Next Article in Journal
Social and Robust Navigation for Indoor Robots Based on Object Semantic Grid and Topological Map
Previous Article in Journal
Precipitation and Calcination of High-Capacity LiNiO2 Cathode Material for Lithium-Ion Batteries
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Curcumin Innovative Delivery Forms: Paving the ‘Yellow Brick Road’ of Antitumoral Phytotherapy

by
Magda Carvalho Henriques
,
Maria Amparo F. Faustino
and
Susana Santos Braga
*
LAQV/REQUIMTE, Chemistry Department, University of Aveiro, 3810-193 Aveiro, Portugal
*
Author to whom correspondence should be addressed.
Present address: Laboratory of Signal Transduction, Department of Medical Sciences, Institute of Biomedicine, University of Aveiro, 3810-193 Aveiro, Portugal.
Appl. Sci. 2020, 10(24), 8990; https://doi.org/10.3390/app10248990
Submission received: 23 November 2020 / Revised: 5 December 2020 / Accepted: 11 December 2020 / Published: 16 December 2020
(This article belongs to the Section Applied Biosciences and Bioengineering)

Abstract

:
This review deals with the various aspects involved in the medicinal action of curcumin, from the photosensitivity and its relevance to storage and shelf-life, to the different routes of administration, which influence the bioavailability. The focus of the review is on the antitumor properties of curcumin and the currently available solutions for their amelioration. The work starts by presenting a brief historical perspective on the origins and uses of curcumin, from early days until the present time. The following sections describe the physico-chemical properties of curcumin and their impact on the biological activity and pharmacokinetics, raising awareness to the need for formulations able to improve the bioavailability. The last section is focused on research efforts being made to circumvent curcumin’s instability and low availability due to the extensive hepatic first pass metabolism, describing innovative scientific advances and new patented formulations and emerging products on the market.

1. Introduction

1.1. Curcumin Overview

The history of curcumin dates back about five thousand years. This strong yellow pigment is the main active component of Curcuma longa, a perennial Zingiberaceae plant native to southwest India, but now grown across the South and Southeast Asia, especially in China and India [1,2,3]. The rhizomes of this plant are dried and powdered to obtain the spice commonly known as turmeric, referenced in Ayurvedic medicine, the characteristic medicinal system of Ancient India, as a home remedy for various diseases [4]. The expansion of the therapeutic use of curcumin to the Western civilizations dates from the time of the Portuguese "State of India", in the XVI century. Garcia de Orta, the physician of the Viceroy of India, mentioned turmeric in his compendium on Indian plants and principles, as “a medicine for jaundice” used widely across Asia [5].
Turmeric dry rhizome is composed mainly of starch, having also, in lesser extent and in varying composition according to geoclimatic factors, carbohydrates, proteins, lipids, fiber, curcuminoid pigments, sesquiterpenes (turmerone, atlantone, zingiberone, turmeronol, germacrone, α-curcumene, β-sesquiphellanderene, bisacurone, curcumenone, dehydrocurdione, procurcumadiol, bis-acumol, curcumenols, zedoaronediol, bisabolene, and curlone), and caffeic acid [6,7]. The curcuminoid content typically varies between 2% and 9%. Curcuminoids are biosynthesized by condensation of malonic acid with cinnamoyl CoA or p-coumaroyl CoA. Curcumin is the most abundant curcuminoid in turmeric, but traces of its precursors, desmethoxycurcumin and bisdemethoxycurcumin (Figure 1), are also present [8]. Curcumin and its analogues can be found in several other Curcuma species as well as in a few species from other genera. Sources of curcuminoids include Curcuma mangga, Curcuma zedoaria, Costus speciosus, Curcuma xanthorrhiza, Curcuma aromatic, Curcuma phaeocaulis, Etlingera elatior, and Zingiber cassumunar [9]. C. mangga, commonly named mango ginger, is indicated as a good dietary source of curcumin, but it is still far from reaching the widespread reputation of turmeric as a dietary supplement and functional food.
In western societies, turmeric consumption is a growing trend due to the recognition of its therapeutic properties against inflammation and cancer [10,11]. Turmeric has been granted the GRAS status (‘Generally Recognized as Safe’) by the FDA [12], with an acceptable daily intake (ADI) limit of 2.5 mg/kg of body weight; for pure curcumin, the ADI is of 0.1 mg/kg weight [13]. The official acknowledgement of turmeric as a safe dietary supplement contributed strongly to its widespread use. Turmeric is, since 2013, the top-ranking herbal supplement in North America, having had, in that year, a most impressive sales growth of 26.2% [14]. The applications of curcumin in commercial products are expanding beyond the field of food supplements and tending to grow into other areas, such as cosmetics. While India is the largest producer of curcumin, North America is the largest curcumin market. Market revenue in 2014 was superior to US$20 million [15]. Future market predictions, from now up to 2022, estimate a shift in the demand towards Europe, with Germany and Denmark expected to grow by 17% and to become the largest market. By 2022, the global curcumin market is predicted to reach $94.3 billion [16].

1.2. The Chemistry of Curcumin

The isolation of the active phytochemicals in turmeric dates back to 1815, when the first crude extract was obtained and described as “a matter of yellow color” [17]. Many of the properties of curcumin could already be observed in this extract: it was insoluble in water, solubilizing upon the addition of alkali to form a reddish-brown solution, and able to react with salts of metals such as lead and tin [17]. The extract was later found to contain a mixture of curcuminoids along with some oils and resins, and only in 1870 was curcumin first purified and isolated in the form of orthorhombic crystals [18]. Its chemical structure was determined in 1910 [19]. Curcumin has the chemical formula C21H20O6 (Mw of 368.38) and it is also referred to as diferuloylmethane, since its IUPAC denomination is quite long: (1E,6E)-1,7-bis(4-hydroxy-3-methoxyphenyl)-1,6-heptadiene-3,5-dione. Its chemical structure comprises two aromatic ring systems with o-methoxy phenol groups connected by a seven-carbon linker consisting of an α,β-unsaturated β-diketone moiety that exhibits keto-enol tautomerism in solution [20]. Due to extended conjugation, the π electron cloud is distributed all along the molecule and thus curcumin is quite hydrophobic, with a log p value of 3.38 and an extremely low solubility in water (1.34 ± 0.02 mg/L) [21]. According to the Biopharmaceutics Classification System (BCS) [22], curcumin is a class IV drug; that is, a compound having low solubility and low permeability. Class IV drugs usually are “not well absorbed over the intestinal mucosa and a high variability [in the absorption profile] is expected”.
Curcumin is reasonably stable in water at pH < 7.0 due to structural stabilization by the conjugated diene; in PBS and at pH > 8 it may degrade rapidly (10 min) [2]. In fact, curcumin possesses three ionizable protons with pKa values of approximately 8.5 (enolic proton) and 10–10.5 (two phenolic protons) [23,24].
Curcumin absorbs light from the near ultraviolet (around 340 nm) to the indigo-blue spectral region (450–460 nm), with absorption peaking at 410–430 nm (violet light) [25]. It presents a fluorescence band between 460 and 560 nm. Furthermore, curcumin is sensitive to ultraviolet radiation and its degradation is accelerated by exposure to sunlight [26,27]. When irradiated with light above 400 nm, curcumin undergoes a self-sensitized photo-decomposition where singlet oxygen is involved, but when reactive oxygen species are not available, other decomposition mechanisms are triggered. Photodegradation products include vanillin, vanillic acid, 4-vinyl-guaiacol, ferulic aldehyde, and ferulic acid [28].

2. Biological Actions of Curcumin

Curcumin is indicated in ayurvedic medicine for an enormous variety of pathologies and ailments [29]. Most of this knowledge is, however, empirical, or it has not been demonstrated by studies on human subjects. Most of studies available in the literature have been conducted either in vitro or in animal models (mostly rodents). They provide information on possible therapeutic indications of curcumin for conditions as varied as viral infections, scleroderma, atherosclerosis, myocardial infarction, brain ischemia, and Alzheimer’s [30], but such activities may not necessarily be manifested in human patients. Only in the latest decades has evidence from clinical trials been gathered on curcumin. This section presents the most relevant results of clinical trials, with highlight on cancer therapy.

2.1. Medicinal Activity in Humans

Turmeric has well-documented anti-inflammatory [31,32], antioxidant [33,34,35,36] and antitumor activities [37,38,39,40] that are associated with the extraordinary performance of its main active component, curcumin. Research conducted over the last decades has demonstrated that curcumin is a strongly pleiotropic molecule, able to modulate the activity of numerous signalling biomolecules, to interfere with different cellular and molecular cascades [41,42] and to interact with transcription factors, growth factors or their receptors, nuclear factors, cytokines, and hormone receptors. Curcumin is even able to regulate the expression of genes associated with the processes of cell proliferation and apoptosis [43]. Details on the different biochemical targets of curcumin are given in the Section 2.2. The complexity of mechanisms underlying the action of curcumin allows understanding its efficacy in the treatment of multi-factorial pathologies such as cancer.

2.1.1. Curcumin against Inflammation and Oxidative Stress

Curcumin acts on inflammation at various levels of the arachidonic acid inflammatory cascade, inhibiting the enzymes phospholipase, cyclooxygenase II, and lipo-oxygenase, and having also effects on cytokines [44,45]. Curcumin was shown to reduce post-operative inflammation in patients having had surgical repair of inguinal hernia and/or hydrocele [46], to ameliorate symptoms of chronic inflammation pathologies such as arthritis [47,48], psoriasis [49], and bowel conditions (IBS, Crown’s disease, and ulcerative colitis) [31,50,51,52,53] and to treat eye inflammations such as the “idiopathic orbital inflammatory syndrome” [54] and uveitis [55].
The antioxidant activity of curcumin is also the result of a multiplicity of actions. Not only does curcumin stabilize superoxide and hydroxyl free radicals due to the electron-donating properties of its phenolic groups [20,56,57,58], but it also induces the expression of antioxidant enzymes. In vitro tests with beta cells of human pancreas islets incubated with curcuminoids have shown increased levels of heme oxygenase 1, gamma-glutamyl-cysteine ligase, and NAD(P)H:quinone oxidoreductase and a consequent increase in glutathione levels [59]. Curcumin protects against oxidative stress caused by advanced glycation end products in patients with diabetes, being under evaluation as a new anti-diabetes drug candidate in a series of clinical and pre-clinical studies [60]. It should also be highlighted that the antioxidant benefits of curcumin do not cease with its metabolization, as many of the metabolites present significant antioxidant properties [61].

2.1.2. Antitumoral Action

The first report on the anticancer activity of curcumin, in 1987 [62], rekindled the interest in this compound and brought it to the spotlight of the western society. Curcumin has been the subject of over 30 clinical trials in the context of cancer, some of them still ongoing. This clearly demonstrates its significant chemo-preventive and anticancer potential.
  • In colorectal cancer, curcumin was studied for both tumor prevention and chemotherapy. In cancer prevention, it was demonstrated to reduce by 40% the formation of aberrant crypt foci in smoking patients (intake of 4 g/day for one month) [63]. In a combination study, curcumin, and quercetin (1440 + 60 mg/day for six months) were shown to reduce the number and size of polyps in patients with familial adenomatous polyposis, a hereditary disorder characterized by the development of hundreds of colorectal adenomas which turn malign when left untreated [64]. In chemotherapy, 1 g/day curcumin for up to one month (prior to surgical removal of the tumor) was shown to improve the patient’s body weight and to increase the apoptosis rates of the patient’s tumor cells [65].
  • In prostate cancer, a trial has demonstrated that curcumin/flavone association reduces the chances of developing cancer by lowering the levels of prostate-specific antigen (PSA). PSA levels are increased due to the presence of chronic inflammation in the prostate, which is one of the most significant causes of tumorigenesis [66]. Association of curcumin (5.4 g/day for seven days around chemotherapy) with docetaxel/prednisone (75 mg/m2 + 24 mg, once every three weeks, for six cycles) demonstrated encouraging results, with a tumor objective response in 40% and a PSA response in 59% of the patients in a group having castration-resistant prostate cancer [67]. There is also preliminary evidence on the ability to reduce the formation of metastases. An association of polyphenols (pomegranate seed, green tea, broccoli, and turmeric), taken over six months, has lowered PSA by 63.8% (compared to placebo) in prostatectomized patients [68]. Note that, since these men have no prostate, PSA is produced only by neoplastic cells, thus being a good indicator of metastasis growth. Curcumin can confer radioprotective effect in patients with prostate cancer who undergo radiation therapy, reducing the severity of radiotherapy related urinary symptoms. Patients were given 3 g of curcuminoids per day (corresponding to ca. 2 g/day of curcumin) for one week before the onset of radiotherapy and until completion of radiotherapy [69,70].
  • In breast cancer, curcumin was used in co-therapy with both chemotherapeutic agents and radiation. A combination therapy with docetaxel and curcumin (in escalating doses of up to 6 g/day) was found to afford better therapeutic results than docetaxel used alone: histological improvements were observed in the fourteen patients under study, all having reduction or elimination of disseminated foci [71]. Curcumin was evaluated in two clinical trials regarding protective action against radiation-induced dermatitis during radiotherapy of breast cancer patients. Despite promising results on a pilot study, with slightly less severe dermatitis in the curcumin group, a second trial on 686 patients showed no significant changes in pain, symptoms, and quality of life of the patients taking curcumin (1.5 g daily) in regard to those taking placebo [72].
  • Pancreatic cancer, in the advanced stage, is a condition with very poor prognosis. In a phase II study with twenty-one patients taking curcumin (8 g/day for up to 18 months), partial regression was observed during the treatment period; after treatment, patient responses varied, one of them having become stable and another having shown a strong tumor response [73]. Another trial evaluated the association of curcuminoids (8 g/day, corresponding to 6.14 g/day of curcumin) with a gemcitabine-based chemotherapeutic treatment. A total of 21 patients was divided into two groups: one, with 2 patients, received gemcitabine monotherapy; the other, with 19 patients, received a combination therapy of gemcitabine and S-1. S-1 is a novel oral antitumor formula based on fluorouracil, comprising three pharmacological agents: (i) tegafur, a prodrug of 5-fluorouracil, (ii) 5-chloro-2,4-dihydroxypyridine, which inhibits dihydropyrimidine dehydrogenase activity; and (iii) potassium oxonate, which reduces gastrointestinal toxicity was also evaluated. Eighty-one percent of the patients died during the study period. In the surviving patients, the treatment was able to stabilize the disease [74].

2.2. Molecular Targets of Curcumin

Curcumin has an immense range of molecular targets, being proven to interact with transcription factors, growth factors and their receptors, cytokines, enzymes, and genes that regulate cell proliferation and apoptosis, as reviewed elsewhere [75,76]. The various effects are summarized in the Table 1 and described with more detail in the following sub-sections.

2.2.1. Curcumin Modulates the Activity of Transcription Factors

Three main families of transcription factors are involved in cell proliferation, cell invasion, metastasis, angiogenesis, and resistance to chemotherapy and radiotherapy. They are:
  • the families of the nuclear factor kappa-light-chain-enhancer of activated B cells (NF-kB) and of the activated protein-1 (AP-1),
  • signal transducers and activators of transcription (STAT), and
  • steroid receptors [77].
NF-kB is involved in cell response to several external agents, including physical strain, oxidative stress (free radicals), and cytokines. NF-kB is usually inactive in the cytoplasm, but once activated by the adequate stimuli it can translocate to the nucleus, inducing expression of apoptosis-suppressing genes to promote cell proliferation and even metastasis. Curcumin was shown to inhibit the activation of the NF-kB pathway by studies in vitro [78] and in vivo [79] and in a phase II clinical trial [80].
AP-1 is involved in the differentiation, proliferation, apoptosis, and oncogenic transformations of the cells [81]. AP-1 can be activated by stimuli of growth factors, cytokines, or bacterial/viral infections. Activated AP-1 induces the expression of several genes that code proteins involved in the angiogenesis and growth of cancer cells, such as cyclin-D1, MMP, and VEGF [82]. Curcumin inhibits this pathway by direct interaction with the AP-1 DNA-binding site, namely in human leukemia cells, transformed keratinocytes and prostate cancer cells [83,84,85,86,87]. The effect of curcumin on the expression of NF-kB and AP-1 members was evaluated in an oral cancer cell line [88]. Nuclear extracts obtained from curcumin-treated cancer cell were evaluated regarding binding of the transcription factors AP-1 and NF-kB, to reveal that binding is reduced in a dose dependent manner and that in cells treated with 100 μM of curcumin, the DNA-binding activities of AP-1 and NF-kB were completely lost. These results confirmed the downregulation of several transcription factors and inhibition of NF-kB and AP-1.
The Janus kinase (JAK) signal transducer of activators of transcription (STAT) pathway signalling pathway is a signaling pathway employed by diverse cytokines, interferons, growth factors, and related molecules, allowing these extracellular factors control gene expression and regulate cell growth and differentiation [89]. In cancer cells, this pathway is consistently active, being involved in metastasis. Inhibition of the JAK/STAT pathway by curcumin was observed in prostate, lung, and glioblastoma cancer cell lines [90,91,92]. Curcumin was shown to inhibit STAT3 phosphorylation and to lower levels of interleukin-6 (IL-6), a pro-inflammatory cytokine involved in cell proliferation and survival. Curcumin also exhibited antineoplastic effects in K652 chronic leukemia, ovarian, and endometrial cancer cells by suppression of JAK/STAT signalling [93,94].
Table 1. Molecular targets and cell processes modulated by curcumin.
Table 1. Molecular targets and cell processes modulated by curcumin.
FamilyMolecular TargetEffectRef
Transcription factorsNF-kB[78,79,80,88,95]
Nrf2[96]
AP-1[83,84,85,86,87,88]
STAT-3[97,98]
STAT-5[99,100]
β-catenin[101,102]
EGR-1[103,104]
HIF-1[105]
Notch-1[106]
Growth factorsEGF[107]
FGF[108]
PDGF[109]
TGF-β[110,111,112,113,114]
VEGF[115,116,117]
Cytokines, pro-inflammatoryTNF-α[95,118,119,120]
IL-1[121]
IL-2[122]
IL-5[123]
IL-6[118]
IL-8[121]
IL-12[124]
IL-18[125]
EnzymesCOX-2[72,80,126,127]
iNOS[127]
Lipoxygenase[128]
MMP-9[78,129,130,131]
KinasesJNK[132]
MAPK[133]
PKC[131]
Akt[134]
CDKs[135]
ReceptorsAR[86]
EGFR[79,119]
Adhesion moleculesICAM-1[95]
VCAM-1[95]
ELAM-1[95]
Antiapoptotic proteinsBcl-2[136,137,138,139,140]
Bcl-xL[136,137,138,141]
Proapoptotic proteinsBax[136,137,138,139,140]
Bak[140]
OthersCyclin D1[142,143]
p53[144,145]

2.2.2. Curcumin Decreases Tumor Angiogenesis

Curcumin has anti-angiogenic properties by inhibition of vascular endothelial growth factor (VEGF) and fibroblast growth factor (FGF) [79]. Several in vitro and in vivo studies have demonstrated the association between the suppression of VEGF expression by curcumin with its inhibitory action on tumor growth [115,116,117]. In vivo studies in mice showed that VEGF expression and angiogenesis suppression is mediated through suppression of the osteopontin gene expression and the NF-κB/ATF-4 pathway [115].

2.2.3. Curcumin Inhibits Inflammatory Cytokines

Tumor necrosis factor (TNF) and interleukins (IL) are two kinds of inflammatory cytokines with an important mediating role in tumorigenesis. Curcumin was shown to have profound effects on TNF inhibition in dendritic cells, macrophages, monocytes, alveolar macrophages, and endothelial and bone marrow cells [118]. The suppression of the TNF-signalling pathway by curcumin is related to the inhibition of NF-kB phosphorylation, as detailed in 2.2.1 [119,120].
Interleukins (ILs) contribute to tumor invasiveness and angiogenesis by induction of the expression of metalloproteinases, adhesion molecules and STATs [146]. Curcumin inhibits the expression of IL-1 [121], IL-2 [122], IL-5 [123], IL-6 [118], IL-8 [122], IL-12 [124], and IL-18 [125], being thus a potent inhibitor of these classes of cytokines.

2.2.4. Curcumin Regulates the Activity of Enzymes with Roles in Inflammation and Cancer

Pro-inflammatory enzymes are linked with various types of cancer. COX-2 is known to participate in uncontrolled cell proliferation and suppression of apoptosis, while inducible nitric oxide synthase (iNOS) and matrix metalloproteinase-9 (MMP-9) are involved in the formation of metastases [147,148]. Several studies, both in vitro and in vivo, demonstrated that the inhibitory action of curcumin on COX-2 expression contributes significantly to its antitumor action [72,80,126]. Curcumin was also shown to inhibit the expression of MMP-9 in orthotopically implanted pancreatic [129] and ovarian [130] tumors in mice and the production of iNOS in chronic colitis [127].

2.2.5. Curcumin and Cell Cycle Regulation

Programmed cell death, or apoptosis, is a mechanism of vital importance in maintaining normal cell growth. Apoptosis is initiated by regulation of protein 53 (p53) and by proteins of the B-cell lymphoma 2 family (Bcl-2). Activated p53 induces activation of two pro-apoptotic proteins, Bcl-2 homologous antagonist killer (Bak), and Bcl-2 associated x protein (Bax), which in turn release cytochrome c into the cytoplasm to activate the caspase cascade. Curcumin is able to induce apoptosis in prostate cancer PC-3, DU-145, and LNCaP cells via p53-dependent mitochondrial pathways [145]. Activation of p53 by curcumin leads to over-expression of Bak, Bax, and several caspase proteins [136,137,138,139,140]. In addition, curcumin inhibits the activity of a few anti-apoptotic proteins, such as Bcl-2 and B-cell lymphoma extra-large protein (Bcl-XL) [136,137,138,139,140,141].
Cyclin-dependent kinases (CDKs) are also involved in the life cycle of cells, being in charge of its progression through the different stages [149]. Malignant cells have thus frequent alterations in CDK expression, with overexpression of cyclins and suppression of CDK inhibitors. Curcumin induces cell cycle arrest in colon cancer cells (HTC116 line) by CDK2-dependent effects [135]. The mechanism underlying cell cycle arrest by curcumin may involve overexpression of CDK inhibitors and blockage of the expression of cyclin E and cyclin D1 [142,143].

3. Pharmacokinetics and Bioavailability of Curcumin from Different Administration Routes

Traditionally used as a spice, curcumin is typically taken orally. Nevertheless, this administration route is, most likely, the least effective way to convey curcumin to the human organism, as evidenced by the bioavailability results detailed in the following subsections.

3.1. Oral Administration

Curcumin has extremely poor bioavailability from oral intake, which poses serious limitations to its widespread use as a therapeutic agent. In fact, serum levels following the ingestion of curcumin are so low that a dose-response profile is difficult to establish. When curcumin is taken in low amounts (30 or 180 mg daily), its levels in plasma or urine are null or within the experimental error [150,151], as evidenced by the first two entries in the Table 2. Ingestion of daily doses of 3.6 g caused half of the patients to show serum levels of 11 nM [152], but for higher levels there is not a straightforward dose-response behavior (see also Table 2). Daily doses of 4 or 6 g corresponded to average serum peaks of 510 and 640 nM [153], while daily doses of 10 or 12 g resulted in average serum peaks of 138 nM and 157 nM, respectively [154]. Taken together, the results demonstrate that curcumin has a quite low absorption rate at the intestine, estimated at less than 1%, meaning that most of it is passed through the feces.
The amount of curcumin that does get absorbed is further decreased due to the hepatic first pass metabolism, involving extensive O-conjugation to form curcumin glucuronides (Figure 2) and curcumin sulphate, and bioreduction to tetra-hydrocurcumin, hexahydro-curcumin, octahydrocurcumin, and hexahydrocurcuminol [155,156]. Seric levels of curcumin peak at around 1–2 h after the ingestion, followed by a gradual decline within 12 h [153]. Curcumin and its conjugates are excreted in the urine [156].
Table 2. Reported maximal serum levels (Cmax) and time (Tmax) in human subjects after ingestion of different dosages of curcumin.
Table 2. Reported maximal serum levels (Cmax) and time (Tmax) in human subjects after ingestion of different dosages of curcumin.
Daily DoseCmax Mean ± S.D. (nM)Tmax (h)Ref.
0.18 gn.d. a[150]
3.6 g11.1 ± 0.61[152]
4.0 g510 ± 1101.67[153]
6.0 g640 ± 602[153]
8.0 g1770 ± 1870 b1.75[153]
10.0 g138 c4[154]
12.0 g157 c2[154]
0.03 g/kg4.9 ± 7.66[151]
2.0 g/kg16.3 ± 13.51[157]
a n.d. means not detected. b The large standard deviation value may be due to the fact that data was collected from only two subjects. c No average available because data is from only one subject; no curcumin was detected in the plasma of the remaining subjects.
Some authors claim that serum levels of free curcumin show a deceivingly low bioavailability because a significant fraction of the compound occurs in plasma as glucuronide and sulfate conjugates [72,158]. In the abovementioned clinical study with patients taking 3.6 g of curcumin, serum levels of the conjugates averaged 0.016 μM for curcumin glucuronide and 0.009 for curcumin sulphate [152]. The real therapeutic properties of these metabolites are, however, different from those of curcumin. Curcumin mono- or di-glucuronides have lower anti-inflammatory and anti-proliferative activities [159], and curcumin sulfate inhibits the activity of prostaglandin E2 very poorly when compared with curcumin [160].

3.2. Parenteral Administration

Intravenous (I.V.) injection was one of the first administration routes for curcumin to be tested on human subjects, in the advent of its introduction to Western medicine (the other tested route was oral administration). The studies were conducted by Albert Oppenheimer from 1932 to 1936 [161]. Curcumin, as a sodium salt, was administered in concentrations of 50 mg/mL (5 mL injections, making a total dose of 250 mg) and its emptying effect on the gallbladder was demonstrated by the cholecystograms of the patients. Seric levels of curcumin were not determined in the study and bioavailability in human subjects following intravenous administration of curcumin remains, until the present day, not fully understood. Studies with rats (Table 3) have demonstrated that an I.V. dose of 10 mg/kg of curcumin given affords a maximum serum level of 0.36 μg/mL [162]. For comparison, the authors also conducted oral administration of the same dose in rats and obtained bioavailability values 1800 times lower.
It should be noted, however, that serum levels of curcumin in rats and in human are not directly comparable [155], and thus further studies are needed to understand bioavailability from this route.

4. Innovative Curcumin Formulations and High-Performance Delivery Forms

Several curcumin supplements are available in the market [164,165,166,167,168]. These products promise a good availability from oral intake of curcumin, based on different strategies: (i) increasing permeability by formulation with nanoparticles, micelles, or self-emulsifying liposomes; (ii) increasing solubility by micronization or molecular encapsulation with γ-cyclodextrin; and (iii) decreasing metabolization by association with piperine, a substance able to block the metabolic degradation of curcumin. A few oral dosage forms are designed for in situ absorption at the mouth cavity or the colon, easily working around hepatic inactivation. Parenteral and transdermal delivery of curcumin also allow good bioavailability. New formulations in these fields rely on the innovations in micro- and nano-scale carriers.
Further solutions for improving curcumin’s biological performance are under research. Many of these involve chemical modifications to afford curcumin prodrugs or metal complexes with superior solubility in aqueous solution and a prolonged stability once inside the human organism.

4.1. Oral Delivery Forms

4.1.1. Marketed Products

There are, at the present day, many curcumin supplements available in the market which promise improved bioavailability when compared with pure curcumin. Albeit these were not specifically developed for (nor tested on) cancer patients, they can still be quite useful for them owing to their claimed superior bioavailability. Optimized for oral delivery, these formulations can offer good systemic action with doses as low as 30 mg/day. Ingestion of such improved formulations is, thus, much more practical than taking pure curcumin, which would theoretically require a patient to have 5–10 g/day, corresponding to 10–20 tablets. The different formulation strategies for curcumin (or curcuminoid mixtures) available from commercial suppliers are systematized in the Table 4 (next page). Many of these products were tested on human subjects, which comprised healthy volunteers in some instances and patients suffering from inflammatory conditions in a few other studies.
Theracurmin is a stable colloidal suspension of curcuminoids formulated as polysaccharide nanoparticles (NPs). It was tested in a dose of 30 mg on seven healthy volunteers, other seven having 30 mg curcumin powder. The nanoparticles were shown to increase the AUC (area under the curve, a measure of the total bioavailability of curcumin over time) by 27-fold (compared to curcuminoid powder), thus pointing to increased bioavailability [151].
In “Biocurcumax”, curcumin is dispersed in turmeric oil containing 45% of ar-turmerone. This formulation, studied in a group of eight volunteers, showed an absorption maximum after 3 h of ingestion and curcumin levels 5–16-fold higher to those of unformulated curcumin [169]. Blending curcumin with turmeric essential oil is also the strategy used in “Biomor”, which is claimed to have bioavailability five times higher than the competitors’ brands of bioavailable curcumin (namely those using piperine as an absorption enhancer). The reported plasma peak after the intake of 4 g of “Biomor” is observed at 4 h and has a value of ca. 1.6 μg/mL [14].
In “Meriva”, curcumin forms a solid dispersion with “phytosomes” and this blend is dosed in the form of gelatin capsules. Phytosome is the tradename given to liposomes rich in phosphatidylcholine [170,171]. When inside the gastrointestinal (GI) tract, these liposomes facilitate the interaction with biological membranes and increase absorption [171]. Meriva increases the average plasma peak levels of curcumin by roughly five times in male Wistar rats [172]. In humans, it allows lowering daily intake values while keeping a safe and effective action. Two clinical studies with 50 [173] and 100 [174] osteoarthritis patients showed improved mobility upon intake of 200 mg of Meriva curcumin. Liposomal curcumin is also found in several other oral liquid formulations from various manufacturers [175,176,177,178].
Molecular encapsulation of curcumin with γ-cyclodextrin affords the inclusion complex denominated as “Cavacurmin”, which features increased solubility and bioavailability over pure curcumin. Taken as a dietary supplement (as capsules), Cavacurmin is claimed to be transported unchanged through the stomach into the upper intestinal tract, where it dissociates, and curcumin is absorbed [179]. A study on 12 healthy volunteers taking cavacurmin showed a Cmax of ca. 190 nM, which is roughly 40-fold higher than pure curcumin.
Micronization of solid curcuminoids to particles with 10 μm average size and blending them with a lipophilic matrix is another simple strategy to improve absorption and achieve sustained release up to 12 h. The product, commercially available as MicroActive® Curcumin, was tested on three healthy volunteers, affording a ten-fold absorption increase [167].
Combination therapy allies curcumin with piperine to address losses in activity caused by glucuronidation, one of the major metabolic pathways responsible for the low oral bioavailability of curcumin. The alkaloid piperine (1-piperoylpiperidine) is one of the major active components of the pepper plants Piper nigra and Piper longa, known to act as a bioavailability enhancer for a variety of drugs by inhibiting glucuronidation both in the liver and in the small intestine [180]. A comparative study with healthy volunteers taking 2 g/kg curcumin, either alone or in combination with 20 mg/kg of piperine, has shown a 20-fold curcumin bioavailability enhancement for the combined therapy, as well as an earlier plasma peak (Tmax of 40 min vs. Tmax of 60 min for curcumin alone) [181].
Table 4. Curcumin oral dosage formulations currently available in the market.
Table 4. Curcumin oral dosage formulations currently available in the market.
Trade NameFormulation StrategyDosage FormProduct CompositionRef.
CurcuminRichColloidal NP w/curcuminoids
(theracurmin)
30 mg capsulesTheracurmin: polysacharide NPs loaded with curcuminoids
Excipients: MCC, magnesium stearate, silica
[151]
BiocurcumaxCurcuminoids w/turmeric essential oil350 mg capsulesCurcuminoids (titrated 95% curcumin)
Excipients: Maltodextrin, magnesium stearate, HPMC
[164]
BiomorCurcuminoids w/turmeric essential oil500 mg capsulesCurcuminoids (titrated 95% curcumin) in a vegetable capsule[14]
CavacurminCyclodextrin inclusion complex
w/curcumin
Dry powderCavacurmin: curcumin, γ-cyclodextrin[179]
MerivaLiposome500 mg capsules
(100 mg curcumin)
Liposome: Turmeric root extract, phosphatidylcholine, phospholipids
Excipients: HPMC, leucine, silicon dioxide and MCC
[165]
Liposomal curcumin mangoLiposomeLiquid,
20 g/L
Liposome: curcumin, phospholipids from sunflower
Excipients: water, xylitol, mango aroma, ascorbic acid,
Preservative: seabuckthorn extract
[177]
Liposomal curcuminLiposomeLiquid, 41.7 g/LLiposome: Turmetic root extract, phospholipids from sunflower
Excipients: water, glucosylsteviosides, flavors, glycerin, xanthan/acacia gums
Preservative: potassium sorbate
[178]
Dr. Mercola
Curcumin Advanced
Micronized curcuminoids
(microactive)
500 mg capsulesMicronized curcumin
Excipients: HPMC, MCC, polyglycerol oleate, silicon dioxide, metolose, medium chain triglycerides and sodium alginate
[167]
Curcumin C3 ComplexCombined therapy of curcuminoids & piperidine500 + 5 mg capsulesComplex: Turmeric root extract (73% curcumin), piperine
Excipients: rice flour, cellulose and ascorbyl palmitate
[166,168]
Abbreviations: HPMC = hydroxypropylmethylcelluloce, MCC = microcrystalline cellulose; NP = nanoparticle.

4.1.2. Emerging Solutions for Oral Delivery

Curcumin availability from oral delivery can be increased by directing absorption towards specific sites of the GI tract. The incorporation of curcumin into self-micro-emulsion-based capsules is claimed to make the absorption occur only at the colon and to increase absorbability [182]. A gastro-resistant liposome carrier for curcumin is also reported, allowing higher accumulation of curcumin along the intestinal track, particularly at the duodenum [183]. Encapsulation of curcumin into nanoparticles made from fibroin, a silk protein, followed by dispersion into a carrier such as chitosan, is claimed to provide a non-toxic formulation for cancer therapy that can be adapted to oral administration or to other administration routes (by changing the dispersive medium) [184].

4.1.3. Buccal Delivery

Absorption of curcumin in the buccal cavity allows good availability for local action and it has been proposed as a suitable strategy to treat a rare variety of cancer occurring in the mouth, squamous-cell carcinoma of the salivary glands. It is, in fact, an excellent target for local treatment with curcumin. This action was demonstrated in patients treated with chewable tablets of curcumin, which evidenced lower activity of the pro-inflammatory enzymes (IKKβ kinases) in salivary cells [185].

4.2. New Injectable forms of Curcumin

Injectable forms of curcumin are, to the best of our knowledge, currently unavailable in the market. Several parenteral formulations of curcumin are described, including the proprietary formulations named “Lipocurc” and “NanoCurc”, but these are still not approved by the FDA. Lipocurc is a liposomal formulation for curcumin [186] with 100 nm mean particle size. It exhibited excellent in vitro antiproliferative activity against several pancreas cancer cell lines (ASPC-1, BxPC-3, Capan-1, Hs766-T, and MiaPaCa2), with half-maximal inhibitory concentration (IC50) values within 2–7 μM, as well as tumor-reducing and anti-angiogenic effects in mice xenografted with human pancreas carcinoma (BxPC-3 and MiaPaCa2) [187]. Clinical trials (phases 1a and 1b) with Lipocurc in 70 human subjects showed no cardiac toxicity and signs of efficacy in end-stage cancer patients who had failed 6–7 prior chemotherapy regimens; results of phase 2 trials should be available soon [188]. Another claimed application of intravenous liposomal curcumin is in the treatment of neurodegenerative and stress disorders, according to a patent filed in 2011 [189].
Nanocurc is a polymeric nanoparticle formulation of curcumin. The nanoparticles have an average size of 50 nm, being thus smaller than the aforementioned liposomes. They are obtained from the micellar aggregates of cross-linked, random co-polymers of N-isopropylacrylamide (NIPAAM), with N-vinyl-2-pyrrolidone (VP) and poly(ethyleneglycol)-monoacrylate (PEG-A) [190]. Nanocurc showed sustained plasma concentrations of curcumin with Tmax of 2.75 ± 1.50 h and Cmax of 46.62 ± 14.05 μM and a good inhibitory effect on Pa03C (human pancreatic cancer) xenografts in mice [191]. It is able to cross the blood-brain barrier [192], which may be useful in addressing brain tumors.
Curcumin-loaded biodegradable micellar nanoparticles based on a co-polymer of PEG and poly(ε-caprolactone) were developed for the treatment of colon cancer [193,194]. These particles have an average size around 28 nm, being even smaller than those of Nanocurc (described above). In vitro studies showed their direct cytotoxic effect on murine colon carcinoma cells (CT26 line) and in vivo studies on xenograft mice further confirmed the anticancer effect, associated with anti-angiogenesis. The therapeutic action and improved pharmacokinetics of curcumin carried by these micelles makes this system an excellent intravenously injectable aqueous formulation of curcumin, one that has potential clinical application in colon cancer therapy.
PLGA was also used to encapsulate curcumin, both in the form of injectable microparticles [195] and of nanoparticles [196]. The curcumin microparticles had an average diameter around 22 μm and featured sustained release properties over four weeks (in mice), inhibiting the growth of human breast cancer cells (MDA-MB-231 line), both in vitro and in vivo in xenograft mice treated by intraperitoneal injection. It should be stressed, however, that PLGA microparticles cause acidity and inflammatory response at the site of injection due to partial degradation to lactic and glycolic acids, even though the inflammation is attenuated by the anti-inflammatory properties of curcumin.
Carrier-free injectable curcumin can also be prepared by a simple process of nanocrystalization. A nanocrystalline suspension (NS) containing only the nano-sized crystals of the active ingredient and minimal amounts of stabilizer (usually a surfactant or polymer), is thus a carrier-free system and an excellent method to formulate BSC class IV drugs such as curcumin [197]. Curcumin NS was produced by high-pressure homogenization and it showed solubility values for curcumin increased by over 600-fold, having lower vascular irritability and hemolytic effect in rabbits [198]. In vitro tests against HeLa (cervix epithelial cancer) and MCF-7 (breast cancer) human cell lines showed good inhibitory action, with IC50 values around 50 and 35 μM, respectively.

4.3. Inhalable Curcumin Formulations

4.3.1. Dry Powder Inhaler for Pulmonary Delivery

Curcumin delivery by absorption through the pulmonary region aims at achieving high bioavailability with reduced side effects. Curcumin inhalable powders were obtained using carbon dioxide as anti-solvent and polyvinylpyrrolidone and hydroxypropyl-β-cyclodextrin (HPβCD) as excipients [199]. The particles presented good aerodynamic properties and cytotoxicity against the H1299 human non-small cell lung carcinoma line, with an IC50 of 17 μM, but also some toxicity on normal lung cells of the MRC-5 line (IC50 of 29 μM).

4.3.2. Intranasal Aerosol for Delivery to the Brain

Delivery of curcumin to the brain through the olfactory pathway by means of an inhalable aerosol was proposed as an alternative to injectable formulations. For this, curcumin had to be modified to be sufficiently hydrophilic to traverse blood circulation while maintaining the lipophilicity that allows it to cross the blood–brain barrier (BBB) [200]. The study involved a perfluoro derivative of curcumin (for fluorescence imaging) dissolved in PBS/Tween 20 (1:6 ratio) and atomized by a cross-flow apparatus. Intranasal exposure in mice showed that distribution to the brain (cortex, hippocampus, and thalamus areas) was achieved, but it also revealed the presence of the compound in major organs such as the lungs, kidney, and liver.

4.4. Transdermal Administration

4.4.1. Curcumin Topical Formulations with Liposomes, Microemulsions, and Polymeric Nanoparticles

The delivery of curcumin across the skin is of enormous interest not only to the cosmetic industry, due to its anti-inflammatory and rejuvenating properties [201], but also in skin cancer prevention, as demonstrated by various studies with skin cancer models [202,203,204,205,206]. In transdermal delivery formulations of curcumin, the liposome strategy is fine-tuned to afford liposome variants such as the “ethosomes”, containing ethanol to convey flexibility and to act as permeability enhancer [207,208], the propylene glycol liposomes (PegLs), also quite flexible but without the volatility issues associated with ethanol [209] and the “invasomes”, containing a small number of terpenes (0.5 to 1.5%). Terpenes are known to increase diffusivity and partitioning into the skin by disturbing its lipid bilayers [210]. A study comparing PegLs, ethosomes and plain liposomes for curcumin delivery to the skin showed that the PegLs were the best formulation, with superior results in the curcumin loading ability (ca. 93%), the transdermal delivery ratio and curcumin local action, followed by ethosomes and lastly by plain liposomes [211]. Curcumin invasomes were prepared using different terpenes (limonene, fenchone, or neradinol) and HPβCD, to assure a high loading (90%) and tested with an ex vivo skin model (from rat abdominal skin), showing that limonene was the most adequate enhancer for the permeation of curcumin [211]. Formulations containing high amounts of terpenes formed microemulsions due to the oily nature of these compounds. The performance of curcumin microemulsions having 5% or more of limonene, 1,2-cineole or α-terpineol was studied on skin excised from neonate pigs [212]. Once more, limonene proved to be the most adequate penetration enhancer for curcumin.
Another strategy for increasing skin permeability of liposomes is to coat their surface with cell-penetrating peptides (CPPs) by means of a conjugation reaction. CPPs are short peptides, generally not exceeding 30 residues, having both the capacity to ubiquitously cross cellular membranes and very limited toxicity. In comparison with permeation of pure curcumin, CPP-coated liposomes carrying curcumin were able to increase the permeation efficacy by 7-fold, whereas the same liposomes without coating only afforded a 2-fold increase. Distribution and deposition of curcumin following its release from the carriers on mouse skin was also studied. CPP-coated liposomes afforded the highest curcumin content in both dermis and epidermis layers (7.0 and 3.5 μg respectively), followed by conventional liposomes and curcumin solution only (Figure 3). The liposomes coated with CPP are thus a good solution for delivering curcumin in high amounts and helping it reach the deep layers of the skin [213].
Curcumin nanogels for transdermal delivery are also reported. Developed from chitin particles with an average size of 70–80 nm, nanogels offer a good transdermal delivery capacity and the promise of a simple topic formulation for the treatment of melanoma [214]. In vitro studies have demonstrated specific inhibition of the growth of human melanoma cells (A375) with low activity against the non-tumoral human dermal fibroblast cell line (HDF).

4.4.2. Dermal Compositions Containing Silver and Curcumin for Combined Therapy

Combining the activity of curcumin with another cytotoxic agent creates opportunity for more effective treatment of skin pathologies, whether these are the result of infection by bacterial microorganisms or of the uncontrolled growth of the body’s skin cancer cells.
Silver-doped hydrogels were tested as vehicles for skin delivery of curcumin and demonstrated good results in vitro, with loading efficiencies up to 80% and in vitro release higher than 70% after 160 h. Furthermore, the curcumin–silver hydrogel demonstrated a cumulative antibacterial action against Escherichia coli, being thus a promising new cytotoxic agent that may also be tested for cancer treatment [215]. Typically used as antimicrobial, silver has demonstrated antitumoral activity in various studies. Colloidal silver inhibits in vitro growth of MCF-7 breast cancer cells at concentrations > 1 nM [216] and in vivo Dalton’s lymphoma and lymphosarcoma xenografts were inhibited by nanoparticles of silver [217] and silver oxide [218], respectively.

4.4.3. Iontophoresis Techniques

Iontophoresis facilitates the migration of charged compounds across a skin area which is subject to a low-intensity continuous electrical current. The administration of charged curcuminoids by this procedure was first patented in 2009 as a therapy for Alzheimer’s disease [219]. Another study aimed at developing a new breast cancer treatment by formulating curcumin into liposomes and promoting their transdermal migration by iontophoresis [220]. In vitro skin permeation studies on newborn pig skin showed that iontophoresis allowed for the permeation of a cumulative amount and flux of liposomes. The cytotoxic action of the curcumin-loaded liposomes was demonstrated by incubation with different cultured lines of breast cancer cells (MCF-7, MDA-MB-231, and MDA-MB-435S), IC50 values ranging from 15 to 20 μM.

4.4.4. Photodynamic Therapy for Increased Antitumoral Action on Skin

Curcumin can also be used as a photosensitizer for photodynamic therapy of tumors (PDT) [221]. It was already demonstrated that curcumin in the presence of oxygen can be photocytotoxic against mammalian cells [222,223]. When irradiated by light, curcumin produces reactive oxygen species (ROS), such as singlet oxygen and reduced forms of molecular oxygen, as well as carbon-centered radicals [222]. Indeed, some reports have shown that the combination of curcumin with light increases significantly the antitumoral action of curcumin, particularly at very short incubation times (1–2 h) [224,225]. On HeLa cells curcumin exhibited IC50 values of 85.4 µM in the dark and of 8.2 µM under irradiation [225]. Similar situation was reported by Verena et al. using four epithelial liver tumor cells lines (HuH6, HepT1, HepG2, and HC-AFW1), for which the IC50 values diminished 9–26 times under blue light irradiation, varying slightly according to the cell density used, but being always significantly lower than those measured under dark conditions [224]. In addition to curcumin, other curcuminoids may be used as photosensitizers, namely, desmethoxycurcumin and bisdesmethoxycurcumin. On the MDA-MB-231 (breast cancer) line, desmethoxycurcumin had IC50 values of 4 µM and 33 µM under blue light and dark conditions, respectively [226]. Increased growth inhibitory action was observed in human head and neck cancer cells (AMC-HN3) following a photofrin-PDT treatment combined with curcumin, which was postulated to result from mitochondrial-dependent apoptosis due to increased ROS generation [227].

4.5. Curcumin Prodrugs and Metal Complexes

Curcumin metabolization occurs, as described in Section 2.2.1., mainly by conjugation at its phenolic hydroxyl groups. This way, curcumin is often functionalized at these two positions to form derivatives with improved physico-chemical, biological, and pharmacokinetic properties [228,229]. A useful strategy is to design curcumin prodrugs that are able to prevent metabolization. This is achieved by protecting the hydroxyls with functional groups such as amino acids, small carboxylic acids [230,231,232] and even nucleotides [233]. The prodrugs curcumin maleoyl valinate and curcumin maleoyl glycinate were tested against human bladder cancer EJ cells, having shown dose- and time-dependent inhibition [231]. The toxic side effects on healthy human renal proximal cells (HKC) were considerably lower in comparison with pure curcumin. Succinate prodrugs of curcumin with good cytotoxicity against colon adenocarcinoma (Caco-2 cell line) were also reported, with highlight to curcumin diethyl disuccinate, having an IC50 of 1.9 μM [232]. Conjugation of curcumin with an oligonucleotidic chain complementary to a sequence of telomerase RNA allows targeting this specific cellular function [234]. Human prostrate tumor cells (DU145) were transfected with this 5′-O-curcumin DNA prodrug but the antiproliferative action was very slow (20% cell survival after 75 days).
Another class or curcumin prodrugs comprises nanocarriers. These are obtained by functionalizing the phenolic hydroxyls with oligomers or polymers using linkers that are hydrolyzed in vivo by enzymes or pH [235,236,237]. The prodrugs aggregate into nanoparticles, forming a carrier system. Curcumin oligo(ethylene glycol) nanoparticles were shown to reduce the size of intraperitoneal SKOV-3 tumors and subcutaneous (mammary fat pad) MDA-MB-468 tumors in xenografted mice [235]. Additionally, nanocarriers made of biocompatible aluminosilicate clays were reported [238]. These materials are named halloysite nanotubes due to their hollow tubular structure. Curcumin is linked to them by a cysteamine arm, hydrolysable at the disulfide bond. This carrier system can thus be considered a prodrug. In vitro studies with two hepatic carcinoma cell lines showed good cytotoxic activity. At 48 h and using a dose of 50 μM, cell viability rates were low, 22% and 16% for Hep3B and HA22T/VGH, respectively, and at the dose of 100 μM zero viability for the Hep3B line was observed (also after 48 h).
Curcumin coordination to a metal center brings advantages to the solubility and stability (both redox and to light). This is very helpful for the biological activity, and positive results have been reported in a varied range of actions, not only with antitumor compounds, but also with reports of better antioxidant and anti-inflammatory agents and even of new complexes to be used in the treatment of Alzheimer’s disease [239].
The preparation of metal complexes with curcumin typically involves deprotonating curcumin in a first step to form a stable enol form [240]; then, an appropriate molar ratio of curcuminate is made react with a metal precursor. Suitable precursors include halogenates of zinc(II) [241,242,243,244], calcium(II) [244] and iron(III) [244,245], acetates of copper(II) [246,247], palladium(II) [248] and nickel(II), or vanadyl bis(acetylacetonate), VO(acac)2 [249]. The VO(curcuminate)2 complex was very successful in inhibiting mouse lymphoma cells (L1210), with an IC50 of ca. 15 μM. It was, however, also toxic towards rat smooth muscle cells (CRL-1444), with an IC50 of 2.9 μM [249]. Curcumin zinc complexes with dinonyl-2,2′-bipyridine (bpy-9) and 4,4′-bis(hydroxymethyl)-2,2′-bipyridine as the spectator ligands presented good inhibitory action on the prostate cancer cell lines DU145, PC3, and LNCaP, and the neuroblastoma SHSY-5Y and SD LAN-5 lines, with IC50 values in the 12–37 μM range at 72 h [242]. [(bpy-9)Zn(curcumin)]BF4, a cationic derivative, was also very active against the SH-SY5Y neuroblastoma cell line [243]. A palladium analogue, [(bpy-9)Zn(curcumin)]CF3SO3, was tested successfully on prostate cancer cells (LNCaP, PC3, and DU145) having IC50 values in the 20–24 μM range [248]. Curcumin palladium complexes bearing cyclohexane-diamine derivatives as the spectator ligand were developed for colon cancer treatment, having IC50 values in the 10–34 μM range against the cell lines DLD-1, HT-29 and CRL-1790 [250].
Curcumin ruthenium(II) organometallic complexes obtained from precursors that share a common “piano-stool” geometry are represented in Figure 4 [251,252,253].
The arene that comprises the “top” of the piano stool was either a hexamethylbenzene or a p-cymene. Curcumin coordinated at the positions of two of the legs and the third leg was a chloride, in some instances replaced by 1,3,5-triaza-7-phospha-adamantane (PTA) for higher aqueous solubility. The complexes with PTA displayed the most promising activities, being able to inhibit the growth of breast cancer A2780 cells (both cisplatin-sensitive and cisplatin-resistant lines) at sub-micromolar concentrations [251]. Their IC50 on these cell lines were 0.4 μM, approximately ten times lower than pure curcumin (IC50 = 4.3 μM). Safety was evaluated on a healthy embryonic kidney cell line, HEK293. Ru(hexamethylbenzene)(curcumin)(PTA) had an IC50 value of 4.3 μM against this cell line while for Ru(p-cymene)(curcumin)(PTA) IC50 was 9.1 μM [251]. The activity of Ru(p-cymene)(curcumin)Cl was inferior or equal to that of curcumin on the tested cancer cell lines [252].
Analogues of curcumin arene complexes can be obtained by replacing the arene with the heterocyclic moiety of 1,4,7-tritiacyclononane (abbreviated as [9]ansS3), to afford complexes with higher polarity and thus better compatibility with aqueous media. The complex [Ruthenium(II)(curcumin([9]aneS3)DMSO]Cl was reported as a strong DNA-binding agent, with higher affinity than pure curcumin [254]. Cytotoxicity against the PC-3 prostate cancer cell line was, however, not observed at concentrations as high as 80 mM. This suggests that cell-penetrating abilities of the complex need to be ameliorated.
Two ruthenium polypyridyl complexes with curcumin, with the formulae [Ruthenium(II)(curcumin)(bpy)(N,N′-Lig)]Cl, where N,N′-Lig = 2,2′bipyridine (bpy) or dppn (benzo-[i]dipyrido [3,2-a:2′,3′-c]phenazine), were recently reported [255]. Despite being cationic, the complexes were lipophylic, with log p values of 0.75 for the one with two bipyridines and 1.06 for the one with dppn. The in vitro inhibitory activity was evaluated on the A549 human non-small-cell lung cancer line, the MCF-7 human breast adenocarcinoma cell line and the SGC7901 human gastric cancer cell line, with the dppn-bearing complex having exhibited higher activity: IC50 values were within the 2.1 ± 0.2 to 2.7 ± 0.2 μM range. Compared to curcumin and cisplatin, which inhibited the same cell lines with IC50 values of 11.4 ± 0.3 to 15.4 ± 1.0, the complex is quite more potent. Additional biomolecular studies on the A549 line have shown that the mode of action of the complexes involves apoptosis.

5. Conclusions

The present review presents a concise description on the history of curcumin as a therapeutic agent. From the digestive and choleretic aid in the past centuries to the exciting antitumoral new agent which attracts the interest and hope of both patients and scientist of the present day, curcumin has been a constant and much appreciated herbal medicine for mankind. As the pharmacokinetics of curcumin became more well-known, its low bioavailability was evidenced. Since then, many efforts to circumvent it were undertaken. Fruitful results are substantiated by the large number of curcumin-based supplements on the market, eight of which are formulated to assure a superior bioavailability (Table 4), and by the countless oral, injectable, intranasal, inhalable, and dermal formulations which are either pending approval or at the R&D stage.
Future directions along this yellow-brick road will need to address the long-term safety of curcumin, a sine-qua-non-requirement to assure its approval as a drug that is safe for the treatment of oncologic and other diseases, be it ingested, administered parenterally or topically or even implanted, to treat bone cancer [256]. Regarding long-term exposure associated with, thus far only a study in animals raised a slight warning, which involved possible iron deficiency [257]. It is our opinion that the large number of people presently incorporating curcumin supplements into their daily dietary routine will contribute to produce reports and feedback that will serve as a natural database of information of curcumin tolerability and safety. This way, the knowledge of curcumin long-term safety is expected to expand tremendously in the years to come, hopefully dissipating concerns and helping to establish it as a recognized therapy for cancer.

Author Contributions

Conceptualization, M.C.H., M.A.F.F. and S.S.B.; writing—original draft preparation, M.H.C., and S.S.B.; writing—review and editing, M.C.H., M.A.F.F. and S.S.B. All authors have read and agreed to the published version of the manuscript.

Funding

The authors acknowledge the University of Aveiro and FCT/MCTES (Fundação para a Ciência e a Tecnologia, Ministério da Ciência, Tecnologia e Ensino Superior) through national funds and, where applicable, co-financed by the FEDER (European Fund for Regional Development) within the PT2020 Partnership Agreement, for general financial support to LAQV-REQUIMTE (UIDB/50006/2020).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Akram, M.; Uddin, S.; Ahmed, A.; Khan, U.; Hannan, A.; Mohihuddin, E.; Asif, M. Curcuma Longa and Curcumin. Rom. J. Biol. Plant. Biol. 2010, 55, 65–70. [Google Scholar]
  2. 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]
  3. Nabavi, S.M.; Daglia, M.; Moghaddam, A.H.; Habtemariam, S. Curcumin and Liver Disease: From Chemistry to Medicine. Compr. Rev. Food Sci. Food Saf. 2013, 13, 62–77. [Google Scholar] [CrossRef]
  4. Mukherjee, P.K.; Wahile, A. Integrated approaches towards drug development from Ayurveda and other Indian system of medicines. J. Ethnopharmacol. 2006, 103, 25–35. [Google Scholar] [CrossRef]
  5. de Orta, G. Colóquios Dos Simples, E Drogas E Coisas Medicinais Da India E Assim de Algumas Frutas Achadas Nela Onde Se Tratam Algumas Coisas Tocantes a Medicina Prática, E Outras Coisas Boas Para Saber; Ioannes de Endem: Goa, India, 1563. [Google Scholar]
  6. Govindarajan, V.; Stahl, W.H. Turmeric—chemistry, technology, and quality. CRC Crit. Rev. Food Sci. Nutr. 1980, 12, 199–301. [Google Scholar] [CrossRef]
  7. Prasad, S.; Aggarwal, B.B. Turmeric, the Golden Spice: From Traditional Medicine to Modern Medicine. In Herbal Medicine: Biomolecular and Clinical Aspects, 2nd ed.; Benzie, I.F.F., Wachtel-Galor, S., Eds.; CRC Press: Boca Raton, FL, USA, 2011. [Google Scholar]
  8. Kita, T.; Imai, S.; Sawada, H.; Kumagai, H.; Seto, H. The Biosynthetic Pathway of Curcuminoid in Turmeric (Curcuma longa) as Revealed by13C-Labeled Precursors. Biosci. Biotechnol. Biochem. 2008, 72, 1789–1798. [Google Scholar] [CrossRef] [Green Version]
  9. Aggarwal, B.B.; Sundaram, C.; Malani, N.; Ichikawa, H. Curcumin: The Indian Solid Gold. In Results and Problems in Cell Differentiation; Springer Science and Business Media LLC: Berlin, Germany, 2007; Volume 595, pp. 1–75. [Google Scholar]
  10. Toda, S.; Miyase, T.; Arichi, H.; Tanizawa, H.; Takino, Y. Natural antioxidants. III. Antioxidative components isolated from rhizome of Curcuma longa L. Chem. Pharm. Bull. 1985, 33, 1725–1728. [Google Scholar] [CrossRef] [Green Version]
  11. Osawa, T.; Sugiyama, Y.; Inayoshi, M.; Kawakishi, S. Antioxidative Activity of Tetrahydrocurcuminoids. Biosci. Biotechnol. Biochem. 1995, 59, 1609–1612. [Google Scholar] [CrossRef]
  12. FDA. Inventory of GRAS Notices, Number 460. Curcuminoids Purified from Turmeric (Curcuma longa L.); Food and Drugs Administration: Silver Spring, MD, USA, 2013. Available online: https://www.cfsanappsexternal.fda.gov/scripts/fdcc/?set=GRASNotices&id=460&sort=GRN_No&order=DESC&startrow=1&type=basic&search=460 (accessed on 15 December 2020).
  13. Joint FAO/WHO Expert Committee on Food Additives (JECFA). Turmeric and curcumin (WHO Food Additives Series 6). Available online: http://www.inchem.org/documents/jecfa/jecmono/v06je29.htm (accessed on 16 November 2020).
  14. Anna’s Tuin & Ruigte. Available online: http://annastuinenruigte.nl/ (accessed on 24 March 2020).
  15. The State of the Curcumin Market. Available online: http://www.naturalproductsinsider.com/articles/2015/12/the-state-of-the-curcumin-market.aspx (accessed on 16 November 2020).
  16. Curcumin Market. Size, Share & Trends Analysis Report by Application (Pharmaceutical, Food, Cosmetics), By Region (North America, Europe, Asia Pacific, Central & South America, Middle East & Africa), And Segment Forecasts, 2020–2027. Available online: http://www.grandviewresearch.com/industry-analysis/turmeric-extract-curcumin-market (accessed on 16 November 2020).
  17. Pelletier, J.; Vogel, A. Examen Chimique de La Racine de Curcuma. J. Pharm. 1815, 1, 289–300. [Google Scholar]
  18. Daube, F.W. Ueber Curcumin, den Farbstoff der Curcumawurzel. Eur. J. Inorg. Chem. 1870, 3, 609–613. [Google Scholar] [CrossRef]
  19. Lampe, V.; Milobedzka, J. Studien Über Curcumin. Berichte der Dtsch. Chem. Gesellschaft 1913, 46, 2235–2240. [Google Scholar] [CrossRef]
  20. Priyadarsini, K.I. The Chemistry of Curcumin: From Extraction to Therapeutic Agent. Molecules 2014, 19, 20091–20112. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Carvalho, D.D.M.; Takeuchi, K.P.; Geraldine, R.M.; De Moura, C.J.; Torres, M.C.L. Production, solubility and antioxidant activity of curcumin nanosuspension. Food Sci. Technol. 2015, 35, 115–119. [Google Scholar] [CrossRef] [Green Version]
  22. Chavda, H.; Patel, C.; Anand, I. Biopharmaceutics classification system. Syst. Rev. Pharm. 2010, 1, 62. [Google Scholar] [CrossRef]
  23. Hatcher, H.; Planalp, R.; Cho, J.; Tortia, F.M.; Tortic, S.V. Curcumin: From ancient medicine to current clinical trials. Cell. Mol. Life Sci. 2008, 65, 1631–1652. [Google Scholar] [CrossRef]
  24. Bernabé-Pineda, M.; Ramírez-Silva, M.T.; Romero-Romo, M.; González-Vergara, E.; Rojas-Hernández, A. Determination of acidity constants of curcumin in aqueous solution and apparent rate constant of its decomposition. Spectrochim. Acta Part. A Mol. Biomol. Spectrosc. 2004, 60, 1091–1097. [Google Scholar] [CrossRef]
  25. Chignell, C.F.; Bilskj, P.; Reszka, K.J.; Motten, A.G.; Sik, R.H.; Dahl, T.A. spectral and photochemical properties of curcumin. Photochem. Photobiol. 1994, 59, 295–302. [Google Scholar] [CrossRef]
  26. 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]
  27. Khurana, A.; Ho, C.-T. High Performance Liquid Chromatographic Analysis of Curcuminoids and Their Photo-oxidative Decomposition Compounds in Curcuma Longa L. J. Liq. Chromatogr. 1988, 11, 2295–2304. [Google Scholar] [CrossRef]
  28. Tønnesen, H.H.; Karlsen, J.; van Henegouwen, G.B. Studies on Curcumin and Curcuminoids VIII. Photochemical Stability of Curcumin. Z. Lebensm. Unters. Forsch. 1986, 183, 116–122. [Google Scholar] [CrossRef]
  29. De Jager, P. Turmeric: The Ayurvedic Spice of Life, 2nd ed.; Pioneer imprints: Maui, HI, USA, 2010. [Google Scholar]
  30. De Oliveira, M.R.; Jardim, F.R.; Setzer, W.N.; Nabavi, S.M. Curcumin, mitochondrial biogenesis, and mitophagy: Exploring recent data and indicating future needs. Biotechnol. Adv. 2016, 34, 813–826. [Google Scholar] [CrossRef] [PubMed]
  31. Holt, P.R.; Katz, S.; Kirshoff, R. Curcumin Therapy in Inflammatory Bowel Disease: A Pilot Study. Dig. Dis. Sci. 2005, 50, 2191–2193. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Shishodia, S.; Sethi, G.; Aggarwal, B.B. Curcumin: Getting Back to the Roots. Ann. N. Y. Acad. Sci. 2005, 1056, 206–217. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Subramanian, M.; Sreejayan; Devasagayam, T.P.; Singh, B. Diminution of singlet oxygen-induced DNA damage by curcmin and related antioxidants. Mutat. Res. Mol. Mech. Mutagen. 1994, 311, 249–255. [Google Scholar] [CrossRef]
  34. Iqbal, M.; Sharma, S.D.; Okazaki, Y.; Fujisawa, M.; Okada, S. Dietary Supplementation of Curcumin Enhances Antioxidant and Phase II Metabolizing Enzymes in ddY Male Mice: Possible Role in Protection against Chemical Carcinogenesis and Toxicity. Pharmacol. Toxicol. 2003, 92, 33–38. [Google Scholar] [CrossRef]
  35. Kuo, M.-L.; Huang, T.-S.; Lin, J.-K. Curcumin, an antioxidant and anti-tumor promoter, induces apoptosis in human leukemia cells. Biochim. et Biophys. Acta (BBA) Mol. Basis Dis. 1996, 1317, 95–100. [Google Scholar] [CrossRef] [Green Version]
  36. Sreejayan; Rao, M.N.A. Curcuminoids as Potent Inhibitors of Lipid Peroxidation. J. Pharm. Pharmacol. 1994, 46, 1013–1016. [Google Scholar] [CrossRef]
  37. Busquets, S.; Carbó, N.; Almendro, V.; Quiles, M.T.; López-Soriano, F.J.; Argilés, J.M. Curcumin, a natural product present in turmeric, decreases tumor growth but does not behave as an anticachectic compound in a rat model. Cancer Lett. 2001, 167, 33–38. [Google Scholar] [CrossRef]
  38. Duvoix, A.; Blasius, R.; Delhalle, S.; Schnekenburger, M.; Morceau, F.; Henry, E.; Dicato, M.; Diederich, M. Chemopreventive and therapeutic effects of curcumin. Cancer Lett. 2005, 223, 181–190. [Google Scholar] [CrossRef]
  39. Garcea, G.; Berry, D.P.; Jones, D.J.L.; Singh, R.; Dennison, A.R.; Farmer, P.B.; Sharma, R.A.; Steward, W.P.; Gescher, A.J. Consumption of the putative chemopreventive agent curcumin by cancer patients: Assessment of curcumin levels in the colorectum and their pharmacodynamic consequences. Cancer Epidem. Biomarkers Prev. 2005, 14, 120–125. [Google Scholar]
  40. Bentham Science Publisher Tzeng-Horng Leu; Bentham Science Publisher Ming-Chei Maa. The Molecular Mechanisms for the Antitumorigenic Effect of Curcumin. Curr. Med. Chem. Agents 2002, 2, 357–370. [Google Scholar] [CrossRef]
  41. Gupta, S.C.; Prasad, S.; Kim, J.H.; Patchva, S.; Webb, L.J.; Priyadarsini, I.K.; Aggarwal, B.B. Multitargeting by Curcumin as Revealed by Molecular Interaction Studies. Nat. Prod. Rep. 2011, 28, 1937. [Google Scholar] [CrossRef] [Green Version]
  42. Ravindran, J.; Prasad, S.; Aggarwal, B.B. Curcumin and Cancer Cells: How Many Ways Can Curry Kill Tumor Cells Selectively? AAPS J. 2009, 11, 495–510. [Google Scholar] [CrossRef] [PubMed]
  43. Strimpakos, A.S.; Sharma, R.A. Curcumin: Preventive and Therapeutic Properties in Laboratory Studies and Clinical Trials. Antioxidants Redox Signal. 2008, 10, 511–546. [Google Scholar] [CrossRef] [PubMed]
  44. Chainani-Wu, N. Safety and Anti-Inflammatory Activity of Curcumin: A Component of Tumeric (Curcuma longa). J. Altern. Complement. Med. 2003, 9, 161–168. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Perrone, D.; Ardito, F.; Giannatempo, G.; Dioguardi, M.; Troiano, G.; Russo, L.L.; De Lillo, A.; Laino, L.; Muzio, L.L. Biological and therapeutic activities, and anticancer properties of curcumin. Exp. Ther. Med. 2015, 10, 1615–1623. [Google Scholar] [CrossRef] [Green Version]
  46. Satoskar, R.R.; Shah, S.J.; Shenoy, S.G. Evaluation of anti-inflammatory property of curcumin (diferuloyl methane) in patients with postoperative inflammation. Int. J. Clin. Pharmacol. Ther. Toxicol. 1986, 24, 651–654. [Google Scholar] [PubMed]
  47. Siviero, A.; Gallo, E.; Maggini, V.; Gori, L.; Mugelli, A.; Firenzuoli, F.; Vannacci, A. Curcumin, a golden spice with a low bioavailability. J. Herb. Med. 2015, 5, 57–70. [Google Scholar] [CrossRef]
  48. Chandran, B.; Goel, A. A Randomized, Pilot Study to Assess the Efficacy and Safety of Curcumin in Patients with Active Rheumatoid Arthritis. Phytotherapy Res. 2012, 26, 1719–1725. [Google Scholar] [CrossRef]
  49. Heng, M.; Song, M.; Harker, J. Drug-induced suppression of phosphorylase kinase activity correlates with resolution of psoriasis as assessed by clinical, histological and immunohistochemical parameters. Br. J. Dermatol. 2000, 143, 937–949. [Google Scholar] [CrossRef]
  50. Hanai, H.; Iida, T.; Takeuchi, K.; Watanabe, F.; Maruyama, Y.; Andoh, A.; Tsujikawa, T.; Fujiyama, Y.; Mitsuyama, K.; Sata, M.; et al. Curcumin Maintenance Therapy for Ulcerative Colitis: Randomized, Multicenter, Double-Blind, Placebo-Controlled Trial. Clin. Gastroenterol. Hepatol. 2006, 4, 1502–1506. [Google Scholar] [CrossRef] [PubMed]
  51. Lahiff, C.; Moss, A.C. Curcumin for clinical and endoscopic remission in ulcerative colitis. Inflamm. Bowel Dis. 2011, 17, E66. [Google Scholar] [CrossRef]
  52. Epstein, J.; Docena, G.; MacDonald, T.T.; Sanderson, I.R. Curcumin Suppresses p38 Mitogen-Activated Protein Kinase Activation, Reduces IL-1beta and Matrix Metalloproteinase-3 and Enhances IL-10 in the Mucosa of Children and Adults with Inflammatory Bowel Disease. Br. J. Nutr. 2010, 103, 824–832. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Bundy, R.; Walker, A.F.; Middleton, R.W.; Booth, J. Turmeric Extract May Improve Irritable Bowel Syndrome Symptomology in Otherwise Healthy Adults: A Pilot Study. J. Altern. Complement. Med. 2004, 10, 1015–1018. [Google Scholar] [CrossRef] [PubMed]
  54. Lal, B.; Kapoor, A.K.; Agrawal, P.K.; Asthana, O.P.; Srimal, R.C. Role of curcumin in idiopathic inflammatory orbital pseudotumours. Phytotherapy Res. 2000, 14, 443–447. [Google Scholar] [CrossRef]
  55. Lal, B.; Kapoor, A.K.; Asthana, O.P.; Agrawal, P.K.; Prasad, R.; Kumar, P.; Srimal, R.C. Efficacy of Curcumin in the Management of Chronic Anterior Uveitis. Phytother. Res. 1999, 13, 318–322. [Google Scholar] [CrossRef]
  56. Itokawa, H.; Shi, Q.; Akiyama, T.; Morris-Natschke, S.L.; Lee, K.-H. Recent advances in the investigation of curcuminoids. Chin. Med. 2008, 3. [Google Scholar] [CrossRef] [Green Version]
  57. Reddy, A.P.; Lokesh, B. Studies on spice principles as antioxidants in the inhibition of lipid peroxidation of rat liver microsomes. Mol. Cell. Biochem. 1992, 111, 117–124. [Google Scholar] [CrossRef] [PubMed]
  58. Noorafshan, A.; Ashkani-Esfahani, S. A Review of Therapeutic Effects of Curcumin. Curr. Pharm. Des. 2013, 19, 2032–2046. [Google Scholar]
  59. Balamurugan, A.; Akhov, L.; Selvaraj, G.; Pugazhenthi, S. Induction of Antioxidant Enzymes by Curcumin and Its Analogues in Human Islets. Pancreas 2009, 38, 454–460. [Google Scholar] [CrossRef]
  60. Pivari, F.; Mingione, A.; Brasacchio, C.; Soldati, L. Curcumin and Type 2 Diabetes Mellitus: Prevention and Treatment. Nutrients 2019, 11, 1837. [Google Scholar] [CrossRef] [Green Version]
  61. Hatcher, H.C.; Torti, F.M.; Torti, S.V. Curcumin, Oxidative Stress, and Cancer Therapy. In Oxidative Stress in Cancer Biology and Therapy; Springer Science and Business Media LLC: Berlin, Germany, 2012; pp. 233–256. [Google Scholar]
  62. Kuttan, R.; Sudheeran, P.; Josph, C. Turmeric and Curcumin as Topical Agents in Cancer Therapy. Tumori J. 1987, 73, 29–31. [Google Scholar] [CrossRef]
  63. Carroll, R.E.; Benya, R.V.; Turgeon, D.K.; Vareed, S.; Neuman, M.; Rodriguez, L.; Kakarala, M.; Carpenter, P.M.; McLaren, C.; Meyskens, F.L.; et al. Phase IIa Clinical Trial of Curcumin for the Prevention of Colorectal Neoplasia. Cancer Prev. Res. 2011, 4, 354–364. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Cruz–Correa, M.; Shoskes, D.A.; Sanchez, P.; Zhao, R.; Hylind, L.M.; Wexner, S.D.; Giardiello, F.M. Combination Treatment With Curcumin and Quercetin of Adenomas in Familial Adenomatous Polyposis. Clin. Gastroenterol. Hepatol. 2006, 4, 1035–1038. [Google Scholar] [CrossRef] [PubMed]
  65. He, Z.; Shi, C.-B.; Wen, H.; Li, F.-L.; Wang, B.-L.; Wang, J. Upregulation of p53 Expression in Patients with Colorectal Cancer by Administration of Curcumin. Cancer Investig. 2011, 29, 208–213. [Google Scholar] [CrossRef]
  66. Ide, H.; Tokiwa, S.; Sakamaki, K.; Nishio, K.; Isotani, S.; Muto, S.; Hama, T.; Masuda, H.; Horie, S. Combined inhibitory effects of soy isoflavones and curcumin on the production of prostate-specific antigen. Prostate 2010, 70, 1127–1133. [Google Scholar] [CrossRef]
  67. Mahammedi, H.; Planchat, E.; Pouget, M.; Durando, X.; Curé, H.; Guy, L.; Van-Praagh, I.; Savareux, L.; Atger, M.; Bayet-Robert, M.; et al. The New Combination Docetaxel, Prednisone and Curcumin in Patients with Castration-Resistant Prostate Cancer: A Pilot Phase II Study. Oncology 2016, 90, 69–78. [Google Scholar] [CrossRef]
  68. Thomas, R.J.; Williams, M.M.A.; Sharma, H.; Chaudry, A.; Bellamy, P. A double-blind, placebo RCT evaluating the effect of a polyphenol-rich whole food supplement on PSA progression in men with prostate cancer: The U.K. National Cancer Research Network (NCRN) Pomi-T study. J. Clin. Oncol. 2013, 31. [Google Scholar] [CrossRef]
  69. Taleban, R.R.F.-A.; Hejazi, J. A Pilot Clinical Trial of Radioprotective Effects of Curcumin Supplementation in Patients with Prostate Cancer. J. Cancer Sci. Ther. 2013, 5, 320–324. [Google Scholar] [CrossRef] [Green Version]
  70. Hejazi, J.; Rastmanesh, R.; Taleban, F.-A.; Molana, S.-H.; Hejazi, E.; Ehtejab, G.; Hara, N. Effect of Curcumin Supplementation During Radiotherapy on Oxidative Status of Patients with Prostate Cancer: A Double Blinded, Randomized, Placebo-Controlled Study. Nutr. Cancer 2016, 68, 77–85. [Google Scholar] [CrossRef]
  71. Bayet-Robert, M.; Kwiatkowski, F.; Leheurteur, M.; Gachon, F.; Planchat, E.; Abrial, C.; Mouret-Reynier, M.A.; Durando, X.; Barthomeuf, C.; Chollet, P. Phase I dose escalation trial of docetaxel plus curcumin in patients with advanced and metastatic breast cancer. Cancer Biol. Ther. 2010, 9, 8–14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Ryan, J.L.; Heckler, C.E.; Guido, J.J.; Peoples, A.R.; Gewandter, J.S.; Ling, M.; Vinciguerra, V.P.; Anderson, T.; Evans, L.; Wade, J.; et al. Oral curcumin for radiation dermatitis: A URCC NCORP study of 686 breast cancer patients. Support. Care Cancer 2017, 26, 1543–1552. [Google Scholar] [CrossRef]
  73. Dhillon, N.; Aggarwal, B.B.; Newman, R.A.; Wolff, R.; Kunnumakkara, A.B.; Abbruzzese, J.L.; Ng, C.S.; Badmaev, V.; Kurzrock, R. Phase II Trial of Curcumin in Patients with Advanced Pancreatic Cancer. Clin. Cancer Res. 2008, 14, 4491–4499. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Kanai, M.; Yoshimura, K.; Asada, M.; Imaizumi, A.; Suzuki, C.; Matsumoto, S.; Nishimura, T.; Mori, Y.; Masui, T.; Kawaguchi, Y.; et al. A phase I/II study of gemcitabine-based chemotherapy plus curcumin for patients with gemcitabine-resistant pancreatic cancer. Cancer Chemother. Pharmacol. 2010, 68, 157–164. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Devi, K.P.; Tamilselvam, R.; Skalicka-Woźniak, K.; Nabavi, S.M.; Daglia, M.; Bishayee, A.; Pazoki-Toroudi, H. Molecular targets of curcumin for cancer therapy: An updated review. Tumor Biol. 2016, 37, 13017–13028. [Google Scholar] [CrossRef]
  76. Gupta, S.C.; Patchva, S.; Koh, W.; Aggarwal, B.B. Discovery of curcumin, a component of golden spice, and its miraculous biological activities. Clin. Exp. Pharmacol. Physiol. 2012, 39, 283–299. [Google Scholar] [CrossRef]
  77. Libermann, T.A.; Zerbini, L.F. Targeting Transcription Factors for Cancer Gene Therapy. Curr. Gene Ther. 2006, 6, 17–33. [Google Scholar] [CrossRef]
  78. Xiang, L.; He, B.; Liu, Q.; Hu, D.; Liao, W.; Li, R.; Peng, X.; Wang, Q.; Zhao, G. Antitumor effects of curcumin on the proliferation, migration and apoptosis of human colorectal carcinoma HCT-116 cells. Oncol. Rep. 2020, 44, 1997–2008. [Google Scholar] [CrossRef]
  79. Juturu, V.; Sahin, K.; Pala, R.; Tuzcu, M.; Ozdemir, O.; Orhan, C.; Sahin, N. Curcumin prevents muscle damage by regulating NF-kB and Nrf2 pathways and improves performance: An in vivo model. J. Inflamm. Res. 2016, 9, 147–154. [Google Scholar] [CrossRef] [Green Version]
  80. Vadhan-Raj, S.; Weber, D.M.; Wang, M.; Giralt, S.A.; Thomas, S.K.; Alexanian, R.; Zhou, X.; Patel, P.; Bueso-Ramos, C.E.; Newman, R.A.; et al. Curcumin Downregulates NF-kB and Related Genes in Patients with Multiple Myeloma: Results of a Phase I/II Study. Blood 2015, 110, 1177. [Google Scholar] [CrossRef]
  81. Wagner, E.F. Functions of AP1 (Fos/Jun) in bone development. Ann. Rheum. Dis. 2002, 61, 40–42. [Google Scholar] [CrossRef] [PubMed]
  82. Dorai, T.; Aggarwal, B.B. Role of chemopreventive agents in cancer therapy. Cancer Lett. 2004, 215, 129–140. [Google Scholar] [CrossRef] [PubMed]
  83. Bierhaus, A.; Zhang, Y.; Quehenberger, P.; Luther, T.; Haase, M.; Müller, M.; Mackman, N.; Ziegler, R.; Nawroth, P.P. The dietary pigment curcumin reduces endothelial tissue factor gene expression by inhibiting binding of AP-1 to the DNA and activation of NF-kappa B. Thromb. Haemost. 1997, 77, 772–782. [Google Scholar] [CrossRef]
  84. Han, S.-S.; Keum, Y.-S.; Seo, H.-J.; Surh, Y.-J. Curcumin Suppresses Activation of NF-kappaB and AP-1 Induced by Phorbol Ester in Cultured Human Promyelocytic Leukemia Cells. J. Biochem. Mol. Biol. 2002, 35, 337–342. [Google Scholar] [PubMed] [Green Version]
  85. Balasubramanian, S.; Eckert, R.L. Curcumin Suppresses AP1 Transcription Factor-dependent Differentiation and Activates Apoptosis in Human Epidermal Keratinocytes. J. Biol. Chem. 2007, 282, 6707–6715. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Nakamura, K.; Yasunaga, Y.; Segawa, T.; Ko, D.; Moul, J.W.; Srivastava, S.; Rhim, J.S. Curcumin down-regulates AR gene expression and activation in prostate cancer cell lines. Int. J. Oncol. 2002, 21, 825–830. [Google Scholar] [CrossRef] [PubMed]
  87. Teiten, M.-H.; Gaascht, F.; Eifes, S.; Dicato, M.; Han, B.W. Chemopreventive potential of curcumin in prostate cancer. Genes Nutr. 2010, 5, 61–74. [Google Scholar] [CrossRef] [Green Version]
  88. Mishra, A.; Kumar, R.; Tyagi, A.; Kohaar, I.; Hedau, S.; Bharti, A.C.; Sarker, S.; Dey, D.; Saluja, D.; Das, B. Curcumin modulates cellular AP-1, NF-kB, and HPV16 E6 proteins in oral cancer. Ecancermedicalscience 2015, 9. [Google Scholar] [CrossRef] [Green Version]
  89. O’Shea, J.J.; Schwartz, D.M.; Villarino, A.V.; Gadina, M.; McInnes, I.B.; Laurence, A. The JAK-STAT Pathway: Impact on Human Disease and Therapeutic Intervention. Annu. Rev. Med. 2015, 66, 311–328. [Google Scholar] [CrossRef] [Green Version]
  90. Yang, C.-L.; Liu, Y.; Ma, Y.-G.; Xue, Y.-X.; Liu, D.-G.; Ren, Y.; Liu, X.-B.; Li, Y.; Li, Z. Curcumin Blocks Small Cell Lung Cancer Cells Migration, Invasion, Angiogenesis, Cell Cycle and Neoplasia through Janus Kinase-STAT3 Signalling Pathway. PLoS ONE 2012, 7, e37960. [Google Scholar] [CrossRef]
  91. Kroon, P.; Berry, P.A.; Stower, M.J.; Rodrigues, G.; Mann, V.M.; Simms, M.; Bhasin, D.; Chettiar, S.; Li, C.; Li, P.-K.; et al. JAK-STAT Blockade Inhibits Tumor Initiation and Clonogenic Recovery of Prostate Cancer Stem-like Cells. Cancer Res. 2013, 73, 5288–5298. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Weissenberger, J.; Priester, M.; Bernreuther, C.; Rakel, S.; Glatzel, M.; Seifert, V.; Kögel, D. Dietary Curcumin Attenuates Glioma Growth in a Syngeneic Mouse Model by Inhibition of the JAK1,2/STAT3 Signaling Pathway. Clin. Cancer Res. 2010, 16, 5781–5795. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Blasius, R.; Reuter, S.; Henry, E.; Dicato, M.; Diederich, M. Curcumin regulates signal transducer and activator of transcription (STAT) expression in K562 cells. Biochem. Pharmacol. 2006, 72, 1547–1554. [Google Scholar] [CrossRef] [PubMed]
  94. Saydmohammed, M.; Joseph, D.; Syed, V. Curcumin suppresses constitutive activation of STAT-3 by up-regulating protein inhibitor of activated STAT-3 (PIAS-3) in ovarian and endometrial cancer cells. J. Cell. Biochem. 2010, 110, 447–456. [Google Scholar] [CrossRef]
  95. Kumar, A.; Dhawan, S.; Hardegen, N.J.; Aggarwal, B.B. Curcumin (Diferuloylmethane) Inhibition of Tumor Necrosis Factor (TNF)-Mediated Adhesion of Monocytes to Endothelial Cells by Suppression of Cell Surface Expression of Adhesion Molecules and of Nuclear Factor-kappaB Activation. Biochem. Pharmacol. 1998, 55, 775–783. [Google Scholar] [CrossRef]
  96. Kang, E.S.; Woo, I.S.; Kim, H.J.; Eun, S.Y.; Paek, K.S.; Chang, K.C.; Lee, J.H.; Lee, H.T.; Kim, J.-H.; Kim, H.J.; et al. Up-regulation of aldose reductase expression mediated by phosphatidylinositol 3-kinase/Akt and Nrf2 is involved in the protective effect of curcumin against oxidative damage. Free. Radic. Biol. Med. 2007, 43, 535–545. [Google Scholar] [CrossRef]
  97. Kahl, G. Nuclear Factor. In The Dictionary of Genomics, Transcriptomics and Proteomics; Wiley: Hoboken, NJ, USA, 2015; Volume 103, p. 1. [Google Scholar]
  98. Bharti, A.C.; Donato, N.; Aggarwal, B.B. Curcumin (Diferuloylmethane) Inhibits Constitutive and IL-6-Inducible STAT3 Phosphorylation in Human Multiple Myeloma Cells. J. Immunol. 2003, 171, 3863–3871. [Google Scholar] [CrossRef]
  99. Chen, W.; Chen, Y.; Cui, G.-H.; Gu, J.; Hu, N.; Li, X.-G. Effect of curcumin on STAT5 signaling pathway in primary CML cells. Zhongguo Shi Yan Xue Ye Xue Za Zhi 2004, 12, 572–576. [Google Scholar]
  100. Chen, W.; Chen, Y.; Gu, J.; He, J. Effect of curcumin on STAT5 signaling molecule in K562 cells. Zhonghua Xue Ye Xue Za Zhi 2004, 25, 151–153. [Google Scholar]
  101. Jaiswal, A.S.; Marlow, B.P.; Gupta, N.; Narayan, S. Beta-Catenin-Mediated Transactivation and Cell-Cell Adhesion Pathways Are Important in Curcumin (Diferuylmethane)-Induced Growth Arrest and Apoptosis in Colon Cancer Cells. Oncogene 2002, 21, 8414–8427. [Google Scholar] [CrossRef] [Green Version]
  102. Park, C.H.; Hahm, E.R.; Park, S.; Kim, H.-K.; Yang, C.H. The inhibitory mechanism of curcumin and its derivative against β-catenin/Tcf signaling. FEBS Lett. 2005, 579, 2965–2971. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Pendurthi, U.R.; Rao, L.M. Suppression of Transcription Factor Egr-1 by Curcumin. Thromb. Res. 2000, 97, 179–189. [Google Scholar] [CrossRef]
  104. Chen, A.; Xu, J.; Johnson, A.C. Curcumin inhibits human colon cancer cell growth by suppressing gene expression of epidermal growth factor receptor through reducing the activity of the transcription factor Egr-1. Oncogene 2006, 25, 278–287. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Bae, M.-K.; Kim, S.-H.; Jeong, J.-W.; Lee, Y.M.; Kim, H.-S.; Kim, S.-R.; Yun, I.; Bae, S.-K.; Kim, K.-W. Curcumin inhibits hypoxia-induced angiogenesis via down-regulation of HIF-1. Oncol. Rep. 2006, 15, 1557–1562. [Google Scholar] [CrossRef]
  106. Wang, Z.; Zhang, Y.; Banerjee, S.; Li, Y.; Sarkar, F.H. Retracted: Notch-1 down-regulation by curcumin is associated with the inhibition of cell growth and the induction of apoptosis in pancreatic cancer cells. Cancer 2006, 106, 2503–2513. [Google Scholar] [CrossRef]
  107. Zhou, Y.; Zheng, S.; Lin, J.; Zhang, Q.-J.; Chen, A. The interruption of the PDGF and EGF signaling pathways by curcumin stimulates gene expression of PPARγ in rat activated hepatic stellate cell in vitro. Lab. Investig. 2007, 87, 488–498. [Google Scholar] [CrossRef]
  108. Mohan, R.; Sivak, J.; Ashton, P.; Russo, L.A.; Pham, B.Q.; Kasahara, N.; Raizman, M.B.; Fini, M.E. Curcuminoids Inhibit the Angiogenic Response Stimulated by Fibroblast Growth Factor-2, Including Expression of Matrix Metalloproteinase Gelatinase B. J. Biol. Chem. 2000, 275, 10405–10412. [Google Scholar] [CrossRef] [Green Version]
  109. Yang, X.; Thomas, D.P.; Zhang, X.; Culver, B.W.; Alexander, B.M.; Murdoch, W.J.; Rao, M.N.; Tulis, D.A.; Ren, J.; Sreejayan, N. Curcumin Inhibits Platelet-Derived Growth Factor–Stimulated Vascular Smooth Muscle Cell Function and Injury-Induced Neointima Formation. Arter. Thromb. Vasc. Biol. 2006, 26, 85–90. [Google Scholar] [CrossRef] [Green Version]
  110. Santibanez, J.F.; Quintanilla, M.; Martínez, J. Genistein and Curcumin Block TGF-β1-Induced u-PA Expression and Migratory and Invasive Phenotype in Mouse Epidermal Keratinocytes. Nutr. Cancer 2000, 37, 49–54. [Google Scholar] [CrossRef]
  111. Gaedeke, J.; Noble, N.A.; Border, W.A. Curcumin blocks multiple sites of the TGF-β signaling cascade in renal cells. Kidney Int. 2004, 66, 112–120. [Google Scholar] [CrossRef] [Green Version]
  112. Hu, Y.; Liang, H.; Du, Y.; Zhu, Y.; Wang, X. Curcumin Inhibits Transforming Growth Factor-β Activity via Inhibition of Smad Signaling in HK-2 Cells. Am. J. Nephrol. 2010, 31, 332–341. [Google Scholar] [CrossRef] [PubMed]
  113. Song, K.; Peng, S.; Sun, Z.; Li, H.; Yang, R. Curcumin suppresses TGF-β signaling by inhibition of TGIF degradation in scleroderma fibroblasts. Biochem. Biophys. Res. Commun. 2011, 411, 821–825. [Google Scholar] [CrossRef] [PubMed]
  114. Zhang, L.; Cheng, X.; Gao, Y.; Zhang, C.; Bao, J.; Guan, H.; Yu, H.; Lu, R.; Xu, Q.; Sun, Y. Curcumin Inhibits Metastasis in Human Papillary Thyroid Carcinoma BCPAP Cells via down-Regulation of the TGF-β/Smad2/3 Signaling Pathway. Exp. Cell Res. 2016, 341, 157–165. [Google Scholar] [CrossRef] [PubMed]
  115. Chakraborty, G.; Jain, S.; Kale, S.; Raja, R.; Kumar, S.; Mishra, R.; Kundu, G.C. Curcumin suppresses breast tumor angiogenesis by abrogating osteopontin-induced VEGF expression. Mol. Med. Rep. 2008, 1, 641–646. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Yoysungnoen, P.; Wirachwong, P.; Changtam, C.; Suksamrarn, A.; Patumraj, S. Anti-cancer and anti-angiogenic effects of curcumin and tetrahydrocurcumin on implanted hepatocellular carcinoma in nude mice. World J. Gastroenterol. 2008, 14, 2003–2009. [Google Scholar] [CrossRef]
  117. Ferreira, L.C.; Arbab, A.S.; Jardim-Perassi, B.V.; Borin, T.F.; Gonçalves, N.N.; Nadimpalli, R.S.V.; Zuccari, D.A.P.D.C. Abstract A02: Effect of curcumin on the tumor growth and angiogenesis of breast cancer. Tumor-Assoc. Blood Vessels Lymph. 2015, 75, A02. [Google Scholar] [CrossRef]
  118. Kunnumakkara, A.B.; Anand, P.; Aggarwal, B.B. Curcumin inhibits proliferation, invasion, angiogenesis and metastasis of different cancers through interaction with multiple cell signaling proteins. Cancer Lett. 2008, 269, 199–225. [Google Scholar] [CrossRef]
  119. Aggarwal, B.B.; Bhatt, I.D.; Ichikawa, H. Curcumin-Biological and Medicinal Properties. In Tumeric: The Genus Curcuma; CRC Press: Boca Raton, FL, USA, 2006; pp. 297–368. [Google Scholar]
  120. Aggarwal, B.B. Activation of Transcription Factor NF-kappaB Is Suppressed by Curcumin (Diferuloylmethane). J. Biol. Chem. 1995, 270, 24995–25000. [Google Scholar]
  121. Cho, J.-W.; Lee, K.-S.; Kim, C.-W. Curcumin Attenuates the Expression of IL-1beta, IL-6, and TNF-Alpha as Well as Cyclin E in TNF-Alpha-Treated HaCaT Cells; NF-kappaB and MAPKs as Potential Upstream Targets. Int. J. Mol. Med. 2007, 19, 469–474. [Google Scholar]
  122. Ranjan, D.; Chen, C.; Johnston, T.D.; Jeon, H.; Nagabhushan, M. Curcumin inhibits mitogen stimulated lymphocyte proliferation, NFκB activation, and IL-2 signaling. J. Surg. Res. 2004, 121, 171–177. [Google Scholar] [CrossRef]
  123. Kobayashi, T.; Hashimoto, S.; Horie, T. Curcumin inhibition of Dermatophagoides farinea-induced interleukin-5 (IL-5) and granulocyte macrophage-colony stimulating factor (GM-CSF) production by lymphocytes from bronchial asthmatics. Biochem. Pharmacol. 1997, 54, 819–824. [Google Scholar] [CrossRef]
  124. Fahey, A.J.; Robins, R.A.; Constantinescu, C.S. Curcumin modulation of IFN-β and IL-12 signalling and cytokine induction in human T cells. J. Cell. Mol. Med. 2007, 11, 1129–1137. [Google Scholar] [CrossRef] [PubMed]
  125. Grandjean-Laquerriere, A.; Antonicelli, F.; Gangloff, S.C.; Guenounou, M.; Le Naour, R. UVB-Induced IL-18 Production in Human Keratinocyte Cell Line NCTC 2544 through NF-κB Activation. Cytokine 2007, 37, 76–83. [Google Scholar] [CrossRef] [PubMed]
  126. Chun, K.-S.; Keum, Y.-S.; Han, S.S.; Song, Y.-S.; Kim, S.-H.; Surh, Y.-J. Curcumin inhibits phorbol ester-induced expression of cyclooxygenase-2 in mouse skin through suppression of extracellular signal-regulated kinase activity and NF- B activation. Carcinogenesis 2003, 24, 1515–1524. [Google Scholar] [CrossRef] [Green Version]
  127. Camacho-Barquero, L.; Villegas, I.; Sánchez-Calvo, J.M.; Talero, E.; Sánchez-Fidalgo, S.; Motilva, V.; De La Lastra, C.A. Curcumin, a Curcuma longa constituent, acts on MAPK p38 pathway modulating COX-2 and iNOS expression in chronic experimental colitis. Int. Immunopharmacol. 2007, 7, 333–342. [Google Scholar] [CrossRef]
  128. Huang, M.T.; Lysz, T.; Ferraro, T.; Abidi, T.F.; Laskin, J.D.; Conney, A.H. Inhibitory effects of curcumin on in vitro lipoxygenase and cyclooxygenase activities in mouse epidermis. Cancer Res. 1991, 51, 813–819. [Google Scholar]
  129. Kunnumakkara, A.B.; Guha, S.; Krishnan, S.; Diagaradjane, P.; Gelovani, J.; Aggarwal, B.B. Curcumin Potentiates Antitumor Activity of Gemcitabine in an Orthotopic Model of Pancreatic Cancer through Suppression of Proliferation, Angiogenesis, and Inhibition of Nuclear Factor-κB–Regulated Gene Products. Cancer Res. 2007, 67, 3853–3861. [Google Scholar] [CrossRef] [Green Version]
  130. Lin, Y.G.; Kunnumakkara, A.B.; Nair, A.S.; Merritt, W.M.; Han, L.Y.; Armaiz-Pena, G.N.; Kamat, A.A.; Spannuth, W.A.; Gershenson, D.M.; Lutgendorf, S.K.; et al. Curcumin Inhibits Tumor Growth and Angiogenesis in Ovarian Carcinoma by Targeting the Nuclear Factor- B Pathway. Clin. Cancer Res. 2007, 13, 3423–3430. [Google Scholar] [CrossRef] [Green Version]
  131. Woo, M.-S.; Jung, S.-H.; Kim, S.-Y.; Hyun, J.-W.; Ko, K.-H.; Kim, W.-K.; Kim, H.-S. Curcumin suppresses phorbol ester-induced matrix metalloproteinase-9 expression by inhibiting the PKC to MAPK signaling pathways in human astroglioma cells. Biochem. Biophys. Res. Commun. 2005, 335, 1017–1025. [Google Scholar] [CrossRef]
  132. Chen, Y.-R.; Tan, T.-H. Inhibition of the c-Jun N-terminal kinase (JNK) signaling pathway by curcumin. Oncogene 1998, 17, 173–178. [Google Scholar] [CrossRef] [Green Version]
  133. Salh, B.; Assi, K.; Templeman, V.; Parhar, K.K.S.; Owen, D.; Gomez-Muñoz, A.; Jacobson, K. Curcumin attenuates DNB-induced murine colitis. Am. J. Physiol. Liver Physiol. 2003, 285, G235–G243. [Google Scholar] [CrossRef] [PubMed]
  134. Hussain, A.R.; Al-Rasheed, M.; Manogaran, P.S.; Al-Hussein, K.A.; Platanias, L.C.; Al Kuraya, K.; Uddin, S. Curcumin induces apoptosis via inhibition of PI3′-kinase/AKT pathway in Acute T cell Leukemias. Apoptosis 2006, 11, 245–254. [Google Scholar] [CrossRef] [PubMed]
  135. Surh, Y.-J.; Lee, S.-Y.; Huang, Z.; Lim, D.Y.; Chen, H.; Jung, S.K.; Bode, A.M.; Lee, K.W.; Dong, Z. Curcumin Suppresses Proliferation of Colon Cancer Cells by Targeting CDK2. Cancer Prev. Res. 2014, 7, 466–474. [Google Scholar] [CrossRef] [Green Version]
  136. Bush, J.A.; Cheung, K.-J.J.; Li, G. Curcumin Induces Apoptosis in Human Melanoma Cells through a Fas Receptor/Caspase-8 Pathway Independent of p53. Exp. Cell Res. 2001, 271, 305–314. [Google Scholar] [CrossRef] [PubMed]
  137. Anto, R.J.; Mukhopadhyay, A.; Denning, K.; Aggarwal, B.B. Curcumin (diferuloylmethane) induces apoptosis through activation of caspase-8, BID cleavage and cytochrome c release: Its suppression by ectopic expression of Bcl-2 and Bcl-xl. Carcinogenesis 2002, 23, 143–150. [Google Scholar] [CrossRef]
  138. 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. 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]
  139. Jiang, A.-J.; Jiang, G.; Li, L.-T.; Zheng, J. Curcumin induces apoptosis through mitochondrial pathway and caspases activation in human melanoma cells. Mol. Biol. Rep. 2014, 42, 267–275. [Google Scholar] [CrossRef]
  140. Li, F.; Chen, X.; Xu, B.; Zhou, H. Curcumin induces p53-independent necrosis in H1299 cells via a mitochondria-associated pathway. Mol. Med. Rep. 2015, 12, 7806–7814. [Google Scholar] [CrossRef] [Green Version]
  141. Shishodia, S.; Amin, H.M.; Lai, R.; Aggarwal, B.B. Curcumin (Diferuloylmethane) Inhibits Constitutive NF-κB Activation, Induces G1/S Arrest, Suppresses Proliferation, and Induces Apoptosis in Mantle Cell Lymphoma. Biochem. Pharmacol. 2005, 70, 700–713. [Google Scholar] [CrossRef]
  142. Choudhuri, T.; Pal, S.; Das, T.; Sa, G. Curcumin Selectively Induces Apoptosis in Deregulated Cyclin D1-expressed Cells at G2Phase of Cell Cycle in a p53-dependent Manner. J. Biol. Chem. 2005, 280, 20059–20068. [Google Scholar] [CrossRef] [Green Version]
  143. Srivastava, R.K.; Chen, Q.; Siddiqui, I.; Sarva, K.; Shankar, S. Linkage of Curcumin-Induced Cell Cycle Arrest and Apoptosis by Cyclin-Dependent Kinase Inhibitor p21/WAF1/CIP1. Cell Cycle 2007, 6, 2953–2961. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Liu, E.; Wu, J.; Cao, W.; Zhang, J.; Liu, W.; Jiang, X.; Zhang, X. Curcumin induces G2/M cell cycle arrest in a p53-dependent manner and upregulates ING4 expression in human glioma. J. Neuro-Oncology 2007, 85, 263–270. [Google Scholar] [CrossRef] [PubMed]
  145. Shankar, S.; Srivastava, R.K. Involvement of Bcl-2 family members, phosphatidylinositol 3’-kinase/AKT and mitochondrial p53 in curcumin (diferulolylmethane)-induced apoptosis in prostate cancer. Int. J. Oncol. 2007, 30, 905–918. [Google Scholar] [CrossRef] [PubMed]
  146. Dinarello, C.A. The paradox of pro-inflammatory cytokines in cancer. Cancer Metastasis Rev. 2006, 25, 307–313. [Google Scholar] [CrossRef]
  147. Williams, C.S.; Mann, M.; Dubois, R.N. The role of cyclooxygenases in inflammation, cancer, and development. Oncogene 1999, 18, 7908–7916. [Google Scholar] [CrossRef] [Green Version]
  148. Kumar, A.; Dhawan, S.; Mukhopadhyay, A.; Aggarwal, B.B. Human Immunodeficiency Virus-1-Tat Induces Matrix Metalloproteinase-9 in Monocytes through Protein Tyrosine Phosphatase-Mediated Activation of Nuclear Transcription Factor NF-kappaB. FEBS Lett. 1999, 462, 140–144. [Google Scholar] [CrossRef] [Green Version]
  149. John, P.C.L.; Mews, M.; Moore, R. Cyclin/cdk complexes: Their involvement in cell cycle progression and mitotic division. Protoplasma 2001, 216, 119–142. [Google Scholar] [CrossRef]
  150. Sharma, R.A.; McLelland, H.R.; Hill, K.A.; Ireson, C.R.; Euden, S.A.; Manson, M.M.; Pirmohamed, M.; Marnett, L.J.; Gescher, A.J.; Steward, W.P. Pharmacodynamic and pharmacokinetic study of oral Curcuma extract in patients with colorectal cancer. Clin. Cancer Res. 2001, 7, 1894–1900. [Google Scholar]
  151. Sasaki, H.; Sunagawa, Y.; Takahashi, K.; Imaizumi, A.; Fukuda, H.; Hashimoto, T.; Wada, H.; Katanasaka, Y.; Kakeya, H.; Fujita, M.; et al. Innovative Preparation of Curcumin for Improved Oral Bioavailability. Biol. Pharm. Bull. 2011, 34, 660–665. [Google Scholar] [CrossRef] [Green Version]
  152. Sharma, R.A.; Euden, S.A.; Platton, S.L.; Cooke, D.N.; Shafayat, A.; Hewitt, H.R.; Marczylo, T.H.; Morgan, B.; Hemingway, D.; Plummer, S.M.; et al. Phase I Clinical Trial of Oral Curcumin: Biomarkers of Systemic Activity and Compliance. Clin. Cancer Res. 2004, 10, 6847–6854. [Google Scholar] [CrossRef] [Green Version]
  153. Cheng, A.L.; Hsu, C.-H.; Lin, J.K.; Hsu, M.-M.; Ho, Y.F.; Shen, T.S.; Ko, J.Y.; Lin, J.T.; Lin, B.R.; Ming-Shiang, W.; et al. Phase I clinical trial of curcumin, a chemopreventive agent, in patients with high-risk or pre-malignant lesions. Anticancer. Res. 2001, 21, 2895–2900. [Google Scholar] [PubMed]
  154. Lao, C.D.; Ruffin, M.T.; Normolle, D.P.; Heath, D.D.; I Murray, S.; Bailey, J.M.; Boggs, M.E.; Crowell, J.; Rock, C.L.; Brenner, D.E. Dose escalation of a curcuminoid formulation. BMC Complement. Altern. Med. 2006, 6, 10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Anand, P.; Kunnumakkara, A.B.; Newman, R.A.; Aggarwal, B.B. Bioavailability of Curcumin: Problems and Promises. Mol. Pharm. 2007, 4, 807–818. [Google Scholar] [CrossRef] [PubMed]
  156. Sharma, R.A.; Steward, W.P.; Gescher, A.J. Pharmacokinetics and pharmacodynamics of curcumin. Adv. Exp. Med. Biol. 2007, 595, 453–470. [Google Scholar] [PubMed] [Green Version]
  157. Shoba, G.; Joy, D.; Joseph, T.; Majeed, M.; Rajendran, R.; Srinivas, P.S.S.R. Influence of Piperine on the Pharmacokinetics of Curcumin in Animals and Human Volunteers. Planta Medica 1998, 64, 353–356. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Asai, A.; Miyazawa, T. Occurrence of orally administered curcuminoid as glucuronide and glucuronide/sulfate conjugates in rat plasma. Life Sci. 2000, 67, 2785–2793. [Google Scholar] [CrossRef]
  159. Pan, M.H.; Lin-Shiau, S.Y.; Lin, J.K. Comparative studies on the suppression of nitric oxide synthase by curcumin and its hydrogenated metabolites through down-regulation of IkappaB kinase and NFkappaB activation in macrophages. Biochem. Pharmacol. 2000, 60, 1665–1676. [Google Scholar] [CrossRef]
  160. Ireson, C.; Orr, S.; Jones, D.J.; Verschoyle, R.; Lim, C.K.; Luo, J.L.; Howells, L.; Plummer, S.; Jukes, R.; Williams, M.; et al. Characterization of metabolites of the chemopreventive agent curcumin in human and rat hepatocytes and in the rat in vivo, and evaluation of their ability to inhibit phorbol ester-induced prostaglandin E2 production. Cancer Res. 2001, 61, 1058–1064. [Google Scholar]
  161. Oppenheimer, A. Turmeric (Currumin) in Billiary Diseases. Lancet 1937, 229, 619–621. [Google Scholar] [CrossRef]
  162. Yang, K.-Y.; Lin, L.-C.; Tseng, T.-Y.; Wang, S.-C.; Tsai, T.-H. Oral bioavailability of curcumin in rat and the herbal analysis from Curcuma longa by LC–MS/MS. J. Chromatogr. B 2007, 853, 183–189. [Google Scholar] [CrossRef]
  163. Pan, M.H.; Huang, T.M.; Lin, J.K. Biotransformation of curcumin through reduction and glucuronidation in mice. Drug Metab. Dispos. 1999, 27, 486–494. [Google Scholar] [PubMed]
  164. Antony, B. A Composition to Enhance the Bioavailability of Curcumin. World Patent 2006129323 A1, 7 December 2006. [Google Scholar]
  165. Giori, A.; Franceschi, F. Phospholipid Complexes of Curcumin Having Improved Bioavailability. World Patent 2007101551, 25 October 2007. [Google Scholar]
  166. Badmaev, V.; Majeed, M.; Rajendran, R. Bioprotectant Composition, Method of Use and Extraction Process of Curcuminoids. World Patent 1997003674 A1, 6 February 1997. [Google Scholar]
  167. Madhavi, D.; Kagan, D. Bioavailability of a Sustained Release Formulation of Curcumin. Integr. Med. (Encinitas, Calif.) 2014, 13, 24–30. [Google Scholar]
  168. Siddiqui, R.A.; Hassan, S.; Harvey, K.A.; Rasool, T.; Das, T.; Mukerji, P.; DeMichele, S. Attenuation of proteolysis and muscle wasting by curcumin c3 complex in MAC16 colon tumour-bearing mice. Br. J. Nutr. 2009, 102, 967–975. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  169. Anthony, B. Composition to Enhance the Bioavailability of Curcumin. U.S. Patent 7879373 B2, 30 May 2005. [Google Scholar]
  170. Bombardelli, E. Phytosome: New cosmetic delivery system. Boll. Chim. Farm. 1991, 130, 431–438. [Google Scholar]
  171. Phytosome Technical Paper. Available online: https://www.indena.com/indena_files/2020/01/wp_phytosome_int.pdf (accessed on 17 November 2020).
  172. Marczylo, T.; Verschoyle, R.D.; Cooke, D.N.; Morazzoni, P.; Steward, W.P.; Gescher, A.J. Comparison of systemic availability of curcumin with that of curcumin formulated with phosphatidylcholine. Cancer Chemother. Pharmacol. 2007, 60, 171–177. [Google Scholar] [CrossRef] [Green Version]
  173. Belcaro, G.; Cesarone, M.R.; Dugall, M.; Pellegrini, L.; Ledda, A.; Grossi, M.G.; Togni, S.; Appendino, G. Product-evaluation registry of Meriva®, a curcumin-phosphatidylcholine complex, for the complementary management of osteoarthritis. Panminerva Med. 2010, 52, 55–62. [Google Scholar]
  174. 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. A J. Clin. Ther. 2010, 15, 337–344. [Google Scholar]
  175. Liposomal Curcumin—16 Fluid Ounces. Available online: http://www.healthyitems.com/curcumin-p/607.htm (accessed on 17 November 2020).
  176. Amy Myers MD—Liposomal Curcumin. Available online: https://store.amymyersmd.com/products/liposomal-curcumin (accessed on 17 November 2020).
  177. Actinovo Liposomal Curcumin (Turmeric, Curcuminoids). Available online: https://www.actinovo.com/en/liposomal-curcumin-turmeric (accessed on 17 November 2020).
  178. Valimenta Liposomal Curcumin. Available online: https://www.valimenta.com/product/liposomal-curcumin/ (accessed on 17 November 2020).
  179. Cavacurmin®: Highly Bioavailable Curcumin. Available online: https://www.wacker.com/cms/en/industries/food/curcumin.jsp (accessed on 5 December 2020).
  180. Atal, C.K.; Dubey, R.K.; Singh, J. Biochemical basis of enhanced drug bioavailability by piperine: Evidence that piperine is a potent inhibitor of drug metabolism. J. Pharmacol. Exp. Ther. 1985, 232, 258–262. [Google Scholar] [PubMed]
  181. Han, H.-K. The effects of black pepper on the intestinal absorption and hepatic metabolism of drugs. Expert Opin. Drug Metab. Toxicol. 2011, 7, 721–729. [Google Scholar] [CrossRef] [PubMed]
  182. Zhai, G.; Zhang, L. Self-Micro-Emulsion Colon Site-Specific Delivery Capsule Comprises Curcumin, Oil Phase, Surfactant, Cosurfactant, and Folic Acid Grease Material. Chinese Patent CN102266287, 7 December 2011. [Google Scholar]
  183. Catalan-Latorre, A.; Ravaghi, M.; Manca, M.L.; Caddeo, C.; Marongiu, F.; Ennas, G.; Escribano-Ferrer, E.; Peris, J.E.; Diez-Sales, O.; Fadda, A.M.; et al. Freeze-dried eudragit-hyaluronan multicompartment liposomes to improve the intestinal bioavailability of curcumin. Eur. J. Pharm. Biopharm. 2016, 107, 49–55. [Google Scholar] [CrossRef] [PubMed]
  184. Mathur, A.B.; Rios, C.N.; Gupta, V.; Aseh, A. Preparation and Methodology of Silk Fibroin Nanoparticles. World Patent WO2010059963 A2, 27 May 2010. [Google Scholar]
  185. Kim, S.G.; Veena, M.S.; Basak, S.K.; Han, E.; Tajima, T.; Gjertson, D.W.; Starr, J.; Eidelman, O.; Pollard, H.B.; Srivastava, M.; et al. Curcumin Treatment Suppresses IKK Kinase Activity of Salivary Cells of Patients with Head and Neck Cancer: A Pilot Study. Clin. Cancer Res. 2011, 17, 5953–5961. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Kurzrock, R.; Li, L.; Mehta, K.; Aggarawal, B.B. Liposomal Curcumin for the Treatment of Cancer. U.S. Patent 20060067998 A1, 30 March 2006. [Google Scholar]
  187. Kurzrock, R.; Li, L. Liposome-encapsulated curcumin: In vitro and in vivo effects on proliferation, apoptosis, signaling, and angiogenesis. J. Clin. Oncol. 2005, 23. [Google Scholar] [CrossRef]
  188. Clinicaltrials.Gov: A Phase IB Dose Escalation Study of Lipocurc in Patients With Cancer. Available online: https://clinicaltrials.gov/ct2/show/NCT02138955 (accessed on 13 November 2019).
  189. Helson, L.; Chiu, S. Intravenous Curcumin and Derivatives for Treatment of Neurodegenerative and Stress Disorders. U.S. Patent 20110229555 A1, 22 September 2011. [Google Scholar]
  190. Bisht, S.; Feldmann, G.; Soni, S.; Ravi, R.; Karikar, C.; Maitra, A.; Maitra, A. Polymeric nanoparticle-encapsulated curcumin (“nanocurcumin”): A novel strategy for human cancer therapy. J. Nanobiotechnol. 2007, 5. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  191. Bisht, S.; Mizuma, M.; Feldmann, G.; Ottenhof, N.A.; Hong, S.-M.; Pramanik, D.; Chenna, V.; Karikari, C.; Sharma, R.; Goggins, M.G.; et al. Systemic Administration of Polymeric Nanoparticle-Encapsulated Curcumin (NanoCurc) Blocks Tumor Growth and Metastases in Preclinical Models of Pancreatic Cancer. Mol. Cancer Ther. 2010, 9, 2255–2264. [Google Scholar] [CrossRef] [Green Version]
  192. Chiu, S.S.; Lui, E.; Majeed, M.; Vishwanatha, J.K.; Ranjan, A.P.; Maitra, A.; Pramanik, D.; Smith, J.A.; Helson, L. Differential distribution of intravenous curcumin formulations in the rat brain. Anticancer. Res. 2011, 31, 907–911. [Google Scholar]
  193. Gou, M.; Men, K.; Shi, H.; Xiang, M.; Zhang, J.; Song, J.; Long, J.; Wan, Y.; Luo, F.; Zhao, X.; et al. Curcumin-loaded biodegradable polymeric micelles for colon cancer therapy in vitro and in vivo. Nanoscale 2011, 3, 1558–1567. [Google Scholar] [CrossRef]
  194. Yang, X.; Li, Z.; Wang, N.; Li, L.; Song, L.; He, T.; Sun, L.; Wang, Z.; Wu, Q.; Luo, N.; et al. Curcumin-Encapsulated Polymeric Micelles Suppress the Development of Colon Cancer In Vitro and In Vivo. Sci. Rep. 2015, 5. [Google Scholar] [CrossRef]
  195. Shahani, K.; Swaminathan, S.K.; Freeman, D.; Blum, A.; Ma, L.; Panyam, J. Injectable Sustained Release Microparticles of Curcumin: A New Concept for Cancer Chemoprevention. Cancer Res. 2010, 70, 4443–4452. [Google Scholar] [CrossRef] [Green Version]
  196. Ranjan, A.P.; Mukerjee, A.; Vishwanatha, J.K. Solid in Oil/Water Emulsion-Diffusion Evaporation Formulation for Preparing Curcumin-Loaded PLGA Nanoparticles. U.S. Patent 20100290982 A1, 18 November 2010. [Google Scholar]
  197. Keck, C.M.; Santini, A. Drug nanocrystals of poorly soluble drugs produced by high pressure homogenisation. Eur. J. Pharm. Biopharm. 2006, 62, 3–16. [Google Scholar] [CrossRef]
  198. Gao, Y.; Li, Z.; Sun, M.; Guo, C.; Yu, A.; Xi, Y.; Cui, J.; Lou, H.; Zhai, G. Preparation and characterization of intravenously injectable curcumin nanosuspension. Drug Deliv. 2010, 18, 131–142. [Google Scholar] [CrossRef]
  199. Kurniawansyah, F.; Duong, H.T.T.; Luu, T.D.; Mammucari, R.; Vittorio, O.; Boyer, C.; Foster, N.R. Inhalable curcumin formulations: Micronization and bioassay. Chem. Eng. J. 2015, 279, 799–808. [Google Scholar] [CrossRef]
  200. McClure, R.; Yanagisawa, D.; Stec, D.; Abdollahian, D.; Koktysh, D.; Xhillari, D.; Jaeger, R.; Stanwood, G.; Chekmenev, E.; Tooyama, I.; et al. Inhalable Curcumin: Offering the Potential for Translation to Imaging and Treatment of Alzheimer’s Disease. J. Alzheimer’s Dis. 2015, 44, 283–295. [Google Scholar] [CrossRef] [Green Version]
  201. Saraf, G.J.S. Topical Delivery of Curcuma longa Extract Loaded Nanosized Ethosomes to Combat Facial Wrinkles Research Article. J. Pharm. Drug Deliv. Res. 2014, 3, 1. [Google Scholar] [CrossRef]
  202. Azuine, M.A.; Bhide, S.V. Chemopreventive effect of turmeric against stomach and skin tumors induced by chemical carcinogens in Swiss mice. Nutr. Cancer 1992, 17, 77–83. [Google Scholar] [CrossRef] [PubMed]
  203. Lü, Y.; Chang, R.L.; Lou, Y.-R.; Huang, M.-T.; Newmark, H.L.; Reuhl, K.R.; Conney, A.H. Effect of curcumin on 12-O-tetradecanoylphorbol-13-acetate- and ultraviolet B light-induced expression of c-Jun and c-Fos in JB6 cells and in mouse epidermis. Carcinogenesis 1994, 15, 2363–2370. [Google Scholar] [CrossRef] [PubMed]
  204. Ishizaki, C.; Oguro, T.; Yoshida, T.; Wen, C.; Sueki, H.; Iijima, M. Enhancing Effect of Ultraviolet A on Ornithine Decarboxylase Induction and Dermatitis Evoked by 12-o-Tetradecanoylphorbol-13-Acetate and Its Inhibition by Curcumin in Mouse Skin. Dermatology 1996, 193, 311–317. [Google Scholar] [CrossRef] [PubMed]
  205. Villase~Or, I.M.; Simon, M.K.B.; Villanueva, A.M.A. Comparative Potencies of Nutraceuticals in Chemically Induced Skin Tumor Prevention. Nutr. Cancer 2002, 44, 66–70. [Google Scholar] [CrossRef]
  206. Phillips, J.M.; Clark, C.; Herman-Ferdinandez, L.; Moore-Medlin, T.; Rong, X.; Gill, J.R.; Clifford, J.L.; Abreo, F.; Nathan, C.-A.O. Curcumin Inhibits Skin Squamous Cell Carcinoma Tumor Growth In Vivo. Otolaryngol. Neck Surg. 2011, 145, 58–63. [Google Scholar] [CrossRef]
  207. Touitou, E.; Dayan, N.; Bergelson, L.; Godin, B.; Eliaz, M. Ethosomes—novel vesicular carriers for enhanced delivery: Characterization and skin penetration properties. J. Control. Release 2000, 65, 403–418. [Google Scholar] [CrossRef]
  208. Crowther, O.; Chua, D.; Eppley, W.; Meyer, B.; Salomon, M.; Driedger, A.; Morgan, M. Lithium-Air Cell Protective Membranes Comprising Polytetrafluroethylene Coated Fiberglass Cloth. U.S. Patent 2011/0177401 A1, 21 July 2011. [Google Scholar]
  209. Zhao, Y.-Z.; Lu, C.-T.; Zhang, Y.; Xiao, J.; Zhao, Y.-P.; Tian, J.-L.; Xu, Y.-Y.; Feng, Z.-G.; Xu, C.-Y. Selection of high efficient transdermal lipid vesicle for curcumin skin delivery. Int. J. Pharm. 2013, 454, 302–309. [Google Scholar] [CrossRef]
  210. Kogan, A.; Garti, N. Microemulsions as transdermal drug delivery vehicles. Adv. Colloid Interface Sci. 2006, 123‒126, 369–385. [Google Scholar] [CrossRef]
  211. Lakshmi, P.K.; Mounica, V.; Manoj Kumar, Y.; Prasanthi, D. Preparation and Evaluation of Curcumin Invasomes. Int. J. Drug Deliv. 2014, 6, 113–120. [Google Scholar]
  212. Liu, C.-H.; Chang, F.-Y.; Hung, D.-K. Terpene microemulsions for transdermal curcumin delivery: Effects of terpenes and cosurfactants. Colloids Surfaces B Biointerfaces 2011, 82, 63–70. [Google Scholar] [CrossRef]
  213. Patra, S.; Roy, E.; Madhuri, R.; Sharma, P.K. Retracted Article: The next generation cell-penetrating peptide and carbon dot conjugated nano-liposome for transdermal delivery of curcumin. Biomater. Sci. 2016, 4, 418–429. [Google Scholar] [CrossRef]
  214. Mangalathillam, S.; Rejinold, N.S.; Nair, A.; Lakshmanan, V.-K.; Nair, S.V.; Jayakumar, R. Curcumin loaded chitin nanogels for skin cancer treatment via the transdermal route. Nanoscale 2012, 4, 239–250. [Google Scholar] [CrossRef]
  215. Nedelcu, I.-A.; Ficai, A.; Sonmez, M.; Ficai, D.; Oprea, O.; Andronescu, E. Silver Based Materials for Biomedical Applications. Curr. Org. Chem. 2014, 18, 173–184. [Google Scholar] [CrossRef]
  216. Franco-Molina, M.A.; Mendoza-Gamboa, E.; Sierra-Rivera, C.A.; Gomez-Flores, R.; Zapata-Benavides, P.; Castillo-Tello, P.; Alcocer-González, J.M.; Miranda-Hernández, D.F.; Tamez-Guerra, R.S.; Rodríguez-Padilla, C. Antitumor activity of colloidal silver on MCF-7 human breast cancer cells. J. Exp. Clin. Cancer Res. 2010, 29. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Sangiliyandi, G.; Sriram, M.I.; Kanth, S.B.M.; Kalishwaralal, K.; Gurunathan, S. Antitumor activity of silver nanoparticles in Dalton’s lymphoma ascites tumor model. Int. J. Nanomed. 2010, 5, 753–762. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  218. Rutberg, F.G.; Dubina, M.V.; Kolikov, V.A.; Moiseenko, F.V.; Ignat’Eva, E.V.; Volkov, N.M.; Snetov, V.N.; Stogov, A.Y. Effect of silver oxide nanoparticles on tumor growth in vivo. Dokl. Biochem. Biophys. 2008, 421, 191–193. [Google Scholar] [CrossRef] [PubMed]
  219. Lilienfeld, S.; Dimauro, T.M. Iontophoretic Delivery of Curcumin and Curcumin Analogs for the Treatment of Alzheimer’s Disease. World Patent 2009158407 A1, 30 December 2009. [Google Scholar]
  220. Chang, C.-C.; Yang, W.-T.; Ko, S.-Y.; Hsu, Y.-C. Liposomal Curcuminoids for Transdermal Delivery: Iontophoresis Potential for Breast Cancer Chemotherapeutics. Dig. J. Nanomater. Biostruct. 2012, 7, 59–71. [Google Scholar]
  221. Pröhl, M.; Schubert, U.S.; Weigand, W.; Gottschaldt, M. Metal complexes of curcumin and curcumin derivatives for molecular imaging and anticancer therapy. Co-Ord. Chem. Rev. 2016, 307, 32–41. [Google Scholar] [CrossRef]
  222. Dahl, T.A.; Bilski, P.; Reszka, K.J.; Chignell, C.F.; Dahll, T.A. photocytotoxicity of curcumin. Photochem. Photobiol. 1994, 59, 290–294. [Google Scholar] [CrossRef] [PubMed]
  223. Dahl, T.A.; McGowan, W.M.; Shand, M.A.; Srinivasan, V.S. Photokilling of bacteria by the natural dye curcumin. Arch. Microbiol. 1989, 151, 183–185. [Google Scholar] [CrossRef] [PubMed]
  224. Ellerkamp, V.; Bortel, N.; Schmid, E.; Kirchner, B.; Armeanu-Ebinger, S.; Fuchs, J. Photodynamic Therapy Potentiates the Effects of Curcumin on Pediatric Epithelial Liver Tumor Cells. Anticancer. Res. 2016, 36, 3363–3372. [Google Scholar] [PubMed]
  225. Banerjee, S.; Prasad, P.; Hussain, A.; Khan, I.; Kondaiah, P.; Chakravarty, A.R. Remarkable photocytotoxicity of curcumin in HeLa cells in visible light and arresting its degradation on oxovanadium(iv) complex formation. Chem. Commun. 2012, 48, 7702. [Google Scholar] [CrossRef]
  226. Lin, H.-Y.; Lin, J.-N.; Ma, J.-W.; Yang, N.-S.; Ho, C.-T.; Kuo, S.-C.; Way, T.-D. Demethoxycurcumin induces autophagic and apoptotic responses on breast cancer cells in photodynamic therapy. J. Funct. Foods 2015, 12, 439–449. [Google Scholar] [CrossRef]
  227. Ahn, J.-C.; Kang, J.-W.; Shin, J.-I.; Chung, P. Combination treatment with photodynamic therapy and curcumin induces mitochondria-dependent apoptosis in AMC-HN3 cells. Int. J. Oncol. 2012, 41, 2184–2190. [Google Scholar] [CrossRef] [Green Version]
  228. Yadav, S.; Singh, A.K.; Agrahari, A.K.; Sharma, K.; Singh, A.S.; Gupta, M.K.; Tiwari, V.K.; Prakash, P. Making of water soluble curcumin to potentiate conventional antimicrobials by inducing apoptosis-like phenomena among drug-resistant bacteria. Sci. Rep. 2020, 10, 1–22. [Google Scholar] [CrossRef]
  229. Ding, L.; Ma, S.; Lou, H.; Sun, L.-R.; Ji, M. Synthesis and Biological Evaluation of Curcumin Derivatives with Water-Soluble Groups as Potential Antitumor Agents: An in Vitro Investigation Using Tumor Cell Lines. Molecules 2015, 20, 21501–21514. [Google Scholar] [CrossRef] [Green Version]
  230. Mishra, S.; Narain, U.; Mishra, R.; Misra, K. Design, Development and Synthesis of Mixed Bioconjugates of Piperic Acid-Glycine, Curcumin-Glycine/alanine and Curcumin-Glycine-Piperic Acid and Their Antibacterial and Antifungal Properties. Bioorganic Med. Chem. 2005, 13, 1477–1486. [Google Scholar] [CrossRef]
  231. Lu, P.; Tong, Q.; Jiang, F.; Zheng, L.; Chen, F.; Zeng, F.; Dong, J.; Du, Y. Preparation of curcumin prodrugs and their in vitro anti-tumor activities. J. Huazhong Univ. Sci. Technol. Med. Sci. 2005, 25, 668–678. [Google Scholar]
  232. Wichitnithad, W.; Nimmannit, U.; Wacharasindhu, S.; Rojsitthisak, P. Synthesis, Characterization and Biological Evaluation of Succinate Prodrugs of Curcuminoids for Colon Cancer Treatment. Molecules 2011, 16, 1888–1900. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  233. Kumar, S.; Narain, U.; Tripathi, S.; Misra, K. Syntheses of Curcumin Bioconjugates and Study of Their Antibacterial Activities against beta-Lactamase-Producing Microorganisms. Bioconjugate Chem. 2001, 12, 464–469. [Google Scholar] [CrossRef] [PubMed]
  234. Mishra, S.; Tripathi, S.; Mishra, R.; Misra, K. Design, Synthesis and Characterisation of a Novel Anticancer Prodrug Having Antiproliferative Activity against Prostrate Tumour. Indian J. Chem. Sect. B Org. Med. Chem. 2005, 44, 2582–2588. [Google Scholar]
  235. Tang, H.; Murphy, C.J.; Zhang, B.; Shen, Y.; Sui, M.; Van Kirk, E.A.; Feng, X.; Murdoch, W.J. Amphiphilic Curcumin Conjugate-Forming Nanoparticles as Anticancer Prodrug and Drug Carriers: In Vitro and in Vivo Effects. Nanomedicine 2010, 5, 855–865. [Google Scholar] [CrossRef]
  236. Wang, Z.; Chen, C.; Zhang, Q.; Gao, M.; Zhang, J.; Kong, D.; Zhao, Y. Tuning the architecture of polymeric conjugate to mediate intracellular delivery of pleiotropic curcumin. Eur. J. Pharm. Biopharm. 2015, 90, 53–62. [Google Scholar] [CrossRef]
  237. Li, M.; Gao, M.; Fu, Y.; Chen, C.; Meng, X.; Fan, A.; Kong, D.; Wang, Z.; Deng, J. Acetal-linked polymeric prodrug micelles for enhanced curcumin delivery. Colloids Surfaces B Biointerfaces 2016, 140, 11–18. [Google Scholar] [CrossRef]
  238. Massaro, M.; Amorati, R.; Cavallaro, G.; Guernelli, S.; Lazzara, G.; Milioto, S.; Noto, R.; Poma, P.; Riela, S. Direct chemical grafted curcumin on halloysite nanotubes as dual-responsive prodrug for pharmacological applications. Colloids Surfaces B Biointerfaces 2016, 140, 505–513. [Google Scholar] [CrossRef]
  239. Shakeri, A.; Panahi, Y.; Johnston, T.P.; Sahebkar, A. Biological properties of metal complexes of curcumin. BioFactors 2019, 45, 304–317. [Google Scholar] [CrossRef]
  240. Wanninger, S.; Lorenz, V.; Subhan, A.; Edelmann, F.T. Metal complexes of curcumin—synthetic strategies, structures and medicinal applications. Chem. Soc. Rev. 2015, 44, 4986–5002. [Google Scholar] [CrossRef] [Green Version]
  241. Sareen, R.; Jain, N.; Dhar, K.L. Curcumin–Zn(II) complex for enhanced solubility and stability: An approach for improved delivery and pharmacodynamic effects. Pharm. Dev. Technol. 2015, 21, 630–635. [Google Scholar] [CrossRef] [PubMed]
  242. Pucci, D.; Bellini, T.; Crispini, A.; D’Agnano, I.; Liguori, P.F.; García-Orduña, P.; Pirillo, S.; Valentini, A.; Zanchetta, G. DNA binding and cytotoxicity of fluorescent curcumin-based Zn(ii) complexes. MedChemComm 2012, 3, 462. [Google Scholar] [CrossRef]
  243. Pucci, D.; Crispini, A.; Mendiguchía, B.S.; Pirillo, S.; Ghedini, M.; Morelli, S.; De Bartolo, L. Improving the bioactivity of Zn(ii)-curcumin based complexes. Dalton Trans. 2013, 42, 9679–9687. [Google Scholar] [CrossRef] [PubMed]
  244. Hieu, T.Q.; Doan, T.T. A novel study on curcumin metal complexes: Solubility improvement, bioactivity, and trial burn wound treatment in rats. New J. Chem. 2020, 44, 13036–13045. [Google Scholar] [CrossRef]
  245. Kühlwein, F.; Polborn, K.; Beck, W. Metallkomplexe von Farbstoffen. VIII Übergangsmetallkomplexe des Curcumins und Seiner Derivate. Zeitschrift für Anorg. und Allg. Chemie 1997, 623, 1211–1219. [Google Scholar] [CrossRef]
  246. Pi, Z.; Wang, J.; Jiang, B.; Cheng, G.; Zhou, S. A curcumin-based TPA four-branched copper(II) complex probe for in vivo early tumor detection. Mater. Sci. Eng. C 2015, 46, 565–571. [Google Scholar] [CrossRef]
  247. Zhao, X.-Z.; Jiang, T.; Wang, L.; Yang, H.; Zhang, S.; Zhou, P. Interaction of curcumin with Zn(II) and Cu(II) ions based on experiment and theoretical calculation. J. Mol. Struct. 2010, 984, 316–325. [Google Scholar] [CrossRef]
  248. Valentini, A.; Conforti, F.; Crispini, A.; De Martino, A.; Condello, R.; Stellitano, C.; Rotilio, G.; Ghedini, M.; Federici, G.; Bernardini, S.; et al. Synthesis, Oxidant Properties, and Antitumoral Effects of a Heteroleptic Palladium(II) Complex of Curcumin on Human Prostate Cancer Cells. J. Med. Chem. 2009, 52, 484–491. [Google Scholar] [CrossRef]
  249. Thompson, K.H.; Böhmerle, K.; Polishchuk, E.; Martins, C.; Toleikis, P.; Tse, J.; Yuen, V.; McNeill, J.H.; Orvig, C. Complementary inhibition of synoviocyte, smooth muscle cell or mouse lymphoma cell proliferation by a vanadyl curcumin complex compared to curcumin alone. J. Inorg. Biochem. 2004, 98, 2063–2070. [Google Scholar] [CrossRef]
  250. Miklášová, N.; Fischer-Fodor, E.; Mikláš, R.; Kucková, L.; Kožíšek, J.; Liptaj, T.; Soritau, O.; Valentová, J.; Tomuleasa, C. Synthesis and characterization of new biologically active palladium(II) complexes with (1E,6E)-1,7-bis(3,4-diethoxyphenyl)-1,6-heptadiene-3,5-dione. Inorg. Chem. Commun. 2014, 46, 229–233. [Google Scholar] [CrossRef]
  251. Pettinari, R.; Marchetti, F.; Condello, F.; Pettinari, C.; Lupidi, G.; Scopelliti, R.; Mukhopadhyay, S.; Riedel, T.; Dyson, P.J. Ruthenium(II)–Arene RAPTA Type Complexes Containing Curcumin and Bisdemethoxycurcumin Display Potent and Selective Anticancer Activity. Organometallics 2014, 33, 3709–3715. [Google Scholar] [CrossRef]
  252. Caruso, F.; Rossi, M.; Benson, A.; Opazo, C.; Freedman, D.; Monti, E.; Gariboldi, M.B.; Shaulky, J.; Marchetti, F.; Pettinari, R.; et al. Ruthenium-arene Complexes of Curcumin: X-Ray and Density Functional Theory Structure, Synthesis, and Spectroscopic Characterization, in Vitro Antitumor Activity, and DNA Docking Studies of (p-Cymene)Ru(curcuminato)chloro. J. Med. Chem. 2012, 55, 1072–1081. [Google Scholar] [CrossRef] [PubMed]
  253. Caruso, F.; Pettinari, R.; Rossi, M.; Monti, E.; Gariboldi, M.B.; Marchetti, F.; Pettinari, C.; Caruso, A.; Ramani, M.V.; Subbaraju, G.V. The in vitro antitumor activity of arene-ruthenium(II) curcuminoid complexes improves when decreasing curcumin polarity. J. Inorg. Biochem. 2016, 162, 44–51. [Google Scholar] [CrossRef] [PubMed]
  254. Henriques, M.C.; Faustino, M.A.F.; Silva, A.M.S.; Felgueiras, J.; Fardilha, M.; Braga, S.S. A ruthenium(II)-trithiacyclononane curcuminate complex: Synthesis, characterization, DNA-interaction, and cytotoxic activity. J. Co-Ord. Chem. 2017, 70, 2393–2408. [Google Scholar] [CrossRef]
  255. Li, S.; Xu, G.; Zhu, Y.; Zhao, J.; Gou, S. Bifunctional ruthenium(ii) polypyridyl complexes of curcumin as potential anticancer agents. Dalton Trans. 2020, 49, 9454–9463. [Google Scholar] [CrossRef]
  256. Ficai, A.; Marques, C.F.; Ferreira, J.M.; Andronescu, E.; Ficai, D.; Sonmez, M. Multifunctional materials for bone cancer treatment. Int. J. Nanomed. 2014, 9, 2713–2725. [Google Scholar] [CrossRef] [Green Version]
  257. Chin, D.; Huebbe, P.; Frank, J.; Rimbach, G.; Pallauf, K. Curcumin may impair iron status when fed to mice for six months. Redox Biol. 2014, 2, 563–569. [Google Scholar] [CrossRef] [Green Version]
Figure 1. The three main curcuminoids found in Curcuma longa.
Figure 1. The three main curcuminoids found in Curcuma longa.
Applsci 10 08990 g001
Figure 2. The metabolite curcumin β-D-glucuronide.
Figure 2. The metabolite curcumin β-D-glucuronide.
Applsci 10 08990 g002
Figure 3. Comparison of the skin penetration rate of curcumin (without any carrier) and curcumin loaded into two different carriers: conventional liposomes and cell-penetrating peptides (CPP)-coated liposomes. Drawn with base on the data reported by Patra et al. [213].
Figure 3. Comparison of the skin penetration rate of curcumin (without any carrier) and curcumin loaded into two different carriers: conventional liposomes and cell-penetrating peptides (CPP)-coated liposomes. Drawn with base on the data reported by Patra et al. [213].
Applsci 10 08990 g003
Figure 4. Synthesis and structure Ru(II) arene curcumin complexes, with R = hexamethylbenzene or p-cymene, and X = Cl or 1,3,5-triaza-7-phospha-adamantane (PTA).
Figure 4. Synthesis and structure Ru(II) arene curcumin complexes, with R = hexamethylbenzene or p-cymene, and X = Cl or 1,3,5-triaza-7-phospha-adamantane (PTA).
Applsci 10 08990 g004
Table 3. Plasma peak levels (Cmax) in rodents after parenteral administration of curcumin.
Table 3. Plasma peak levels (Cmax) in rodents after parenteral administration of curcumin.
SpeciesDoseRouteCmax Mean ± S.D. (nM)Ref.
Rat10 mg/kgI.V.977 ± 27[162]
Mouse100 mg/kgI.P.6107 ± (n.a.)[163]
Note: I.P. = intraperitoneal, I.V. = intravenous, n.a. = not available (S.D. was not calculated).
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Carvalho Henriques, M.; Faustino, M.A.F.; Santos Braga, S. Curcumin Innovative Delivery Forms: Paving the ‘Yellow Brick Road’ of Antitumoral Phytotherapy. Appl. Sci. 2020, 10, 8990. https://doi.org/10.3390/app10248990

AMA Style

Carvalho Henriques M, Faustino MAF, Santos Braga S. Curcumin Innovative Delivery Forms: Paving the ‘Yellow Brick Road’ of Antitumoral Phytotherapy. Applied Sciences. 2020; 10(24):8990. https://doi.org/10.3390/app10248990

Chicago/Turabian Style

Carvalho Henriques, Magda, Maria Amparo F. Faustino, and Susana Santos Braga. 2020. "Curcumin Innovative Delivery Forms: Paving the ‘Yellow Brick Road’ of Antitumoral Phytotherapy" Applied Sciences 10, no. 24: 8990. https://doi.org/10.3390/app10248990

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