360. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. 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]
  188. Huang, W.J.; Zhang, X.; Chen, W.W. Role of oxidative stress in Alzheimer’s disease (review). Biomed. Rep. 2016, 4, 519–522. [Google Scholar] [CrossRef] [Green Version]
  189. Kattoor, A.J.; Pothineni, N.V.K.; Palagiri, D.; Mehta, J.L. Oxidative Stress in Atherosclerosis. Curr. Atheroscler. Rep. 2017, 19, 42. [Google Scholar] [CrossRef]
  190. Zia, A.; Farkhondeh, T.; Pourbagher-Shahri, A.M.; Samarghandian, S. The role of curcumin in aging and senescence: Molecular mechanisms. Biomed. Pharmacother. 2021, 134, 111119. [Google Scholar] [CrossRef]
  191. Menon, V.P.; Sudheer, A.R. Antioxidant and anti-inflammatory properties of curcumin. Adv. Exp. Med. Biol. 2007, 595, 105–125. [Google Scholar] [CrossRef]
  192. Khopde, S.M.; Priyadarsini, K.I.; Venkatesan, P.; Rao, M.N.A. Free radical scavenging ability and antioxidant efficiency of curcumin and its substituted analogue. Biophys. Chem. 1999, 80, 85–91. [Google Scholar] [CrossRef]
  193. Ak, T.; Gülçin, I. Antioxidant and radical scavenging properties of curcumin. Chem. Biol. Interact. 2008, 174, 27–37. [Google Scholar] [CrossRef]
  194. Jovanovic, S.V.; Steenken, S.; Boone, C.W.; Simic, M.G. H-Atom Transfer Is A Preferred Antioxidant Mechanism of Curcumin. J. Am. Chem. Soc. 1999, 121, 9677–9681. [Google Scholar] [CrossRef]
  195. Wright, J.S. Predicting the antioxidant activity of curcumin and curcuminoids. J. Mol. Struct. Theochem 2002, 591, 207–217. [Google Scholar] [CrossRef]
  196. Priyadarsini, K.I.; Maity, D.K.; Naik, G.H.; Kumar, M.S.; Unnikrishnan, M.K.; Satav, J.G.; Mohan, H. Role of phenolic O-H and methylene hydrogen on the free radical reactions and antioxidant activity of curcumin. Free Radic. Biol. Med. 2003, 35, 475–484. [Google Scholar] [CrossRef]
  197. Sharma, R.A.; Gescher, A.J.; Steward, W.P. Curcumin: The story so far. Eur. J. Cancer 2005, 41, 1955–1968. [Google Scholar] [CrossRef]
  198. Faten, R.A.; Ibrahim, A.E.; Khaled, A.E. Protective and modulatory effects of Curcumin and L-Carnitine against Methotrexate-induced Oxidative stress in albino rats. Res. J. Pharm. Biol. Chem. Sci. 2013, 4, 744–754. [Google Scholar]
  199. Balamurugan, A.N.; Akhov, L.; Selvaraj, G.; Pugazhenthi, S. Induction of antioxidant enzymes by curcumin and its analogues in human islets: Implications in transplantation. Pancreas 2009, 38, 454–460. [Google Scholar] [CrossRef]
  200. Meghana, K.; Sanjeev, G.; Ramesh, B. Curcumin prevents streptozotocin-induced islet damage by scavenging free radicals: A prophylactic and protective role. Eur. J. Pharmacol. 2007, 577, 183–191. [Google Scholar] [CrossRef]
  201. Balogun, E.; Hoque, M.; Gong, P.; Killeen, E.; Green, C.J.; Foresti, R.; Alam, J.; Motterlini, R. Curcumin activates the haem oxygenase-1 gene via regulation of Nrf2 and the antioxidant-responsive element. Biochem. J. 2003, 371, 887–895. [Google Scholar] [CrossRef] [Green Version]
  202. Forouzanfar, F.; Barreto, G.; Majeed, M.; Sahebkar, A. Modulatory effects of curcumin on heat shock proteins in cancer: A promising therapeutic approach. BioFactors 2019, 45, 631–640. [Google Scholar] [CrossRef] [PubMed]
  203. Muhammad, I.; Wang, H.; Sun, X.; Wang, X.; Han, M.; Lu, Z.; Cheng, P.; Hussain, M.A.; Zhang, X. Dual role of dietary curcumin through attenuating AFB1-induced oxidative stress and liver injury via modulating liver phase-I and phase-II enzymes involved in AFB1 bioactivation and detoxification. Front. Pharmacol. 2018, 9, 1–10. [Google Scholar] [CrossRef] [Green Version]
  204. Garg, R.; Gupta, S.; Maru, G.B. Dietary curcumin modulates transcriptional regulators of phase I and phase II enzymes in benzo[a]pyrene-treated mice: Mechanism of its anti-initiating action. Carcinogenesis 2008, 29, 1022–1032. [Google Scholar] [CrossRef] [Green Version]
  205. Nayak, S.; Sashidhar, R.B. Metabolic intervention of aflatoxin B1 toxicity by curcumin. J. Ethnopharmacol. 2010, 127, 641–644. [Google Scholar] [CrossRef]
  206. Limaye, A.; Yu, R.C.; Chou, C.C.; Liu, J.R.; Cheng, K.C. Protective and detoxifying effects conferred by dietary selenium and curcumin against AFB1-mediated toxicity in livestock: A review. Toxins 2018, 10, 25. [Google Scholar] [CrossRef] [Green Version]
  207. Pignanelli, C.; Ma, D.; Noel, M.; Ropat, J.; Mansour, F.; Curran, C.; Pupulin, S.; Larocque, K.; Wu, J.; Liang, G.; et al. Selective Targeting of Cancer Cells by Oxidative Vulnerabilities with Novel Curcumin Analogs. Sci. Rep. 2017, 7, 1105. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  208. Nakamae, I.; Morimoto, T.; Shima, H.; Shionyu, M.; Fujiki, H.; Yoneda-Kato, N.; Yokoyama, T.; Kanaya, S.; Kakiuchi, K.; Shirai, T.; et al. Curcumin derivatives verify the essentiality of ROS upregulation in tumor suppression. Molecules 2019, 24, 4067. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  209. Larasati, Y.A.; Yoneda-Kato, N.; Nakamae, I.; Yokoyama, T.; Meiyanto, E.; Kato, J.Y. Curcumin targets multiple enzymes involved in the ROS metabolic pathway to suppress tumor cell growth. Sci. Rep. 2018, 8, 2039. [Google Scholar] [CrossRef] [PubMed]
  210. Araveti, P.B.; Srivastava, A. Curcumin induced oxidative stress causes autophagy and apoptosis in bovine leucocytes transformed by Theileria annulata. Cell Death Discov. 2019, 5, 100. [Google Scholar] [CrossRef] [Green Version]
  211. Kunwar, A.; Barik, A.; Mishra, B.; Rathinasamy, K.; Pandey, R.; Priyadarsini, K.I. Quantitative cellular uptake, localization and cytotoxicity of curcumin in normal and tumor cells. Biochim. Biophys. Acta Gen. Subj. 2008, 1780, 673–679. [Google Scholar] [CrossRef]
  212. Contreras, J.L.; Smyth, C.A.; Bilbao, G.; Eckstein, C.; Young, C.J.; Thompson, J.A.; Curiel, D.T.; Eckhoff, D.E. Coupling endoplasmic reticulum stress to cell death program in isolated human pancreatic islets: Effects of gene transfer of Bcl 2. Transpl. Int. 2003, 16, 537–542. [Google Scholar] [CrossRef]
  213. Ye, M.; Qiu, H.; Cao, Y.; Zhang, M.; Mi, Y.; Yu, J.; Wang, C. Curcumin improves palmitate-induced insulin resistance in human umbilical vein endothelial cells by maintaining proteostasis in endoplasmic reticulum. Front. Pharmacol. 2017, 8, 148. [Google Scholar] [CrossRef] [Green Version]
  214. Wang, N.; Wang, F.; Gao, Y.; Yin, P.; Pan, C.; Liu, W.; Zhou, Z.; Wang, J. Curcumin protects human adipose-derived mesenchymal stem cells against oxidative stress-induced inhibition of osteogenesis. J. Pharmacol. Sci. 2016, 132, 192–200. [Google Scholar] [CrossRef] [Green Version]
  215. Li, Y.; Li, J.; Li, S.; Li, Y.; Wang, X.; Liu, B.; Fu, Q.; Ma, S. Curcumin attenuates glutamate neurotoxicity in the hippocampus by suppression of ER stress-associated TXNIP/NLRP3 inflammasome activation in a manner dependent on AMPK. Toxicol. Appl. Pharmacol. 2015, 286, 53–63. [Google Scholar] [CrossRef] [PubMed]
  216. Back, S.H.; Kaufman, R.J. Endoplasmic reticulum stress and type 2 diabetes. Annu. Rev. Biochem. 2012, 81, 767–793. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Rashid, K.; Sil, P.C. Curcumin enhances recovery of pancreatic islets from cellular stress induced inflammation and apoptosis in diabetic rats. Toxicol. Appl. Pharmacol. 2015, 282, 297–310. [Google Scholar] [CrossRef] [PubMed]
  218. Huang, A.C.; Chang, C.L.; Yu, C.S.; Chen, P.Y.; Yang, J.S.; Ji, B.C.; Lin, T.P.; Chiu, C.F.; Yeh, S.P.; Huang, Y.P. Induction of apoptosis by curcumin in murine myelomonocytic leukemia WEHI 3 cells is mediated via endoplasmic reticulum stress and mitochondria dependent pathways. Environ. Toxicol. 2013, 28, 255–266. [Google Scholar] [CrossRef]
  219. Kocyigit, A.; Guler, E.M. Curcumin induce DNA damage and apoptosis through generation of reactive oxygen species and reducing mitochondrial membrane potential in melanoma cancer cells. Cell. Mol. Biol. 2017, 63, 97–105. [Google Scholar] [CrossRef]
  220. Lin, S.-S.; Huang, H.-P.; Yang, J.-S.; Wu, J.-Y.; Hsai, T.-C.; Lin, C.-C.; Lin, C.-W.; Kuo, C.-L.; Wood, W.G.; Chung, J.-G. DNA damage and endoplasmic reticulum stress mediated curcumin-induced cell cycle arrest and apoptosis in human lung carcinoma A-549 cells through the activation caspases cascade-and mitochondrial-dependent pathway. Cancer Lett. 2008, 272, 77–90. [Google Scholar] [CrossRef]
  221. Pae, H.-O.; Jeong, S.-O.; Jeong, G.-S.; Kim, K.M.; Kim, H.S.; Kim, S.-A.; Kim, Y.-C.; Kang, S.-D.; Kim, B.-N.; Chung, H.-T. Curcumin induces pro-apoptotic endoplasmic reticulum stress in human leukemia HL-60 cells. Biochem. Biophys. Res. Commun. 2007, 353, 1040–1045. [Google Scholar] [CrossRef]
  222. Qin, J.; Li, R.; Raes, J.; Arumugam, M.; Burgdorf, K.S.; Manichanh, C.; Nielsen, T.; Pons, N.; Levenez, F.; Yamada, T. A human gut microbial gene catalogue established by metagenomic sequencing. Nature 2010, 464, 59–65. [Google Scholar] [CrossRef] [Green Version]
  223. Booijink, C.C.G.M.; El-Aidy, S.; Rajilić-Stojanović, M.; Heilig, H.G.H.J.; Troost, F.J.; Smidt, H.; Kleerebezem, M.; De Vos, W.M.; Zoetendal, E.G. High temporal and inter-individual variation detected in the human ileal microbiota. Environ. Microbiol. 2010, 12, 3213–3227. [Google Scholar] [CrossRef]
  224. Zoetendal, E.G.; Raes, J.; Van Den Bogert, B.; Arumugam, M.; Booijink, C.C.G.M.; Troost, F.J.; Bork, P.; Wels, M.; De Vos, W.M.; Kleerebezem, M. The human small intestinal microbiota is driven by rapid uptake and conversion of simple carbohydrates. ISME J. 2012, 6, 1415–1426. [Google Scholar] [CrossRef]
  225. 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 (BBA) General Subj. 2017, 1861, 1801–1812. [Google Scholar] [CrossRef] [PubMed]
  226. Zhang, Z.; Chen, Y.; Xiang, L.; Wang, Z.; Xiao, G.G.; Hu, J. Effect of curcumin on the diversity of gut microbiota in ovariectomized rats. Nutrients 2017, 9, 1146. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  227. 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] [Green Version]
  228. Rashmi, R.; Kumar, T.R.S.; Karunagaran, D. Human colon cancer cells differ in their sensitivity to curcumin-induced apoptosis and heat shock protects them by inhibiting the release of apoptosis-inducing factor and caspases. FEBS Lett. 2003, 538, 19–24. [Google Scholar] [CrossRef]
  229. Ohno, M.; Nishida, A.; Sugitani, Y.; Nishino, K.; Inatomi, O.; Sugimoto, M.; Kawahara, M.; Andoh, A. Nanoparticle curcumin ameliorates experimental colitis via modulation of gut microbiota and induction of regulatory T cells. PLoS ONE 2017, 12, e0185999. [Google Scholar] [CrossRef] [Green Version]
  230. Khoury, T.; Rmeileh, A.A.; Yosha, L.; Benson, A.A.; Daher, S.; Mizrahi, M. Drug induced liver injury: Review with a focus on genetic factors, tissue diagnosis, and treatment options. J. Clin. Transl. Hepatol. 2015, 3, 99–108. [Google Scholar] [PubMed]
  231. Epstein, J.; Docena, G.; MacDonald, T.T.; Sanderson, I.R. Curcumin suppresses p38 mitogen-activated protein kinase activation, reduces IL-1 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]
  232. Vecchi Brumatti, L.; Marcuzzi, A.; Tricarico, P.M.; Zanin, V.; Girardelli, M.; Bianco, A.M. Curcumin and inflammatory bowel disease: Potential and limits of innovative treatments. Molecules 2014, 19, 21127–21153. [Google Scholar] [CrossRef] [Green Version]
  233. Shakibaei, M.; John, T.; Schulze-Tanzil, G.; Lehmann, I.; Mobasheri, A. Suppression of NF- B activation by curcumin leads to inhibition of expression of cyclo-oxygenase-2 and matrix metalloproteinase-9 in human articular chondrocytes: Implications for the treatment of osteoarthritis. Biochem. Pharmacol. 2007, 73, 1434–1445. [Google Scholar] [CrossRef]
  234. Martelli, L.; Ragazzi, E.; Di Mario, F.; Martelli, M.; Castagliuolo, I.; Dal Maschio, M.; Palù, G.; Maschietto, M.; Scorzeto, M.; Vassanelli, S. A potential role for the vanilloid receptor TRPV1 in the therapeutic effect of curcumin in dinitrobenzene sulphonic acid-induced colitis in mice. Neurogastroenterol. Motil. 2007, 19, 668–674. [Google Scholar] [CrossRef]
  235. Karthikeyan, A.; Young, K.N.; Moniruzzaman, M.; Beyene, A.M.; Do, K.; Kalaiselvi, S.; Min, T. Curcumin and Its Modified Formulations on Inflammatory Bowel Disease (IBD): The Story So Far and Future Outlook. Pharmaceutics 2021, 13, 484. [Google Scholar] [CrossRef] [PubMed]
  236. Rudrappa, T.; Bais, H.P. Curcumin, a known phenolic from Curcuma longa, attenuates the virulence of Pseudomonas aeruginosa PAO1 in whole plant and animal pathogenicity models. J. Agric. Food Chem. 2008, 56, 1955–1962. [Google Scholar] [CrossRef] [PubMed]
  237. Ammayappan, L.; Moses, J.J. Study of antimicrobial activity of aloevera, chitosan, and curcumin on cotton, wool, and rabbit hair. Fibers Polym. 2009, 10, 161–166. [Google Scholar] [CrossRef]
  238. Zorofchian Moghadamtousi, S.; Abdul Kadir, H.; Hassandarvish, P.; Tajik, H.; Abubakar, S.; Zandi, K. A review on antibacterial, antiviral, and antifungal activity of curcumin. Biomed Res. Int. 2014, 2014, 186864. [Google Scholar] [CrossRef]
  239. Chen, D.Y.; Shien, J.H.; Tiley, L.; Chiou, S.S.; Wang, S.Y.; Chang, T.J.; Lee, Y.J.; Chan, K.W.; Hsu, W.L. Curcumin inhibits influenza virus infection and haemagglutination activity. Food Chem. 2010, 119, 1346–1351. [Google Scholar] [CrossRef]
  240. Ingolfsson, H.I.; Koeppe, R.E.; Andersen, O.S. Curcumin is a Modulator of Bilayer Material Properties. Biochemistry 2007, 46, 10384–10391. [Google Scholar] [CrossRef] [PubMed]
  241. Colpitts, C.C.; Schang, L.M.; Rachmawati, H.; Frentzen, A.; Pfaender, S.; Behrendt, P.; Brown, R.J.P.; Bankwitz, D.; Steinmann, J.; Ott, M. Turmeric curcumin inhibits entry of all hepatitis C virus genotypes into human liver cells. Gut 2014, 63, 1137–1149. [Google Scholar]
  242. Chen, T.-Y.; Chen, D.-Y.; Wen, H.-W.; Ou, J.-L.; Chiou, S.-S.; Chen, J.-M.; Wong, M.-L.; Hsu, W.-L. Inhibition of enveloped viruses infectivity by curcumin. PLoS ONE 2013, 8, e62482. [Google Scholar] [CrossRef] [Green Version]
  243. Kim, H.J.; Yoo, H.S.; Kim, J.C.; Park, C.S.; Choi, M.S.; Kim, M.; Choi, H.; Min, J.S.; Kim, Y.S.; Yoon, S.W.; et al. Antiviral effect of Curcuma longa Linn extract against hepatitis B virus replication. J. Ethnopharmacol. 2009, 124, 189–196. [Google Scholar] [CrossRef]
  244. Narayanan, A.; Kehn-Hall, K.; Senina, S.; Lundberg, L.; Van Duyne, R.; Guendel, I.; Das, R.; Baer, A.; Bethel, L.; Turell, M.; et al. Curcumin inhibits rift valley fever virus replication in human cells. J. Biol. Chem. 2012, 287, 33198–33214. [Google Scholar] [CrossRef] [Green Version]
  245. Ferreira, V.H.; Nazli, A.; Dizzell, S.E.; Mueller, K.; Kaushic, C. The anti-inflammatory activity of curcumin protects the genital mucosal epithelial barrier from disruption and blocks replication of HIV-1 and HSV-2. PLoS ONE 2015, 10, e0124903. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  246. Sui, Z.; Salto, R.; Li, J.; Craik, C.; de Montellano, P.R.O. 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]
  247. Ali, A.; Banerjea, A.C. Curcumin inhibits HIV-1 by promoting Tat protein degradation. Sci. Rep. 2016, 6, 27539. [Google Scholar] [CrossRef] [PubMed]
  248. Kumari, N.; Kulkarni, A.A.; Lin, X.; McLean, C.; Ammosova, T.; Ivanov, A.; Hipolito, M.; Nekhai, S.; Nwulia, E. Inhibition of HIV-1 by curcumin A, a novel curcumin analog. Drug Des. Devel. Ther. 2015, 9, 5051–5060. [Google Scholar]
  249. Devadas, K.; Hewlett, I.K.; Dhawan, S. Lipopolysaccharide suppresses HIV 1 replication in human monocytes by protein kinase C dependent heme oxygenase 1 induction. J. Leukoc. Biol. 2010, 87, 915–924. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  250. Sharma, R.K.; Cwiklinski, K.; Aalinkeel, R.; Reynolds, J.L.; Sykes, D.E.; Quaye, E.; Oh, J.; Mahajan, S.D.; Schwartz, S.A. Immunomodulatory activities of curcumin-stabilized silver nanoparticles: Efficacy as an antiretroviral therapeutic. Immunol. Investig. 2017, 46, 833–846. [Google Scholar] [CrossRef]
  251. Wen, C.-C.; Kuo, Y.-H.; Jan, J.-T.; Liang, P.-H.; Wang, S.-Y.; Liu, H.-G.; Lee, C.-K.; Chang, S.-T.; Kuo, C.-J.; Lee, S.-S. Specific plant terpenoids and lignoids possess potent antiviral activities against severe acute respiratory syndrome coronavirus. J. Med. Chem. 2007, 50, 4087–4095. [Google Scholar] [CrossRef] [Green Version]
  252. Ferreira, L.L.C.; Abreu, M.P.; Costa, C.B.; Leda, P.O.; Behrens, M.D.; Dos Santos, E.P. Curcumin and Its Analogs as a Therapeutic Strategy in Infections Caused by RNA Genome Viruses. Food Environ. Virol. 2022, 14, 120–137. [Google Scholar] [CrossRef]
  253. Avasarala, S.; Zhang, F.; Liu, G.; Wang, R.; London, S.D.; London, L. Curcumin modulates the inflammatory response and inhibits subsequent fibrosis in a mouse model of viral-induced acute respiratory distress syndrome. PLoS ONE 2013, 8, e57285. [Google Scholar] [CrossRef] [Green Version]
  254. Xu, X.; Cai, Y.; Yu, Y. Effects of a novel curcumin derivative on the functions of kidney in streptozotocin-induced type 2 diabetic rats. Inflammopharmacology 2018, 26, 1257–1264. [Google Scholar] [CrossRef] [Green Version]
  255. Sun, X.; Liu, Y.; Li, C.; Wang, X.; Zhu, R.; Liu, C.; Liu, H.; Wang, L.; Ma, R.; Fu, M. Recent advances of curcumin in the prevention and treatment of renal fibrosis. Biomed Res. Int. 2017, 2017, 2418671. [Google Scholar] [CrossRef] [PubMed]
  256. Dai, J.; Gu, L.; Su, Y.; Wang, Q.; Zhao, Y.; Chen, X.; Deng, H.; Li, W.; Wang, G.; Li, K. Inhibition of curcumin on influenza A virus infection and influenzal pneumonia via oxidative stress, TLR2/4, p38/JNK MAPK and NF-κB pathways. Int. Immunopharmacol. 2018, 54, 177–187. [Google Scholar] [CrossRef] [PubMed]
  257. Han, S.; Xu, J.; Guo, X.; Huang, M. Curcumin ameliorates severe influenza pneumonia via attenuating lung injury and regulating macrophage cytokines production. Clin. Exp. Pharmacol. Physiol. 2018, 45, 84–93. [Google Scholar] [CrossRef]
  258. Richart, S.M.; Li, Y.-L.; Mizushina, Y.; Chang, Y.-Y.; Chung, T.-Y.; Chen, G.-H.; Tzen, J.T.-C.; Shia, K.-S.; Hsu, W.-L. Synergic effect of curcumin and its structural analogue (Monoacetylcurcumin) on anti-influenza virus infection. J. Food Drug Anal. 2018, 26, 1015–1023. [Google Scholar] [CrossRef] [Green Version]
  259. Lai, Y.; Yan, Y.; Liao, S.; Li, Y.; Ye, Y.; Liu, N.; Zhao, F.; Xu, P. 3D-quantitative structure—Activity relationship and antiviral effects of curcumin derivatives as potent inhibitors of influenza H1N1 neuraminidase. Arch. Pharm. Res. 2020, 43, 489–502. [Google Scholar] [CrossRef]
  260. Shaikh, J.; Ankola, D.D.; Beniwal, V.; Singh, D.; Kumar, M.N.V.R. Nanoparticle encapsulation improves oral bioavailability of curcumin by at least 9-fold when compared to curcumin administered with piperine as absorption enhancer. Eur. J. Pharm. Sci. 2009, 37, 223–230. [Google Scholar] [CrossRef]
  261. Balasubramanian, A.; Pilankatta, R.; Teramoto, T.; Sajith, A.M.; Nwulia, E.; Kulkarni, A.; Padmanabhan, R. Inhibition of dengue virus by curcuminoids. Antiviral Res. 2019, 162, 71–78. [Google Scholar] [CrossRef]
  262. Padilla, S.L.; Rodríguez, A.; Gonzales, M.M.; Gallego, G.J.C.; Castaño, O.J.C. Inhibitory effects of curcumin on dengue virus type 2-infected cells in vitro. Arch. Virol. 2014, 159, 573–579. [Google Scholar] [CrossRef]
  263. Huang, H.-I.; Chio, C.-C.; Lin, J.-Y. Inhibition of EV71 by curcumin in intestinal epithelial cells. PLoS ONE 2018, 13, e0191617. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  264. Lin, C.J.; Chang, L.; Chu, H.W.; Lin, H.J.; Chang, P.C.; Wang, R.Y.L.; Unnikrishnan, B.; Mao, J.Y.; Chen, S.Y.; Huang, C.C. High amplification of the antiviral activity of curcumin through transformation into carbon quantum dots. Small 2019, 15, 1902641. [Google Scholar] [CrossRef]
  265. Martins, C.V.B.; Da Silva, D.L.; Neres, A.T.M.; Magalhães, T.F.F.; Watanabe, G.A.; Modolo, L.V.; Sabino, A.A.; De Fátima, Â.; De Resende, M.A. Curcumin as a promising antifungal of clinical interest. J. Antimicrob. Chemother. 2009, 63, 337–339. [Google Scholar] [CrossRef]
  266. Chen, C.; Long, L.; Zhang, F.; Chen, Q.; Chen, C.; Yu, X.; Liu, Q.; Bao, J.; Long, Z. Antifungal activity, main active components and mechanism of Curcuma longa extract against Fusarium graminearum. PLoS ONE 2018, 13, 1–19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  267. Rex, J.H.; Rinaldi, M.G.; Pfaller, M.A. Resistance of Candida species to fluconazole. Antimicrob. Agents Chemother. 1995, 39, 1–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  268. White, T.C.; Marr, K.A.; Bowden, R.A.; Apisariyakul, A.; Vanittanakom, N.; Buddhasukh, D.; Rex, J.H.; Rinaldi, M.G.; Pfaller, M.A.; Dovigo, L.N.; et al. Clinical, cellular, and molecular factors that contribute to antifungal drug resistance. Photochem. Photobiol. 1995, 49, 382–402. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  269. Fisher, M.C.; Hawkins, N.J.; Sanglard, D.; Gurr, S.J. Worldwide emergence of resistance to antifungal drugs challenges human health and food security. Science 2018, 742, 739–742. [Google Scholar] [CrossRef] [Green Version]
  270. Pancholi, V.; Smina, T.P.; Kunnumakkara, A.B.; Maliakel, B.; Krishnakumar, I.M. Safety assessment of a highly bioavailable curcumin-galactomannoside complex (CurQfen) in healthy volunteers, with a special reference to the recent hepatotoxic reports of curcumin supplements: A 90-days prospective study. Toxicol. Rep. 2021, 8, 1255–1264. [Google Scholar] [CrossRef] [PubMed]
  271. Gupta, S.C.; Patchva, S.; Aggarwal, B.B. Therapeutic roles of curcumin: Lessons learned from clinical trials. AAPS J. 2013, 15, 195–218. [Google Scholar] [CrossRef] [Green Version]
  272. Sharifi-Rad, J.; Rayess, Y.E.; Rizk, A.A.; Sadaka, C.; Zgheib, R.; Zam, W.; Sestito, S.; Rapposelli, S.; Neffe-Skocińska, K.; Zielińska, D.; et al. Turmeric and Its Major Compound Curcumin on Health: Bioactive Effects and Safety Profiles for Food, Pharmaceutical, Biotechnological and Medicinal Applications. Front. Pharmacol. 2020, 11, 1–23. [Google Scholar] [CrossRef]
  273. Sharma, M.; Manoharlal, R.; Negi, A.S.; Prasad, R. Synergistic anticandidal activity of pure polyphenol curcumin i in combination with azoles and polyenes generates reactive oxygen species leading to apoptosis. FEMS Yeast Res. 2010, 10, 570–578. [Google Scholar] [CrossRef]
  274. Lee, W.; Lee, D.G. An antifungal mechanism of curcumin lies in membrane-targeted action within Candida albicans. IUBMB Life 2014, 66, 780–785. [Google Scholar] [CrossRef]
  275. Khan, N.; Shreaz, S.; Bhatia, R.; Ahmad, S.I.; Muralidhar, S.; Manzoor, N.; Khan, L.A. Anticandidal activity of curcumin and methyl cinnamaldehyde. Fitoterapia 2012, 83, 434–440. [Google Scholar] [CrossRef] [PubMed]
  276. Garcia-Gomes, A.S.; Curvelo, J.A.R.; Soares, R.M.A.; Ferreira-Pereira, A. Curcumin acts synergistically with fluconazole to sensitize a clinical isolate of Candida albicans showing a MDR phenotype. Med. Mycol. 2012, 50, 26–32. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  277. Dovigo, L.N.; Pavarina, A.C.; Ribeiro, A.P.D.; Brunetti, I.L.; Costa, C.A.D.S.; Jacomassi, D.P.; Bagnato, V.S.; Kurachi, C. Investigation of the photodynamic effects of curcumin against Candida albicans. Photochem. Photobiol. 2011, 87, 895–903. [Google Scholar] [CrossRef]
  278. Brasch, J.; Freitag-Wolf, S.; Beck-Jendroschek, V.; Huber, M. Inhibition of dermatophytes by photodynamic treatment with curcumin. Med. Mycol. 2017, 55, 754–762. [Google Scholar] [CrossRef] [PubMed]
  279. Al-Asmari, F.; Mereddy, R.; Sultanbawa, Y. A novel photosensitization treatment for the inactivation of fungal spores and cells mediated by curcumin. J. Photochem. Photobiol. B Biol. 2017, 173, 301–306. [Google Scholar] [CrossRef] [PubMed]
  280. Shinobu-Mesquita, C.S.; Bertoni, T.A.; Guilhermetti, E.; Svidzinski, T.I.E. Antifungal activity of the extract of Curcuma zedoaria against yeasts of the genus Candida isolated from the oral cavity of patients infected with the human immunodeficiency virus. Rev. Bras. Farmacogn. 2011, 21, 128–132. [Google Scholar] [CrossRef] [Green Version]
  281. Fernandes, L.D.S.; Amorim, Y.M.; da Silva, E.L.; Silva, S.C.; Santos, A.J.A.; Peixoto, F.N.; Neves Pires, L.M.; Sakamoto, R.Y.; Horta Pinto, F.D.C.; Scarpa, M.V.C.; et al. Formulation, stability study and preclinical evaluation of a vaginal cream containing curcumin in a rat model of vulvovaginal candidiasis. Mycoses 2018, 61, 723–730. [Google Scholar] [CrossRef]
  282. 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]
  283. Wang, Y.; Lu, Z.; Wu, H.; Lv, F. Study on the antibiotic activity of microcapsule curcumin against foodborne pathogens. Int. J. Food Microbiol. 2009, 136, 71–74. [Google Scholar] [CrossRef]
  284. Basniwal, R.K.; Buttar, H.S.; Jain, V.K.; Jain, N. Curcumin nanoparticles: Preparation, characterization, and antimicrobial study. J. Agric. Food Chem. 2011, 59, 2056–2061. [Google Scholar]
  285. Song, J.; Choi, B.; Jin, E.-J.; Yoon, Y.; Choi, K.-H. Curcumin suppresses Streptococcus mutans adherence to human tooth surfaces and extracellular matrix proteins. Eur. J. Clin. Microbiol. Infect. Dis. 2012, 31, 1347–1352. [Google Scholar] [CrossRef] [PubMed]
  286. Betts, J.W.; Wareham, D.W. In vitro activity of curcumin in combination with epigallocatechin gallate (EGCG) versus multidrug-resistant Acinetobacter baumannii. BMC Microbiol. 2014, 14, 172. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  287. Moghaddam, K.M.; Iranshahi, M.; Yazdi, M.C.; Shahverdi, A.R. The combination effect of curcumin with different antibiotics against Staphylococcus aureus. Int. J. Green Pharm 2009, 3, 141. [Google Scholar]
  288. Mun, S.-H.; Joung, D.-K.; Kim, Y.-S.; Kang, O.-H.; Kim, S.-B.; Seo, Y.-S.; Kim, Y.-C.; Lee, D.-S.; Shin, D.-W.; Kweon, K.-T. Synergistic antibacterial effect of curcumin against methicillin-resistant Staphylococcus aureus. Phytomedicine 2013, 20, 714–718. [Google Scholar] [CrossRef] [PubMed]
  289. Rai, D.; Singh, J.K.; Roy, N.; Panda, D. Curcumin inhibits FtsZ assembly: An attractive mechanism for its antibacterial activity. Biochem. J. 2008, 410, 147–155. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  290. Goel, A.; Kunnumakkara, A.B.; Aggarwal, B.B. Curcumin as “Curecumin”: From kitchen to clinic. Biochem. Pharmacol. 2008, 75, 787–809. [Google Scholar] [CrossRef] [Green Version]
  291. Priyadarsini, K.I. Photophysics, photochemistry and photobiology of curcumin: Studies from organic solutions, bio-mimetics and living cells. J. Photochem. Photobiol. C Photochem. Rev. 2009, 10, 81–95. [Google Scholar] [CrossRef]
  292. 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. Phase I clinical trial of oral curcumin: Biomarkers of systemic activity and compliance. Clin. Cancer Res. 2004, 10, 6847–6854. [Google Scholar] [CrossRef] [Green Version]
  293. Burgos Morón, E.; Calderón Montaño, J.M.; Salvador, J.; Robles, A.; López Lázaro, M. The dark side of curcumin. Int. J. Cancer 2010, 126, 1771–1775. [Google Scholar] [CrossRef]
  294. Kurien, B.T.; Dillon, S.P.; Dorri, Y.; D’souza, A.; Scofield, R.H. Curcumin does not bind or intercalate into DNA and a note on the grey side of curcumin. Int. J. Cancer 2011, 128, 242–245. [Google Scholar] [CrossRef]
  295. Danafar, H.; Salehiabar, M.; Barsbay, M.; Rahimi, H.; Ghaffarlou, M.; Arbabi Zaboli, K.; Faghfoori, M.H.; Kaboli, S.; Nosrati, H.; Faghfoori, Z. Curcumin delivery by modified biosourced carbon-based nanoparticles. Nanomedicine 2022, 17, 95–105. [Google Scholar] [CrossRef] [PubMed]
  296. Shahbaz, S.K.; Koushki, K.; Sathyapalan, T.; Majeed, M.; Sahebkar, A. PLGA-Based Curcumin Delivery System: An Interesting Therapeutic Approach in the Treatment of Alzheimer’s Disease. Curr. Neuropharmacol. 2022, 20, 309–323. [Google Scholar] [CrossRef]
  297. Prabhu, A.; Jose, J.; Kumar, L.; Salwa, S.; Vijay Kumar, M.; Nabavi, S.M. Transdermal Delivery of Curcumin-Loaded Solid Lipid Nanoparticles as Microneedle Patch: An In Vitro and In Vivo Study. AAPS PharmSciTech 2022, 23, 49. [Google Scholar] [CrossRef]
  298. Zeng, X.; Zhang, Y.; Xu, X.; Chen, Z.; Ma, L.; Wang, Y.; Guo, X.; Li, J.; Wang, X. Construction of pH-sensitive targeted micelle system co-delivery with curcumin and dasatinib and evaluation of anti-liver cancer. Drug Deliv. 2022, 29, 792–806. [Google Scholar] [CrossRef] [PubMed]
  299. Hassan, F.; Rehman, M.S.; Khan, M.S.; Ali, M.A.; Javed, A.; Nawaz, A.; Yang, C. Curcumin as an alternative epigenetic modulator: Mechanism of action and potential effects. Front. Genet. 2019, 10, 514. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  300. Ramalingam, P.; Yoo, S.W.; Ko, Y.T. Nanodelivery systems based on mucoadhesive polymer coated solid lipid nanoparticles to improve the oral intake of food curcumin. Food Res. Int. 2016, 84, 113–119. [Google Scholar] [CrossRef]
  301. Sarkar, H.; Li, F.; Wang, Y.; Padhye, Z.S. Lesson learned from nature for the development of novel anti-cancer agents: Implication of isoflavone, curcumin, and their synthetic analogs. Curr. Pharm. Des. 2010, 16, 1801–1812. [Google Scholar] [CrossRef]
  302. Kidd, P.M. Bioavailability and activity of phytosome complexes from botanical polyphenols: The silymarin, curcumin, green tea, and grape seed extracts. Altern. Med. Rev. 2009, 14, 226–246. [Google Scholar]
  303. Wen, C.; Fu, L.; Huang, J.; Dai, Y.; Wang, B.; Xu, G.; Wu, L.; Zhou, H. Curcumin reverses doxorubicin resistance via inhibition the efflux function of ABCB4 in doxorubicin-resistant breast cancer cells. Mol. Med. Rep. 2019, 19, 5162–5168. [Google Scholar] [CrossRef] [Green Version]
  304. Mocanu, V.; Zhang, Z.; Deehan, E.C.; Kao, D.H.; Hotte, N.; Karmali, S.; Birch, D.W.; Samarasinghe, K.K.; Walter, J.; Madsen, K.L. Fecal microbial transplantation and fiber supplementation in patients with severe obesity and metabolic syndrome: A randomized double-blind, placebo-controlled phase 2 trial. Nat. Med. 2021, 27, 1272–1279. [Google Scholar] [CrossRef]
  305. Banerjee, Y.; Taranikanti, V.; Bayoumi, R. Triglyceride-mediated pathways and coronary heart disease. Lancet 2010, 376, 956. [Google Scholar] [CrossRef]
Figure 1. A flow diagram summarizing the plausible mechanism of action of curcumin.
Figure 1. A flow diagram summarizing the plausible mechanism of action of curcumin.
Metabolites 12 00639 g001
Figure 2. (a) The turmeric plant, (b) The turmeric rhizome with a yellow-orange interior, (c) The powdered form of turmeric, (d) The chemical structure of curcumin. (e) Demethoxycurcumin and (f) Bis-demethoxycurcumin.
Figure 2. (a) The turmeric plant, (b) The turmeric rhizome with a yellow-orange interior, (c) The powdered form of turmeric, (d) The chemical structure of curcumin. (e) Demethoxycurcumin and (f) Bis-demethoxycurcumin.
Metabolites 12 00639 g002
Figure 3. Role of curcumin in the inflammatory signaling pathway.
Figure 3. Role of curcumin in the inflammatory signaling pathway.
Metabolites 12 00639 g003
Table 1. Methods for curcumin extraction.
Table 1. Methods for curcumin extraction.
Method NameYieldAdvantage(s)Drawback(s)Reference
Microwave-assisted
extraction (MAE)
The yield obtained using MAE is 4.98%. The Curcuma longa plant was soaked in methanol and extracted using acetone via a dual heating mechanism under microwave energy.Safe and cost-effective method. Retains the biological activity of the extracted compounds.Time-consuming, as the vessel needs to be slowly cooled to avoid the loss of volatile components. Involves an initial high cost for setting up the equipment on a large scale.[28]
Enzyme-assisted
extraction (EAE)
The yield obtained using EAE is 5.73%. Turmeric was pretreated using alpha-amylase and amyloglucosidase enzymes. Extraction was performed using N,N- dipropyl ammonium N,N’-dipropylcarbamate.Mild reaction conditions are cost-effective, eco-friendly and feasible.Low purity of the final product.[29]
Supercritical fluid
extraction (SFE)
The yield obtained using SFE is 4.3%. A two-step extraction was performed using SFE, followed by pressurized liquid extraction using ethanol as solvent.Short extraction time and mild operating temperature. The manufacturing costs are lower than those of conventional methods.High technical complexity, and many operational parameters need to be optimized before the initiation of the extraction process.[30]
Ultrasound-assisted
extraction (UAE)
The yield obtained using UAE is 1.03%. Optimal conditions were a 60% amplitude and a 3/1 (s/s) pulsed interval. Ethanol was the extraction solvent.Greater solvent penetration into the samples increases the contact surface area, which increases efficiency.Ultrasound can lead to the degradation of the final purified product.[31]
Ionic liquid-based
extraction (ILE)
The yield obtained using ILE is 6.18%. The ionic liquid used for extraction was an anionic [Omin][Br-] aqueous solution.Eco-friendly and increases extraction efficiency.High cost involved in the preparation and use.[32]
Table 2. Activity of curcumin against cancer.
Table 2. Activity of curcumin against cancer.
Cancer TypeCurcumin Conc.Signaling Molecules Up/DownregulatedOverviewDelivery ModesRef.
Prostate cancer10–100 μMDownregulates
NF-κB, AP-1, Cyclin D1, CXCL-1 and CXCL-2, Bcl-2, Bcl-xL and XIAP.
-Curcumin is a potent inhibitor of NF-κB in both ADPC and AIPC cells, thereby preventing cell proliferation and inducing apoptosis.
-Curcumin restores the response of AIPC cells to anti-androgen treatment.
-Prevents metastasis in AIPC cells.
-Free curcumin or in combination with chemotherapeutic agents such as TRAIL.
-Curcumin in poly(lactic-co-glycolic) acid.
-Using the nanoparticle formulation of curcumin.
-Curcumin-loaded liposomes.
[84,85,86,87,88,89,90,91,92,93,94,95]
Breast
cancer
10–40 μMDownregulates Bcl-2, CXCL-1, CXCL-2, MMP-9, urokinase plasminogen activator, intercellular adhesion molecule 1 and chemokine receptor 4, PECAM-1, Cyclin D1 and p65.-Curcumin can suppress ODC activity and inhibit cell proliferation.
-Inhibits the MPA-induced secretion of pro-angiogenic factors such as VEGF.
-Hyaluronic acid-modified mesoporous silica nanoparticles.
-Chitosan nanoparticles.
-Zinc oxide nanoparticles.
-In combination with niclosamide using PLGA nanoparticles.
-PEGylated PLGA nanoparticles
-Nanovesicles.
[21,96,97,98,99,100,101,102,103,104,105,106,107]
Colon
cancer
10–50 µMDownregulates TNFα, JNK activation, miR-21 and COX-2.-Curcumin inhibits the activation of the TLR4/MyD88/NF-κB signaling axis.
-Reduces IκB kinase activity and inhibits the degradation of IkBα.
-Inhibits the production of TNF-α, IL-6 and IL-12.
-Inhibits Foxp3 expression and enhanced interferon-gamma secretion in regulatory T cells.
-Curcumin in silica nanoparticle.
-Polymeric nanocarrier.
-Curcumin-loaded thiolated chitosan nanoparticle.
-Dendrosomal carrier.
-Curcumin-loaded micelles.
-Curcumin–PLGA nanoparticles.
[108,109,110,111,112,113,114,115,116,117,118,119]
Pancreatic10–50 µMDownregulates EGFR, COX-2, NF-κb, AKT and Prostaglandin E2.-Inhibited cell survival and enhanced apoptosis in pancreatic adenocarcinoma cell lines.
-Suppressed tumor growth by inhibiting the NF-KB pathway.
-Induced apoptosis via ATM/Chk1.
-Anti-proliferative activity by suppressing Sp1 and disrupting NF-κB translocation to the nucleus.
-Curcumin analogues PEGylated Curcumin, [Dlys6]-LHRH and its analog called L49H37.
-CDF and PEGylated curcumin.
-Liposomal Curcumin.
-Curcumin analogues with the hydroxyl group.
-Magnetic particles that were used to encapsulate curcumin.
-Ester-mediated conjugations of curcumin to cholesteryl-hyaluronic acid nanogel.
[120,121,122,123,124,125,126,127,128,129,130,131]
Gastric10–100 µMDownregulates the Akt pathway, BCL-2, COX-2 and cyclin D1.-Induced apoptosis by activating caspase-3,PARP; reduction in Bcl-XL levels.
-Curcumin also activates the Fas pathway by stimulating the activity of caspase-8.
-Activated Bax protein expression and inhibited the Bcl-2 protein
-Suppressed the transition of the cancer cells from the G(1) to S phase.
-Cyclodextrin complexes with curcumin.
-Nanoparticles such as polymer-encapsulated ZnO nanoparticles.
-Microsponges using polymers such as ethyl cellulose and polyvinyl alcohol.
-Curcumin-loaded nanoemulsion.
-Cationic polysaccharides such as chitosan.
[132,133,134,135,136,137,138,139,140,141,142]
Lung5–50 µMDownregulates the Akt pathway, BRCA pathway, Beta-catenin signaling and MMP-2 and upregulates caspase-3, Bax and p53.Induces apoptotic cell death by activating caspase-7 and 3.
-Enhances PARP cleavage and stimulates ER stress.
-Enhances ROS production to cause apoptosis.
-Increases the sensitivity of cancer cells to chemotherapy.
-Induces DNA damage and prevents the migration of cancer cells.
-Lipid-based liposome.
-Polymeric carrier and micelle.
-Chitosan microsphere.
-Polymeric and lipid nanoparticle.
-Nanocrystal.
[143,144,145,146,147,148,149,150,151,152,153,154]
Oral10–100 µM-Prevents cell proliferation and promotes apoptosis.-Curcumin reduces the migration and progression of TSCC cells, promotes apoptosis and inhibits tumorigenesis.
-Suppresses the CAF (cancer-associated fibroblast)-mediated proliferation and tumorigenicity of Cal27 by inhibiting TSCC CAFs.
-Nanohybrid formulation.
-Lozenges.
-Silica nanoparticles.
-Mucoadhesive nanogel system.
[155,156,157,158,159,160]
Skin10–50µM-Shows effective anti-proliferative activity.-Antiproliferative effect, as they effectively inhibit the clonogenic ability in melanoma cells.-Cationic liposomes.
-Ethosomal nanocarriers.
[161,162]
Table 4. Effect of curcumin on gut microbiota and its mechanism of action.
Table 4. Effect of curcumin on gut microbiota and its mechanism of action.
Curcumin DosesEffect on Gut MicrobiotaMolecular MechanismsModelRef.
Curcumin at a low dose (1 g/day)Curcumin shifted the structure of gut microbiotaCurcumin enervated the Western diet-induced development of atherosclerosis and type 2 diabetes mellitusSprague
Dawley rats
[225]
100 mg/kg/dayLowers the increasing abundance of the genera Anaerotruncus and Helicobacter in the gut microbiotaDecreases the estrogen level, resulting in an increase in body weightWistar rats[226]
100 mg/kg/dayCurcumin affected the presence of Prevotellaceae, Bacteroidacea, and Rikenellaceae in gut microbiotaCurcumin possesses anticancer activity in vitro and in preclinical animal models via the activation of caspases 3, 8 and 9 in the colon cancer cell linesFecal sample[227]
8000 mg per dayIncrease in Lactobacillus and decrease in CoriobacteralesInduction of apoptosis through the COX-2 and non-COX-2 pathways. It targets cancer stem cells (CSC) through direct or indirect influences on the CSC self-renewal pathways.Colon cancer cell lines, SW480 and SW62[228]
0.2% (w/w) nanoparticles of curcuminIncrease in butyrate-producing bacteria and the fecal butyrate levelMucosal mRNA expression of inflammatory mediators and the activation of NF-κB in colonic epithelial cells were suppressed by curcumin nanoparticlesBALB/c mice[229]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Sivani, B.M.; Azzeh, M.; Patnaik, R.; Pantea Stoian, A.; Rizzo, M.; Banerjee, Y. Reconnoitering the Therapeutic Role of Curcumin in Disease Prevention and Treatment: Lessons Learnt and Future Directions. Metabolites 2022, 12, 639. https://doi.org/10.3390/metabo12070639

AMA Style

Sivani BM, Azzeh M, Patnaik R, Pantea Stoian A, Rizzo M, Banerjee Y. Reconnoitering the Therapeutic Role of Curcumin in Disease Prevention and Treatment: Lessons Learnt and Future Directions. Metabolites. 2022; 12(7):639. https://doi.org/10.3390/metabo12070639

Chicago/Turabian Style

Sivani, Bala Mohan, Mahmoud Azzeh, Rajashree Patnaik, Anca Pantea Stoian, Manfredi Rizzo, and Yajnavalka Banerjee. 2022. "Reconnoitering the Therapeutic Role of Curcumin in Disease Prevention and Treatment: Lessons Learnt and Future Directions" Metabolites 12, no. 7: 639. https://doi.org/10.3390/metabo12070639

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