Hostname: page-component-7c8c6479df-xxrs7 Total loading time: 0 Render date: 2024-03-28T08:08:00.380Z Has data issue: false hasContentIssue false

Curcumin as a therapeutic agent: the evidence from in vitro, animal and human studies

Published online by Cambridge University Press:  26 January 2010

Jenny Epstein*
Affiliation:
Centre for Digestive Diseases, Institute of Cell and Molecular Science, Barts and the London School of Medicine, Queen Mary, University of London, 4 Newark Street, LondonE1 2AT, UK
Ian R. Sanderson
Affiliation:
Centre for Digestive Diseases, Institute of Cell and Molecular Science, Barts and the London School of Medicine, Queen Mary, University of London, 4 Newark Street, LondonE1 2AT, UK
Thomas T. MacDonald
Affiliation:
Centre for Digestive Diseases, Institute of Cell and Molecular Science, Barts and the London School of Medicine, Queen Mary, University of London, 4 Newark Street, LondonE1 2AT, UK
*
*Corresponding author: Jenny Epstein, fax +44 2078822187, email j.epstein@qmul.ac.uk
Rights & Permissions [Opens in a new window]

Abstract

Curcumin is the active ingredient of turmeric. It is widely used as a kitchen spice and food colorant throughout India, Asia and the Western world. Curcumin is a major constituent of curry powder, to which it imparts its characteristic yellow colour. For over 4000 years, curcumin has been used in traditional Asian and African medicine to treat a wide variety of ailments. There is a strong current public interest in naturally occurring plant-based remedies and dietary factors related to health and disease. Curcumin is non-toxic to human subjects at high doses. It is a complex molecule with multiple biological targets and different cellular effects. Recently, its molecular mechanisms of action have been extensively investigated. It has anti-inflammatory, antioxidant and anti-cancer properties. Under some circumstances its effects can be contradictory, with uncertain implications for human treatment. While more studies are warranted to further understand these contradictions, curcumin holds promise as a disease-modifying and chemopreventive agent. We review the evidence for the therapeutic potential of curcumin from in vitro studies, animal models and human clinical trials.

Type
Review Article
Copyright
Copyright © The Authors 2009

For thousands of years, humankind has used plants for therapeutics. Recent years have seen the development of highly targeted biological treatments and synthetic therapies, some with serious side effects. At the same time, there is renewed public interest in complementary therapies, naturally occurring treatments with minimal toxicity and diets related to health and disease.

Curcumin is a constituent of the spice turmeric, one of the principal ingredients in curry powder. Turmeric is prepared from the root of the Curcuma longa plant, a member of the ginger family. It is native to India and Southeast Asia, where fresh turmeric root is widely used in a similar way to ginger; in the West, turmeric is much more commonly available as a dried powder. It has been used to treat a broad range of common ailments in Indian Ayurvedic medicine for at least 4000 years, as well as in Chinese, Arabic and other traditional medicines. Curcumin is in modern use worldwide as a cooking spice, flavouring agent and colorant. Dishes traditionally made with turmeric include dahls and most other curries, as well as pickles, relishes and chutneys. It is widely used to colour mustards, mayonnaises and margarines and has been designated as international food additive E100. Because of its resemblance to saffron, curcumin is sometimes referred to as ‘Indian saffron’ and used as a (much less expensive) substitute.

Chemistry

The active ingredient of curcumin is diferuloylmethane, a hydrophobic polyphenol with a characteristic yellow colour. In chemical terms it is bis-α, β-unsaturated β-diketone, a linear diarylheptanoid compound, where two oxy-substituted aryl moieties are linked together through a seven carbon chain (Fig. 1). The aryl rings may be substituted by varying numbers of hydroxy or methoxy groups in a symmetrical or asymmetrical fashion to produce analogues of curcumin or curcuminoids. Curcumin is the most abundantly occurring natural analogue at 77 %(Reference Anand, Thomas and Kunnumakkara1), followed by demethoxycurcumin (17 %) in which one methoxy group is absent, then bis-demethoxycurcumin (3 %) in which the methoxy group is absent from both the aryl rings (Fig. 1).

Fig. 1 Chemical structure of curcumin (diferuloylmethane), its natural analogues and principal metabolites.

There is no explicit evidence that correlates the molecular or stoichiometric properties of curcumin or its analogues with their biological effects. While several groups have studied the differential bioactivities of these different analogues, no single curcuminoid shows overall highest potency. Differential efficacy varies widely according to the cell type, function, disease system and organism in question(Reference Anand, Thomas and Kunnumakkara1). Thus, there is no consensus as to the most effective preparation for human use. Commercially available curcumin preparations are largely derived from natural curcumin sources and therefore contain the three main curcuminoids in approximately the afore mentioned proportions. Indeed, some data suggest that such a mixture of curcuminoids have synergistically greater activity than any of their individual elements(Reference Sandur, Pandey and Sung2).

Dose and safety

The safety, tolerability and non-toxicity of curcumin at high doses are well established. Oral doses up to 12 g/d are well tolerated in human subjects(Reference Lao, Ruffin and Normolle3), although dosing diet regimen above 8 g may be difficult to achieve due to the bulky nature of this quantity of compound(Reference Cheng, Hsu and Lin4). However, drug delivery is a problem and the bioavailability of oral curcumin is low(Reference Anand, Kunnumakkara and Newman5, Reference Sharma, Steward and Gescher6) due to a combination of efficient first pass metabolism, poor gastrointestinal absorption, rapid elimination and poor aqueous solubility. Elimination is largely via hepatic glucuronidation and sulphation. Glucuronidation of curcuminoids preferentially occurs on the phenolic hydroxyl group, when incubated with rat or human liver microsomes(Reference Pfeiffer, Hoehle and Walch7). This produces a strong lipophilic conjugate that is less stable than its unconjugated form and is excreted through stool. Whether such conjugates have pharmacological activity is uncertain(Reference Pfeiffer, Hoehle and Walch7Reference Aggarwal and Harikumar9). However, other, potentially active, metabolites have been identified (Fig. 1), perhaps the most important and intensively studied of which is tetrahydrocurcumin, a reduction metabolite. It lacks the yellow colour and hydrophobicity of curcumin and does not occur in natural curcumin sources. While it has less anti-inflammatory activity than curcumin in terms of its ability to inhibit NF-κB(Reference Sandur, Pandey and Sung2, Reference Pan, Lin-Shiau and Lin10), it exhibits greater antioxidant potency than curcumin in a number of different models(Reference Osawa, Sugiyama and Inayoshi11Reference Okada, Wangpoengtrakul and Tanaka13).

After oral curcumin dosing, serum concentrations peak at 1–2 h and are undetectable by 12 h(Reference Cheng, Hsu and Lin4). Some investigators report that serum curcumin is undetectable below oral doses of about 4 g(Reference Lao, Ruffin and Normolle3, Reference Cheng, Hsu and Lin4); however, others have detected curcumin not only in serum, but also in urine, at much lower doses(Reference Sharma, Euden and Platton14). Some studies demonstrate the presence of curcumin in colorectal tissue at oral doses of 3·6 g(Reference Garcea, Berry and Jones15), so the gut may represent a promising local clinical target for curcumin. The pharmacokinetic profile of its major metabolites may also be relevant to the biological effects of curcumin. Most curcumin conjugates produced by in vivo human metabolism are glucuronides (less commonly sulphates), and these are detectable in plasma at greater concentrations than free curcumin with a peak at 4 h after oral dosing(Reference Vareed, Kakarala and Ruffin8).

Thus, the apparent discrepancies in pharmacodynamics observed in different in vivo studies of curcumin may be explained by its high rate of conjugation. Additionally they may relate to the differing formulations used. Curcumin constitutes about 5 % of turmeric root(Reference Ammon and Wahl16, Reference Strimpakos and Sharma17); the remainder is made up of carbohydrates, proteins and essential oils. Preparations used for human consumption are either naturally produced from purified turmeric extract, which contain varying proportions of the different curcuminoids, or are synthetically produced, containing only pure chemically synthesised curcumin. Strategies have also been employed to improve bioavailability based on changes in drug formulation, such as the use of nanoparticles to reduce particle size delivery and micelles to counter hydrophobicity. Recently, it has been reported that heat treatment improves the water solubility of curcumin(Reference Kurien, Singh and Matsumoto18).

In human trials, only minor side effects of curcumin, namely diarrhoea(Reference Sharma, Euden and Platton14), have been reported, and it is considered safe and well tolerated. As a caveat, however, these trials have usually examined short-term outcomes. There is some evidence that long-term, high-dose curcumin administration in rodents can be tumourigenic(19, Reference Somasundaram, Edmund and Moore20). It has also been shown that curcumin's predominant activity switches from antioxidant to pro-oxidant with increasing concentration(Reference Sandur, Ichikawa and Pandey21), which may provide an explanation for its seeminglyopposing biological effects in vivo. These apparent contradictory roles of curcumin, as both anti-cancer and pro-carcinogenic agent, are as yet unexplained, and epitomise the complexity and paradoxical nature of the compound. Nevertheless, there is good evidence from India, at a population level, about the safety of lifelong curcumin ingestion up to about 100 mg/d(Reference Chainani-Wu22), and it is classified ‘Generally Recognised As Safe’ by the United States Food and Drug Administration.

In vitro studies

A wide variety of cellular properties of curcumin have been demonstrated, including antioxidant, anti-inflammatory, anti-proliferative, pro-apoptotic, anti-bacterial and anti-cancer activities (Table 1 and Fig. 2).

Table 1 Molecular targets of curcumin in cell line studies

Fig. 2 Cellular activities of curcumin and molecular mechanisms of action. ROS, reactive oxygen species; HO-1, haem oxygenase; Bax, BCL2-associated X protein; MAPK, mitogen-activated protein kinase; MMP, matrix metalloproteinase; STAT, signal transducer and activator of transcription; NOS, NO synthase; COX2, cyclo-oxygenase-2. , Activation by curcumin; ↕, effects vary; , suppression by curcumin.

Transcription factors

NF-κB

NF-κB is one of the key transcription factors responsive to curcumin. In human myeloid ML-1a cells, curcumin suppresses NF-κB activation induced by TNF-α, phorbol ester and hydrogen peroxide(Reference Singh and Aggarwal23). The mechanism appears to be via reduced IκBα phosphorylation and degradation(Reference Aggarwal, Ichikawa and Takada24), suggesting that curcumin acts at a step above IκB kinase (IKK) in the NF-κB activation pathway. Many of the observed biological effects of curcumin involve processes that are NF-κB-dependent. Therefore, examination of NF-κB signalling was a natural focus and its inhibition by curcumin is a consistent finding in a number of different models. For example, in four different human mantle cell lymphoma lines (an aggressive non-Hodgkin's B cell lymphoma), curcumin down-regulated NF-κB, inhibited IKK and reduced IκBα phosphorylation, leading to cell cycle arrest, apoptosis and suppression of proliferation(Reference Shishodia, Amin and Lai25). The reproducible finding of inhibition of IKK by curcumin suggests that curcumin acts at or above the level of IKK in the NF-κB pathway. Investigators have shown modulation by curcumin of the serine/threonine protein kinase Akt, a ubiquitous cell signalling molecule, which is known to activate NF-κB. Curcumin suppresses both Akt activation and Akt–IKK association(Reference Aggarwal, Ichikawa and Takada24), and thus its effects on NF-κB may be a downstream consequence of true targets that lie higher upstream. The identification not only of NF-κB but also multiple other signalling molecules and transcription factors which are modulated by curcumin further suggests that an upstream direct target (or targets) of curcumin common to these pathways may exist.

Signal transducer and activator of transcription

Signal transducer and activator of transcription (STAT)3 is a transcriptional activator with a ubiquitous role in tumourigenesis. It is involved in dysregulation of cell growth, invasion, angiogenesis, metastasis and resistance to apoptosis(Reference Aggarwal, Sethi and Ahn26). Aberrant STAT3 signalling is an important process in the development and progression of cancer, thus agents that block its activation have therapeutic potential. Curcumin reversibly inhibits STAT3 activation in human multiple myeloma cells and by this mechanism suppresses IL-6-induced cell proliferation(Reference Bharti, Donato and Aggarwal27). It also inhibits STAT3 activation in five different human Hodgkin and Reed-Sternberg lymphoma cell lines(Reference Mackenzie, Queisser and Wolfson28). The down-regulation by curcumin of proteins involved in cell cycling and apoptosis such as cyclin D1 and bcl-XL(Reference Bharti, Donato and Aggarwal27, Reference Han, Chung and Robertson29) may be secondary manifestations of curcumin's inhibitory effects upon STAT3 and NF-κB, which are known to regulate expression of both of these genes(Reference Fujio, Kunisada and Hirota30, Reference Sinibaldi, Wharton and Turkson31).

PPAR-γ

PPAR-γ is a nuclear receptor and transcription factor involved in cell cycle control, proliferation and differentiation, exerting anti-inflammatory, anti-cancer and insulin-sensitising actions. It is highly expressed in adipose tissue and colonic mucosa, where tight control of proliferation, differentiation and apoptosis is vital for homeostasis and prevention of oncogenesis, and here PPAR-γ may have tumour suppressor functions(Reference Sarraf, Mueller and Smith32). It is activated by PG products of the eicosanoid cascade(Reference Nakamura and Omaye33, Reference Ondrey34) and possibly by dietary components such as linolenic and linoleic acids. Curcumin induces and activates PPAR-γ in rat hepatic stellate cells, a liver cell type responsible for fibrosis in liver injury, which contributes to chronic liver damage and cirrhosis. PPAR-γ inhibited the proliferation of stellate cells, and curcumin greatly enhanced this effect(Reference Xu, Fu and Chen35).

PPAR-γ activity in Moser cells (a human colon cancer cell line) is also enhanced by curcumin, interrupting the cell cycle through reduced expression of cyclin D1 and inhibition of epidermal growth factor signalling(Reference Chen and Xu36). Both of these effects were PPAR-γ-dependent. However, the anti-cancer effect of curcumin occurs through multiple mechanisms, and this is supported by the finding that a different human colon cancer cell line, HT-29 cells, despite being more sensitive to curcumin-induced growth suppression, is less responsive to specific PPAR-γ antagonism than Moser cells(Reference Chen and Xu36). These data reflect both that PPAR-γ function is one of the many mechanisms involved in the generation of cancer, and that curcumin exerts its anti-cancer effects through multiple pathways.

Mitogen-activated protein kinase signalling pathways

The mitogen-activated protein kinase (MAPK) cascade is activated by a large number of different types of receptor, including cytokine, growth and toll-like receptors and receptors sensitive to environmental stressors. The precise mechanisms of activation are incompletely understood(Reference Wetzker and Bohmer37). Curcumin modulates MAPK signalling in several different in vitro models, although the data are somewhat contradictory. Under some circumstances, curcumin inhibits MAPK activation, as in a recent study in primary human intestinal microvascular endothelial cells, where curcumin inhibited p38 MAPK activation in response to vascular endothelial growth factor, as well as cyclo-oxygenase (COX)-2 and PGE2 production(Reference Binion, Otterson and Rafiee38). These anti-angiogenic properties of curcumin are of potential clinical benefit in gut inflammation and cancer. Further evidence that curcumin inhibits MAPK pathways includes its inhibition of c-Jun N-terminal kinase activation by a number of different agonists in Jurkat T cells (a human T cell line)(Reference Chen and Tan39). Here the investigators provide evidence that the target of inhibition lies proximally within the pathway, at the level of MAPK kinase kinase or above. Other investigators paradoxically show activation of MAPK by curcumin, for example c-Jun N-terminal kinase in HCT116 cells, a human colon cancer cell line(Reference Collett and Campbell40) and p38 MAPK in primary human neutrophils(Reference Hu, Du and Vancurova41).

Is it feasible that curcumin can both activate and inhibit MAPK signalling? Where MAPK is activated, the biological consequence seen is apoptosis; where MAPK is inhibited, the consequences are anti-inflammatory and anti-angiogenic. The mechanism for the opposing actions of curcumin on MAPK is unexplained, and in both cases its final effects are demonstrably anti-neoplastic and anti-inflammatory. Assuming the primary molecular targets of curcumin lie elsewhere, the MAPK signals observed experimentally may merely represent intermediary pathways by which its ultimate biological effects are mediated. Alternatively, where MAPK activation is seen, it is possible that this has been due to ubiquitous experimental contaminants such as lipopolysaccharide masking the true inhibitory effect of curcumin. In support of this explanation, curcumin-mediated apoptosis (suggested to occur via p38 MAPK activation) was not abrogated by the specific p38 MAPK inhibitor, SB203580(Reference Collett and Campbell40).

Tumour suppressor gene p53

Mutation of the tumour suppressor p53 plays an important role in the evolution of many different human cancers. Once again, the role of curcumin is complex. In an early study of the effects of curcumin on BKS-2 and WEHI-231 cells (both immature B cell lymphoma mouse cell lines), proliferation was inhibited(Reference Han, Chung and Robertson29). Interestingly, and with obvious potential clinical benefit in cancer chemotherapy, this inhibitory effect was much less marked on normal B cells. The investigators demonstrated (unexpected) inhibition of expression of p53 by curcumin, as well as inhibition of various other genes involved in growth, proliferation and transcriptional activation, including early growth response factor (egr)-1, the proto-oncogene c-myc and the transmembrane anti-apoptotic bcl-XL. The finding of reduced p53 activity was confirmed in RKO cells (a colon cancer cell line), where curcumin impairs the post-translational folding of p53 required for its function(Reference Moos, Edes and Mullally42), and in myeloid leukaemic cells, where it induces p53 degradation(Reference Tsvetkov, Asher and Reiss43).

Conversely, other experiments show induction of p53 by curcumin, for example in human epithelial breast cancer, prostate cancer and B cell lymphoma cell lines(Reference Choudhuri, Pal and Das44) and in HT-29 cells (a human colon adenocarcinoma cell line), where it induced apoptosis(Reference Song, Mao and Cai45). In the former work, once again the authors show differential sensitivity of cancer cells compared with healthy cells to curcumin. While some investigators have shown anti-proliferative effects despite inhibition of the tumour suppressor p53(Reference Han, Chung and Robertson29), established precedents exist where an agent that is cancer-preventative in one system can be carcinogenic in another, for example tamoxifen (therapeutic in breast; pro-neoplastic in uterus)(Reference Fisher, Costantino and Wickerham46). Curcumin may be a clinically useful chemopreventive agent, and this might relate specifically to certain types of cancer and not others. Alternatively, it may confer cancer risk that is inseparable from its benefits. These cautions must be borne in mind when considering its human use.

Angiogenesis

There is strong evidence that curcumin is anti-angiogenic. Angiogenesis (the growth of new blood vessels) is required for the development of both inflammation and cancer, where it is crucial for the survival of tumours beyond a certain size. It is also integral to the generation of diabetic eye disease, which is characterised by growth of abnormal vessels across the retina, a major cause of blindness worldwide. In an early study in both primary bovine and immortalised mouse endothelial cells, curcumin inhibited endothelial cell proliferation(Reference Arbiser, Klauber and Rohan47). Curcumin inhibits angiogenesis in response to vascular endothelial growth factor in the human intestinal microvascular endothelium(Reference Binion, Otterson and Rafiee38) and inhibits the angiogenic differentiation of human umbilical vein endothelial cells(Reference Thaloor, Singh and Sidhu48, Reference Bae, Kim and Jeong49). Also in human umbilical vein endothelial cells, curcumin binds to and irreversibly inhibits aminopeptidase N(Reference Shim, Kim and Cho50), a membrane-bound matrix metalloproteinase (MMP), which increases tumour invasiveness and is involved in retinal neovascularisation and tumour angiogenesis(Reference Pasqualini, Koivunen and Kain51). Finally, curcumin decreases hypoxia-inducible factor-1α, an angiogenic transcriptional activator, in human hepatocellular carcinoma cells(Reference Bae, Kim and Jeong49). In this work, curcumin also inhibited the transcriptional action of hypoxia-inducible factor-1α, down-regulating the expression of vascular endothelial growth factor, a potent hypoxia-induced angiogenic factor.

Inflammatory cytokines

Several studies demonstrate the suppression of downstream pro-inflammatory and pro-neoplastic mediators by curcumin. Recent examples include reduced expression of IL-6 and IL-8 in response to acid exposure in a human oesophageal epithelial cell line(Reference Rafiee, Nelson and Manley52) and reduced spontaneous expression of IL-6 and IL-8 in four different head and neck squamous carcinoma cell lines(Reference Cohen, Veena and Srivatsan53). These observations may be secondary to the suppression by curcumin of intermediary signalling pathways such as NF-κB, and some investigators provide evidence to this effect(Reference Cohen, Veena and Srivatsan53). Even if curcumin-mediated cytokine suppression is a later consequence of proximal event(s), this remains a potentially useful clinical application, and the one which is consistently reproduced in pre-clinical models.

Cyclo-oxygenase

COX2 is an inducible form of PGH synthase. It is an early response gene induced by cytokines, growth factors and toxins. COX2 mediates inflammation through production of PG and plays an important role in colon cancer. Over-expression of COX2 in colonic epithelium appears to promote tumour development(Reference Nishisho, Nakamura and Miyoshi54) and non-steroidal anti-inflammatory drugs that inhibit COX2, reduce the risk of colon cancer(Reference Thun, Namboodiri and Heath55). Curcumin inhibits COX2 production in a primary human intestinal microvascular endothelial cell line(Reference Binion, Otterson and Rafiee38) and inhibits COX2 induction in human colonic epithelial cells(Reference Plummer, Holloway and Manson56). In this latter work, the authors note that the COX2 gene promoter contains two NF-κB binding sites and show evidence that the effect of curcumin on COX2 is due to inhibition of NF-κB binding. In agreement with other investigators(Reference Aggarwal, Ichikawa and Takada24), the level of impact of curcumin upon the NF-κB pathway appears to be at or above IKK. Inhibition of COX2 by curcumin, which lacks the adverse effects of chronic aspirin or non-steroidal anti-inflammatory drugs ingestion, holds considerable promise for long-term bowel cancer prevention in human subjects.

Matrix metalloproteinases

In health, fibroblasts produce low levels of MMP that remain largely in latent form and mediate physiological extracellular matrix turnover. In inflammatory disease, MMP are over-expressed and become activated in cascades causing unchecked tissue destruction, fibrosis and further increasing immune cell activation and homing(Reference Pender and MacDonald57). MMP also play a key role in tumour progression, since matrix dissolution is an important step in the conversion of a pre-malignant cell into a frankly malignant one, as well as in tumour growth, invasion, metastasis and angiogenesis(Reference Vihinen, Ala-aho and Kahari58). There are over twenty different types of MMP, which are sub-classified according to the primary stromal substrate upon which they act. Curcumin down-regulates MMP production in various cell types. In human fibrosarcoma cells, it decreases invasion, migration and production of MMP-2 and MMP-9(Reference Yodkeeree, Garbisa and Limtrakul59), and in human and rabbit peripheral blood mononuclear cells, it reduces MMP-9(Reference Saja, Babu and Karunagaran60). Recently, it has been shown to reduce MMP-9 in human intestinal epithelial cells(Reference Claramunt, Bouissane and Cabildo61), and our group has shown dose-dependent inhibition of MMP-3 production by curcumin in primary human colonic myofibroblasts from patients with inflammatory bowel disease (IBD)(Reference Epstein, Docena and MacDonald62). Thus, inhibition of MMP by curcumin is a consistent finding under a range of different cellular conditions. The clinical implications for prevention and treatment of inflammation and cancer are wide ranging.

p300 Acetyl transferase

Lastly, curcumin is a known inhibitor of acetylation, acting on the enzyme p300 acetyl transferase(Reference Lee, Lin and Lin63, Reference Balasubramanyam, Varier and Altaf64). Acetylation modifies proteins when an acetyl group binds to a lysine residue, altering the protein's shape, charge and biological fate in the cell. Traditionally the study of acetylation has examined how the acetylation of histones changes their conformation, loosens their interactions with DNA and thus opens out the nucleosome, exposing DNA for gene transcription(Reference Roth, Denu and Allis65, Reference Sanderson and Naik66). However, recent work shows that other (non-histone) regulatory proteins within the cell are also subjected to acetylation, initiating separate cellular events that regulate for example transforming growth factor-β signalling(Reference Monteleone, Del Vecchio Blanco and Monteleone67) and insulin-like growth factor binding protein-3 expression(Reference Ongeri, Verderame and Hammond68, Reference White, Mulligan and King69). Such events are important in inflammation and cellular proliferation. Another important such non-histone example is the tumour suppressor gene p53 whose capacity to activate transcription and therefore DNA repair is altered by p300 status(Reference Yang70, Reference Brooks and Gu71), and indeed mutations in p300 have been found in several different types of cancer specimen, particularly in gut cancers(Reference Muraoka, Konishi and Kikuchi-Yanoshita72).

p300 Acetyl transferase, as a potent catalyst of acetylation, plays a role in a wide variety of gene transcription and other cellular events. Several effects of curcumin resulting from its p300 inhibitor activity are documented, including inhibition of inflammatory responses in human tracheal smooth muscle cells(Reference Lee, Lin and Lin63), suppression of HIV proliferation(Reference Balasubramanyam, Varier and Altaf64) and inhibition of proliferation of Raji cells (a non-Hodgkin's B cell lymphoma line)(Reference Chen, Shu and Chen73), reflecting once again a broad spectrum of potential clinical applications that might be developed.

Animal models: inflammatory bowel disease

While curcumin has shown benefits in a number of different models of inflammatory disease, particular interest has focused on its use in the gut. IBD (Crohn's disease (CD) and ulcerative colitis (UC)) is a source of considerable morbidity, and its incidence is increasing worldwide. Currently available treatments such as steroids, 5-aminosalicylic acids and immunomodulators do not offer cure, but CD responds well to polymeric or elemental feed that brings about remission in 80 % of paediatric patients(Reference Heuschkel, Menache and Megerian74, Reference Bannerjee, Camacho-Hubner and Babinska75). IBD is less common in developing countries than in the industrialised world(Reference Goh and Xiao76), and individuals emigrating from East to West take on the Western disease risk(Reference Goh and Xiao76, Reference Montgomery, Morris and Pounder77). This holds further relevance to the importance of diet in IBD, and there is keen interest to develop nutritional therapies.

Several studies in various rodent disease models provide strong pre-clinical evidence for the benefit of curcumin(Reference Jian, Mai and Wang78Reference Venkataranganna, Rafiq and Gopumadhavan81). For example, in multidrug resistance gene-deficient mice, which spontaneously develop colitis, the addition of curcumin to their diet significantly reduced intestinal inflammation(Reference Nones, Knoch and Dommels80). Other investigators used 2,4-dinitrochlorobenzene-induced colitis in rats and showed a dose-dependent improvement in disease activity parameters with dietary curcumin of equal potency to sulfasalazine treatment(Reference Venkataranganna, Rafiq and Gopumadhavan81). Curcumin treatment was associated with a reduction in colonic NF-κB, inducible NO synthase and various measures of oxidative stress, for example myeloperoxidase and lipid peroxidation.

The efficacy of curcumin in IBD may differ according to inflammatory circumstances and dose. For example, trinitrobenzene sulphonic acid colitis in NKT-deficient SJL/J mice exhibits a classic T helper cell (Th)1-type response, while BALB/c mice with trinitrobenzene sulphonic acid colitis exhibit a mixed Th1/Th2 profile(Reference Billerey-Larmonier, Uno and Larmonier82). Curcumin caused improvement in all disease activity parameters only in the BALB/c mice. In simple terms, Th1-type inflammation relates more closely to CD and Th2 to UC, although in real terms the situation is probably more complex with a degree of overlap. The reason for the differential efficacy of curcumin in these two models is unclear. The IL-10 knockout mouse develops spontaneous Th1-type inflammation in large and small bowel, which is dependent on gut bacteria, making it a good model of CD. The protective effect of curcumin in this model (by colon morphology and colonic interferon γ and IL-12/23p40 mRNA) was modest, and paradoxically occurred only at the lowest dietary concentration of 0·1 %(Reference Larmonier, Uno and Lee83). In vivo NF-κB activation in the gut was unaffected by curcumin at any concentration, but curcumin acted synergistically with IL-10 on epithelial cells to decrease NF-κB activity. These data raise once again the suggestion that curcumin can have paradoxically opposing effects at different concentrations, and when clinical studies take place, a wide range of dosages are warranted.

Animal models: cancer

Chemoprevention

The molecular targets of curcumin include many pathways and processes involved in the generation and propagation of cancer. The observation that many common cancers (including colon, breast, prostate and lung) are commoner in the Western world than in countries such as India, where there is high natural dietary curcumin consumption(Reference Chainani-Wu22), while not indicative of cause and effect, is intriguing. Curcumin has been investigated as both chemotherapeutic and chemopreventive agent in many different animal (largely rodent) models of carcinogenesis. Its chemopreventive efficacy for colon cancer is particularly well established(Reference Rao, Simi and Reddy84, Reference Kim, Araki and Kim85). Other gastrointestinal cancers against which curcumin has shown protective effects include oesophageal(Reference Ushida, Sugie and Kawabata86), stomach(Reference Ikezaki, Nishikawa and Furukawa87), liver(Reference Chuang, Cheng and Lin88) and oral(Reference Azuine and Bhide89); all in rodent models. Curcumin also shows chemopreventive properties in rodent models of various extra-intestinal cancers, including breast(Reference Huang, Lou and Xie90), lung(Reference Hecht, Kenney and Wang91), kidney(Reference Okazaki, Iqbal and Okada92), bladder(Reference Sindhwani, Hampton and Baig93), blood(Reference Huang, Lou and Xie90) and skin(Reference Limtrakul, Lipigorngoson and Namwong94) (Table 2).

Table 2 Animal models in which curcumin has chemopreventive efficacy

Chemotherapy

Curcumin inhibits tumour growth and metastasis, and has chemosensitising and radiosensitising properties. One of the earliest examples of the ability of curcumin to inhibit tumour growth is that of lymphoma cells in a mouse ascites model, when it was administered intraperitoneally at 50 mg/kg(Reference Kuttan, Bhanumathy and Nirmala95). Curcumin also has anti-tumour efficacy against human melanoma cell xenografts if given intraperitoneally(Reference Odot, Albert and Carlier96). Also in xenograft models, sub-cutaneous delivery of curcumin suppresses growth of head and neck squamous carcinoma cells(Reference LoTempio, Veena and Steele97), and when given orally it inhibits proliferation and angiogenesis and induces apoptosis in prostate cancer cells(Reference Dorai, Cao and Dorai98).

Curcumin also suppresses proliferation and angiogenesis and enhances apoptosis in pancreatic cancer; both when given orally in combination with gemcitabine in an orthotopic model(Reference Kunnumakkara, Guha and Krishnan99), and in a xenograft model when given intravenously in a liposomal formulation(Reference Li, Braiteh and Kurzrock100). The same group have also used an intravenous liposomal curcumin preparation in luminal gastrointestinal cancers, where it has chemosensitising properties against colorectal cancer in a mouse xenograft model(Reference Li, Ahmed and Mehta101). In this work, tumour growth and angiogenesis were inhibited and apoptosis enhanced in combination with oxaliplatin. In an orthotopic implantation model of hepatocellular carcinoma, curcumin also prevented intrahepatic metastasis(Reference Ohashi, Tsuchiya and Koizumi102).

Finally, in recent work, oral curcumin has shown efficacy in preventing breast cancer metastasis to lung in orthotopic models, both as chemosensitiser in conjunction with paclitaxel(Reference Aggarwal, Shishodia and Takada103) and in the prevention of its haematogenous spread in immunodeficient mice(Reference Bachmeier, Nerlich and Iancu104). Curcumin given intraperitoneally in combination with docetaxel inhibits tumour growth and angiogenesis in an orthotopic nude mouse model of ovarian cancer(Reference Lin, Kunnumakkara and Nair105).

Human trials

The wealth of in vitro and pre-clinical data has provided a strong basis from which to progress to the trialling of curcumin in human subjects. Many of the molecular efficacies of curcumin demonstrated in cell culture systems and animal models are comparable to those seen in human subjects (Fig. 3). The anti-inflammatory targets of curcumin including reduction of NF-κB, COX2 and pro-inflammatory cytokines such as IL-1, IL-6 and TNF-α, translate into clinical anti-inflammatory efficacy with improvement of rheumatoid arthritis(Reference Kobelt106, Reference Deodhar, Sethi and Srimal107), psoriasis(Reference Heng, Song and Harker108), post-operative inflammation(Reference Satoskar, Shah and Shenoy109), chronic anterior uveitis(Reference Lal, Kapoor and Agrawal110) and orbital inflammatory pseudo-tumours(Reference Lal, Kapoor and Asthana111). Concordant with the finding that high concentrations of curcumin are achievable in gastrointestinal tissue, curcumin shows clinical benefit in irritable bowel syndrome(Reference Bundy, Walker and Middleton112), tropical pancreatitis(Reference Durgaprasad, Pai and Vasanthkumar113), gall bladder and biliary motility(Reference Niederau and Gopfert114Reference Rasyid, Rahman and Jaalam116), gastric ulceration(Reference Prucksunand, Indrasukhsri and Leethochawalit117) and familial adenomatous polyposis coli(Reference Cruz-Correa, Shoskes and Sanchez118). The in vitro findings of enhanced PPAR-γ expression and modulation of NOS, glutathione and other antioxidant activities are supported by the clinical potency of curcumin to lower serum cholesterol(Reference Soni and Kuttan119) and improve endothelial function in type 2 diabetes mellitus(Reference Usharani, Mateen and Naidu120). Curcumin also enhances early post-transplant renal graft function(Reference Shoskes, Lapierre and Cruz-Correa121), presumably through multiple mechanisms.

Fig. 3 Clinical effects of curcumin: results from human trials (with references).

Consistent with the strong pre-clinical evidence of benefit in animal models of IBD, curcumin is showing early promise as a treatment for CD and UC in human subjects. In a small open-label study of five patients with CD and five with ulcerative proctitis, improvements in clinical and laboratory parameters with reduction in need for concomitant medications were observed in nine out of ten cases(Reference Holt, Katz and Kirshoff122). Further encouraging results came from a larger multicentre, randomised, double-blind, controlled trial of eighty-nine patients with quiescent UC, in which two out of forty-three patients (5 %) taking oral curcumin had relapsed by 6 months compared with eight out of thirty-nine (21 %) in the placebo group(Reference Hanai, Iida and Takeuchi123). The investigators also showed significant clinical and endoscopic improvements in the curcumin-treated group.

There is a strong foundation of evidence from both in vitro and animal models that curcumin has anti-cancer actions, including its pro-apoptotic and anti-angiogenic effects and its modulation of the cell cycle, growth factor expression and signal transduction pathways. Building upon this foundation, curcumin appears to prevent and treat cancer in human subjects. Results from a trial of twenty-five patients with various different pre-malignant or high-risk lesions suggested that oral curcumin may have chemopreventive effects in progression of these lesions(Reference Cheng, Hsu and Lin4). While two of the twenty-five patients progressed to frank cancer, seven regressed; a remarkably high proportion considering the high-grade nature of the lesions (bladder cancer, oral leukoplakia, gastric intestinal metaplasia, cervical intraepithelial metaplasia and Bowen's disease). In another uncontrolled study of fifteen patients with advanced colorectal cancer refractory to standard treatments, the lymphocytic biomarker glutathione S transferase showed a 59 % reduction in activity with low-dose (440 mg daily) oral curcuma extract, and five patients maintained radiologically stable disease over the 2- to 4-month study period(Reference Sharma, McLelland and Hill124). Once again there is a suggestion here that curcumin exhibits paradoxical efficacy at low v. high dose, since this effect was not observed at higher doses. In an interesting, but also uncontrolled, study of sixty-two patients with oral cancerous lesions, topical curcumin application reduced symptoms in the majority (70 %) and caused tumour shrinkage in 10 %(Reference Kuttan, Sudheeran and Josph125). Of twenty-one patients with advanced, normally rapidly fatal, pancreatic cancer treated with high-dose oral curcumin, encouragingly four showed disease stability or regression(Reference Dhillon, Aggarwal and Newman126).

These preliminary data hold promise, and interest in curcumin as a therapeutic agent continues to grow. There are several clinical trials currently ongoing, some involving larger numbers of patients and with a more rigorous, randomised, controlled design. A search on clinicaltrials.gov currently reveals thirty-one human trials using curcumin, of which fourteen are investigating its chemopreventive or chemotherapeutic potential in cancer or pre-malignant conditions. As in the data already reviewed, there is a preponderance of gut cancers; six are in colorectal cancer, two in familial adenomatous polyposis coli, one in UC and three in pancreatic cancer. A novel area of interest is in Alzheimer's disease and cognitive impairment. The first clinical trial failed to show benefit, but this may have been due to an unexpected lack of cognitive decline in the placebo group(Reference Baum, Lam and Cheung127). Three current ongoing trials of curcumin are further assessing its efficacy in age-related cognitive impairment. Interest also continues in systemic inflammatory conditions, and there are two ongoing trials of curcumin in arthritis and one in psoriasis.

Summary and conclusions

Since ancient times, curcumin has been used in a wide range of inflammatory, neoplastic and other conditions. In recent years, the molecular basis for its efficacy has been extensively investigated. Many cellular and molecular targets have been identified and many questions still remain. In complex multifactorial illnesses such as systemic inflammatory diseases and cancer, an agent that acts at a number of different cellular levels offers perhaps a better chance of effective prophylaxis or treatment. Its non-toxicity and good tolerability in human subjects, in combination with strong promising results from cell line, animal and early human clinical studies, support the ongoing research and development of curcumin as a preventive and disease-modifying agent.

Acknowledgements

This work was supported by the Crohn's in Childhood Research Association. There are no conflicts of interest. J. E. prepared and wrote the manuscript and designed the tables and figures. I. R. S. contributed to the preparation of the manuscript and was the Principal Investigator on the grant which funded the work. T. T. M. contributed to the preparation of the manuscript and advised on its structure and organisation. All the authors read and approved the manuscript.

References

1Anand, P, Thomas, SG, Kunnumakkara, AB, et al. (2008) Biological activities of curcumin and its analogues (congeners) made by man and mother nature. Biochem Pharmacol 76, 15901611.CrossRefGoogle ScholarPubMed
2Sandur, SK, Pandey, MK, Sung, B, et al. (2007) Curcumin, demethoxycurcumin, bisdemethoxycurcumin, tetrahydrocurcumin and tumerones differentially regulate anti-inflammatory and anti-proliferative responses through a ROS-independent mechanism. Carcinogenesis 28, 17651773.CrossRefGoogle ScholarPubMed
3Lao, CD, Ruffin, MTt, Normolle, D, et al. (2006) Dose escalation of a curcuminoid formulation. BMC Complement Altern Med 6, 10.CrossRefGoogle ScholarPubMed
4Cheng, AL, Hsu, CH, Lin, JK, et al. (2001) Phase I clinical trial of curcumin, a chemopreventive agent, in patients with high-risk or pre-malignant lesions. Anticancer Res 21, 28952900.Google ScholarPubMed
5Anand, P, Kunnumakkara, AB, Newman, RA, et al. (2007) Bioavailability of curcumin: problems and promises. Mol Pharm 4, 807818.CrossRefGoogle Scholar
6Sharma, RA, Steward, WP & Gescher, AJ (2007) Pharmacokinetics and pharmacodynamics of curcumin. Adv Exp Med Biol 595, 453470.CrossRefGoogle ScholarPubMed
7Pfeiffer, E, Hoehle, SI, Walch, SG, et al. (2007) Curcuminoids form reactive glucuronides in vitro. J Agric Food Chem 55, 538544.CrossRefGoogle ScholarPubMed
8Vareed, SK, Kakarala, M, Ruffin, MT, et al. (2008) Pharmacokinetics of curcumin conjugate metabolites in healthy human subjects. Cancer Epidemiol Biomarkers 17, 14111417.CrossRefGoogle ScholarPubMed
9Aggarwal, BB & Harikumar, KB (2009) Potential therapeutic effects of curcumin, the anti-inflammatory agent, against neurodegenerative, cardiovascular, pulmonary, metabolic, autoimmune and neoplastic diseases. Int J Biochem Cell Biol 41, 4059.CrossRefGoogle ScholarPubMed
10Pan, M-H, Lin-Shiau, S-Y & Lin, J-K (2000) Comparative studies on the suppression of nitric oxide synthase by curcumin and its hydrogenated metabolites through down-regulation of IκB kinase and NF-κB activation in macrophages. Biochem Pharmacol 60, 16651676.CrossRefGoogle ScholarPubMed
11Osawa, T, Sugiyama, Y, Inayoshi, M, et al. (1995) Antioxidative activity of tetrahydrocurcuminoids. Biosci Biotechnol Biochem 59, 16091612.CrossRefGoogle ScholarPubMed
12Naito, M, Wu, X, Nomura, H, et al. (2002) The protective effects of tetrahydrocurcumin on oxidative stress in cholesterol-fed rabbits. J Atheroscler Thromb 9, 243250.CrossRefGoogle ScholarPubMed
13Okada, K, Wangpoengtrakul, C, Tanaka, T, et al. (2001) Curcumin and especially tetrahydrocurcumin ameliorate oxidative stress-induced renal injury in mice. J Nutr 131, 20902095.CrossRefGoogle ScholarPubMed
14Sharma, RA, Euden, SA, Platton, SL, et al. (2004) Phase I clinical trial of oral curcumin: biomarkers of systemic activity and compliance. Clin Cancer Res 10, 68476854.CrossRefGoogle ScholarPubMed
15Garcea, G, Berry, DP, Jones, DJ, et al. (2005) Consumption of the putative chemopreventive agent curcumin by cancer patients: assessment of curcumin levels in the colorectum and their pharmacodynamic consequences. Cancer Epidemiol Biomarkers Prev 14, 120125.CrossRefGoogle ScholarPubMed
16Ammon, HP & Wahl, MA (1991) Pharmacology of Curcuma longa. Planta Med 57, 17.CrossRefGoogle ScholarPubMed
17Strimpakos, AS & Sharma, RA (2008) Curcumin: preventive and therapeutic properties in laboratory studies and clinical trials. Antioxid Redox Signal 10, 511545.CrossRefGoogle ScholarPubMed
18Kurien, BT, Singh, A, Matsumoto, H, et al. (2007) Improving the solubility and pharmacological efficacy of curcumin by heat treatment. Assay Drug Dev Technol 5, 567576.CrossRefGoogle ScholarPubMed
19National Toxicology Program (1993) NTP Toxicology and Carcinogenesis Studies of Turmeric Oleoresin (CAS No. 8024-37-1) (Major Component 79 %–85 % Curcumin, CAS No. 458-37-7) in F344/N Rats and B6C3F1 Mice (Feed Studies). Natl Toxicol Program Tech Rep Ser 427, 1275.Google Scholar
20Somasundaram, S, Edmund, NA, Moore, DT, et al. (2002) Dietary curcumin inhibits chemotherapy-induced apoptosis in models of human breast cancer. Cancer Res 62, 38683875.Google ScholarPubMed
21Sandur, SK, Ichikawa, H, Pandey, MK, et al. (2007) Role of pro-oxidants and antioxidants in the anti-inflammatory and apoptotic effects of curcumin (diferuloylmethane). Free Radic Biol Med 43, 568580.CrossRefGoogle ScholarPubMed
22Chainani-Wu, N (2003) Safety and anti-inflammatory activity of curcumin: a component of tumeric (Curcuma longa). J Altern Complement Med 9, 161168.CrossRefGoogle ScholarPubMed
23Singh, S & Aggarwal, BB (1995) Activation of transcription factor NF-kappa B is suppressed by curcumin (diferuloylmethane) [corrected]. J Biol Chem 270, 2499525000.CrossRefGoogle ScholarPubMed
24Aggarwal, S, Ichikawa, H, Takada, Y, et al. (2006) Curcumin (diferuloylmethane) down-regulates expression of cell proliferation and antiapoptotic and metastatic gene products through suppression of IkappaBalpha kinase and Akt activation. Mol Pharmacol 69, 195206.CrossRefGoogle ScholarPubMed
25Shishodia, S, Amin, HM, Lai, R, et al. (2005) Curcumin (diferuloylmethane) inhibits constitutive NF-kappaB activation, induces G1/S arrest, suppresses proliferation, and induces apoptosis in mantle cell lymphoma. Biochem Pharmacol 70, 700713.CrossRefGoogle ScholarPubMed
26Aggarwal, BB, Sethi, G, Ahn, KS, et al. (2006) Targeting signal-transducer-and-activator-of-transcription-3 for prevention and therapy of cancer: modern target but ancient solution. Ann N Y Acad Sci 1091, 151169.CrossRefGoogle ScholarPubMed
27Bharti, AC, Donato, N & Aggarwal, BB (2003) Curcumin (diferuloylmethane) inhibits constitutive and IL-6-inducible STAT3 phosphorylation in human multiple myeloma cells. J Immunol 171, 38633871.CrossRefGoogle ScholarPubMed
28Mackenzie, GG, Queisser, N, Wolfson, ML, et al. (2008) Curcumin induces cell-arrest and apoptosis in association with the inhibition of constitutively active NF-kappaB and STAT3 pathways in Hodgkin's lymphoma cells. Int J Cancer 123, 5665.CrossRefGoogle ScholarPubMed
29Han, SS, Chung, ST, Robertson, DA, et al. (1999) Curcumin causes the growth arrest and apoptosis of B cell lymphoma by downregulation of egr-1, c-myc, bcl-XL, NF-kappa B, and p53. Clin Immunol 93, 152161.CrossRefGoogle ScholarPubMed
30Fujio, Y, Kunisada, K, Hirota, H, et al. (1997) Signals through gp130 upregulate bcl-x gene expression via STAT1-binding cis-element in cardiac myocytes. J Clin Invest 99, 28982905.CrossRefGoogle ScholarPubMed
31Sinibaldi, D, Wharton, W, Turkson, J, et al. (2000) Induction of p21WAF1/CIP1 and cyclin D1 expression by the Src oncoprotein in mouse fibroblasts: role of activated STAT3 signaling. Oncogene 19, 54195427.CrossRefGoogle ScholarPubMed
32Sarraf, P, Mueller, E, Smith, WM, et al. (1999) Loss-of-function mutations in PPAR gamma associated with human colon cancer. Mol Cell 3, 799804.CrossRefGoogle ScholarPubMed
33Nakamura, YK & Omaye, ST (2009) Conjugated linoleic acid isomers' roles in the regulation of PPAR-gamma and NF-kappaB DNA binding and subsequent expression of antioxidant enzymes in human umbilical vein endothelial cells. Nutrition 25, 800811.CrossRefGoogle ScholarPubMed
34Ondrey, F (2009) Peroxisome proliferator-activated receptor gamma pathway targeting in carcinogenesis: implications for chemoprevention. Clin Cancer Res 15, 28.CrossRefGoogle Scholar
35Xu, J, Fu, Y & Chen, A (2003) Activation of peroxisome proliferator-activated receptor-gamma contributes to the inhibitory effects of curcumin on rat hepatic stellate cell growth. Am J Physiol 285, G20G30.Google Scholar
36Chen, A & Xu, J (2005) Activation of PPAR{gamma} by curcumin inhibits Moser cell growth and mediates suppression of gene expression of cyclin D1 and EGFR. Am J Physiol 288, G447G456.Google ScholarPubMed
37Wetzker, R & Bohmer, FD (2003) Transactivation joins multiple tracks to the ERK/MAPK cascade. Nat Rev Mol Cell Biol 4, 651657.CrossRefGoogle Scholar
38Binion, DG, Otterson, MF & Rafiee, P (2008) Curcumin inhibits VEGF-mediated angiogenesis in human intestinal microvascular endothelial cells through COX-2 and MAPK inhibition. Gut 57, 15091517.CrossRefGoogle ScholarPubMed
39Chen, YR & Tan, TH (1998) Inhibition of the c-Jun N-terminal kinase (JNK) signaling pathway by curcumin. Oncogene 17, 173178.CrossRefGoogle ScholarPubMed
40Collett, GP & Campbell, FC (2004) Curcumin induces c-jun N-terminal kinase-dependent apoptosis in HCT116 human colon cancer cells. Carcinogenesis 25, 21832189.CrossRefGoogle ScholarPubMed
41Hu, M, Du, Q, Vancurova, I, et al. (2005) Proapoptotic effect of curcumin on human neutrophils: activation of the p38 mitogen-activated protein kinase pathway. Crit Care Med 33, 25712578.CrossRefGoogle ScholarPubMed
42Moos, PJ, Edes, K, Mullally, JE, et al. (2004) Curcumin impairs tumor suppressor p53 function in colon cancer cells. Carcinogenesis 25, 16111617.CrossRefGoogle ScholarPubMed
43Tsvetkov, P, Asher, G, Reiss, V, et al. (2005) Inhibition of NAD(P)H:quinone oxidoreductase 1 activity and induction of p53 degradation by the natural phenolic compound curcumin. Proc Natl Acad Sci U S A 102, 55355540.CrossRefGoogle ScholarPubMed
44Choudhuri, T, Pal, S, Das, T, et al. (2005) Curcumin selectively induces apoptosis in deregulated cyclin D1-expressed cells at G2 phase of cell cycle in a p53-dependent manner. J Biol Chem 280, 2005920068.CrossRefGoogle Scholar
45Song, G, Mao, YB, Cai, QF, et al. (2005) Curcumin induces human HT-29 colon adenocarcinoma cell apoptosis by activating p53 and regulating apoptosis-related protein expression. Braz J Med Biol Res 38, 17911798.CrossRefGoogle ScholarPubMed
46Fisher, B, Costantino, JP, Wickerham, DL, et al. (1998) Tamoxifen for prevention of breast cancer: report of the National Surgical Adjuvant Breast and Bowel Project P-1 Study. J Natl Cancer Inst 90, 13711388.CrossRefGoogle ScholarPubMed
47Arbiser, JL, Klauber, N, Rohan, R, et al. (1998) Curcumin is an in vivo inhibitor of angiogenesis. Mol Med 4, 376383.CrossRefGoogle Scholar
48Thaloor, D, Singh, AK, Sidhu, GS, et al. (1998) Inhibition of angiogenic differentiation of human umbilical vein endothelial cells by curcumin. Cell Growth Differ 9, 305312.Google ScholarPubMed
49Bae, MK, Kim, SH, Jeong, JW, et al. (2006) Curcumin inhibits hypoxia-induced angiogenesis via down-regulation of HIF-1. Oncol Rep 15, 15571562.Google ScholarPubMed
50Shim, JS, Kim, JH, Cho, HY, et al. (2003) Irreversible inhibition of CD13/aminopeptidase N by the antiangiogenic agent curcumin. Chem Biol 10, 695704.CrossRefGoogle ScholarPubMed
51Pasqualini, R, Koivunen, E, Kain, R, et al. (2000) Aminopeptidase N is a receptor for tumor-homing peptides and a target for inhibiting angiogenesis. Cancer Res 60, 722727.Google Scholar
52Rafiee, P, Nelson, VM, Manley, S, et al. (2009) Effect of curcumin on acidic pH-induced expression of IL-6 and IL-8 in human esophageal epithelial cells (HET-1A): role of PKC, MAPKs, and NF-kappaB. Am J Physiol 296, G388G398.Google ScholarPubMed
53Cohen, AN, Veena, MS, Srivatsan, ES, et al. (2009) Suppression of interleukin 6 and 8 production in head and neck cancer cells with curcumin via inhibition of Ikappa beta kinase. Arch Otolaryngol Head Neck Surg 135, 190197.CrossRefGoogle ScholarPubMed
54Nishisho, I, Nakamura, Y, Miyoshi, Y, et al. (1991) Mutations of chromosome 5q21 genes in FAP and colorectal cancer patients. Science 253, 665669.CrossRefGoogle ScholarPubMed
55Thun, MJ, Namboodiri, MM & Heath, CW Jr (1991) Aspirin use and reduced risk of fatal colon cancer. N Engl J Med 325, 15931596.CrossRefGoogle ScholarPubMed
56Plummer, SM, Holloway, KA, Manson, MM, et al. (1999) Inhibition of cyclo-oxygenase 2 expression in colon cells by the chemopreventive agent curcumin involves inhibition of NF-kappaB activation via the NIK/IKK signalling complex. Oncogene 18, 60136020.CrossRefGoogle ScholarPubMed
57Pender, SL & MacDonald, TT (2004) Matrix metalloproteinases and the gut – new roles for old enzymes. Curr Opin Pharmacol 4, 546550.CrossRefGoogle ScholarPubMed
58Vihinen, P, Ala-aho, R & Kahari, VM (2005) Matrix metalloproteinases as therapeutic targets in cancer. Curr Cancer Drug Targets 5, 203220.CrossRefGoogle ScholarPubMed
59Yodkeeree, S, Garbisa, S & Limtrakul, P (2008) Tetrahydrocurcumin inhibits HT1080 cell migration and invasion via downregulation of MMPs and uPA. Acta Pharmacol Sin 29, 853860.CrossRefGoogle ScholarPubMed
60Saja, K, Babu, MS, Karunagaran, D, et al. (2007) Anti-inflammatory effect of curcumin involves downregulation of MMP-9 in blood mononuclear cells. Int Immunopharmacol 7, 16591667.CrossRefGoogle ScholarPubMed
61Claramunt, RM, Bouissane, L, Cabildo, MP, et al. (2009) Synthesis and biological evaluation of curcuminoid pyrazoles as new therapeutic agents in inflammatory bowel disease: effect on matrix metalloproteinases. Bioorg Med Chem 17, 12901296.CrossRefGoogle ScholarPubMed
62Epstein, J, Docena, G, MacDonald, TT, et al. (2010) Curcumin suppresses p38 mitogen-activated protein kinase activation, reduces IL-1 beta and matrix metalloproteinase-3 and enhances IL-10 in the mucosa of children and adults with inflammatory bowel disease. Br J Nutr 103, 824832.CrossRefGoogle ScholarPubMed
63Lee, CW, Lin, WN, Lin, CC, et al. (2006) Transcriptional regulation of VCAM-1 expression by tumor necrosis factor-alpha in human tracheal smooth muscle cells: involvement of MAPKs, NF-kappaB, p300, and histone acetylation. J Cell Physiol 207, 174186.CrossRefGoogle ScholarPubMed
64Balasubramanyam, K, Varier, RA, Altaf, M, et al. (2004) Curcumin, a novel p300/CREB-binding protein-specific inhibitor of acetyltransferase, represses the acetylation of histone/nonhistone proteins and histone acetyltransferase-dependent chromatin transcription. J Biol Chem 279, 5116351171.CrossRefGoogle ScholarPubMed
65Roth, SY, Denu, JM & Allis, CD (2001) Histone acetyltransferases. Annu Rev Biochem 70, 81120.CrossRefGoogle ScholarPubMed
66Sanderson, IR & Naik, S (2000) Dietary regulation of intestinal gene expression. Annu Rev Nutr 20, 311338.CrossRefGoogle ScholarPubMed
67Monteleone, G, Del Vecchio Blanco, G, Monteleone, I, et al. (2005) Post-transcriptional regulation of Smad7 in the gut of patients with inflammatory bowel disease. Gastroenterology 129, 14201429.CrossRefGoogle ScholarPubMed
68Ongeri, EM, Verderame, MF & Hammond, JM (2005) Follicle-stimulating hormone induction of ovarian insulin-like growth factor-binding protein-3 transcription requires a TATA box-binding protein and the protein kinase A and phosphatidylinositol-3 kinase pathways. Mol Endocrinol 19, 18371848.CrossRefGoogle ScholarPubMed
69White, NR, Mulligan, P, King, PJ, et al. (2006) Sodium butyrate-mediated Sp3 acetylation represses human insulin-like growth factor binding protein-3 expression in intestinal epithelial cells. J Pediatr Gastroenterol Nutr 42, 134141.CrossRefGoogle ScholarPubMed
70Yang, XJ (2004) The diverse superfamily of lysine acetyltransferases and their roles in leukemia and other diseases. Nucleic Acids Res 32, 959976.CrossRefGoogle ScholarPubMed
71Brooks, CL & Gu, W (2003) Ubiquitination, phosphorylation and acetylation: the molecular basis for p53 regulation. Curr Opin Cell Biol 15, 164171.CrossRefGoogle ScholarPubMed
72Muraoka, M, Konishi, M, Kikuchi-Yanoshita, R, et al. (1996) p300 Gene alterations in colorectal and gastric carcinomas. Oncogene 12, 15651569.Google ScholarPubMed
73Chen, Y, Shu, W, Chen, W, et al. (2007) Curcumin, both histone deacetylase and p300/CBP-specific inhibitor, represses the activity of nuclear factor kappa B and Notch 1 in Raji cells. Basic Clin Pharmacol Toxicol 101, 427433.CrossRefGoogle ScholarPubMed
74Heuschkel, RB, Menache, CC, Megerian, JT, et al. (2000) Enteral nutrition and corticosteroids in the treatment of acute Crohn's disease in children. J Pediatr Gastroenterol Nutr 31, 815.CrossRefGoogle ScholarPubMed
75Bannerjee, K, Camacho-Hubner, C, Babinska, K, et al. (2004) Anti-inflammatory and growth-stimulating effects precede nutritional restitution during enteral feeding in Crohn disease. J Pediatr Gastroenterol Nutr 38, 270275.Google ScholarPubMed
76Goh, K & Xiao, SD (2009) Inflammatory bowel disease: a survey of the epidemiology in Asia. J Dig Dis 10, 16.CrossRefGoogle ScholarPubMed
77Montgomery, SM, Morris, DL, Pounder, RE, et al. (1999) Asian ethnic origin and the risk of inflammatory bowel disease. Eur J Gastroenterol Hepatol 11, 543546.CrossRefGoogle ScholarPubMed
78Jian, YT, Mai, GF, Wang, JD, et al. (2005) Preventive and therapeutic effects of NF-kappaB inhibitor curcumin in rats colitis induced by trinitrobenzene sulfonic acid. World J Gastroenterol 11, 17471752.CrossRefGoogle ScholarPubMed
79Sugimoto, K, Hanai, H, Tozawa, K, et al. (2002) Curcumin prevents and ameliorates trinitrobenzene sulfonic acid-induced colitis in mice. Gastroenterology 123, 19121922.CrossRefGoogle ScholarPubMed
80Nones, K, Knoch, B, Dommels, YE, et al. (2009) Multidrug resistance gene deficient (mdr1a( − / − )) mice have an altered caecal microbiota that precedes the onset of intestinal inflammation. J Appl Microbiol 107, 557566.CrossRefGoogle Scholar
81Venkataranganna, MV, Rafiq, M, Gopumadhavan, S, et al. (2007) NCB-02 (standardized Curcumin preparation) protects dinitrochlorobenzene- induced colitis through down-regulation of NFkappa-B and iNOS. World J Gastroenterol 13, 11031107.CrossRefGoogle ScholarPubMed
82Billerey-Larmonier, C, Uno, JK, Larmonier, N, et al. (2008) Protective effects of dietary curcumin in mouse model of chemically induced colitis are strain dependent. Inflamm Bowel Dis 14, 780793.CrossRefGoogle ScholarPubMed
83Larmonier, CB, Uno, JK, Lee, KM, et al. (2008) Limited effects of dietary curcumin on Th-1 driven colitis in IL-10 deficient mice suggest an IL-10-dependent mechanism of protection. Am J Physiol 295, G1079G1091.Google ScholarPubMed
84Rao, CV, Simi, B & Reddy, BS (1993) Inhibition by dietary curcumin of azoxymethane-induced ornithine decarboxylase, tyrosine protein kinase, arachidonic acid metabolism and aberrant crypt foci formation in the rat colon. Carcinogenesis 14, 22192225.CrossRefGoogle ScholarPubMed
85Kim, JM, Araki, S, Kim, DJ, et al. (1998) Chemopreventive effects of carotenoids and curcumins on mouse colon carcinogenesis after 1,2-dimethylhydrazine initiation. Carcinogenesis 19, 8185.CrossRefGoogle ScholarPubMed
86Ushida, J, Sugie, S, Kawabata, K, et al. (2000) Chemopreventive effect of curcumin on N-nitrosomethylbenzylamine-induced esophageal carcinogenesis in rats. Jpn J Cancer Res 91, 893898.CrossRefGoogle ScholarPubMed
87Ikezaki, S, Nishikawa, A, Furukawa, F, et al. (2001) Chemopreventive effects of curcumin on glandular stomach carcinogenesis induced by N-methyl-N′-nitro-N-nitrosoguanidine and sodium chloride in rats. Anticancer Res 21, 34073411.Google ScholarPubMed
88Chuang, SE, Cheng, AL, Lin, JK, et al. (2000) Inhibition by curcumin of diethylnitrosamine-induced hepatic hyperplasia, inflammation, cellular gene products and cell-cycle-related proteins in rats. Food Chem Toxicol 38, 991995.CrossRefGoogle ScholarPubMed
89Azuine, MA & Bhide, SV (1992) Protective single/combined treatment with betel leaf and turmeric against methyl (acetoxymethyl) nitrosamine-induced hamster oral carcinogenesis. Int J Cancer 51, 412415.CrossRefGoogle ScholarPubMed
90Huang, MT, Lou, YR, Xie, JG, et al. (1998) Effect of dietary curcumin and dibenzoylmethane on formation of 7,12-dimethylbenz[a]anthracene-induced mammary tumors and lymphomas/leukemias in Sencar mice. Carcinogenesis 19, 16971700.CrossRefGoogle Scholar
91Hecht, SS, Kenney, PM, Wang, M, et al. (1999) Evaluation of butylated hydroxyanisole, myo-inositol, curcumin, esculetin, resveratrol and lycopene as inhibitors of benzo[a]pyrene plus 4-(methylnitrosamino)-1-(3-pyridyl)-1-butanone-induced lung tumorigenesis in A/J mice. Cancer Lett 137, 123130.CrossRefGoogle Scholar
92Okazaki, Y, Iqbal, M & Okada, S (2005) Suppressive effects of dietary curcumin on the increased activity of renal ornithine decarboxylase in mice treated with a renal carcinogen, ferric nitrilotriacetate. Biochim Biophys Acta 1740, 357366.CrossRefGoogle ScholarPubMed
93Sindhwani, P, Hampton, JA, Baig, MM, et al. (2001) Curcumin prevents intravesical tumor implantation of the MBT-2 tumor cell line in C3H mice. J Urol 166, 14981501.CrossRefGoogle ScholarPubMed
94Limtrakul, P, Lipigorngoson, S, Namwong, O, et al. (1997) Inhibitory effect of dietary curcumin on skin carcinogenesis in mice. Cancer Lett 116, 197203.CrossRefGoogle ScholarPubMed
95Kuttan, R, Bhanumathy, P, Nirmala, K, et al. (1985) Potential anticancer activity of turmeric (Curcuma longa). Cancer Lett 29, 197202.CrossRefGoogle ScholarPubMed
96Odot, J, Albert, P, Carlier, A, et al. (2004) In vitro and in vivo anti-tumoral effect of curcumin against melanoma cells. Int J Cancer 111, 381387.CrossRefGoogle ScholarPubMed
97LoTempio, MM, Veena, MS, Steele, HL, et al. (2005) Curcumin suppresses growth of head and neck squamous cell carcinoma. Clin Cancer Res 11, (19 Pt 1), 69947002.CrossRefGoogle ScholarPubMed
98Dorai, T, Cao, YC, Dorai, B, et al. (2001) Therapeutic potential of curcumin in human prostate cancer. III. Curcumin inhibits proliferation, induces apoptosis, and inhibits angiogenesis of LNCaP prostate cancer cells in vivo. Prostate 47, 293303.CrossRefGoogle ScholarPubMed
99Kunnumakkara, AB, Guha, S, Krishnan, S, et al. (2007) Curcumin potentiates antitumor activity of gemcitabine in an orthotopic model of pancreatic cancer through suppression of proliferation, angiogenesis, and inhibition of nuclear factor-kappaB-regulated gene products. Cancer Res 67, 38533861.CrossRefGoogle Scholar
100Li, L, Braiteh, FS & Kurzrock, R (2005) Liposome-encapsulated curcumin: in vitro and in vivo effects on proliferation, apoptosis, signaling, and angiogenesis. Cancer 104, 13221331.CrossRefGoogle ScholarPubMed
101Li, L, Ahmed, B, Mehta, K, et al. (2007) Liposomal curcumin with and without oxaliplatin: effects on cell growth, apoptosis, and angiogenesis in colorectal cancer. Mol Cancer Ther 6, 12761282.CrossRefGoogle ScholarPubMed
102Ohashi, Y, Tsuchiya, Y, Koizumi, K, et al. (2003) Prevention of intrahepatic metastasis by curcumin in an orthotopic implantation model. Oncology 65, 250258.CrossRefGoogle Scholar
103Aggarwal, BB, Shishodia, S, Takada, Y, et al. (2005) Curcumin suppresses the paclitaxel-induced nuclear factor-kappaB pathway in breast cancer cells and inhibits lung metastasis of human breast cancer in nude mice. Clin Cancer Res 11, 74907498.CrossRefGoogle ScholarPubMed
104Bachmeier, B, Nerlich, AG, Iancu, CM, et al. (2007) The chemopreventive polyphenol curcumin prevents hematogenous breast cancer metastases in immunodeficient mice. Cell Physiol Biochem 19, 137152.CrossRefGoogle ScholarPubMed
105Lin, YG, Kunnumakkara, AB, Nair, A, et al. (2007) Curcumin inhibits tumor growth and angiogenesis in ovarian carcinoma by targeting the nuclear factor-kappaB pathway. Clin Cancer Res 13, 34233430.CrossRefGoogle ScholarPubMed
106Kobelt, G (2006) Health economic issues in rheumatoid arthritis. Scand J Rheumatol 35, 415425.CrossRefGoogle ScholarPubMed
107Deodhar, SD, Sethi, R & Srimal, RC (1980) Preliminary study on antirheumatic activity of curcumin (diferuloyl methane). Indian J Med Res 71, 632634.Google Scholar
108Heng, MC, Song, MK, Harker, J, et al. (2000) Drug-induced suppression of phosphorylase kinase activity correlates with resolution of psoriasis as assessed by clinical, histological and immunohistochemical parameters. Br J Dermatol 143, 937949.CrossRefGoogle ScholarPubMed
109Satoskar, RR, Shah, SJ & Shenoy, SG (1986) Evaluation of anti-inflammatory property of curcumin (diferuloyl methane) in patients with postoperative inflammation. Int J Clin Pharmacol Ther Toxicol 24, 651654.Google ScholarPubMed
110Lal, B, Kapoor, AK, Agrawal, PK, et al. (2000) Role of curcumin in idiopathic inflammatory orbital pseudotumours. Phytother Res 14, 443447.3.0.CO;2-V>CrossRefGoogle ScholarPubMed
111Lal, B, Kapoor, AK, Asthana, OP, et al. (1999) Efficacy of curcumin in the management of chronic anterior uveitis. Phytother Res 13, 318322.3.0.CO;2-7>CrossRefGoogle ScholarPubMed
112Bundy, R, Walker, AF, Middleton, RW, et al. (2004) Turmeric extract may improve irritable bowel syndrome symptomology in otherwise healthy adults: a pilot study. J Altern Complement Med 10, 10151018.CrossRefGoogle ScholarPubMed
113Durgaprasad, S, Pai, CG & Vasanthkumar, (2005) A pilot study of the antioxidant effect of curcumin in tropical pancreatitis. Indian J Med Res 122, 315318.Google ScholarPubMed
114Niederau, C & Gopfert, E (1999) The effect of chelidonium- and turmeric root extract on upper abdominal pain due to functional disorders of the biliary system. Results from a placebo-controlled double-blind study. Med Klin (Munich) 94, 425430.CrossRefGoogle ScholarPubMed
115Rasyid, A & Lelo, A (1999) The effect of curcumin and placebo on human gall-bladder function: an ultrasound study. Aliment Pharmacol Ther 13, 245249.CrossRefGoogle ScholarPubMed
116Rasyid, A, Rahman, AR, Jaalam, K, et al. (2002) Effect of different curcumin dosages on human gall bladder. Asia Pac J Clin Nutr 11, 314318.CrossRefGoogle ScholarPubMed
117Prucksunand, C, Indrasukhsri, B, Leethochawalit, M, et al. (2001) Phase II clinical trial on effect of the long turmeric (Curcuma longa Linn) on healing of peptic ulcer. Southeast Asian J Trop Med Public Health 32, 208215.Google Scholar
118Cruz-Correa, M, Shoskes, DA, Sanchez, P, et al. (2006) Combination treatment with curcumin and quercetin of adenomas in familial adenomatous polyposis. Clin Gastroenterol Hepatol 4, 10351038.CrossRefGoogle ScholarPubMed
119Soni, KB & Kuttan, R (1992) Effect of oral curcumin administration on serum peroxides and cholesterol levels in human volunteers. Indian J Physiol Pharmacol 36, 273275.Google ScholarPubMed
120Usharani, P, Mateen, AA, Naidu, MU, et al. (2008) Effect of NCB-02, atorvastatin and placebo on endothelial function, oxidative stress and inflammatory markers in patients with type 2 diabetes mellitus: a randomized, parallel-group, placebo-controlled, 8-week study. Drugs R D 9, 243250.CrossRefGoogle ScholarPubMed
121Shoskes, D, Lapierre, C, Cruz-Correa, M, et al. (2005) Beneficial effects of the bioflavonoids curcumin and quercetin on early function in cadaveric renal transplantation: a randomized placebo controlled trial. Transplantation 80, 15561559.CrossRefGoogle ScholarPubMed
122Holt, PR, Katz, S & Kirshoff, R (2005) Curcumin therapy in inflammatory bowel disease: a pilot study. Dig Dis Sci 50, 21912193.CrossRefGoogle ScholarPubMed
123Hanai, H, Iida, T, Takeuchi, K, et al. (2006) Curcumin maintenance therapy for ulcerative colitis: randomized, multicenter, double-blind, placebo-controlled trial. Clin Gastroenterol Hepatol 4, 15021506.CrossRefGoogle ScholarPubMed
124Sharma, RA, McLelland, HR, Hill, KA, et al. (2001) Pharmacodynamic and pharmacokinetic study of oral Curcuma extract in patients with colorectal cancer. Clin Cancer Res 7, 18941900.Google ScholarPubMed
125Kuttan, R, Sudheeran, PC & Josph, CD (1987) Turmeric and curcumin as topical agents in cancer therapy. Tumori 73, 2931.CrossRefGoogle ScholarPubMed
126Dhillon, N, Aggarwal, BB, Newman, RA, et al. (2008) Phase II trial of curcumin in patients with advanced pancreatic cancer. Clin Cancer Res 14, 44914499.CrossRefGoogle ScholarPubMed
127Baum, L, Lam, CW, Cheung, SK, et al. (2008) Six-month randomized, placebo-controlled, double-blind, pilot clinical trial of curcumin in patients with Alzheimer disease. J Clin Psychopharmacol 28, 110113.CrossRefGoogle ScholarPubMed
128Liang, G, Zhou, H, Wang, Y, et al. (2009) Inhibition of LPS-induced production of inflammatory factors in the macrophages by mono-carbonyl analogues of Curcumin. J Cell Mol Med.CrossRefGoogle ScholarPubMed
129Weber, WM, Hunsaker, LA, Gonzales, AM, et al. (2006) TPA-induced up-regulation of activator protein-1 can be inhibited or enhanced by analogs of the natural product curcumin. Biochem Pharmacol 72, 928940.CrossRefGoogle ScholarPubMed
130Olszanecki, R, Gebska, A & Korbut, R (2007) The role of haem oxygenase-1 in the decrease of endothelial intercellular adhesion molecule-1 expression by curcumin. Basic Clin Pharmacol Toxicol 101, 411415.CrossRefGoogle ScholarPubMed
131Srinivasan, M, Rajendra Prasad, N & Menon, VP (2006) Protective effect of curcumin on gamma-radiation induced DNA damage and lipid peroxidation in cultured human lymphocytes. Mutat Res 611, 96103.CrossRefGoogle ScholarPubMed
132Balogun, E, Hoque, M, Gong, P, et al. (2003) Curcumin activates the haem oxygenase-1 gene via regulation of Nrf2 and the antioxidant-responsive element. Biochem J 371 (Pt 3), 887895.CrossRefGoogle ScholarPubMed
133Liao, YF, Hung, HC, Hour, TC, et al. (2008) Curcumin induces apoptosis through an ornithine decarboxylase-dependent pathway in human promyelocytic leukemia HL-60 cells. Life Sci 82, 367375.CrossRefGoogle ScholarPubMed
134Wang, JB, Qi, LL, Zheng, SD, et al. (2009) Curcumin induces apoptosis through the mitochondria-mediated apoptotic pathway in HT-29 cells. J Zhejiang Univ Sci 10, 93102.CrossRefGoogle ScholarPubMed
135Hussain, AR, Al-Rasheed, M, Manogaran, PS, et al. (2006) Curcumin induces apoptosis via inhibition of PI3′-kinase/AKT pathway in acute T cell leukemias. Apoptosis 11, 245254.CrossRefGoogle ScholarPubMed
136Liu, JY, Lin, SJ & Lin, JK (1993) Inhibitory effects of curcumin on protein kinase C activity induced by 12-O-tetradecanoyl-phorbol-13-acetate in NIH 3T3 cells. Carcinogenesis 14, 857861.CrossRefGoogle ScholarPubMed
137Pendurthi, UR & Rao, LV (2000) Suppression of transcription factor Egr-1 by curcumin. Thromb Res 97, 179189.CrossRefGoogle ScholarPubMed
138Huang, MT, Lou, YR, Ma, W, et al. (1994) Inhibitory effects of dietary curcumin on forestomach, duodenal, and colon carcinogenesis in mice. Cancer Res 54, 58415847.Google ScholarPubMed
139Collett, GP, Robson, CN, Mathers, JC, et al. (2001) Curcumin modifies Apc(min) apoptosis resistance and inhibits 2-amino 1-methyl-6-phenylimidazo[4,5-b]pyridine (PhIP) induced tumour formation in Apc(min) mice. Carcinogenesis 22, 821825.CrossRefGoogle ScholarPubMed
140Rao, CV, Rivenson, A, Simi, B, et al. (1995) Chemoprevention of colon cancer by dietary curcumin. Ann N Y Acad Sci 768, 201204.CrossRefGoogle ScholarPubMed
141Kawamori, T, Lubet, R, Steele, VE, et al. (1999) Chemopreventive effect of curcumin, a naturally occurring anti-inflammatory agent, during the promotion/progression stages of colon cancer. Cancer Res 59, 597601.Google ScholarPubMed
142Pereira, MA, Grubbs, CJ, Barnes, LH, et al. (1996) Effects of the phytochemicals, curcumin and quercetin, upon azoxymethane-induced colon cancer and 7,12-dimethylbenz[a]anthracene-induced mammary cancer in rats. Carcinogenesis 17, 13051311.CrossRefGoogle Scholar
143Kwon, Y, Malik, M & Magnuson, BA (2004) Inhibition of colonic aberrant crypt foci by curcumin in rats is affected by age. Nutr Cancer 48, 3743.CrossRefGoogle ScholarPubMed
144Shpitz, B, Giladi, N, Sagiv, E, et al. (2006) Celecoxib and curcumin additively inhibit the growth of colorectal cancer in a rat model. Digestion 74, 140144.CrossRefGoogle ScholarPubMed
145Volate, SR, Davenport, DM, Muga, SJ, et al. (2005) Modulation of aberrant crypt foci and apoptosis by dietary herbal supplements (quercetin, curcumin, silymarin, ginseng and rutin). Carcinogenesis 26, 14501456.CrossRefGoogle ScholarPubMed
146Kwon, Y & Magnuson, BA (2007) Effect of azoxymethane and curcumin on transcriptional levels of cyclooxygenase-1 and -2 during initiation of colon carcinogenesis. Scand J Gastroenterol 42, 7280.CrossRefGoogle ScholarPubMed
147Devasena, T, Menon, VP & Rajasekharan, KN (2006) Prevention of 1,2-dimethylhydrazine-induced circulatory oxidative stress by bis-1,7-(2-hydroxyphenyl)-hepta-1,6-diene-3,5-dione during colon carcinogenesis. Pharmacol Rep 58, 229235.Google Scholar
148Xu, G, Huang, W, Zhang, WM, et al. (2005) Effects of combined use of curcumin and catechin on cyclooxygenase-2 mRNA expression in dimethylhydrazine-induced rat colon carcinogenesis. Di Yi Jun Yi Da Xue Xue Bao 25, 4852.Google ScholarPubMed
149Kim, SJ & Hellerstein, MK (2007) Pharmacological doses of dietary curcumin increase colon epithelial cell proliferation in vivo in rats. Phytother Res 21, 995998.CrossRefGoogle ScholarPubMed
150Azuine, MA & Bhide, SV (1992) Chemopreventive effect of turmeric against stomach and skin tumors induced by chemical carcinogens in Swiss mice. Nutr Cancer 17, 7783.CrossRefGoogle ScholarPubMed
151Singh, SV, Hu, X, Srivastava, SK, et al. (1998) Mechanism of inhibition of benzo[a]pyrene-induced forestomach cancer in mice by dietary curcumin. Carcinogenesis 19, 13571360.CrossRefGoogle Scholar
152Nagabhushan, M & Bhide, SV (1992) Curcumin as an inhibitor of cancer. J Am Coll Nutr 11, 192198.CrossRefGoogle ScholarPubMed
153Swamy, MV, Citineni, B, Patlolla, JM, et al. (2008) Prevention and treatment of pancreatic cancer by curcumin in combination with omega-3 fatty acids. Nutr Cancer 60, Suppl. 1, 8189.CrossRefGoogle ScholarPubMed
154Tanaka, T, Makita, H, Ohnishi, M, et al. (1994) Chemoprevention of 4-nitroquinoline 1-oxide-induced oral carcinogenesis by dietary curcumin and hesperidin: comparison with the protective effect of beta-carotene. Cancer Res 54, 46534659.Google ScholarPubMed
155Lin, CC, Lu, YP, Lou, YR, et al. (2001) Inhibition by dietary dibenzoylmethane of mammary gland proliferation, formation of DMBA-DNA adducts in mammary glands, and mammary tumorigenesis in Sencar mice. Cancer Lett 168, 125132.CrossRefGoogle ScholarPubMed
156Singletary, K, MacDonald, C, Wallig, M, et al. (1996) Inhibition of 7,12-dimethylbenz[a]anthracene (DMBA)-induced mammary tumorigenesis and DMBA-DNA adduct formation by curcumin. Cancer Lett 103, 137141.CrossRefGoogle Scholar
157Deshpande, SS, Ingle, AD & Maru, GB (1998) Chemopreventive efficacy of curcumin-free aqueous turmeric extract in 7,12-dimethylbenz[a]anthracene-induced rat mammary tumorigenesis. Cancer Lett 123, 3540.CrossRefGoogle Scholar
158Inano, H & Onoda, M (2002) Radioprotective action of curcumin extracted from Curcuma longa LINN: inhibitory effect on formation of urinary 8-hydroxy-2′-deoxyguanosine, tumorigenesis, but not mortality, induced by gamma-ray irradiation. Int J Radiat Oncol Biol Phys 53, 735743.CrossRefGoogle Scholar
159Inano, H, Onoda, M, Inafuku, N, et al. (1999) Chemoprevention by curcumin during the promotion stage of tumorigenesis of mammary gland in rats irradiated with gamma-rays. Carcinogenesis 20, 10111018.CrossRefGoogle ScholarPubMed
160Lin, CC, Ho, CT & Huang, MT (2001) Mechanistic studies on the inhibitory action of dietary dibenzoylmethane, a beta-diketone analogue of curcumin, on 7,12-dimethylbenz[a]anthracene-induced mammary tumorigenesis. Proc Natl Sci Counc Repub China 25, 158165.Google Scholar
161Ishizaki, C, Oguro, T, Yoshida, T, et al. (1996) 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 193, 311317.CrossRefGoogle ScholarPubMed
162Huang, MT, Smart, RC, Wong, CQ, et al. (1988) Inhibitory effect of curcumin, chlorogenic acid, caffeic acid, and ferulic acid on tumor promotion in mouse skin by 12-O-tetradecanoylphorbol-13-acetate. Cancer Res 48, 59415946.Google ScholarPubMed
163Lu, YP, Chang, RL, Huang, MT, et al. (1993) Inhibitory effect of curcumin on 12-O-tetradecanoylphorbol-13-acetate-induced increase in ornithine decarboxylase mRNA in mouse epidermis. Carcinogenesis 14, 293297.CrossRefGoogle ScholarPubMed
164Huang, MT, Ma, W, Lu, YP, et al. (1995) Effects of curcumin, demethoxycurcumin, bisdemethoxycurcumin and tetrahydrocurcumin on 12-O-tetradecanoylphorbol-13-acetate-induced tumor promotion. Carcinogenesis 16, 24932497.CrossRefGoogle ScholarPubMed
165Huang, MT, Ma, W, Yen, P, et al. (1997) Inhibitory effects of topical application of low doses of curcumin on 12-O-tetradecanoylphorbol-13-acetate-induced tumor promotion and oxidized DNA bases in mouse epidermis. Carcinogenesis 18, 8388.CrossRefGoogle ScholarPubMed
166Soudamini, KK & Kuttan, R (1989) Inhibition of chemical carcinogenesis by curcumin. J Ethnopharmacol 27, 227233.CrossRefGoogle ScholarPubMed
167Huang, MT, Deschner, EE, Newmark, HL, et al. (1992) Effect of dietary curcumin and ascorbyl palmitate on azoxymethanol-induced colonic epithelial cell proliferation and focal areas of dysplasia. Cancer Lett 64, 117121.CrossRefGoogle ScholarPubMed
168William, BM, Goodrich, A, Peng, C, et al. (2008) Curcumin inhibits proliferation and induces apoptosis of leukemic cells expressing wild-type or T315I-BCR-ABL and prolongs survival of mice with acute lymphoblastic leukemia. Hematology 13, 333343.CrossRefGoogle ScholarPubMed
169Sung, B, Kunnumakkara, AB, Sethi, G, et al. (2009) Curcumin circumvents chemoresistance in vitro and potentiates the effect of thalidomide and bortezomib against human multiple myeloma in nude mice model. Mol Cancer Ther 8, 959970.CrossRefGoogle ScholarPubMed
170Iqbal, M, Okazaki, Y & Okada, S (2009) Curcumin attenuates oxidative damage in animals treated with a renal carcinogen, ferric nitrilotriacetate (Fe-NTA): implications for cancer prevention. Mol Cell Biochem 324, 157164.CrossRefGoogle ScholarPubMed
171Purkayastha, S, Berliner, A, Fernando, SS, et al. (2009) Curcumin blocks brain tumor formation. Brain Res.CrossRefGoogle ScholarPubMed
172Narayanan, NK, Nargi, D, Randolph, C, et al. (2009) Liposome encapsulation of curcumin and resveratrol in combination reduces prostate cancer incidence in PTEN knockout mice. Int J Cancer 125, 18.CrossRefGoogle ScholarPubMed
173Manoharan, S, Balakrishnan, S, Menon, VP, et al. (2009) Chemopreventive efficacy of curcumin and piperine during 7,12-dimethylbenz[a]anthracene-induced hamster buccal pouch carcinogenesis. Singapore Med J 50, 139146.Google Scholar
Figure 0

Fig. 1 Chemical structure of curcumin (diferuloylmethane), its natural analogues and principal metabolites.

Figure 1

Table 1 Molecular targets of curcumin in cell line studies

Figure 2

Fig. 2 Cellular activities of curcumin and molecular mechanisms of action. ROS, reactive oxygen species; HO-1, haem oxygenase; Bax, BCL2-associated X protein; MAPK, mitogen-activated protein kinase; MMP, matrix metalloproteinase; STAT, signal transducer and activator of transcription; NOS, NO synthase; COX2, cyclo-oxygenase-2. , Activation by curcumin; ↕, effects vary; , suppression by curcumin.

Figure 3

Table 2 Animal models in which curcumin has chemopreventive efficacy

Figure 4

Fig. 3 Clinical effects of curcumin: results from human trials (with references).