Hostname: page-component-7c8c6479df-r7xzm Total loading time: 0 Render date: 2024-03-27T22:55:45.808Z Has data issue: false hasContentIssue false

Examining the potential clinical value of curcumin in the prevention and diagnosis of Alzheimer’s disease

Published online by Cambridge University Press:  14 December 2015

K. G. Goozee
Affiliation:
McCusker KARVIAH Research Centre, Anglican Retirement Villages, Sydney, NSW 2154, Australia School of Medical Sciences, Centre of Excellence for Alzheimer’s Disease Research and Care, Edith Cowan University, Joondalup, WA 6027, Australia School of Psychiatry and Clinical Neurosciences, University of Western Australia, Crawley, WA 6009, Australia
T. M. Shah
Affiliation:
School of Medical Sciences, Centre of Excellence for Alzheimer’s Disease Research and Care, Edith Cowan University, Joondalup, WA 6027, Australia Sir James McCusker Alzheimer’s Disease Research Unit, Hollywood Private Hospital, Nedlands, WA 6009, Australia
H. R. Sohrabi
Affiliation:
School of Medical Sciences, Centre of Excellence for Alzheimer’s Disease Research and Care, Edith Cowan University, Joondalup, WA 6027, Australia Sir James McCusker Alzheimer’s Disease Research Unit, Hollywood Private Hospital, Nedlands, WA 6009, Australia
S. R. Rainey-Smith
Affiliation:
School of Medical Sciences, Centre of Excellence for Alzheimer’s Disease Research and Care, Edith Cowan University, Joondalup, WA 6027, Australia Sir James McCusker Alzheimer’s Disease Research Unit, Hollywood Private Hospital, Nedlands, WA 6009, Australia
B. Brown
Affiliation:
School of Medical Sciences, Centre of Excellence for Alzheimer’s Disease Research and Care, Edith Cowan University, Joondalup, WA 6027, Australia Sir James McCusker Alzheimer’s Disease Research Unit, Hollywood Private Hospital, Nedlands, WA 6009, Australia
G. Verdile
Affiliation:
School of Medical Sciences, Centre of Excellence for Alzheimer’s Disease Research and Care, Edith Cowan University, Joondalup, WA 6027, Australia Sir James McCusker Alzheimer’s Disease Research Unit, Hollywood Private Hospital, Nedlands, WA 6009, Australia School of Psychiatry and Clinical Neurosciences, University of Western Australia, Crawley, WA 6009, Australia School of Biomedical Sciences, Curtin Health Innovation Research Institute Biosciences, Curtin University, Bentley, WA 6102, Australia
R. N. Martins*
Affiliation:
McCusker KARVIAH Research Centre, Anglican Retirement Villages, Sydney, NSW 2154, Australia School of Medical Sciences, Centre of Excellence for Alzheimer’s Disease Research and Care, Edith Cowan University, Joondalup, WA 6027, Australia Sir James McCusker Alzheimer’s Disease Research Unit, Hollywood Private Hospital, Nedlands, WA 6009, Australia School of Psychiatry and Clinical Neurosciences, University of Western Australia, Crawley, WA 6009, Australia
*
*Corresponding author: Professor R. N. Martins, fax +61 8 9347 4299, email ralph.n.martins@gmail.com
Rights & Permissions [Opens in a new window]

Abstract

Curcumin derived from turmeric is well documented for its anti-carcinogenic, antioxidant and anti-inflammatory properties. Recent studies show that curcumin also possesses neuroprotective and cognitive-enhancing properties that may help delay or prevent neurodegenerative diseases, including Alzheimer’s disease (AD). Currently, clinical diagnosis of AD is onerous, and it is primarily based on the exclusion of other causes of dementia. In addition, phase III clinical trials of potential treatments have mostly failed, leaving disease-modifying interventions elusive. AD can be characterised neuropathologically by the deposition of extracellular β amyloid (Aβ) plaques and intracellular accumulation of tau-containing neurofibrillary tangles. Disruptions in Aβ metabolism/clearance contribute to AD pathogenesis. In vitro studies have shown that Aβ metabolism is altered by curcumin, and animal studies report that curcumin may influence brain function and the development of dementia, because of its antioxidant and anti-inflammatory properties, as well as its ability to influence Aβ metabolism. However, clinical studies of curcumin have revealed limited effects to date, most likely because of curcumin’s relatively low solubility and bioavailability, and because of selection of cohorts with diagnosed AD, in whom there is already major neuropathology. However, the fresh approach of targeting early AD pathology (by treating healthy, pre-clinical and mild cognitive impairment-stage cohorts) combined with new curcumin formulations that increase bioavailability is renewing optimism concerning curcumin-based therapy. The aim of this paper is to review the current evidence supporting an association between curcumin and modulation of AD pathology, including in vitro and in vivo studies. We also review the use of curcumin in emerging retinal imaging technology, as a fluorochrome for AD diagnostics.

Type
Full Papers
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
Copyright © The Authors 2015

With the ageing of many populations worldwide, it is predicted that over the next few decades there will be a marked increase in the number of people with dementia. Current estimations show that 35·6 million people worldwide have dementia, which is predicted to more than triple to 115 million by 2050( Reference Prince and Jackson 1 ). Of all the dementia sub-types, Alzheimer’s disease (AD) is the most common. AD is a neurodegenerative disease, which is characterised clinically by the progressive loss of memory and cognitive functioning. Major pathological features of an AD brain include the accumulation of extracellular plaques and fibrils, intracellular neurofibrillary tangles (NFT), as well as chronic inflammation and widespread synaptic and neuronal loss, leading to brain atrophy and dysfunction. The deposition of amyloid plaques is suggested as a defining feature of the AD brain, as NFT are featured in other neurodegenerative diseases( Reference Lee, Goedert and Trojanowski 2 , Reference Frost, Kanagasingam and Macaulay 3 ) (although plaques have also been reported in cases of non-AD dementias). Nevertheless, hyper-phosphorylated tau protein, the major component of NFT, may have a critical role in the progression of AD, as it acts together with the major protein component of amyloid plaques, β amyloid (Aβ peptides), driving neurodegeneration( Reference Ittner and Götz 4 , Reference Ittner, Ke and Delerue 5 ). The Aβ peptide is generated from its parent molecule, amyloid precursor protein (APP), via sequential proteolytic processing by the enzymes β-APP-cleaving enzyme-1 (BACE1) and γ-secretase( Reference Krishnaswamy, Verdile and Groth 6 ), to generate multiple Aβ forms of varying amino acid lengths. Aβ peptides aggregate readily into oligomers and fibrils, and small oligomers of the longer, more easily aggregating 42-amino-acid form (Aβ1-42) are considered to be the most neurotoxic Aβ species in the AD brain. Amyloid deposition is thought to occur early in the disease process( Reference Villemagne, Burnham and Bourgeat 7 ), and the accumulation of small Aβ aggregates (‘oligomers’) is thought to have a critical role in early pathogenic events that include tau hyperphosphorylation and accumulation, oxidative stress and inflammatory processes that lead to neurodegeneration in the AD brain( Reference Ittner and Götz 4 , Reference Walsh and Teplow 8 , Reference O’Malley, Oktaviani and Zhang 9 ).

With no current effective disease-modifying treatments available, finding pharmacological/non-pharmacological strategies to halt or slow disease progression is of significant importance. The failure of potential pharmaceuticals in human clinical trials has highlighted the need for research into early diagnosis of AD. This is because of the considerable synaptic loss, neuronal loss and brain shrinkage already present by the time AD clinical symptoms emerge, with treatments aimed at slowing the progress of the disease more likely to be effective before onset of symptoms, preferably at the earliest pre-clinical stage. The continuing lack of effective pharmaceutical drugs has also prompted the evaluation of alternative therapeutics, such as nutraceuticals. Curcumin is one example where, because of its properties as an anti-inflammatory, antioxidant, Aβ-lowering agent and Aβ aggregation inhibitor, it shows potential as a therapeutic for AD. In addition, because of its ability to fluoresce and bind Aβ, curcumin has potential as an imaging agent for diagnostics. This review outlines in vitro, in vivo and human studies that have evaluated the therapeutic potential of curcumin in AD, and it discusses recent research that has assessed curcumin as a diagnostic tool through its use in emerging retinal imaging technologies. All human studies identified in this review met current National Institute of Health and the Alzheimer’s Association diagnostic guidelines( Reference McKhann, Knopman and Chertkow 10 ).

Beneficial properties of curcumin – historical perspective

Curcumin is extracted from turmeric, a spice that is derived from the rhizomes of Curcuma Longa and which belongs to the Zingiberaceae (ginger) family. Turmeric is a perennial herb, native to the monsoon forests of south-east Asia, and it is commonly used in Indian, Asian and Middle Eastern foods. In addition to being used as a culinary spice, turmeric (Sanskrit Haridra, meaning that which is yellow) has been a frequently prescribed herbal medicine. Reputed for its blood-purifying abilities( Reference Majeed, Badmaev and Murrary 11 Reference Goel, Kunnumakkara and Aggarwal 13 ), Ayurveda medicine and traditional Persian and Chinese medicine have prescribed curcumin for centuries for its body-cleansing properties, as well as for pain associated with inflammation of the skin and muscles. Curcumin has also been prescribed for asthma, bronchial hyperreactivity, allergy, anorexia, coryza, cough, sinusitis and hepatic disease( Reference Ammon and Wahl 14 ).

Only 3–5 % of turmeric comprises the yellow-pigmented chemically active curcuminoids, being curcumin (diferuloylmethane), demethoxycurcumin (DMC) and bisdemethoxycurcumin (BDMC)( Reference Begum, Jones and Lim 15 ). Curcumin, considered the most therapeutic of the three curcuminoids, was first isolated in 1815 by Vogel and Pelletier( Reference Zhou, Beevers and Huang 16 ), although its chemical structure was not confirmed until almost a century later( Reference Milobedeska, Kostanecki and Lampe 17 ).

The twenty-first century has witnessed renewed interest in curcumin’s reputed therapeutic effects, which has resulted in considerable scientific enquiry and review( Reference Goel, Kunnumakkara and Aggarwal 13 , Reference Zhou, Beevers and Huang 16 , Reference Hamaguchi, Ono and Yamada 18 Reference Basnet and Skalko-Basnet 21 ). Cell studies( Reference Pan, Lin-Shiau and Lin 22 Reference Zhang, Browne and Child 24 ) report curcumin to possess powerful anti-inflammatory properties, whereas further research in a variety of inflammatory conditions demonstrates its potential. For example, animal and cell culture studies show that curcumin reduces inflammation in arthritis( Reference Joe, Rao and Lokesh 25 , Reference Jackson, Higo and Hunter 26 ); human cell line studies show that curcumin is effective in the management of irritable bowel syndrome( Reference Hanai and Sugimoto 27 ) and in human clinical trials for psoriasis and other skin disorders( Reference Kurd, Smith and VanVoorhees 28 , Reference Thangapazham, Sharma and Maheshwari 29 ). Anti-proliferative and anti-angiogenic influences of curcumin have also been demonstrated, and its therapeutic benefits are shown in human cancer cell and tissue culture, including prostate( Reference Tsui, Feng and Lin 30 ), breast( Reference Liu and Chen 31 ), pancreatic( Reference Friedman, Lin and Ball 32 ) and bowel cancer( Reference Lim, Lee and Huang 33 ), as well as head and neck squamous cell carcinoma( Reference Jackson-Bernitsas, Ichikawa and Takada 34 ).

Curcumin is also considered a powerful antioxidant, reported to be several times more potent than vitamin E as a free-radical scavenger( Reference Zhao, Li and He 35 ). Curcumin’s anti-inflammatory and antioxidant properties have more recently been investigated with respect to AD, as it is now well established that oxidative stress( Reference Martins, Harper and Stokes 36 ) and chronic inflammation are central in the early pathogenic stages of AD( Reference Hardy and Selkoe 37 ). However, in addition to curcumin altering AD development through anti-inflammatory and antioxidant properties, curcumin’s ability to bind to Aβ, influence deposition and aggregation, while possibly also modulating tau processing, has attracted considerable interest in AD research laboratories( Reference Huang and Jiang 38 Reference Mutsuga, Chambers and Uchida 41 ).

Extracellular Aβ plaques and intraneuronal hyper-phosphorylated tau are recognised as hallmark neuropathological features of AD( Reference Kosik, Joachim and Selkoe 42 Reference Masters, Multhaup and Simms 44 ) in addition to oxidative stress and inflammation, and it is believed that abnormal Aβ metabolism, resulting in high levels of toxic Aβ oligomers, combined with oxidative stress and inflammation form an AD pathogenic cycle of neurodegeneration. While the initiating step of this neurodegeneration remains to be elucidated, these changes are thought to begin decades before clinical diagnosis; in fact, the accumulation of Aβ has been shown in radiological imaging to start 20 years or more before the first clinical signs of AD( Reference Villemagne and Rowe 45 ). Aβ accumulation is reported to be associated with impaired synaptic function( Reference Haass and Selkoe 46 ), reduced neurite outgrowth( Reference Manczak, Mao and Calkins 47 ), cerebral atrophy( Reference Cash, Liang and Ryan 48 , Reference Chetelat, Villemagne and Villain 49 ) and reduced cognitive performance, particularly when deposited within the temporal region( Reference Chetelat, Villemagne and Pike 50 ). Synaptic/neuronal loss and NFT load have been shown to correlate positively with cerebral atrophy and cognitive decline, whereas cerebral Aβ load also correlates with cognitive decline, although to a lesser extent( Reference Villemagne, Burnham and Bourgeat 7 ). However, there is also a large body of evidence suggesting that small oligomers of Aβ are particularly toxic to neurons, causing membrane damage, Ca2+ leakage, oxidative damage, disruptions to insulin signalling pathways and synaptic function, as well as mitochondrial damage( Reference Reddy, Tripathi and Troung 51 Reference Zhao, Luo and Jang 53 ). As mentioned above, Aβ-induced changes are believed to occur early in the disease process, and the findings indicate that interventions that can interrupt the production of Aβ or Aβ oligomers, or facilitate their removal from the central nervous system, are highly desirable. Modelled projections suggest that delaying the onset of dementia by even 1 year may reduce the worldwide burden of cases in people over 60 years by as much as approximately 10 %( Reference Johnson, Brookmeyer and Ziegler-Graham 54 ), whereas the introduction of an intervention that delays the onset of dementia by 5 years could reduce the incidence by almost half( Reference Brookmeyer, Gray and Kawas 55 , Reference Vickland, Morris and Draper 56 ). Therefore, early pre-clinical prevention therapy, which could influence the accumulation or clearance of cerebral Aβ and tau pathology, and/or reduce oxidative stress and chronic inflammation, thus slowing or reversing these pathological changes, would be highly significant for the reduction of AD prevalence.

Potentially neuroprotective properties of curcumin: animal studies and in vitro anti-β amyloid activity of curcumin prevents β amyloid aggregation

The Aβ peptide aggregates readily, first into small aggregates of Aβ known as Aβ oligomers and then these oligomers aggregate further to form fibrils, larger fibrils and ultimately plaques of Aβ. Although plaques and large fibrils are the easiest to detect immunohistochemically, these are considered to be relatively inert: as mentioned above, there is now considerable evidence that small Aβ oligomers are the main neurotoxic species( Reference Bieschke, Herbst and Wiglenda 57 ). Therefore, it is interesting that substantial data from in vitro studies indicate that curcumin can bind to Aβ and influence its aggregation. For example, curcumin has been shown to inhibit fibril formation and extension, as well as to destabilise pre-formed fibrils in a dose-dependent manner, effective at concentrations about 0·1–1·0 µm ( Reference Ono, Hasegawa and Naiki 58 ). Later studies have similarly shown that curcumin can inhibit the formation of small Aβ aggregates (Aβ oligomers) in a dose-dependent manner( Reference Reinke and Gestwicki 59 , Reference Yang, Lim and Begum 60 ).

Studies have investigated how curcumin influences Aβ aggregation, and different theories have emerged – for example, one theory involves curcumin binding to metal ions. Biometals such as Cu (Cu(II)) and Zn (Zn(II)) are found in abundance in the brain, particularly at synapses. Dysregulation of metal homeostasis can lead to the binding of these particular metal ions to Aβ, and many studies have shown that this binding accelerates Aβ aggregation. In fact, elevated levels of certain metal ions have been associated with AD( Reference Ghalebani, Wahlstrom and Danielsson 61 ), and considerable research has been undertaken to understand the normal roles of these ions in the brain, as well as the roles the ions may have in disease pathogenesis, particularly the roles of Cu(II) and Zn(II) on Aβ aggregation( Reference Faller and Hureau 62 ). Some recent studies have suggested that curcumin complexes with Cu(II) and/or Zn(II) and that this inhibits the transition from less structured oligomer to β-sheet-rich Aβ protofibrils, which in turn act as seeding factors for further Aβ fibrillisation( Reference Banerjee 63 ). Recent studies looking at the effect of curcumin and curcumin derivatives on metal-induced Aβ aggregation have shown that Gd-linked curcumin (Gd-Cur, a potential Aβ imaging agent), compared with curcumin and Cur-S, a water-soluble form of curcumin, could modulate Cu-induced Aβ aggregation to a greater extent( Reference Kochi, Lee and Vithanarachchi 64 ), supporting the concept of therapeutic and diagnostic uses for the Gd-Cur compounds.

Other theories do not involve metal ions; instead, they suggest that curcumin’s ability to bind Aβ and inhibit its aggregation is because of curcumin’s three structural features: a hydroxyl substitution on the aromatic end group, a rigid linker region between 8 and 16 Å in length and a second terminal phenyl group( Reference Reinke and Gestwicki 59 ). More recent studies using atomic force microscopy have found that curcumin (and another small-molecule inhibitor resveratrol) binds to the N terminus (residues 5–20) of Aβ1-42 monomers and prevents oligomers of 1–2 nm in size from becoming larger 3–5 nm oligomers( Reference Fu, Aucoin and Ahmed 65 ). Yet another recent study has used NMR spectroscopy to investigate the structural modifications of Aβ1-42 aggregates induced by curcumin, and found that curcumin induces major structural changes in the Asp-23–Lys-28 salt bridge region and near the Aβ1-42 C terminus( Reference Mithu, Sarkar and Bhowmik 66 ). The study also used electron microscopy to show that the Aβ1-42 fibrils are disrupted by curcumin. Interestingly, in a Drosophila AD model, curcumin-fed flies showed accelerated conversion of pre-fibrillar to fibrillar Aβ, thereby reducing the neurotoxicity of pre-fibrillar Aβ ( Reference Caesar, Jonson and Nilsson 67 ). Overall, curcumin effects are not limited to modulation of Aβ aggregation, and further studies are needed to determine which effect(s) are the most relevant in promoting brain health in pathological cognitive decline.

Curcumin influences β amyloid production

In vitro studies have shown that Aβ production is influenced by curcumin, as curcumin has been found to inhibit the production of Aβ peptides by altering APP trafficking through the secretory pathway( Reference Zhang, Browne and Child 24 ). Zhang et al. treated mouse primary cortical neurons with different concentrations of curcumin (1–20 μm) for 24 h and found that both Aβ1-40 and Aβ1-42 levels significantly decreased compared with controls. It was suggested that curcumin could stabilise an immature form of APP and reduce the amount reaching the cell surface, thus being available for endocytosis – a process necessary for Aβ production. In an APP-transfected human embryonic kidney cell culture model (SwAPP HEK293), BDMC was shown to reduce BACE1 messenger RNA (mRNA) and protein levels, whereas DMC only affected BACE1 mRNA expression( Reference Liu, Li and Qiu 68 ). Furthermore, in other studies using a neuronal cell line (pheochromocytoma cells – the PC12 cell line) 3–30 µm curcumin suppressed Aβ-induced BACE1 up-regulation( Reference Shimmyo, Kihara and Akaike 69 , Reference Li, Zhang and Si 70 ). Most recently, in studies of an AD Drosophila model, it was found that the curcumin component BDMC was the most effective at rescuing the flies from the morphological and behavioural defects caused by the overexpression of APP and BACE1 ( Reference Wang, Kim and Lee 71 ), possibly via inhibition of the BACE1 enzyme. Recognising curcumin’s ability to reduce Aβ production, by reducing BACE1 mRNA and its corresponding protein( Reference Sathya, Premkumar and Karthick 72 ), curcumin has been used as a potent positive control in the analysis of other compounds/drugs that target not only BACE1 but also metal chelation, Aβ aggregation and oxidation( Reference Jiaranaikulwanitch, Govitrapong and Fokin 73 ). In support of curcumin’s metal chelation properties, curcumin was shown to prevent the up-regulation of APP and BACE1 induced by supraphysiological levels of the metal ions Cu(II) and Mn( Reference Lin, Chen and Li 74 ).

Curcumin can inhibit β amyloid-induced toxicity

Previous studies support the notion that curcumin can reduce Aβ-induced toxicity. A study by Park et al.( Reference Park, Kim and Cho 75 ) used PC12 cell cultures pre-treated with 10 μg/ml curcumin before Aβ exposure. Compared with controls, pre-treated cells had a significant reduction in oxidative stress, as well as lower Ca influx, resulting in protection against DNA damage and cell death. Curcumin (1–30 μm) has also been shown to attenuate the production of Aβ-induced radical O2 species in neuronal cell cultures, and 20 μm curcumin has been shown to prevent structural changes in Aβ towards β-sheet-rich secondary structures( Reference Shimmyo, Kihara and Akaike 69 ). Furthermore, the protection curcumin provided to PC12 cells and to human umbilical vein endothelial cells against Aβ1-42-induced injury was attributed by Kim et al. ( Reference Kim, Park and Kim 76 ) to antioxidant mechanisms of curcuminoids. More recently, in vitro studies of microglia have shown that curcumin can dampen inflammatory pathways that promote neurodegeneration( Reference Shi, Zheng and Li 77 ). In this study, curcumin dose-dependently improved viability against Aβ-42-induced inflammation, as it abolishes Aβ-42-induced IL-1β, IL-6 and TNF-α production. Curcumin was also shown to reduce ERK1/2 and p38 phosphorylation, which was then shown to reduce cytokine production by the microglia( Reference Shi, Zheng and Li 77 ).

Curcumin’s neuroprotective properties may also be attributed to its role in cell signalling. Cell signalling by the Wnt pathways, via the transcription co-activator β-catenin, controls embryonic development, cellular proliferation and neurogenesis. Disruptions to this pathway have been shown to have a significant role in the pathogenesis of diverse diseases such as cancer, metabolic diseases, osteoporosis, epilepsy, as well as AD. In studies of APP-transfected neuroblastoma cells, curcumin was found to activate the Wnt/β-catenin signalling pathway by inhibiting the activity of glycogen synthase kinase 3β (GSK-3β)( Reference Zhang, Yin and Shi 78 ). GSK-3β is a negative regulator of Wnt, and thus lowering its activity will activate Wnt. However, just as importantly, constitutively active GSK-3β contributes to aberrant tau phosphorylation and NFT formation, which are hallmark pathological changes in AD( Reference Olivia, Vargas and Inestrosa 79 ), and thus curcumin-induced inhibition of GSK-3β may also reduce NFT formation. However, the benefits of curcumin in attenuating tau phosphorylation and accumulation have yet to be investigated thoroughly. Interestingly, Aβ oligomers have also been shown to inactivate Wnt in hippocampal slices, by inducing the Wnt antagonist Dickkopf-1( Reference Purro, Dickins and Salinas 80 ). These studies collectively suggest that curcumin can influence GSK-3β and Wnt/β-catenin signalling, which are both key factors in AD pathogenesis( Reference Wan, Xia and Kalionis 81 ). Furthermore, it has been shown recently that activation of the Wnt/β-catenin signalling pathway inhibits the transcription of BACE1 by binding of T-cell factor-4 to the BACE1 promoter gene, thereby reducing the generation of Aβ ( Reference Parr, Mirzaei and Christian 82 ). Other recent studies using molecular modelling software programs have identified curcumin and rosmarinic acid as promising ligands that mimic the inhibitory action of peptidyl inhibitors of caspase-3( Reference Khan, Akhtar and Sharma 83 ). The relationships between AD-related proteins and pathways discussed above provide further indication of curcumin’s therapeutic potential for the prevention of AD.

Curcumin and the clearance of β amyloid

One anti-AD therapeutic approach involves enhancing the clearance of Aβ from the brain. Several mechanisms have been proposed to assist normal clearance of Aβ from the brain, such as enhancing receptor- or apo-mediated transport across the blood–brain barrier (BBB), efflux of Aβ from the brain being the basis of the peripheral sink hypothesis, targeted immune responses to Aβ, dissolution of amyloid fibrils and microglial activation resulting in phagocytosis of amyloid plaques, as reviewed by Bates et al.( Reference Bates, Verdile and Li 84 ). The significance of the innate immune system in Aβ clearance remains pertinent to intervention and treatment modalities. Although previous attempts to use vaccines to augment the immune response were halted because of the incidence of sterile encephalitis( Reference Foster, Verdile and Bates 85 ), interest in this area remains strong. Macrophage activity and phagocytosis of Aβ has been reported to be deficient in AD, suggesting a contributory factor to Aβ accumulation( Reference Fiala, Lin and Ringman 86 ). Furthermore, a later study that pre-treated the AD macrophages with curcuminoids resulted in increased Aβ uptake in 50 % of the macrophages( Reference Zhang, Fiala and Cashman 87 ).

Curcumin’s ability to reduce oxidative damage and amyloid pathology in AD transgenic mice, demonstrated by Garcia-Alloza et al.( Reference Garcia-Alloza, Borrelli and Rozkalne 40 ), also suggests that curcumin can influence amyloid-induced cytopathology, or macrophage processing of amyloid. Garcia-Alloza et al. used multi-photon and in vivo imaging to reveal a marked amyloid clearance effect, with 30 % plaque size reduction and slowed plaque development, in animals receiving curcumin for 7 d via intravenous tail injections. Fiala et al. ( Reference Fiala, Lin and Ringman 86 ) examined curcumin’s effect on enhancing phagocytosis of Aβ at a molecular level, and found that curcumin restored the normal Aβ-induced up-regulation of the transcription of β-1,4-mannosyl-glycoprotein 4β-N-acetylglucosaminyltransferase (MGAT3), an enzyme thought to be involved in phagocytosis. Other proteins such as toll-like receptors were also up-regulated. These results indicate that curcumin may correct immune defects in AD patients, suggesting a novel approach to AD immunotherapy( Reference Fiala, Liu and Espinosa-Jeffrey 88 ). In more recent studies by the same group, it was found that 1α,25(OH)2-vitamin D3 (1,25D3) could restore the defective Aβ phagocytosis in AD macrophages, and that a nuclear vitamin D receptor antagonist could block this phagocytosis. All phagocytes seemed to respond to 1,25D3, yet only a subset responded to curcuminoids by up-regulating MGAT3. Nevertheless, in those who did respond, further studies demonstrated that the 1,25D3 bound to a pocket of the vitamin D3 receptor that influences genomic events, and curcuminoids bound to a non-genomic pocket( Reference Masoumi, Goldenson and Ghirmai 89 , Reference Mizwicki, Menegaz and Barrientos-Durán 90 ), produced an additive effect.

Curcumin effects on lipid metabolism

Early research( Reference Soni and Kuttan 91 Reference Sreejayan and Rao 93 ) reported that curcumin had cholesterol-lowering ability, supported by Peschel et al. ( Reference Peschel, Koerting and Nass 94 ) who reported that curcumin has a hypocholesterolaemic effect, based on its effect on hepatic gene expression. Feng et al.( Reference Feng, Ohlsson and Duan 95 ) also found curcumin to lower cholesterol levels through suppression of Niemann Pick C1-like 1 protein, which is responsible for the uptake of cholesterol through vesicular endocytosis within the intestine. Another potential mechanism for the hypocholesterolaemic effect of curcumin was revealed in studies of rats fed a high-fat diet, in which curcumin was found to decrease significantly the serum levels of TAG, total cholesterol and LDL-cholesterol, when compared with a control group: curcumin was found to up-regulate mRNA levels of cholesterol 7α-hydroxylase (CYP7A1), a rate-limiting enzyme in the biosynthesis of bile acid from cholesterol( Reference Kim and Kim 96 ). More recently, treatment of similar high-fat diet-fed rats with curcumin combined with piperine was found to produce similar changes to the serum lipid profiles of the rats and increased HDL levels, resulting in significant up-regulation of the activities and gene expression of apo A-I, lecithin–cholesterol acyltransferase, CYP7A1 and the LDL receptor( Reference Tu, Sun and Zeng 97 ). As hypercholesterolaemia continues to be considered a likely contributor to AD risk( Reference Zambón, Quintana and Mata 98 , Reference Refolo, Pappola and Malester 99 ), the use of curcumin if proven to lower cholesterol could represent another approach, adding to the armoury for AD risk reduction.

Curcumin and telomerase

Xiao et al.( Reference Xiao, Zhang and Lin 100 ), investigating the role of telomerase (a ribonuclear protein complex that synthesises and elongates telomeric DNA) in the neuroprotective efficacy of curcumin, found curcumin to be protective against Aβ-induced oxidative stress and cell toxicity. This neuroprotection was lost when telomerase was inhibited by telomerase RT small interfering RNA, indicating that the neuroprotection provided by curcumin was dependent on the presence of telomerase.

Focusing on findings in animal studies

Several in vivo studies have found that Aβ deposition and plaque burden in AD-model transgenic mice is decreased following treatment with curcumin( Reference Garcia-Alloza, Borrelli and Rozkalne 40 , Reference Yang, Lim and Begum 60 , Reference Lim, Chu and Yang 101 , Reference Wang, Thomas and Zhong 102 ). Curcumin has also been found to inhibit Aβ-induced tau phosphorylation( Reference Ma, Yang and Rosario 103 ), to reduce microglial activation, indicating a reduction in inflammation( Reference Wang, Thomas and Zhong 102 , Reference Frautschy, Hu and Kim 104 ), and reduced oxidative damage( Reference Frautschy, Hu and Kim 104 ). Other studies of transgenic mouse models of AD have shown that curcumin can reduce genomic instability events( Reference Fenech and Thomas 105 ).

In the study by Lim et al. ( Reference Lim, Chu and Yang 101 ), AD-model Tg2576 mice aged 10 months old were fed a curcumin diet (160 parts per million (ppm)) for 6 months. The results showed that the curcumin diet significantly lowered the levels of oxidised proteins, the inflammatory cytokine, IL-1β, the astrocyte marker glial fibrillary acidic protein (GFAP), soluble and insoluble Aβ and also plaque burden. The study found that the reduction in GFAP was localised, such that increased activity was shown in areas around plaques, demonstrating a stimulatory effect of curcumin on the phagocytosis of plaques by microglia. Frautschy et al. ( Reference Frautschy, Hu and Kim 104 ), using Sprague–Dawley rats infused with Aβ40 and Aβ42 to induce neurodegeneration and Aβ deposits, found that dietary curcumin (2000 ppm (5·43 μmol/g)) suppressed Aβ-induced oxidative damage and memory impairment, and increased microglial labelling within areas adjacent to Aβ deposits. They also found that curcumin reversed changes in synaptophysin and post-synaptic density 95 (PSD-95) levels, associated with brain plasticity, as well as improved rat performance in length and latency within the water maze test( Reference Frautschy, Hu and Kim 104 ). In similar studies of aged Tg2576 AD-model mice by Yang et al. ( Reference Yang, Lim and Begum 60 ), it was demonstrated that curcumin injected peripherally (via the carotid artery) can cross the BBB and bind to amyloid plaques and inhibit the formation of Aβ oligomers and fibrils( Reference Yang, Lim and Begum 60 ). Later, Begum et al.( Reference Begum, Jones and Lim 15 ) showed similar results and suggested that the dienone bridge present in the chemical structure of curcumin is necessary for this reduction in plaque deposition and the lower protein oxidation observed in the curcumin-treated Tg2576 mice.

More recently, Belviranli et al.( Reference Belviranli, Okudan and Atalik 106 ) showed that aged female rats supplemented with curcumin for 12 d demonstrated improved spatial memory (Morris water maze test), and their brains exhibited reduced oxidative damage. In other studies using an Aβ-infused male Sprague–Dawley rat model of AD, the effects of combined curcuminoids, as well as the individual curcumin constituents, were examined in relation to genes related to synaptic plasticity. The genes that were investigated included actin, Ca/calmodulin-dependent protein kinase type IV, PSD-95 and synaptophysin, and significant effects were noted; for example, a significant increase in synaptophysin expression was found following treatment of the hippocampus with curcuminoids, and both DMC and curcumin were found to increase PSD-95 expression several-fold( Reference Ahmed, Enam and Gilani 107 ), demonstrating results similar to the earlier rat study carried out by Frautschy et al.( Reference Frautschy, Hu and Kim 104 ).

Curcumin and neurogenesis

Curcumin has also been found to stimulate proliferation of embryonic neural progenitor cells and neurogenesis in the adult hippocampus, demonstrating other potentially beneficial effects on neuroplasticity( Reference Kim, Son and Park 108 ). In this study, intraperitoneally administered curcumin activated extracellular signal-regulated kinases (ERK) and p38 kinases, which function in cellular signal transduction pathways that are known to be involved in the regulation of neuronal plasticity and stress responses. More recently, a hybrid compound of curcumin and melatonin (5-(4-hydroxy-phenyl)-3-oxo-pentanoic acid (2-(5-methoxy-1H-indol-3-yl)-ethyl)-amide), known as Z-CM-I-1, developed with the aim of improving neuroprotective properties and BBB permeability, was found to reduce Aβ accumulation in the hippocampus and cortex of APP/PS1 transgenic mice, and to increase the expression of the synaptic markers synaptophysin and PSD-95, following oral administration at a dose of 50 mg/kg for 3 months( Reference Gerenu, Liu and Chojnacki 109 ), encouraging further development of this hybrid compound. APP/PS1 mice were also used in another recent study, which tested the effect of 6 months’ dietary supplementation with the curcumin derivative 1,7-bis(4'-hydroxy-3'-trifluoromethoxyphenyl)-4-methoxycarbonylethyl-1,6-heptadiene-3,5-dione (FMeC1). It was found that FMeC1 supplementation resulted in less Aβ deposits, glial cell activity and cognitive deficits, when compared with untreated, curcumin-treated or FMeC2-treated mice, suggesting that FMeC1 has potential in the treatment of AD( Reference Yanagisawa, Ibrahim and Taguchi 110 ). In rat studies, genetic transcriptional responses along with enhanced hippocampal neurogenesis was seen, following 12 weeks of administration of a curcumin-containing diet, as compared with 6 weeks of this diet, or a control diet( Reference Dong, Zeng and Mitchell 111 ), providing further evidence that curcumin may be beneficial through the promotion of neuronal cell growth.

Curcumin and the blood–brain barrier

Studies of rat and mouse models of ischaemic damage found that curcumin could protect the BBB, most likely because of anti-inflammatory and antioxidant effects( Reference Ghoneim, Abdel-Naim and Khalifa 112 , Reference Thiyagarajan and Sharma 113 ), and later studies of cerebral ischaemia in rats found that a single intravenous injection of curcumin could reduce infarct volume and neurological deficit, possibly because of inhibition of inducible nitric oxide synthase( Reference Jiang, Wang and Sun 114 ). More recent studies suggest that curcumin up-regulates heme oxygenase-1 expression to reduce damage and permeability of the BBB( Reference Wang, Gu and Qin 115 ). Encouragingly, an in vivo rat study, using nanoparticulation of curcumin, was able to demonstrate increased organ, as well as the brain, perfusion by curcumin, by prolonging retention time in the hippocampus by 83 % and in the cerebral cortex by 96 %( Reference Tsai, Chien and Lin 116 ). More recently, another group produced a highly stable nanocurcumin formulation (particule size <80 nm) for use within an in vitro and in an AD transgenic mouse model. The study showed higher concentrations of the nanoformulation in plasma and within the brain compared with non-capuslated curcumin or placebo, while demonstrating significant improvements in working and cued memory function( Reference Cheng, Yeung and Ho 117 ).

Curcumin analogues with similar biological activity to curcumin, yet with improved pharmacokinetic characteristics, including increased bioavailability and water solubility, are continuing to be developed( Reference Anand, Nair and Sung 118 , Reference Zona and La Ferla 119 ), while new synthetic products are also emerging( Reference Zona and La Ferla 119 ). Nanotechnology is particularly promising, whereby nanoencapsulation may be able to achieve a synergistic drug delivery system( Reference Re, Gregori and Masserini 120 ). Encouraging results have already been reported in a study examining curcumin for use in breast cancer chemoprevention, which used injectable polymeric micro particles in mice, achieving both sustained blood curcumin levels for almost a month while maximising its BBB penetration, as well as inhibiting tumour vasculature( Reference Shahani, Swaminathan and Freeman 121 ). Other studies exploring nanoparticle technology( Reference Chiu, Lui and Majeed 122 , Reference Shaikh, Ankola and Beniwal 123 ) have been equally promising. For example, following the intravenous administration of liposomal curcumin, polymeric nanocurcumin and poly-lactic-co-glycolic acid co-polymer-curcumin in rats, all of these compounds were found to cross the BBB( Reference Chiu, Lui and Majeed 122 ), whereas in another study nanoparticles containing curcumin were shown to increase oral bioavailability 9-fold( Reference Shaikh, Ankola and Beniwal 123 ). Further evidence of BBB penetration has been obtained in animal models using labelled curcumin derivatives( Reference Zona and La Ferla 119 ).

Curcumin and acetylcholinesterase

In addition to the effects above, curcumin has also been shown to influence acetylcholinesterase activities( Reference Ahmed and Gilani 124 ), following the same pathway as the commonly prescribed pharmaceuticals, acetylcholinesterase inhibitors, which are considered first-line management in AD( Reference Hamaguchi, Ono and Yamada 18 , Reference Gauthier and Molinuevo 125 ). The administration of acetylcholinesterase inhibitors has been found in certain circumstances to slow the progression of AD symptoms or even reduce AD symptoms for a 12-month period, by inhibiting the breakdown of acetylcholine, a major neurotransmitter, depleted in the AD brain. Using in vitro and ex vivo models of acetylcholinesterase activity, Ahmed et al. ( Reference Ahmed, Enam and Gilani 107 ) investigated whether curcumin had an influence on acetylcholinesterase mechanisms, and recorded dose-dependent inhibitory effects in the frontal cortex and in the hippocampal tissue; curcuminoids also demonstrated significant attenuation of scopolamine-induced amnesia. Furthermore, Ahmed et al.( Reference Ahmed, Enam and Gilani 107 ) examined the influence of curcumin on spatial memory in amyloid-infused AD rat models, reporting increased PSD-95 and synaptophysin expression in the hippocampus and a memory-enhancing effect. In studies of streptozotocin-induced diabetic rats, curcumin has been shown to prevent cholinergic-mediated cortical dysfunctions, which are induced by diabetes( Reference Peeyush, Antony and Sonan 126 ), and in mice treated with okadaic acid to induce memory impairment orally administered curcumin has been found to improve cholinergic function and reduce inflammation, among other beneficial effects( Reference Rajasekar, Dwivedi and Tota 127 ). Furthermore, curcumin has been shown to reverse alcohol-induced cognitive deficits in the adult rat brain, partly by preventing the alcohol-induced activation of acetylcholinesterase; curcumin also reduces signs of neuroinflammation in these rats( Reference Tiwari and Chopra 128 ). Another rat study indicated that curcumin may inhibit acetylcholinesterase activity in As- and Al-induced toxicity models( Reference Orhan 129 ).

Despite all these animal studies, the influence of curcumin on acetylcholinesterase has not yet been investigated in human clinical studies. Furthermore, mechanisms underlying many of the effects described above are still being characterised. However, there is now evidence of curcumin derivatives influencing proteasomal function and Aβ degradation, as described below.

Curcumin, proteasome function and β amyloid degradation

Proteasomal activity and its role in the degradation of most oxidised proteins is linked with the processes of cell ageing; it is also believed that age-related decreases in proteasome activity weakens a cell’s capacity to remove oxidatively modified proteins and therefore encourages the development of diseases( Reference Bulteau, Moreau and Saunois 130 ). Curcumin has been demonstrated to have a stimulatory effect on proteasomal activity, causing a 46 % increase in activity at doses of 1 µm in vitro, whereas higher doses, not likely to be achieved in vivo, led to decreased activity( Reference Cole, Teter and Frautschy 131 ). More recent studies have shown that a synthetic derivative of curcumin, CNB-001, can stimulate Aβ degradation through both proteasomes and lysosomes, and the experimental inhibition of the proteasome pathway redirects clearance through lysosomes. Other recent CNB-001 studies have provided a link between the findings of several other AD-related biochemical changes. These include the findings that levels of the enzyme 5-lipoxygenase (5-LOX) are elevated in AD( Reference Ikonomovic, Abrahamson and Uz 132 ) and that disruption of this enzyme and some phospholipases can reduce AD pathology( Reference Qu, Uz and Manev 133 , Reference Firuzi, Zhuo and Chinnici 134 ); as well as that chronic stress can cause cell signalling/over-activation of regulatory kinases, which in turn leads to the phosphorylation of the eukaryotic initiation factor-2α (eIF2α) and disrupts the translation activation of several mRNA, and detected in neurodegenerative diseases including AD( Reference Ohno 135 ). The CNB-001 studies found that CNB-001 could inhibit 5-LOX, which induces the eIF2α phosphorylation. Furthermore, when fed to AD transgenic mice, CNB-001 was found to increase eIF2α phosphorylation (as well as heat shock protein 90 and activating transcription factor 4 levels), improve Aβ clearance and therefore limit the accumulation of soluble Aβ and ubiquitinated aggregated proteins. CNB-001 has also been found to maintain the expression of synapse-associated proteins and to improve memory in the mice( Reference Valera, Dargusch and Maher 136 ). These studies indicate that the curcuminoid derivative’s inhibition of 5-LOX has potential as a therapeutic approach.

Overall, cell culture and animal studies have indicated that curcumin has considerable potential as an inhibitor of Aβ aggregation, as an antioxidant, an anti-inflammatory and as an inhibitor of BACE1. Curcumin, among its modalities of action, has also shown promise in facilitating Aβ clearance/degradation, inhibiting tau phosphorylation, promoting neurogenesis and modulating synaptic plasticity (Fig. 1). Despite these benefits, there is a paucity of population-based studies examining the protective role of curcumin on cognition.

Fig. 1 Curcumin: reported mechanisms of action. BACE1, β-APP-cleaving enzyme-1; Aβ, β amyloid; APP, amyloid precursor protein.

Effects of curcumin on human cognition

Only a handful of clinical studies have been carried out to evaluate the cognitive enhancing potential of curcumin in AD patients( Reference Baum, Lam and Cheung 137 , Reference Ringman, Frautschy and Cole 138 ); however, these have not been particularly successful. Reasons could be because of the low bioavailability of curcumin( Reference Begum, Jones and Lim 15 ), thereby markedly reducing its potential to reach the brain at sufficient concentrations to provide benefits. Alternatively, the subjects may have been treated at a stage of pathology that is too advanced for curcumin to provide benefits. Nevertheless, there are epidemiological data that support the concept that curcumin can reduce the risk of AD. For example, India, with an estimated average daily consumption of curcumin being 80–200 mg( Reference Basnet and Skalko-Basnet 21 , Reference Commandeur and Vermeulen 139 ), has been reported to have a lower incidence and prevalence of AD( Reference Ganguli, Chandra and Kamboh 140 Reference Shaji, Bose and Verghese 142 ), although under-reporting and clinician access may be a contributor. Nevertheless, a study of 1010 cognitively intact Asian participants aged 60–93 years has found that those who consumed curry (which contains turmeric) more often, compared with those who ate curry very rarely or never, performed better on the Mini Mental State Examination (MMSE)( Reference Ng, Chiam and Lee 143 ). These observations, among others, support the concept that turmeric, and in particular its curcumin particle, may possess valuable neuroprotective or cognitive-enhancing properties. However, as mentioned above, clinical trials examining the efficacy of curcumin in patients with cognitive decline have been disappointing; however, more recently, with studies using improved formulations and more appropriate cohorts, encouraging signs are emerging( Reference Cox, Pipingas and Scholey 144 ). Table 1 represents a list of ongoing/completed clinical trials that have used curcumin for the diagnosis, prevention or treatment of AD. These trials are discussed further below.

Table 1 Studies using curcumin in Alzheimer’s disease (AD): diagnosis, prevention and treatment

Aβ, β amyloid; MMSE, Mini Mental State Examination; PET, positron emission tomography; FDG, fluorodeoxyglucose; MCI, mild cognitive impairment; NPIQ, Neuro-Psychiatric Inventory-Brief Questionnaire; fMRI, functional MRI.

Baum et al.( Reference Baum, Lam and Cheung 137 ) randomised AD patients (n 34) to receive 1 g (plus 3 g placebo), 4 g (plus 3 g placebo) or 0 g of oral curcumin (plus 4 g of placebo), once daily. Participants were given the choice of formulation, being either powder or capsule. The intervention group did not demonstrate significant differences in MMSE scores or plasma Aβ-40 levels between 0 and 6 months; however, it was suggested that the outcome measures were not sensitive or specific enough to demonstrate effects( Reference Baum, Lam and Cheung 137 ). Ringman et al.( Reference Ringman, Frautschy and Teng 145 ) conducted a 24-week, randomised, double-blind, placebo-controlled study evaluating the efficacy of two dosages of curcumin (2 and 4 g/d) in patients with mild-to-moderate AD, with an open-label extension for 48 weeks. This was the first study to include measurement of cerebrospinal fluid (CSF) biomarkers. The preliminary results showed no significant differences in cognitive function, in plasma or CSF Aβ-40/Aβ-42 or tau, between placebo and intervention groups; however, bioavailability was again reported as a limitation, although the dosing was well tolerated.

The above studies included AD-diagnosed participants, in whom significant neurodegeneration and AD pathology already exists. Given that the pathological changes begin two decades or more before the first recognisable symptoms( Reference Villemagne and Rowe 45 ), targeting healthy older cohorts or those in the pre-clinical or prodromal AD phase would more likely provide benefits through slowing the pathogenic mechanisms. Considerable synaptic and neuronal loss has already occurred by the time symptoms appear, and the antioxidant, anti-inflammatory and Aβ-lowering and anti-Aβ aggregation properties of curcumin are most likely to be of benefit in the early stages, for the prevention of AD pathogenesis. However, curcumin treatment of AD patients may still provide many benefits, and it warrants further clinical evaluation.

More recent studies have evaluated curcumin’s effects under normal physiological conditions. In a placebo-controlled study targeting healthy middle-aged subjects (n 38, 40–60 years), 80 mg of curcumin (400 mg of Longvida-optimised curcumin) was given orally for 4 weeks to assess the health-promoting effects of curcumin( Reference DiSilvestro, Joseph and Zhao 146 ). This study, because of the diverse health claims of curcumin, investigated several blood and saliva biomarkers, to examine the effect of curcumin on markers associated with lipids, inflammation, liver function, immunity and stress, as well as Aβ levels. Cognitive measures were not included in their study design. Statistically significant results were shown for a number of these markers including increased catalase, nitric oxide and antioxidant status, with lowered plasma alanine aminotransferase and TAG, but not total cholesterol. In addition, curcumin was found to lower plasma Aβ levels. Another interesting finding was that salivary amylase was also significantly lowered, which is an enzyme associated with adrenergic activity during stress( Reference van Stegerena, Rohlederb and Everaerda 147 ).

A number of studies of curcumin supplementation in healthy older subjects are still in progress. However, one such study has been completed: a randomised, double-blind placebo-controlled study (n 60, 60–85 years) using the same 80 mg/d curcumin formulation as used by DiSilvestro et al. ( Reference DiSilvestro, Joseph and Zhao 146 ) (400 mg of Longvida-optimised curcumin). The authors reported acute (1 h post dose) and chronic (1-month duration) effects of curcumin intake on cognition, mood and blood biomarkers( Reference Cox, Pipingas and Scholey 144 ). Benefits on attention and working memory were reported following the acute administration of curcumin, whereas at the 1-month time point, working memory and mood improved. Alertness and contentedness also improved after acute-on-chronic treatment.

Although the results above are encouraging, alternate mechanisms, including modulation of the stress response, may have played a part. Amylase, shown to be lowered in an earlier study using curcumin( Reference DiSilvestro, Joseph and Zhao 146 ), is a recognised biomarker of β-adrenergic stimulation( Reference van Stegerena, Rohlederb and Everaerda 147 Reference Chatterton, Vogelsong and Lu 149 ), and improved attention, working memory and contentedness may be linked to this mode of action. No alterations in blood levels of Aβ-40 or Aβ-42 levels were detected, although these are not thought to be reliable biomarkers on their own. Differences in cognitive performance, as demonstrated in serial 7-s and delayed recall, were not significant( Reference Cox, Pipingas and Scholey 144 ). Nevertheless, curcumin did enhance the lipid profile by lowering LDL, and in contrast to the results found by DiSilvestro et al. ( Reference DiSilvestro, Joseph and Zhao 146 ), Cox et al.( Reference Cox, Pipingas and Scholey 144 ) also recorded a reduction in plasma total cholesterol. Long-term lipid changes such as these may have some effect on AD risk: as mentioned earlier, chronic conditions linked to abnormal lipid profiles such as obesity and diabetes are linked with a higher risk of AD. Interestingly, a case study reported by Hishikawa et al.( Reference Hishikawa, Takahashi and Amakusa 150 ) found that three severe-stage AD patients treated with 100 mg/d oral curcumin for 12 weeks (in addition to their already prescribed acetylcholinesterase inhibitor, donepezil) showed a reduction in agitation, anxiety and irritability; one patient also showed improvement in MMSE score. This may suggest a role for curcumin as a concurrent intervention, and it supports the concept that curcumin may provide benefits, even in advanced stages of AD; however, further research would be required to support these suggestions.

As mentioned earlier, several other clinical studies are still underway, or results have not yet been published; thus, with so few clinical studies having been completed, it is not possible to make any conclusions concerning the clinical significance of curcumin in enhancing cognition.

Epigenetics, Alzheimer’s disease and curcumin

Epigenetic alterations have been reported to occur in AD( Reference Mastroeni, Grover and Delvaux 151 Reference Chouliaras, Rutten and Kenis 154 ). As epigenetic alterations are dynamic, these alterations have been proposed as a target area for AD prevention. Epigenetic alterations include changes in DNA methylation, histone modifications or changes in miRNA expression. Studies including some clinical studies of other conditions have shown that curcumin has the potential to induce epigenetic changes( Reference Teiten, Dicato and Diederich 155 , Reference Reuter, Gupta and Park 156 ). For example, epigenetic effects of curcumin have been shown in patients with breast cancer and advanced pancreatic cancer, and also in people at risk of stroke, by providing vascular protection( Reference Du, Xie and Wu 157 ). Curcumin has been shown to inhibit DNA methyltransferase, histone acetyltransferase and histone deacetylase, and to modulate miRNA, for example down-regulating microRNA-134 and microRNA-124 in cultured hippocampal slices, which are associated with an increase in the brain-derived neurotropic factor (BDNF)( Reference Sezgin and Dincer 158 ). BDNF has been shown to increase hippocampal neuronal survival and to enhance synaptic plasticity. Furthermore, variants of the BDNF gene have been linked to several mental disorders such as major depressive disorder (MDD) and schizophrenia, and low levels of BDNF protein are thought to contribute to the pathology of MDD. Interestingly, antidepressants have also been found to increase blood levels of BDNF( Reference Li, Xu and Gao 159 ). Depression is a major risk factor for AD, and thus this BDNF-modifying property of curcumin is of significant interest. Other recent studies have also found that curcumin can have a significant effect on depression( Reference Lopresti, Maes and Marker 160 ), supported to some extent by the studies described above by Cox et al. ( Reference Cox, Pipingas and Scholey 144 )

Many other studies have discovered the beneficial epigenetic effects of curcumin in relation to various cancers and rheumatoid arthritis, and now similar benefits are being discovered that may have an impact on the risk and severity of AD( Reference Davinelli, Calabrese and Zella 161 ). For example, DNA methylation in neurodegenerative diseases (and many other conditions such as CVD and stroke) has been linked to high homocysteine levels, which occur with ageing and with vitamin B12 or folate deficiencies( Reference Ansari, Mahta and Mallack 162 , Reference Mattson and Shea 163 ). Chronically high homocysteine levels lead to an abnormally high DNA methylation( Reference Fux, Kloor and Hermes 164 ), which requires DNA methyltransferase, and as mentioned above curcumin inhibits DNA methyltransferase. However, rat studies of homocysteine effects suggest that curcumin may be neuroprotective, and may improve learning and memory deficits, by reducing lipid peroxidation and high malondialdehyde levels, both of which are induced by high homocysteine levels( Reference Ataie, Sabetkasaei and Haghparast 165 ). The relative significance of these potential benefits of curcumin dietary supplementation are clearly still not known, and thus further clinical studies are required to evaluate the neuroprotective role of curcumin induced by epigenetic regulation for the prevention of cognitive decline.

Curcumin safety profile, tolerability, bioavailability and mode of administration

As curcumin is a component of the spice turmeric, it is not surprising that curcumin has been reported to be a very safe nutraceutical with a low side-effect profile. However, it should be noted that while curcumin has been reported to be safe and well tolerated at doses of upto 8 g/d( Reference Cheng, Hsu and Lin 166 ), studies have not gone beyond 3 months, and thus the long-term effects of high doses of curcumin are not known. In addition, with enhanced bioavailability and absorption now possible with new formulations, the risk of increased toxicity is higher, particularly for populations taking medications metabolised by the liver or for those with existing liver impairment( Reference Baum, Lam and Cheung 137 ). Nevertheless, although curcumin may not have been tested widely for the purposes of reducing neurodegeneration, it has been clinically tested in patients with various conditions and pro-inflammatory diseases including cancer, CVD, arthritis, Crohn’s disease, ulcerative colitis, irritable bowel disease, tropical pancreatitis, peptic ulcer, psoriasis, atherosclerosis, diabetes, diabetic nephropathy and renal conditions, among others, and has resulted in minimal side effects and many health benefits. Curcumin has also provided protection against hepatic conditions, chronic arsenic exposure and alcohol intoxication( Reference Gupta, Kismali and Aggarwal 167 ).

Curcumin’s pleiotropic effects explain its wide variety of applications; for example, it has been tested for the purpose of stent coating, as curcumin has advantageous anti-coagulant properties( Reference Pan, Tang and Weng 168 ). In addition, it can inhibit the generation of blood clotting factors Xa and thrombin via the extrinsic and intrinsic pathways( Reference Dong-Chan, Sae-Kwang and Jong-Sup 169 ). These properties do indicate that one should be cautious when prescribing curcumin in combination with other blood-thinning preparations. Although the safety of curcumin has been demonstrated( Reference Baum, Lam and Cheung 137 , Reference Ringman, Frautschy and Teng 145 ), in humans, oral ingestion of existing formulations has presented challenges concerning absorption and bioavailability( Reference Ringman, Frautschy and Teng 145 , Reference Belkacemi, Doggui and Dao 170 ). The literature reports that oral curcumin has efficient first-pass metabolism and some degree of intestinal metabolism, including glucuronidation and sulphation (although this occurs mostly in the liver); however, it is excreted largely unconjugated via the intestine. Curcumin is also unstable at neutral and alkaline pH. There appears to be minimal distribution of curcumin to the liver or other tissues beyond the gastrointestinal tract( Reference Sharma, Steward and Gescher 171 ). For example, in rat studies, an oral dose of 500 mg/kg resulted in a peak plasma concentration of only 1·8 ng/ml( Reference Ireson, Orr and Jones 172 ), with the major metabolites being curcumin sulphate and curcumin glucuronide, whereas a clinical study found that oral doses of 4, 6 and 8 g of curcumin daily for 3 months yielded serum curcumin concentrations of only 0·51 (sd 0·11), 0·63 (sd 0·06) and 1·77 (sd 1·87) μm, respectively, with peak levels at 1–2 h post dosing( Reference Cheng, Hsu and Lin 166 ). Unmodified curcumin is reported to be retained in the blood for 2–5 h in humans, whereas retention of a modified form of curcumin – Biocurcumax-95 (BCM-95) – is reported as exceeding 8 h( Reference Benny and Anthony 173 ). In order for the curcumin to elicit a greater nutraceutical benefit, it is critical that more of it is able to enter the bloodstream, it must have a longer half-life and also cross the BBB to be of significant benefit in AD.

To date, most human curcumin studies have used oral formulations. Absorption and bioavailability have continued to be a hindrance, not aided by the variation of formulations available. However, as technology has advanced and new delivery approaches have emerged, the use of adjuvant therapies, isomerisation, liposomes, micelles, phospholipids and nanotechnology also increase. One of the potential therapeutic results of increasing blood levels of curcumin in humans can hopefully be anticipated from AD transgenic mouse studies, in which the intravenous administration of curcumin (7·7 mg/kg per d) for 7 d resulted in significant clearance of cerebral Aβ load( Reference Garcia-Alloza, Borrelli and Rozkalne 40 ). Alternate routes of delivering curcumin are already being used for other disorders, such as the use of topical eye drops, recommended for the treatment of a variety of ophthalmic disorders( Reference Pescosolido, Giannotti and Plateroti 174 ), and transdermal application using encapsulated curcumin as a nanoemulsion, for the treatment of arthritis( Reference Rachmawati, Budiputra and Mauludin 175 ). Recently, the curcumin derivative FMeC1, originally produced as an MRI probe, has been produced in aerosol form for inhalation. A study in 5XFAD transgenic mice suggested improved distribution in the brain, and immunohistochemical studies demonstrated that FMeC1 absorbed following aerosol delivery did bind to amyloid plaques in the mouse brains( Reference McClure, Yanagisawa and Stec 176 ). This technique may also be useful for Aβ imaging studies; however, further studies are needed to validate this notion. The absorption and bioavailability of curcumin is highly relevant, and the formulation, dose and mode of delivery are each important factors. Multiple over-the-counter brands are available, and most of them claim increased bioavailability compared with unformulated curcumin; however, independent comparative analysis is essential. Two formulations BCM-95( Reference Antony, Merina and Iyer 177 , Reference Merina and Antony 178 ) and Longvida( Reference Ringman, Frautschy and Teng 145 ) currently have the strongest independent data available in human trials. For a review of the molecular structure of curcumin and its derivatives, FMeC1 and FMeC2, see Yanagisawa et al.( Reference Yanagisawa, Ibrahim and Taguchi 110 ), and differences between the properties of CNB-001 can be examined as previously published( Reference Jayaraj, Elangovan and Dhanalakshmi 179 Reference Liu, Dargusch and Maher 181 ).

To summarise, curcumin has been trialled at doses as high as 8 g/d, and found to be well tolerated and safe. However, as new formulations are emerging that are showing promise of increasing bioavailability, BBB permeability and longer half-lives, these formulations also need to be evaluated in future safety and tolerability trials. Furthermore, as any curcumin therapy is likely to be long-term in nature, much longer treatment times need to be trialled. The animal and clinical studies that have investigated the role of curcumin have applied a variety of administration modes, including oral, subcutaneous, intraperitoneal, intravenous, topical and the nasal route( Reference Prasad, Tyagi and Aggarwal 182 ). Human trials investigating curcumin’s neuroprotective mechanisms have mostly used the oral route; however, future studies should explore other routes of administration.

Curcumin as a fluorochrome/radioligand in Alzheimer’s disease diagnosis

Turmeric has been used as a colouring agent since ancient times. In 1989, Stockert et al.( Reference Stockert, Del Castillo and Gomez 183 ) identified curcumin as a potential fluorochrome, as curcumin was found to fluoresce yellow/green under a violet/blue (436 nm) light, and it was noted to bind to DNA and chromosomes, as treatment of tissue samples and cell samples with deoxyribonuclease or TCA prevented the chromatin staining. More recently, these innate fluorescent qualities (curcumin absorbs light at about 420 nm and emits fluorescence at about 530 nm in aqueous solutions( Reference Wang, Wu and Wang 184 )), and curcumin’s natural affinity to bind with Aβ, prompted curcumin to be tested as a safe plaque-labelling fluorochrome. A mouse study investigated novel derivatives of curcumin and measured their binding affinities for Aβ aggregates( Reference Ryu, Choe and Lee 185 ). The derivative with the highest affinity was then (18F)-radiolabelled for testing as a radioligand probe for Aβ plaque imaging; the compound also had suitable lipophilicity, good brain uptake and was metabolically stable in the brain. In another study conducted on transgenic AD mice, multiphoton microscopy was used to demonstrate that curcumin crossed the BBB and labelled Aβ plaques and cerebrovascular amyloid angiopathy( Reference Garcia-Alloza, Borrelli and Rozkalne 40 ). Curcumin has since been used in the labelling of neuronal fibrillar tau inclusions in human brain samples of AD and progressive supranuclear palsy( Reference Mohorko, Repovš and Popovic 186 ). More recently, other studies of Aβ imaging used an 19F-containing curcumin derivative injected peripherally in AD-model mice to detect Aβ plaques in the brains, using MRI( Reference Yanagisawa, Ibrahim and Taguchi 110 ). Furthermore, Koronyo-Hamaoui et al.( Reference Koronyo-Hamaoui, Koronyo and Ljubimov 187 ) demonstrated curcumin’s value as a staining agent for Aβ by detecting plaques in human postmortem retinal tissue, and also as a brain and retinal Aβ plaque tracer administered intravenously in transgenic mice. Importantly, the pathology in the retina was detected before the stage at which pathology in the brain could be detected, indicating that curcumin may have potential as a pre-clinical AD biomarker. The research also supports the previous observation that curcumin has the ability to cross the BBB, which is essential for its therapeutic efficacy. Preliminary data from a pilot study (n 40) conducted by our group undertaking retinal imaging using curcumin as a fluorochrome had a 100 % sensitivity and 80·6 % specificity for AD diagnosis( Reference Frost, Kanagasingam and Macaulay 3 ). Another study recruited mild cognitive impairment (MCI) patients (n 30) and administered 80 mg of curcumin (Meriva) twice daily for 3 d and found abnormal deposits in different retinal layers believed to be related to neurodegeneration( Reference Kayabasi, Sergott and Rispoli 188 ). The study reported that curcumin caused patchy hypofluorescent spots; however, it did not quantify the retinal amyloid plaques. The findings were primarily based on the direct perception of the deposits via ocular imaging. Recent studies have again used MRI, this time to detect magnetic nanoparticles made of superparamagnetic iron oxide conjugated with curcumin, which were found to bind to Aβ aggregates in ex vivo AD-model mouse brains, after injection with the curcumin conjugate( Reference Cheng, Chan and Fan 189 ). Other recent studies have produced a novel nanoimaging agent: poly(β-l-malic acid) containing covalently attached (1,4,7,10-tetraazacyclododecane-1,4,7,10-tetraacetic acid) gadolinium and curcumin. The all-in-one agent selectively binds to Aβ plaques and can be detected by MRI( Reference Patil, Gangalum and Wagner 190 ), thus providing another promising Aβ plaque imaging agent.

Curcumin, being non-toxic, accessible and economical, thus becomes highly attractive for both diagnostics and therapeutic research. As discussed above, retinal imaging using curcumin fluorescence is currently being examined in our research centre as part of the Australian Imaging, Biomarkers and Lifestyle flagship study of ageing, and it has been found to be well tolerated by participants. This approach, combined with examination of the retinal vascular features( Reference Frost, Martins and Kanagasingam 191 ), may deliver a novel diagnostic tool that provides a more reliable indication of early AD changes, which is economical, relatively non-invasive and widely applicable.

Limitations of curcumin and future directions

If all the positive results observed in in vitro and in vivo animal studies could be translated into human studies, the significance of curcumin in AD prevention and treatment would be considerable. The recent study by Cox et al.( Reference Cox, Pipingas and Scholey 144 ) provided some encouraging outcomes by investigating cognitive markers, but only acute changes in attention, working memory and mood were significant. No effects were reported in long-term memory or executive function; however, the short duration of the study may have been a limitation. The evidence that curcumin can influence Aβ aggregation and Aβ clearance, support innate immune systems, reduce oxidative stress, enhance cognition and impede the onset of AD in humans remains elusive. However, not to be overlooked is the application of curcumin as a diagnostic fluorochrome, potentially assisting in the earlier identification of pre-clinical stages of AD, during retinal scanning. The use of curcumin as a fluorochrome within retinal amyloid imaging, combined with examination of retinal vascular features, offers a novel diagnostic approach to AD. While retinal imaging is acknowledged as a non-invasive, economical and easily translated technology, further validation is required before it can be adopted as an early AD marker.

Enhanced oral formulations of curcumin are emerging, potentially negating the prior challenges of absorption and bioavailability. However, considering the differences in product formulation, and the multitude of curcumin products already available, comparative analysis would be useful. As the primary focus of AD treatment has turned to primary prevention, the point of intervention is also crucial: interventions introduced early (>10 years before the onset of AD clinical symptoms) may present difficulty establishing statistically significant changes, whereas interventions introduced 1–2 years before the onset may be less effective or ineffective as the disease pathology may already be too advanced. Prevention studies designed with longer duration are highly desirable; however, as nutraceuticals generally do not attract commercial opportunity, realising these type of studies will be difficult to accomplish.

The use of curcumin as an adjunct therapy to cholinesterase inhibitors, particularly in the early stages of AD, offers a potentially new area of research. As anxiety and stress are common co-morbidities in AD, and research has shown curcumin to have effect in these areas, curcumin may offer an appealing alternative to antidepressant and antipsychotic therapies, while potentially offering other synergies, including influencing the underlying neuropathology and enhancing cholinergic activity. In recent years, the focus on curcumin as a compound of interest for the prevention of AD has been intensifying. The results of ongoing clinical trials will hopefully shed more light on the benefits of curcumin in the prevention of AD. In line with curcumin’s complex modes of action, outcome measures should be expanded to include not just cognitive changes; extensive blood biomarker assays should also be carried out, as well as imaging (e.g. measuring cerebral amyloid load and potentially retinal markers) to characterise curcumin’s effects more fully over time, in a pre-clinical population.

Conclusion

To date, AD clinical trials have not been able to generate the anticipated benefits of curcumin; however, this has been broadly attributed to difficulties with absorption, bioavailability and arguably the timing and length of intervention. As reviewed in this article, there is significant evidence that curcumin can act on multiple pathways identified in the pathogenesis of AD. It is possible, however, that sporadic AD in humans with the associated cerebral atrophy and neuronal death may be less responsive to curcumin than the AD induced in transgenic animal models of the disease.

As discussed in this review, increasing the bioavailability, BBB penetration and sustaining the half-life of curcumin remains a major focus in relation to its dose–response. To achieve the same degree of efficacy in human studies as compared with animal studies, closer scrutiny of the administration route and also formulation are required, as increasing bioavailability and BBB penetration is critical. Research analysing the different oral formulations is lacking, and this is an area for further investigation. Furthermore, as the pre-clinical signs of AD are present decades before its clinical onset and most of the late-stage AD clinical trials have recently failed, intervention must be focused at preventing or delaying AD onset. It is reasonable to include healthy community-dwelling older adults and those with subjective memory complaints, in intervention studies with curcumin, for a longer duration with longitudinal follow-up. Last, inclusion of AD-related biomarkers and neuroimaging would add to the clinical significance of curcumin’s efficacy in the prevention of AD and associated cognitive decline.

Acknowledgements

The authors gratefully acknowledge the combined support of the McCusker Alzheimer’s Research Foundation and the Anglican Retirement Villages (ARV).

K. G. G. is supported by a grant through the Foundation for Aged Care, ARV, and a scholarship from the Co-operative Research Centre for Mental Health. R. N. M., T. M. S., H. R. S., S. R. R.-S., B. B. and G. V are supported by the Edith Cowan University and the McCusker Alzheimer’s Research Foundation. H. R. S. has received renumeration from Pfizer and Takeda. R. N. M. is the founder and chief scientific officer of the biotech company, Alzhyme. G. V. is supported by the Curtin University Senior Research Fellowship (CRF140196) and the NHMRC (APP1045507).

All authors contributed to the literature search, analysis of the data published, manuscript writing and revisions of the article.

The authors declare no conflicts of interest arising from the conclusions of this research.

References

1. Prince, M & Jackson, J (editors) (2009) Alzheimer’s Disease International: World Alzheimer Report. Executive Summary. London: Alzheimer’s Disease International.Google Scholar
2. Lee, VM, Goedert, M & Trojanowski, JQ (2001) Neurodegenerative tauopathies. Annu Rev Neurosci 24, 11211159.Google Scholar
3. Frost, S, Kanagasingam, Y, Macaulay, L, et al. (2014) Retinal amyloid fluorescence imaging predicts cerebral amyloid burden and Alzheimer’s disease (oral presentation). Alzheimers Dement 10, 234235.CrossRefGoogle Scholar
4. Ittner, LM & Götz, J (2011) Amyloid-β and tau – a toxic pas de deux in Alzheimer’s disease. Nat Rev Neurosci 12, 6772.Google Scholar
5. Ittner, LM, Ke, YD, Delerue, F, et al. (2010) Dendritic function of tau mediates amyloid-beta toxicity in Alzheimer’s disease mouse models. Cell 142, 387397.Google Scholar
6. Krishnaswamy, K, Verdile, G, Groth, DM, et al. (2009) The structure and function of Alzheimer’s gamma secretase enzyme complex. Crit Rev Clin Lab Sci 46, 282301.Google Scholar
7. Villemagne, VL, Burnham, S, Bourgeat, P, et al. (2013) Amyloid beta deposition, neurodegeneration, and cognitive decline in sporadic Alzheimer’s disease: a prospective cohort study. Lancet Neurol 12, 357367.CrossRefGoogle ScholarPubMed
8. Walsh, DM & Teplow, DB (2012) Alzheimer’s disease and the amyloid beta-protein. Progr Mol Biol Transl Sci 107, 101124.Google Scholar
9. O’Malley, T, Oktaviani, N, Zhang, D, et al. (2014) A beta dimers differ from monomers in structural propensity, aggregation paths and population of synaptotoxic assemblies. Biochem J 461, 413426.Google Scholar
10. McKhann, GM, Knopman, DS, Chertkow, H, et al. (2011) The diagnosis of dementia due to Alzheimer’s disease: recommendations from the National Institute on Aging-Alzheimer’s Association workgroups on diagnostic guidelines for Alzheimer’s disease. Alzheimers Dement 7, 263269.Google Scholar
11. Majeed, M, Badmaev, V & Murrary, F (1996) Turmeric and the Healing Curcuminoids. New Canaan, CT: Keats Publishing Inc.Google Scholar
12. Kelloff, GJ, Crowell, JA, Steele, VE, et al. (2000) Progress in cancer chemoprevention: development of diet-derived chemopreventive agents. J Nutr 130, 467S471S.Google Scholar
13. Goel, A, Kunnumakkara, AB & Aggarwal, BB (2008) Curcumin as ‘Curecumin’: from kitchen to clinic. Biochem Pharmacol 75, 787809.Google Scholar
14. Ammon, HP & Wahl, MA (1991) Pharmacology of Curcuma longa . Plata Med 57, 107.Google Scholar
15. Begum, AN, Jones, MR, Lim, GP, et al. (2008) Curcumin structure-function, bioavailability, and efficacy in models of neuroinflammation and Alzheimer’s disease. J Pharmacol Exp Ther 326, 196208.Google Scholar
16. Zhou, H, Beevers, CS & Huang, S (2011) Targets of curcumin. Curr Drug Targets 12, 332347.Google Scholar
17. Milobedeska, J, Kostanecki, V & Lampe, V (1910) Structure of curcumin. Berichte der Deutschen Chemischen Gesellschaft 43, 21632170.Google Scholar
18. Hamaguchi, T, Ono, K & Yamada, M (2010) Review: curcumin and Alzheimer’s disease. CNS Neurosci Ther 16, 285297.CrossRefGoogle ScholarPubMed
19. Agarwal, DK & Mishra, PK (2010) Curcumin and its analogues: potential anticancer agents. Med Res Rev 30, 818860.Google Scholar
20. Mishra, S & Palanivelu, K (2008) The effect of curcumin (turmeric) on Alzheimer’s disease: an overview. Ann Indian Acad Neurol 11, 1319.CrossRefGoogle ScholarPubMed
21. Basnet, P & Skalko-Basnet, N (2011) Curcumin: an anti-inflammatory molecule from a curry spice on the path to cancer treatment. Molecules 16, 45674598.Google Scholar
22. Pan, MH, Lin-Shiau, SY & Lin, JK (2000) 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 60, 16651676.Google Scholar
23. Xu, YX, Pindolia, KR, Janakiraman, N, et al. (1997) Curcumin inhibits IL1 alpha and TNF-alpha induction of AP-1 and NF-kB DNA-binding activity in bone marrow stromal cells. Hematopathol Mol Hematol 11, 4962.Google Scholar
24. Zhang, C, Browne, A, Child, D, et al. (2010) Curcumin decreases amyloid-beta peptide levels by attenuating the maturation of amyloid-beta precursor protein. J Biol Chem 285, 2847228480.Google Scholar
25. Joe, B, Rao, UJSP & Lokesh, BR (1997) Presence of an acidic glycoprotein in the serum of arthritic rats: modulation by capsaicin and curcumin. Mol Cell Biochem 169, 125134.Google Scholar
26. Jackson, JK, Higo, T, Hunter, WL, et al. (2006) The antioxidants curcumin and quercetin inhibit inflammatory processes associated with arthritis. Inflamm Res 55, 168175.Google Scholar
27. Hanai, H & Sugimoto, K (2009) Curcumin has bright prospects for the treatment of inflammatory bowel disease. Curr Pharm Des 15, 20872094.Google Scholar
28. Kurd, SK, Smith, N, VanVoorhees, A, et al. (2008) Oral curcumin in the treatment of moderate to severe psoriasis vulgaris: a prospective clinical trial. J Am Acad Dermatol 58, 625631.Google Scholar
29. Thangapazham, RL, Sharma, A & Maheshwari, RK (2007) Beneficial role of curcumin in skin diseases. Adv Exp Med Biol 595, 343357.CrossRefGoogle ScholarPubMed
30. Tsui, KH, Feng, TH, Lin, CM, et al. (2008) Curcumin blocks the activation of androgen and interlukin-6 on prostate-specific antigen expression in human prostatic carcinoma cells. J Androl 29, 661668.Google Scholar
31. Liu, D & Chen, Z (2013) The effect of curcumin on breast cancer cells. J Breast Cancer 16, 133137.Google Scholar
32. Friedman, L, Lin, L, Ball, S, et al. (2009) Curcumin analogues exhibit enhanced growth suppressive activity in human pancreatic cancer cells. Anticancer Drugs 20, 444449.Google Scholar
33. Lim, TG, Lee, SY, Huang, Z, et al. (2014) Curcumin suppresses proliferation of colon cancer cells by targeting CDK2. Cancer Prev Res (Phila) 7, 466474.Google Scholar
34. Jackson-Bernitsas, DG, Ichikawa, H, Takada, Y, et al. (2007) Evidence that TNF-TNFR1-TRADD-TRAF2-RIP-TAK1-IKK pathway mediates constitutive NF-kappaB activation and proliferation in human head and neck squamous cell carcinoma. Oncogene 26, 13851397.Google Scholar
35. Zhao, BL, Li, XJ, He, RG, et al. (1989) Scavenging effect of extracts of green tea and natural antioxidants on active oxygen radicals. Cell Biophys 14, 175185.Google Scholar
36. Martins, RN, Harper, CG, Stokes, GB, et al. (1986) Increased cerebral glucose-6-phosphate dehydrogenase activity in Alzheimer’s disease may reflect oxidative stress. J Neurochem 46, 10421045.Google Scholar
37. Hardy, J & Selkoe, DJ (2002) The amyloid hypothesis of Alzheimer’s disease: progress and problems on the road to therapeutics. Science 297, 353356.Google Scholar
38. Huang, HC & Jiang, ZF (2009) Accumulated amyloid-β peptide and hyperphosphorylated tau protein: relationship and links in Alzheimer’s disease. J Alzheimers Dis 16, 1527.CrossRefGoogle ScholarPubMed
39. Huang, HC, Chang, P, Dai, XL, et al. (2012) Protective effects of curcumin on amyloid-beta-induced neuronal oxidative damage. Neurochem Res 37, 15841597.Google Scholar
40. Garcia-Alloza, M, Borrelli, LA, Rozkalne, A, et al. (2007) Curcumin labels amyloid pathology in vivo, disrupts existing plaques, and partially restores distorted neurites in an Alzheimer mouse model. J Neurochem 102, 10951104.Google Scholar
41. Mutsuga, M, Chambers, JK, Uchida, K, et al. (2012) Binding of curcumin to senile plaques and cerebral amyloid angiopathy in the aged brain of various animals and to neurofibrillary tangles in Alzheimer’s brain. J Vet Med Sci 74, 5157.Google Scholar
42. Kosik, KS, Joachim, CL & Selkoe, DJ (1986) Microtubule-associated protein tau (tau) is a major antigenic component of paired helical filaments in Alzheimer disease. Proc Natl Acad Sci U S A 83, 40444048.Google Scholar
43. Glenner, GG & Wong, CW (1984) Alzheimer’s disease: initial report of the purification and characterization of a novel cerevrovascular amyloid protein. Biochem Biophys Res Commun 120, 885890.Google Scholar
44. Masters, CL, Multhaup, G, Simms, G, et al. (1985b) Neuronal origin of a cerebral amyloid: neurofibrillary tangles of Alzheimer’s disease contain the same protein as the amyloid of plaque cores and blood vessels. EMBO J 4, 27572763.CrossRefGoogle ScholarPubMed
45. Villemagne, VL & Rowe, CC (2013) Long night’s journey into the day: amyloid-beta imaging in Alzheimer’s disease. J Alzheimers Dis 33, Suppl. 1, S349S359.Google Scholar
46. Haass, C & Selkoe, DJ (2007) Soluble protein oligomers in neurodegeneration: lessons from the Alzheimer’s amyloid beta-peptide. Nat Rev Mol Cell Biol 8, 101112.CrossRefGoogle ScholarPubMed
47. Manczak, M, Mao, P, Calkins, MJ, et al. (2010) Mitochondria-targeted antioxidants protect against amyloid-beta toxicity in Alzheimer’s disease neurons. J Alzheimers Dis 20, Suppl. 2, S609S631.Google Scholar
48. Cash, DM, Liang, Y, Ryan, NS, et al. (2013) The pattern of atrophy in familial alzheimer disease: volumetric MRI results from the DIAN study. Neurology 81, 14251433.Google Scholar
49. Chetelat, G, Villemagne, VL, Villain, N, et al. (2012) Accelerated cortical atrophy in cognitively normal elderly with high beta-amyloid deposition. Neurology 78, 477484.Google Scholar
50. Chetelat, G, Villemagne, VL, Pike, KE, et al. (2011) Independent contribution of temporal beta-amyloid deposition to memory decline in the pre-dementia phase of Alzheimer’s disease. Brain 134, 798807.Google Scholar
51. Reddy, PH, Tripathi, R, Troung, Q, et al. (2012) Abnormal mitochondrial dynamics and synaptic degeneration as early events in Alzheimer’s disease: implications to mitochondria-targeted antioxidant therapeutics. Biochim Biophys Acta 1822, 639649.Google Scholar
52. Ferreira, ST & Klein, WL (2011) The Aβ oligomer hypothesis for synapse failure and memory loss in Alzheimer’s disease. Neurobiol Learn Mem 96, 529543.Google Scholar
53. Zhao, J, Luo, Y, Jang, H, et al. (2012) Probing ion channel activity of human islet amyloid polypeptide (amylin). Biochim Biophys Acta 1818, 31213130.Google Scholar
54. Johnson, E, Brookmeyer, R & Ziegler-Graham, K (2007) Modeling the effect of Alzheimer’s disease on mortality. Int J Biostat 3, Article 13, 12–21.Google Scholar
55. Brookmeyer, R, Gray, S & Kawas, C (1998) Projections of Alzheimer’s disease in the United States and the public health impact of delaying disease onset. Am J Public Health 88, 13371342.Google Scholar
56. Vickland, V, Morris, T, Draper, B, et al. (2012) Modelling the Impact of Interventions to Delay the Onset of Dementia in Australia. Report for Alzheimers Australia. Sydney, Australia: Alzheimer’s Australia Inc.Google Scholar
57. Bieschke, J, Herbst, M, Wiglenda, T, et al. (2012) Small-molecule conversion of toxic oligomers to nontoxic β-sheet-rich amyloid fibrils. Nat Chem Biol 8, 93101.Google Scholar
58. Ono, K, Hasegawa, K, Naiki, H, et al. (2004) Curcumin has potent anti-amyloidogenic effects for Alzheimer’s beta-amyloid fibrils in vitro . J Neurosci Res 75, 742750.Google Scholar
59. Reinke, AA & Gestwicki, JE (2007) Structure-activity of amyloid betta-aggregation inhibitors based on curcumin: influence of linker length and flexibility. Chem Biol Drug Des 70, 206215.Google Scholar
60. Yang, F, Lim, GP, Begum, AN, et al. (2005) Curcumin inhibits formation of amyloid beta oligomers and fibrils, binds plaques, and reduces amyloid in vivo . J Biol Chem 280, 58925901.Google Scholar
61. Ghalebani, L, Wahlstrom, A, Danielsson, J, et al. (2012) pH-dependence of the specific binding of Cu(II) and Zn(II) ions to the amyloid-β peptide. Biochem Biophys Res Commun 421, 554560.Google Scholar
62. Faller, P & Hureau, C (2012) A bioinorganic view of Alzheimers disease: when misplaced metal ions (Re)direct the electrons to the wrong target. Chemistry 18, 1591015920.CrossRefGoogle Scholar
63. Banerjee, R (2014) Effect of curcumin on the metal ion induced fibrillization of amyloid-β peptide. Spectrochim Acta A Mol Biomol Spectrosc 117, 798800.Google Scholar
64. Kochi, A, Lee, HJ, Vithanarachchi, SM, et al. (2015) Inhibitory activity of curcumin derivatives towards metal-free and metal-induced amyloid-beta aggregation. Curr Alzheimer Res 12, 415423.CrossRefGoogle ScholarPubMed
65. Fu, Z, Aucoin, D, Ahmed, M, et al. (2014) Capping of aβ42 oligomers by small molecule inhibitors. Biochemistry 53, 78937903.Google Scholar
66. Mithu, VS, Sarkar, B, Bhowmik, D, et al. (2014) Curcumin alters the salt bridge-containing turn region in amyloid β(1-42) aggregates. J Biol Chem 289, 1112211131.Google Scholar
67. Caesar, I, Jonson, M, Nilsson, K, et al. (2012) Curcumin promotes Aβ fibrillation and reduces neurotoxicity in transgenic drosophila (reduced neurotoxicity by promoted fibrillation). PLOS ONE 7, e31424.Google Scholar
68. Liu, H, Li, Z, Qiu, D, et al. (2010) The inhibitory effects of different curcuminoids on β-amyloid protein, β-amyloid precursor protein and β-site amyloid precursor protein cleaving enzyme 1 in swAPP HEK293 cells. Neurosci Lett 485, 8388.Google Scholar
69. Shimmyo, Y, Kihara, T, Akaike, A, et al. (2008) Epigallocatechin-3-gallate and curcumin suppress amyloid beta-induced beta-site APP cleaving enzyme-1 upregulation. Neuroreport 19, 13291333.Google Scholar
70. Li, Y, Zhang, X & Si, L (2009) Curcumin reduces A beta generation by PPAR gamma activation and BACE1 inhibition in vitro . J Neurochem 110, 61.Google Scholar
71. Wang, X, Kim, JR, Lee, SB, et al. (2014) Effects of curcuminoids identified in rhizomes of Curcuma longa on BACE-1 inhibitory and behavioral activity and lifespan of Alzheimer’s disease Drosophila models. BMC Complement Altern Med 14, 8888.Google Scholar
72. Sathya, M, Premkumar, P, Karthick, C, et al. (2012) BACE1 in Alzheimer’s disease. Clin Chim Acta 414, 171178.Google Scholar
73. Jiaranaikulwanitch, J, Govitrapong, P, Fokin, VV, et al. (2012) From BACE1 inhibitor to multifunctionality of tryptoline and tryptamine triazole derivatives for Alzheimer’s disease. Molecules 17, 83128333.CrossRefGoogle ScholarPubMed
74. Lin, R, Chen, X, Li, W, et al. (2008) Exposure to metal ions regulates mRNA levels of APP and BACE1 in PC12 cells: blockage by curcumin. Neurosci Lett 440, 344347.Google Scholar
75. Park, SY, Kim, HS, Cho, EK, et al. (2008) Curcumin protected PC12 cells against beta-amyloid-induced toxicity through the inhibition of oxidative damage and tau hyperphosphorylation. Food Chem Toxicol 46, 28812887.Google Scholar
76. Kim, DS, Park, SY & Kim, JK (2001) Curcuminoids from Curcuma longa L (Zingiberaceae) that protect PC12 rat pheochromocytoma and normal human umbilical vein endothelial cells from beta A(1-42) insult. Neurosci Lett 303, 5761.Google Scholar
77. Shi, X, Zheng, Z, Li, J, et al. (2015) Curcumin inhibits Abeta-induced microglial inflammatory responses in vitro: involvement of ERK1/2 and p38 signaling pathways. Neurosci Lett 594, 105110.Google Scholar
78. Zhang, X, Yin, WK, Shi, XD, et al. (2011) Curcumin activates Wnt/β-catenin signaling pathway through inhibiting the activity of GSK-3B in APPswe transfected SY5Y cells. Eur J Pharm Sci 42, 540546.Google Scholar
79. Olivia, CA, Vargas, JY & Inestrosa, NC (2013) Wnt signaling: role in LTP, neural networks and memory. Ageing Res Rev 12, 786800.Google Scholar
80. Purro, SA, Dickins, EM & Salinas, PC (2012) The secreted Wnt antagonist Dickkopf-1 is required for amyloid B-medicated synaptic loss. J Neurosci 32, 34923498.Google Scholar
81. Wan, W, Xia, S, Kalionis, B, et al. (2014) The role of Wnt signaling in the development of Alzheimer’s disease: a potential therapeutic target. Biomed Res Int 2014, 19.Google Scholar
82. Parr, C, Mirzaei, N & Christian, M (2015) Activation of the Wnt/beta-catenin pathway represses the transcription of the beta-amyloid precursor protein cleaving enzyme (BACE-1) via binding of T-cell factor-4 to BACE1 promoter. FASEB J 29, 623635.CrossRefGoogle Scholar
83. Khan, MA, Akhtar, N, Sharma, V, et al. (2015) Product development studies on sonocrystallized curcumin for the treatment of gastric cancer. Pharmaceutics 7, 4363.CrossRefGoogle ScholarPubMed
84. Bates, KA, Verdile, G, Li, QX, et al. (2009) Clearance mechanisms of Alzheimer’s amyloid-beta peptide: implications for therapeutic design and diagnostic tests. Mol Psychiatry 14, 469486.Google Scholar
85. Foster, JK, Verdile, G, Bates, KA, et al. (2009) Immunization in Alzheimer’s disease: naive hope or realistic clinical potential? Mol Psychiatry 14, 239251.Google Scholar
86. Fiala, M, Lin, J, Ringman, J, et al. (2005) Ineffective phagocytosis of amyloid-beta by macrophages of Alzheimer’s disease patients. J Alzheimers Dis 7, 221232.Google Scholar
87. Zhang, L, Fiala, M, Cashman, J, et al. (2006) Curcuminoids enhance amyloid-beta uptake by macrophages of Alzheimer’s disease patients. J Alzheimers Dis 10, 17.Google Scholar
88. Fiala, M, Liu, PT, Espinosa-Jeffrey, A, et al. (2007) Innate immunity and transcription of MGAT-II and toll-like receptors in Alzheimer’s disease patients are improved by bisdemethoxycurcumin. Proc Natl Acad Sci U S A 104, 1284912854.Google Scholar
89. Masoumi, A, Goldenson, B, Ghirmai, S, et al. (2009) 1 alpha,25-Dihydroxyvitamin D3 interacts with curcuminoids to stimulate amyloid-beta clearance by macrophages of Alzheimer’s disease patients. J Alzheimers Dis 17, 703717.Google Scholar
90. Mizwicki, MT, Menegaz, D, Barrientos-Durán, A, et al. (2012) Genomic and nongenomic signaling induced by 1α,25(OH) 2-vitamin D3 promotes the recovery of amyloid-β phagocytosis by Alzheimer’s disease macrophages. J Alzheimers Dis 29, 5162.Google Scholar
91. Soni, 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 Scholar
92. Soudamini, KK, Unnikrishnan, MC, Soni, KB, et al. (1992) Inhibition of lipid peroxidation and cholesterol levels in mice by curcumin. Indian J Physiol Pharmacol 36, 239243.Google Scholar
93. Sreejayan, N & Rao, MNA (1994) Curcuminoids as potent inhibitors of lipid peroxidation. J Pharm Pharmacol 46, 10131016.Google Scholar
94. Peschel, D, Koerting, R & Nass, N (2007) Curcumin induces changes in expression of genes involved in cholesterol homeostasis. J Nutr Biochem 18, 113119.Google Scholar
95. Feng, D, Ohlsson, L & Duan, RD (2010) Curcumin inhibits cholesterol uptake in Caco-2 cells by down-regulation of NPC1L1 expression. Lipids Health Dis 9, 4045.Google Scholar
96. Kim, M & Kim, Y (2010) Hypocholesterolemic effects of curcumin via up-regulation of cholesterol 7a-hydroxylase in rats fed a high fat diet. Nutr Res Pract 4, 191195.Google Scholar
97. Tu, Y, Sun, D, Zeng, X, et al. (2014) Piperine potentiates the hypocholesterolemic effect of curcumin in rats fed on a high fat diet. Exp Ther Med 8, 260266.Google Scholar
98. Zambón, D, Quintana, M, Mata, P, et al. (2010) Higher incidence of mild cognitive impairment in familial hypercholesterolemia. Am J Med 123, 267274.Google Scholar
99. Refolo, LM, Pappola, MA, Malester, B, et al. (2000) Hypercholesterolemia accelerates the Alzheimer’s amyloid pathology in a transgenic mouse model. Neurobiol Dis 7, 690691.Google Scholar
100. Xiao, Z, Zhang, A, Lin, J, et al. (2014) Telomerase: a target for therapeutic effects of curcumin and a curcumin derivative in Abeta insult in vitro . PLOS ONE 9, e101251.Google Scholar
101. Lim, GP, Chu, T, Yang, F, et al. (2001) The curry spice curcumin reduces oxidative damage and amyloid pathology in an Alzheimer transgenic mouse. J Neurosci 21, 83708377.Google Scholar
102. Wang, YJ, Thomas, P, Zhong, JH, et al. (2009) Consumption of grape seed extract prevents amyloid-β deposition and attenuates inflammation in brain of an Alzheimer’s disease mouse. Neurotox Res 15, 314.Google Scholar
103. Ma, QL, Yang, F, Rosario, ER, et al. (2009) Beta-amyloid oligomers induce phosphorylation of tau and inactivation of insulin receptor substrate via c-Jun N-terminal kinase signaling: suppression by omega-3 fatty acids and curcumin. J Neurosci 29, 90789089.Google Scholar
104. Frautschy, SA, Hu, W, Kim, P, et al. (2001) Phenolic anti-inflammatory antioxidant reversal of Abeta-induced cognitive deficits and neuropathology. Neurobiol Aging 22, 9931005.Google Scholar
105. Fenech, M & Thomas, P (2010) Grape seed polyphenols and curcumin reduce genomic instability events in a transgenic mouse model for Alzheimer’s disease. Alzheimers Dement 6, S70S72.Google Scholar
106. Belviranli, M, Okudan, N, Atalik, K, et al. (2013) Curcumin improves spatial memeory and decreases oxidative damage in aged female rats. Biogerontology 14, 187196.Google Scholar
107. Ahmed, T, Enam, SA & Gilani, AH (2010) Curcuminoids enhance memory in an amyloid-infused rat model of Alzheimer’s disease. Neuroscience 169, 12961306.Google Scholar
108. Kim, SJ, Son, TG, Park, HR, et al. (2008) Curcumin stimulates proliferation of embryonic neural progenitor cells and neurogenesis in the adult hippocampus. J Biol Chem 283, 1449714505.Google Scholar
109. Gerenu, G, Liu, K, Chojnacki, JE, et al. (2015) Curcumin/melatonin hybrid 5-(4-hydroxy-phenyl)-3-oxo-pentanoic acid [2-(5-methoxy-1H-indol-3-yl)-ethyl]-amide ameliorates AD-like pathology in the APP/PS1 mouse model. ACS Chem Neurosci 6, 13931399.Google Scholar
110. Yanagisawa, D, Ibrahim, NF, Taguchi, H, et al. (2015) Curcumin derivative with the substitution at C-4 position, but not curcumin, is effective against amyloid pathology in APP/PS1 mice. Neurobiol Aging 36, 201210.Google Scholar
111. Dong, S, Zeng, Q, Mitchell, ES, et al. (2012) Curcumin enhances neurogenesis and cognition in aged rats: implications for transcriptional interactions related to growth and synaptic plasticity. PLOS ONE 7, e31211.Google Scholar
112. Ghoneim, AI, Abdel-Naim, AB, Khalifa, AE, et al. (2002) Protective effects of curcumin against ischaemia/reperfusion insult in rat forebrain. Pharmacol Res 46, 273279.Google Scholar
113. Thiyagarajan, M & Sharma, SS (2004) Neuroprotective effect of curcumin in middle cerebral artery occlusion induced focal cerebral ischemia in rats. Life Sci 74, 969985.Google Scholar
114. Jiang, J, Wang, W, Sun, YJ, et al. (2007) Neuroprotective effect of curcumin on focal cerebral ischemic rats by preventing blood-brain barrier damage. Eur J Pharmacol 561, 5462.Google Scholar
115. Wang, YF, Gu, YT, Qin, GH, et al. (2013) Curcumin ameliorates the permeability of the blood–brain barrier during hypoxia by upregulating heme oxygenase-1 expression in brain microvascular endothelial cells. J Mol Neurosci 51, 344351.Google Scholar
116. Tsai, YM, Chien, CF, Lin, LC, et al. (2011) Curcumin and its nano-formulation: the kinetics of tissue distribution and blood-brain barrier penetration. Int J Pharm 416, 331338.Google Scholar
117. Cheng, KK, Yeung, CF, Ho, SW, et al. (2013) Highly stabilized curcumin nanoparticles tested in an in vitro blood-brain barrier model and in Alzheimer’s disease Tg2576 mice. AAPS J 15, 324336.Google Scholar
118. Anand, P, Nair, HB, Sung, B, et al. (2010) Design of curcumin loaded PLGA nanoparticles formulation with enhanced cellular uptake, and increased bioactivity in vitro and superior bioavailability in vivo . Biochem Pharmacol 79, 330338.CrossRefGoogle ScholarPubMed
119. Zona, C & La Ferla, B (2011) Synthesis of labeled curcumin derivatives as tools for in vitro blood brain barrier trafficking studies. J Label Compd Radiopharm 54, 629632.Google Scholar
120. Re, F, Gregori, M & Masserini, M (2012) Nanotechnology for neurodegenerative disorders. Nanomed Nanotechnol 8, S51S58.Google Scholar
121. Shahani, K, Swaminathan, SK, Freeman, D, et al. (2010) Injectable sustained release microparticles of curcumin: a new concept for cancer chemoprevention. Cancer Res 70, 44434452.Google Scholar
122. Chiu, SS, Lui, E, Majeed, M, et al. (2011) Differential distribution of intravenous curcumin formulations in the rat brain. Anticancer Res 31, 907911.Google Scholar
123. Shaikh, J, Ankola, DD, Beniwal, V, et al. (2009) Nanoparticle encapsulation improves oral bioavailability of curcumin by at least 9-fold when compared to curcumin administered with piperine as absorption enhancer. Eur J Pharm Sci 37, 223230.Google Scholar
124. Ahmed, T & Gilani, AH (2009) Inhibitory effect of curcuminoids on acetylcholinesterase activity and attenuation of scopolamine-induced amnesia may explain medicinal use of turmeric in Alzheimer’s disease. Pharmocol Biochem Behav 91, 554559.Google Scholar
125. Gauthier, S & Molinuevo, JL (2013) Benefits of combined cholinesterase inhibitor and memantine treatment in moderate-severe AD. Alzheimers Dement 9, 326331.Google Scholar
126. Peeyush, KT, Antony, S, Sonan, S, et al. (2011) Role of curcumin in the prevention of cholinergic mediated cortical dysfunctions in streptozotocin-induced diabetic rats. Mol Cell Endocrinol 331, 110.Google Scholar
127. Rajasekar, N, Dwivedi, S, Tota, SK, et al. (2013) Neuroprotective effect of curcumin on okadaic acid induced memory impairment in mice. Eur J Pharmacol 715, 381394.Google Scholar
128. Tiwari, V & Chopra, K (2013) Protective effect of curcumin against chronic alcohol-induced cognitive deficits and neuroinflammation in the adult rat brain. Neuroscience 6, 147158.Google Scholar
129. Orhan, IE (2013) Nature: a substantial source of auspicious substances with acetylcholinesterase inhibitory action. Curr Neuropharmacol 11, 379387.Google Scholar
130. Bulteau, AL, Moreau, M, Saunois, A, et al. (2006) Algae extract-mediated stimulation and protection of proteasome activity within human keratinocytes exposed to UVA and UVB irradiation. Antioxid Redox Signal 8, 136143.Google Scholar
131. Cole, GM, Teter, B & Frautschy, SA (2007) Neuroprotective effects of curcumin. Adv Exp Med Biol 595, 197212.Google Scholar
132. Ikonomovic, MD, Abrahamson, EE, Uz, T, et al. (2008) Increased 5-lopoxygenase immunoreactivity in the hippocampus of patients with Alzheimer’s disease. J Histochem Cytochem 56, 10651073.Google Scholar
133. Qu, J, Uz, T & Manev, H (2000) Inflammatory 5-LOX mRNA and protein are increased in brain of aging rats. Neurobiol Aging 21, 647652.Google Scholar
134. Firuzi, O, Zhuo, J, Chinnici, CM, et al. (2008) 5-Lipoxygenase gene disruption reduces amyloid-β pathology in a mouse model of Alzheimer’s disease. FASEB J 22, 11691178.Google Scholar
135. Ohno, M (2014) Roles of eIF2α kinases in the pathogenesis of Alzheimer’s disease. Front Mol Neurosci 7, 18.Google Scholar
136. Valera, E, Dargusch, R, Maher, PA, et al. (2013) Modulation of 5-lipoxygenase in proteotoxicity and Alzheimer’s disease. J Neurosci 33, 1051211052.CrossRefGoogle ScholarPubMed
137. Baum, L, Lam, C, Cheung, S, et al. (2008) Six-month randomized, placebo-controlled, double-blind, pilot clinical trial of curcumin in patients with Alzheimer disease. J Clin Psychopharm 28, 110114.Google Scholar
138. Ringman, JM, Frautschy, SA, Cole, GM, et al. (2005) A potential role of the curry spice curcumin in Alzheimer’s disease. Curr Alzheimer Res 2, 131136.Google Scholar
139. Commandeur, J & Vermeulen, N (1996) Cytotoxicity and cytoprotective activities of natural compounds. The case of curcumin. Xenobiotica 26, 667680.Google Scholar
140. Ganguli, M, Chandra, V, Kamboh, MI, et al. (2000) Apolipoprotein E polymorphism and Alzheimer disease: the Indo-US cross-national dementia study. Arch Neurol 57, 824830.Google Scholar
141. Chandra, V, Pandav, R, Dodge, H, et al. (2001) Incidence of Alzheimer’s disease in a rural community in India The Indo-US study. Neurology 57, 985989.Google Scholar
142. Shaji, S, Bose, S & Verghese, A (2005) Prevalence of dementia in an urban population in Kerala, India. B J Psychiatry 186, 136140.Google Scholar
143. Ng, TP, Chiam, PC, Lee, T, et al. (2006) Curry consumption and cognitive function in the elderly. Am J Epidemiol 164, 898906.Google Scholar
144. Cox, KH, Pipingas, A & Scholey, AB (2015) Investigation of the effects of solid lipid curcumin on cognition and mood in a healthy older population. J Psychopharmacol 29, 642651.Google Scholar
145. Ringman, JM, Frautschy, SA, Teng, E, et al. (2012) Oral curcumin for Alzheimer’s disease: tolerability and efficacy in a 24-week randomized, double blind, placebo-controlled study. Alzheimer’s Res Ther 4, 4351.Google Scholar
146. DiSilvestro, RA, Joseph, E, Zhao, S, et al. (2012) Diverse effects of a low dose supplement of lipidated curcumin in healthy middle aged people. Nutr J 11, 7987.Google Scholar
147. van Stegerena, A, Rohlederb, N, Everaerda, W, et al. (2006) Salivary alpha amylase as marker for adrenergic activity during stress: effect of betablockade. Psychoneuroendocrinology 31, 137141.Google Scholar
148. Gallacher, DV & Petersen, OH (1983) Stimulus-secretion coupling in mamalian salivary glands. Intern Rev Physiol 28, 152.Google Scholar
149. Chatterton, RT, Vogelsong, KM, Lu, YC, et al. (1996) Salivary alpha-amylase as a measure of endogenous adrenergic activity. Clin Physiol 16, 433448.Google Scholar
150. Hishikawa, N, Takahashi, Y, Amakusa, Y, et al. (2012) Effects of turmeric on Alzheimer’s disease with behavioral and psychological symptoms of dementia. Ayu 33, 499504.Google Scholar
151. Mastroeni, D, Grover, A, Delvaux, E, et al. (2011) Epigenetic mechanisms in Alzheimer’s disease. J Neurobiol Aging 32, 11611180.Google Scholar
152. Daniilidou, M, Koutroumani, M & Tsolaki, M (2011) Epigenetic mechanisms in Alzheimer’s disease. Curr Med Chem 18, 17511756.Google Scholar
153. Balazs, R, Vernon, J & Hardy, J (2011) Epigenetic mechanisms in Alzheimer’s disease:progress but much to do. Neurobiol Aging 32, 11811187.Google Scholar
154. Chouliaras, L, Rutten, BPF, Kenis, G, et al. (2010) Epigenetic regulation in the pathophysiology of Alzheimer’s disease. Prog Neurobiol 90, 498510.Google Scholar
155. Teiten, MH, Dicato, M & Diederich, M (2013) Curcumin as a regulator of epigenetic events. Mol Nutr Food Res 57, 16191629.Google Scholar
156. Reuter, S, Gupta, SC, Park, B, et al. (2011) Epigenetic changes induced by curcumin and other natural compounds. Genes Nutr 6, 93108.Google Scholar
157. Du, L, Xie, Z, Wu, LC, et al. (2012) Reactivation of RASSF1A in breast cancer cells by curcumin. Nutr Cancer 64, 12281235.Google Scholar
158. Sezgin, Z & Dincer, Y (2014) Alzheimer’s disease and epigenetic diet. Neurochem Int 78, 105116.Google Scholar
159. Li, YJ, Xu, M, Gao, ZH, et al. (2013) Alterations of serum levels of BDNF-related miRNAs in patients with depression. PLOS ONE 8, e63648.Google Scholar
160. Lopresti, AL, Maes, M, Marker, GL, et al. (2014) Curcumin for the treatment of major depression: a randomised, double-blind, placebo controlled study. J Affect Disord 167, 368375.Google Scholar
161. Davinelli, S, Calabrese, V, Zella, D, et al. (2014) Epigenetic nutraceutical diets in Alzheimer’s disease. J Nutr Health Aging 18, 800805.Google Scholar
162. Ansari, R, Mahta, A, Mallack, E, et al. (2014) Hyperhomocysteinemia and neurologic disorders: a review. J Clin Neurol 10, 281288.Google Scholar
163. Mattson, MP & Shea, TB (2003) Folate and homocysteine metabolism in neural plasticity and neurodegenerative disorders. Trends Neurosci 26, 137146.Google Scholar
164. Fux, R, Kloor, D, Hermes, M, et al. (2005) Effect of acute hyperhomocysteinemia on methylation potential of erthrocytes and on DNA methylation of lymphocytes in healthy male volunteers. Am J Renal Physiol 289, F786F792.Google Scholar
165. Ataie, A, Sabetkasaei, M, Haghparast, A, et al. (2010) Curcumin exerts neuroprotective effects against homocysteine intracerebroventricular injection-induced cognitive impariement and oxidative stress in rat brain. J Med Food 13, 821826.Google Scholar
166. Cheng, AL, Hsu, CH, Lin, CH, et al. (2001) Phase I clinical trial of curcumin, a chemopreventative agent, in patients with high-risk or pre-malignant lesions. Anticancer Res 21, 28952900.Google Scholar
167. Gupta, SC, Kismali, G & Aggarwal, BB (2013) Curcumin, a component of turmeric: from farm to pharmacy. Biofactors 39, 213.Google Scholar
168. Pan, CJ, Tang, JJ, Weng, YJ, et al. (2006) Preparation, characterization and anticoagulation of curcumin-eluting controlled biodgradable coating stents. J Control Release 116, 249.Google Scholar
169. Dong-Chan, K, Sae-Kwang, K & Jong-Sup, B (2012) Anticoagulant activities of curcumin and its derivative. BMB Rep 45, 221226.Google Scholar
170. Belkacemi, A, Doggui, S, Dao, L, et al. (2011) Challenges associated with curcumin therapy in Alzheimer disease. Exp Rev Mol Med 13, e34.Google Scholar
171. Sharma, RA, Steward, WP & Gescher, AJ (2007) The molecular targets and therapeutic uses of curcumin in health and disease. Adv Exp Med Biol 595, 453470.Google Scholar
172. Ireson, C, Orr, S & Jones, DJL (2001) Characterization of metabolites of the chemopreventative agent curcumin in human and rat hepatocytes and in the rat in vivo, and evaluation of their ability to inhibit phorbol ester-induced prosaglandin E2 production. Cancer Res 61, 10581064.Google Scholar
173. Benny, M & Anthony, B (2006) Bioavailability of Biocurcumax™(BCM-95™). Research and Development Laboratory, Arjuna Natural Extracts Ltd, Binanipuram.Google Scholar
174. Pescosolido, N, Giannotti, R, Plateroti, AM, et al. (2014) Curcumin: therapeutical potential in ophthalmology. Planta Med 80, 249254.Google Scholar
175. Rachmawati, H, Budiputra, DK & Mauludin, R (2014) Curcumin nanoemulsion for transdermal application: formulation and evaluation. Drug Dev Ind Pharm 41, 560566.Google Scholar
176. McClure, R, Yanagisawa, D, Stec, D, et al. (2015) Inhalable curcumin: offering the potential for translation to imaging and treatment of Alzheimer’s disease. J Alzheimers Dis 44, 283295.Google Scholar
177. Antony, B, Merina, B, Iyer, VS, et al. (2008) A pilot cross-over study to evaluate human oral bioavailability of BCM-95CG (Biocurcumax), a novel bioenhanced preparation of curcumin. Indian J Pharm Sci 70, 445449.Google Scholar
178. Merina, B & Antony, B (2006) Bioavailability of Biocurcumax (BCM-095). Spice India 2, 1116.Google Scholar
179. Jayaraj, RL, Elangovan, N, Dhanalakshmi, C, et al. (2014) CNB-001, a novel pyrazole derivative mitigates motor impairments associated with neurodegeneration via suppression of neuroinflammatory and apoptotic response in experimental Parkinson’s disease mice. Chem Biol Interact 220, 149157.Google Scholar
180. Maher, P, Akaishi, T, Schubert, D, et al. (2010) A pyrazole derivative of curcumin enhances memory. Neurobiol Aging 31, 706709.Google Scholar
181. Liu, Y, Dargusch, R, Maher, P, et al. (2008) A broadly neuroprotective derivative of curcumin. J Neurochem 105, 13361345.Google Scholar
182. Prasad, S, Tyagi, AK & Aggarwal, BB (2014) Recent developments in delivery, bioavailability, absorption and metabolism of curcumin: the golden pigment from golden spice. Cancer Res Treat 46, 218.Google Scholar
183. Stockert, JC, Del Castillo, P, Gomez, A, et al. (1989) Fluorescence reaction of chromatin by curcumin. Z Naturforsch C 44, 327329.Google Scholar
184. Wang, F, Wu, X, Wang, F, et al. (2006) The sensitive fluorimetric method for the determination of curcumin using the enhancement of mixed micelle. J Fluoresc 16, 5359.Google Scholar
185. Ryu, EK, Choe, YS, Lee, KH, et al. (2006) Curcumin and dehydrozingerone derivatives: synthesis, radiolabeling, and evaluation for β-amyloid plaque imaging. J Med Chem 49, 61116119.Google Scholar
186. Mohorko, N, Repovš, G, Popovic, M, et al. (2010) Curcumin labeling of neuronal fibrillar tau inclusions in human brain samples. J Neuropathol Exp Neurol 69, 405414.Google Scholar
187. Koronyo-Hamaoui, M, Koronyo, Y, Ljubimov, AV, et al. (2011) Identification of amyloid plaques in retinas from Alzheimer’s patients and noninvasive in vivo optical imaging of retinal plaques in a mouse model. Neuroimage, S204S217.Google Scholar
188. Kayabasi, U, Sergott, RC & Rispoli, M (2014) Retinal examination for the diagnosis of Alzheimer’s disease. Int J Ophthalmol Clin Res 1, 14.Google Scholar
189. Cheng, KK, Chan, PS, Fan, S, et al. (2015) Curcumin-conjugated magnetic nanoparticles for detecting amyloid plaques in Alzheimer’s disease mice using magnetic resonance imaging (MRI). Biomaterials 44, 155172.Google Scholar
190. Patil, R, Gangalum, PR, Wagner, S, et al. (2015) Curcumin targeted, polymalic acid-based MRI contrast agent for the detection of abeta plaques in Alzheimer’s disease. Macromol Biosci 15, 12121217.Google Scholar
191. Frost, S, Martins, RN & Kanagasingam, Y (2010) Ocular biomarkers for early detection of Alzheimer’s disease. J Alzheimers Dis 22, 116.Google Scholar
Figure 0

Fig. 1 Curcumin: reported mechanisms of action. BACE1, β-APP-cleaving enzyme-1; Aβ, β amyloid; APP, amyloid precursor protein.

Figure 1

Table 1 Studies using curcumin in Alzheimer’s disease (AD): diagnosis, prevention and treatment