Hostname: page-component-8448b6f56d-gtxcr Total loading time: 0 Render date: 2024-04-23T19:30:42.079Z Has data issue: false hasContentIssue false

Diet, nutrients and metabolism: cogs in the wheel driving Alzheimer's disease pathology?

Published online by Cambridge University Press:  10 April 2015

Rhona Creegan*
Affiliation:
Centre of Excellence for Science, Seafood and Health, Curtin University, 7 Parker Place, Technology Park, WA6102, Australia Curtin University, GPO Box U1987, Perth, WA6845, Australia
Wendy Hunt
Affiliation:
Centre of Excellence for Science, Seafood and Health, Curtin University, 7 Parker Place, Technology Park, WA6102, Australia Curtin University, GPO Box U1987, Perth, WA6845, Australia
Alexandra McManus
Affiliation:
Centre of Excellence for Science, Seafood and Health, Curtin University, 7 Parker Place, Technology Park, WA6102, Australia Curtin University, GPO Box U1987, Perth, WA6845, Australia
Stephanie R. Rainey-Smith
Affiliation:
Centre of Excellence for Alzheimer's Disease Research and Care, School of Medical Sciences, Edith Cowan University, 270 Joondalup Drive, Joondalup, WA6027, Australia Sir James McCusker Alzheimer's Disease Research Unit (Hollywood Private Hospital), 115 Monash Avenue, Nedlands, WA6009, Australia
*
*Corresponding author: Dr R. Creegan, fax +61 8 9341 1615, email rhobru@iinet.net.au
Rights & Permissions [Opens in a new window]

Abstract

Alzheimer's disease (AD), the most common form of dementia, is a chronic, progressive neurodegenerative disease that manifests clinically as a slow global decline in cognitive function, including deterioration of memory, reasoning, abstraction, language and emotional stability, culminating in a patient with end-stage disease, totally dependent on custodial care. With a global ageing population, it is predicted that there will be a marked increase in the number of people diagnosed with AD in the coming decades, making this a significant challenge to socio-economic policy and aged care. Global estimates put a direct cost for treating and caring for people with dementia at $US604 billion, an estimate that is expected to increase markedly. According to recent global statistics, there are 35·6 million dementia sufferers, the number of which is predicted to double every 20 years, unless strategies are implemented to reduce this burden. Currently, there is no cure for AD; while current therapies may temporarily ameliorate symptoms, death usually occurs approximately 8 years after diagnosis. A greater understanding of AD pathophysiology is paramount, and attention is now being directed to the discovery of biomarkers that may not only facilitate pre-symptomatic diagnosis, but also provide an insight into aberrant biochemical pathways that may reveal potential therapeutic targets, including nutritional ones. AD pathogenesis develops over many years before clinical symptoms appear, providing the opportunity to develop therapy that could slow or stop disease progression well before any clinical manifestation develops.

Type
Review Article
Copyright
Copyright © The Authors 2015 

Alzheimer's disease (AD), the most common form of dementia, is a chronic, progressive neurodegenerative disease that manifests clinically as a slow global decline in cognitive function, including deterioration of memory, reasoning, abstraction, language and emotional stability, culminating in a patient with end-stage disease, totally dependent on custodial care( Reference Blennow, de Leon and Zetterberg 1 , Reference Roses, Alberts and Strittmatter 2 ). The greatest risk factor for AD is age and so with an ageing population, it is predicted that there will be a marked increase in the number of AD cases in the coming decades, in both the developed and developing world( Reference Brookmeyer, Johnson and Ziegler-Graham 3 , Reference Ferri, Prince and Brayne 4 ).

In the UK, 850 000 people are expected to have dementia by 2015, and recent estimates from the Alzheimer's society (UK) state that only 44 % of dementia cases receive a diagnosis. In Australia, approximately 300 000 people are suffering from dementia, with this number expected to be close to 1 million by 2050( 5 ). Current estimates put a direct cost of $AUD3·2 billion to the healthcare system in Australia, with a predicted increase to $AUD6 billion within the next 5 years( 6 ). This represents a major economic burden to an already stretched healthcare budget. According to the Alzheimer's Association in the USA, unpaid care by family members and caregivers in 2009 represented $144 billion. Globally, it is expected that the incidence of AD will quadruple by 2050; according to these estimates, one in eighty-five individuals will be suffering from AD by 2050( Reference Roses, Alberts and Strittmatter 2 ).

Currently, there is no cure for AD; while current therapies may temporarily ameliorate symptoms, death usually occurs approximately 8 years after diagnosis. These alarming statistics emphasise the importance of gaining a greater understanding of the pathophysiology of AD. Attention is now being directed to the discovery of biomarkers, which may not only facilitate pre-symptomatic diagnosis, but also provide an insight into aberrant biochemical pathways that may reveal potential therapeutic targets, including nutritional targets, which could slow or stop disease progression well before any clinical symptoms manifest. The present review provides a detailed discussion of the myriad of integrated nutritional factors probably contributing to the pathogenesis of AD. Specific emphasis is placed on abnormal lipid metabolism.

Hallmarks of disease

The causes and factors leading to disease progression are poorly understood; however, AD pathology is characterised by the presence of β-amyloid (Aβ) deposits and neurofibrillary tangles in the cerebral cortex and sub-cortical grey matter. There is much debate in the literature as to which of these hallmarks are causative and which result from other pathophysiological processes. Multiple factors have been implicated in the aetiology and pathogenesis of AD. These factors include genetic defects, abnormal lipid metabolism, energy metabolism deficits and mitochondrial defects, inflammation, abnormal amyloid precursor protein (APP) processing, deficiency of neurotrophic factors, glutamate excitotoxicity, free radical-induced neuron degeneration, and trace element toxicity, all of which are influenced by diet and nutrition. It is unlikely that any of these factors are isolated, but when combined with imbalances in neurotransmitters and hormones as well as impaired hepatic metabolic and detoxification pathways, a complex web of dysfunctional biochemistry is established, resulting in characteristic AD pathology( Reference Kidd 7 ).

Amyloid plaques

Amyloid plaques are a characteristic feature of AD and contain aggregates of Aβ, a 4 kDa peptide of thirty-nine to forty-three amino acids normally found in the brain, albeit at low concentrations( Reference Verdile, Fuller and Atwood 8 ). Aβ is a protein fragment cleaved by the action of secretase enzymes from its larger parent protein called APP. APP is located in the plasma (outer) membrane of brain cells, including neurons, glial cells and the endothelial cells of the blood–brain barrier (BBB). APP is an abundant protein in the central nervous system (CNS), which is also ubiquitously expressed in the peripheral tissues such as epithelium, muscle and circulating cells( Reference Borroni, Akkawi and Martini 9 ).

APP undergoes proteolytic cleavage by enzymes termed BACE (β-site amyloid precursor protein-cleaving enzyme) and α- and γ-secretases. These enzymes act via two competing pathways that depend on the localisation of APP to specific membrane domains and feedback loops controlling gene transcription for the appropriate enzymes (reviewed in Krishnaswamy et al. ( Reference Krishnaswamy, Verdile and Groth 10 )).

One product of APP cleavage is Aβ, of which two main isoforms are secreted; Aβ40 and Aβ42 consisting of forty and forty-two amino acids, respectively. Aβ42 accounts for approximately 10 % of secreted Aβ, but is the main form found in amyloid plaques, suggesting a more pathological role for Aβ42 due to its ability to aggregate and polymerise into amyloid fibrils more readily than Aβ40 ( Reference Verdile, Fuller and Atwood 8 ). An overproduction or reduced clearance of Aβ is considered a key component of AD, and alterations in Aβ kinetics may contribute to the induction of inflammation, oxidation and neurotoxicity, turning a physiologically relevant protein into a potentially toxic one( Reference De Strooper 11 ). Whether Aβ is the cause of AD or its deposition merely represents a protective response to some other pathophysiological process is an intense subject of debate( Reference Castellani, Lee and Siedlak 12 ).

Neurofibrillary tangles

Affected areas of the AD brain have significantly lower neuronal numbers, and the remaining neurons possess reduced numbers of dendrite branches and synaptic densities. Neurofibrillary tangles arise within individual neurons as deposits of abnormal fibrils, the primary constituent of which is the microtubule-associated protein tau. The tau protein in neurons is normally bound to microtubules, which provide structural integrity and are involved in intracellular transport( Reference Ballatore, Lee and Trojanowski 13 ). Phosphorylation of tau is part of the normal process of assembly of microtubules, conferring stability. However, in AD, tau becomes hyper-phosphorylated or glycosylated, thereby weakening its affinity for microtubules. Once dissociation occurs, tau forms the filamentous double helical structures that characterise the paired helical filaments of neurofibrillary tangles. The tangles adversely affect neuronal function and result in loss of intracellular communication( Reference Small and Duff 14 Reference Iqbal, Liu and Gong 16 ). The phosphorylation and dephosphorylation of tau, as with other proteins, is tightly regulated by various kinases and phosphatases. Dysregulation of these processes, as occurs with defects in insulin and other signalling pathways, may contribute to the accumulation of neurofibrillary tangles.

Abnormal lipid metabolism

Abnormal lipid metabolism is emerging as a very important pathophysiological process in the development of AD. This is logical as the adverse effects of abnormal lipid metabolism on neuronal biochemistry are numerous, affecting membrane lipid composition and a myriad of cellular signals generated by lipid mediators. The link between AD and lipid metabolism was firmly established when carriage of the APOEɛ4 allele was identified as a major risk factor for AD( Reference Corder, Saunders and Strittmatter 17 ). Lipidomic studies have identified specific lipid changes in AD, including phospholipids, sphingomyelins, cholesterol and ceramides. Also, lipid metabolism has been intimately connected to processing of APP, leading to increased generation of Aβ( Reference Di Paolo and Kim 18 Reference Han and Gross 20 ). These studies have also demonstrated the relevance of particular lipid alterations, such as ceramides, to general metabolic dysfunction and insulin resistance (IR)( Reference Frisardi, Panza and Seripa 21 ). Diet and lifestyle factors generate signals to a complex network of hormones and transcription factors that orchestrate metabolism by sensing cellular energy requirements and influencing anti-ageing and pro-survival proteins, such as the sirtuins and AMP kinase. There are many integrated factors that derail normal lipid metabolism such as abnormal hormone signalling, inflammation, oxidation, altered neurotransmitters and abnormal hepatic metabolic pathways, all of which are profoundly influenced by diet nutrient status.

In recent years, numerous reports have highlighted the strong relationship between dementia and metabolic disorders that include dyslipidaemia, obesity, diabetes, CVD and hypertension (reviewed in Farooqui et al. ( Reference Farooqui, Farooqui and Panza 22 ), Frisardi & Imbimbo( Reference Frisardi and Imbimbo 23 ), Craft( Reference Craft 24 ), Merlo et al. ( Reference Merlo, Spampinato and Canonico 25 ), Luchsinger( Reference Luchsinger 26 ) and Moreira( Reference Moreira 27 )). While the mechanisms that underpin this association are beyond the scope of this article, it is important to emphasise that the aforementioned conditions rarely occur in isolation; the complex network of metabolic dysfunction that is associated with these conditions has been shown to influence many aspects of AD pathogenesis and neurodegeneration.

Insulin resistance

Poor diet and lifestyle choices can result in the development of IR, which often precedes the development of metabolic disease. IR is the failure of insulin to affect glucose disposal in muscle and adipose tissue as well as the failure to inhibit gluconeogenesis in the liver, resulting in elevations of blood glucose. Due to the tight relationship between glucose and lipid homeostasis, lipid abnormalities will also co-exist. When IR develops, increased circulating NEFA result from increased adipose tissue lipolysis as insulin normally inhibits hormone-sensitive lipase. The increased NEFA delivered to the liver eventually lead to increased hepatic VLDL secretion (elevated plasma TAG) and increased plasma cholesterol. A ‘lipotoxic’ state results from abnormal accumulation of TAG and fatty acids in the muscle and liver, which exacerbates IR by inhibiting insulin receptor substrate cascades( Reference Kim, Fillmore and Chen 28 Reference Yu, Chen and Cline 30 ). Additionally, the increased expression of the inflammatory cytokine TNF-α by adipose tissue in obesity both impairs phosphorylation of the insulin receptor and enhances the release of NEFA( Reference Borst 31 , Reference Ruan and Lodish 32 ). Insulin has also been shown to induce the expression of fatty acid desaturase in monocytes, and may provide an explanation of the increased inflammation promoted by postprandial hyperinsulinaemia when the dietary n-6:n-3 ratio is high, resulting in the production of inflammatory eicosanoids from n-6 fatty acids( Reference Arbo, Halle and Malik 33 ). Systemic inflammation is a key contributor to AD, particularly with respect to vasculature and integrity of the BBB( Reference Lopez-Ramirez, Wu and Pryce 34 ).

Insulin is required for blood glucose regulation in both the periphery and the brain. When peripheral IR develops, metabolic dysfunction results and brain insulin signalling is altered, affecting glucose utilisation, Aβ and tau pathology, vasculature, mitochondrial function, inflammation, oxidation, neuronal maintenance and plasticity( Reference Craft 24 , Reference Craft 35 Reference Shoelson, Lee and Goldfine 37 ). A neurodegenerative cycle consisting of AD pathogenesis, IR and inflammation has often been described. The components of this cycle have elevated ceramide levels as a common factor, suggesting that these toxic lipids may represent the link between all these pieces of the AD puzzle( Reference Chavez and Summers 38 , Reference Summers 39 ) (see Fig. 1).

Fig. 1 Ceramides – the toxic intermediate linking metabolic dysfunction, inflammatory cytokines and insulin resistance? When adipose tissue exceeds its storage capacity, adipokines increase inflammation that increases ceramides. This inhibits insulin signalling, further increasing lipolysis and increasing the release of fatty acids for ceramide synthesis. Ceramide promotes apoptosis and elevated SFA inhibit the B-cell lymphoma 2 (Bcl2) anti-apoptotic protein family of anti-apoptotic proteins. iNOS, inducible nitric oxide synthase; IRS-1, insulin receptor substrate-1; PI3-K, phosphatidylinositol-3 kinase; Akt/PKB, Akt also known as protein kinase B, a serine/threonine-specific protein kinase. A colour version of this figure can be found online at http://www.journals.cambridge.org/bjn

Diet and nutrition

Dietary habits and nutritional status are emerging as key components of chronic degenerative diseases, including AD. This is further compounded by the fact that AD is an age-related disorder with a general decline in digestive function, absorptive capacity and assimilation of nutrients.

Atrophic gastritis is common in older people and can affect absorption of key nutrients such as vitamin B12 and folate( Reference van Asselt, de Groot and van Staveren 40 , Reference Krasinski, Russell and Samloff 41 ), which have been associated with hyperhomocysteinaemia and increased risk of AD( Reference Obeid and Herrmann 42 , Reference Tucker, Qiao and Scott 43 ). Atrophic gastritis can also affect protein absorption and, therefore, the supply of essential amino acids, such as tryptophan and tyrosine, which are required for the synthesis of neurotransmitters such as serotonin and catecholamines( Reference Cropper, Smith and Groff 44 ).

There is an age-related decline in metabolic rate, coupled with a decrease in physical activity; elderly people tend to eat less and may consume inadequate micronutrients. Additionally, co-morbidities often exist, requiring pharmaceutical agents that may further compromise nutrient status. The well-established risk factors for developing AD are elevated cholesterol and dyslipidaemia, obesity, diabetes, hypertension, depression, CVD and cerebrovascular disease (reviewed in Polidori et al. ( Reference Polidori, Pientka and Mecocci 45 ) and Patterson et al. ( Reference Patterson, Feightner and Garcia 46 )), all of which are biochemically connected and are influenced by diet and possible nutrient deficiencies.

An increase in AD incidence has been reported in populations who have a high intake of saturated fat, trans-fatty acids, refined sugar, processed foods, and total energy( Reference Luchsinger and Mayeux 47 ). High fish consumption is inversely correlated with the development of dementia, and moderate alcohol consumption appears to offer a protective effect( Reference Luchsinger and Mayeux 47 Reference Schiepers, de Groot and Jolles 49 ).

Many studies have confirmed the link between high-fat/high-sugar diets and declining cognitive function, strongly suggesting a role for IR and diet-induced endocrine abnormalities( Reference Craft 35 , Reference Brayne, Gao and Matthews 50 Reference Greenwood and Winocur 52 ). Diets that contain high saturated fat, cholesterol, added sugar, including high-fructose maize syrup, and high-glycaemic load foods contribute to dyslipidaemia( Reference Brand-Miller, Hayne and Petocz 53 , Reference Brand-Miller 54 ). Animal studies have shown that a diet high in fat and refined sugar influences brain structure and function via the regulation of neurotrophins( Reference Molteni, Barnard and Ying 55 ). Until recently, it was assumed that the adverse effects of such diets on AD were a direct result of the negative influence on insulin sensitivity, metabolism and cardiovascular health. While these are major contributing factors, these studies have shown a direct effect of such diets on the brain. Recent studies have also shown that hippocampal neurogenesis is adversely affected by high intakes of SFA( Reference Stangl and Thuret 56 , Reference Kanoski and Davidson 57 ).

Neuronal plasticity is the brain's ability to compensate for challenges by influencing synapse formation and neurite growth. A high-fat/high-sugar diet has been shown in animals to decrease brain plasticity via the regulation of brain-derived neurotrophic factor (BDNF)( Reference Molteni, Barnard and Ying 55 ), a major mediator of neuronal plasticity and a contributor to learning and memory capabilities( Reference Poo 58 , Reference Castren, Berninger and Leingartner 59 ). Animals that learn a spatial memory task faster have been shown to have more BDNF mRNA and protein in the hippocampus, and after feeding a high-fat/high-sugar diet for 2 months, the hippocampal levels of BDNF were reduced and the animals demonstrated reduced spatial learning performance( Reference Molteni, Barnard and Ying 55 ).

In contrast to the standard Western diet, characterised by high fat, high refined sugar, low fibre and high salt, the Mediterranean diet has been shown to reduce the risk for AD( Reference Scarmeas, Stern and Tang 60 , Reference Gu, Nieves and Stern 61 ). This eating style consists of high intakes of nutrient-dense, high-fibre foods such as vegetables, legumes, fruits and cereals. The intakes of high-fibre, non-refined carbohydrates contribute to a low-glycaemic load eating plan that helps to prevent IR and dyslipidaemia. Intakes of meat, dairy products and poultry (and, therefore, SFA and cholesterol) are low to moderate, and intake of fish and seafood is moderately high. Unsaturated fat intake in the form of monounsaturated fats from olive oil and n-3 fats from fish and seafood is high. Dairy products are consumed mainly in the form of cheese or yogurt, and there is regular consumption of red wine, usually with meals( Reference Scarmeas, Stern and Tang 60 , Reference Solfrizzi, Panza and Capurso 62 ).

The Mediterranean-style diet also contains numerous nutrients and plant phytochemicals that are anti-inflammatory, antioxidant and beneficial to health. For example, low dietary intake of antioxidants may contribute to increased oxidative stress, a feature of AD( Reference Luchsinger and Mayeux 47 , Reference Morris, Evans and Bienias 63 ). Some of these nutrients and plant phytochemicals are now being shown to have a powerful influence on gene transcription factors, such as SIRT1 (silent mating type information regulation 2 homologue 1) that influence energy homeostasis, lipid metabolism and possibly longevity( Reference Son, Camandola and Mattson 64 Reference Mattson 68 ). A meta-analysis involving twelve studies with a total of over 1·5 million people, followed for a period of 3–18 years, has shown that a greater adherence to the Mediterranean eating style was associated with a reduced risk of mortality and morbidity, including AD( Reference Sofi, Cesari and Abbate 69 ). The beneficial effects were observed in many markers of coagulation and inflammation, including homocysteine, C-reactive protein, IL-6, white cell count and fibrinogen. Blood lipids and blood pressure were also positively affected, all of which are risk factors for both CVD and AD( Reference Panagiotakos, Dimakopoulou and Katsouyanni 70 ).

Key macronutrients

Dietary fatty acids

Apart from dietary fatty acids being an energy-generating nutrient, the amount and type of fat is important for determining disease risk due to the diverse function of lipids in general; the structure and function of which is influenced by the fatty acids they contain. Fatty acids consumed affect plasma cholesterol levels differently due to the impact on the LDL receptor, the activity of which is regulated by the sterol content of the cell via sterol regulatory element-binding proteins (SREBP)( Reference Cropper, Smith and Groff 44 ). High saturated fat intake, particularly palmitic acid, is strongly associated with the development of dyslipidaemia, IR, obesity, diabetes, vascular disease and metabolic abnormalities, which as mentioned previously are all risk factors for AD( Reference Dietschy 71 Reference Woollett, Spady and Dietschy 73 ). Saturated and trans-fatty acid intakes are positively correlated with increased cholesterol levels and unfavourable shifts in LDL:HDL ratios( Reference Caggiula and Mustad 74 Reference Kris-Etherton, Yu and Etherton 76 ). Excess palmitic acid can also up-regulate ceramide production, which is considered a major contributor to dyslipidaemia and IR( Reference Chavez and Summers 38 , Reference Gill and Sattar 77 ). Emerging evidence suggests an association between IR and ceramides, where toxic ceramides may be the intermediate that links excess dietary SFA and inflammatory cytokines with IR( Reference Summers 39 ). This proposed relationship is depicted in Fig. 1. Additionally, studies have shown elevated ceramides in AD brain tissue( Reference Han 78 , Reference Han, Holtzman and McKeel 79 ). Moreover, the ganglioside monosialotetrahexosylganglioside (a sialylated glycoceramide) is thought to accelerate Aβ pathology by promoting the formation of insoluble fibrils( Reference Matsuzaki 80 ), and abnormal sphingolipid metabolism has been shown to enhance tau pathology( Reference He, Huang and Li 81 , Reference Grimm, Haupenthal and Rothhaar 82 ).

IR results in an atherogenic lipid profile that also includes a decrease in LDL particle size, reduced HDL and a postprandial accumulation of TAG-rich remnant lipoproteins. This has implications for vascular abnormalities in the brain.

Trans-fatty acids

Dietary trans-fatty acids are known to contribute to dyslipidaemia and associated increased disease risk( Reference Ascherio, Katan and Zock 83 ), and have been linked to brain ageing and impaired cognition( Reference Morris, Evans and Bienias 84 , Reference Bowman, Silbert and Howieson 85 ). Trans-fatty acids alter membrane fluidity and responses of various membrane receptors through their incorporation into membrane phospholipids. As fatty acids are ligands for nuclear receptors, such as PPAR, liver X receptor and SREBP, regulation of gene transcription can be altered( Reference Khan and Vanden Heuvel 86 Reference Vanden Heuvel 88 ), directly modulating metabolic and inflammatory responses in an adverse way.

PUFA

Unsaturated fatty acids are competitive substrates for the enzymes involved in PUFA metabolism( Reference Cropper, Smith and Groff 44 ). PUFA are converted via the action of cyclo-oxygenase and lipoxygenase enzymes to prostaglandins, leukotrienes, thromboxanes and other metabolites that are important mediators of cellular function; the signalling molecules generated, therefore, are partly influenced by dietary intakes( Reference Horrocks and Farooqui 89 ). Signalling molecules derived from n-6 PUFA are more inflammatory, atherogenic and pro-thrombotic than those derived from the n-3 series( Reference Cropper, Smith and Groff 44 ). α-Linolenic acid is the precursor for the important EPA and DHA; DHA is the most abundant PUFA in the CNS( Reference Sastry 90 ).

Deficiency of essential fatty acids is rare; however, conversion of α-linolenic acid by the action of elongase and desaturase enzymes to the longer-chain PUFA is affected by hormone imbalances, diet and nutrient deficiencies( Reference Bordoni, Hrelia and Lorenzini 91 Reference Horrobin 93 ). High intakes of n-6 fatty acids, including arachidonic acid (AA), typically found in meat can elevate pro-inflammatory eicosanoids and up-regulate pro-inflammatory cytokines. Levels of non-enzymatically derived isoprostanes, which are vasoconstrictive, are also elevated in AA-enriched diets( Reference Montuschi, Barnes and Roberts 94 ). Conversely, diets enriched in DHA from fish and fish oil are more anti-inflammatory, anti-thrombotic and vasodilatory, and have neuroprotective effects in terms of synaptic function and plasticity via the generation of docosanoids( Reference Horrocks and Farooqui 89 , Reference Oster and Pillot 95 ). The ratio of dietary AA:DHA would appear to influence several diseases, including AD, and may be important in designing nutrition-based strategies for disease prevention( Reference Simopoulos 96 ). The brain relies on a supply of AA and DHA from the periphery, which are delivered via plasma lipoproteins and lysophospholipids. Unesterified AA and DHA can also enter the brain where they are esterified into the sn-2 structural position of phospholipids following activation by long-chain fatty acyl-CoA synthase( Reference Cropper, Smith and Groff 44 ).

Low DHA levels may reflect inadequate intakes and/or the presence of abnormal biochemical pathways involving this compound. While DHA can be obtained from dietary sources, an adequate supply to the brain also relies on peroxisomal production. The elongation and desaturation of fatty acids occurs in the endoplasmic reticulum or microsomes; however, the conversion of EPA to DHA requires a final additional step that only occurs in peroxisomes( Reference Lord and Bralley 97 ). EPA undergoes two further elongation steps and a final δ-6 desaturation step occurs in the endoplasmic reticulum to yield tetracosahexaenoic acid (THA). This very-long-chain fatty acid is transported to the peroxisome for a final β-oxidation step to remove two carbons and yield DHA( Reference Infante and Huszagh 98 ). This additional step is suggested as a reason for inefficient conversion of the essential fatty acid α-linolenic acid to DHA, and why additional dietary intakes of DHA may be required to meet the needs of the brain. A recent study of AD patients has shown deficient liver biosynthesis of DHA; higher levels of THA, and lower expression of peroxisomal D-bifunctional protein (required for the conversion of THA to DHA) were detected in these individuals( Reference Astarita, Jung and Berchtold 99 ). Any accumulation of very-long-chain fatty acids such as THA can adversely affect mitochondrial function( Reference Kou, Kovacs and Hoftberger 100 ), with mitochondrial dysfunction and oxidative stress thought to be central to AD pathogenesis. Abnormal elongase and desaturase enzyme activity could be present in AD, resulting in a lower than normal production of longer-chain PUFA. This may then have multiple consequences including reduced membrane fluidity and an imbalance in PUFA-derived signalling molecules such as eicosanoids and docosanoids. Δ6-Desaturase acts as a gateway for the flow of fatty acids into the elongation and desaturation pathways; the activity of this enzyme may be influenced by dietary intakes of fatty acids, various nutrients such as Zn, Mg and vitamin B6 ( Reference Bordoni, Hrelia and Lorenzini 91 ) and metabolic hormones( Reference Cunnane 101 , Reference Cunnane 102 ). Furthermore, insulin is known to affect the activity of both Δ5- and Δ6-desaturases( Reference Das 92 ) and by reducing the supply of long-chain PUFA to the brain, may represent an additional mechanism by which metabolic dysfunction and associated IR could contribute to AD.

DHA and EPA are important in maintaining normal plasma TAG levels as they regulate the activity of various nuclear receptors resulting in a repartitioning of fatty acids away from storage as TAG and towards oxidation. These receptors include liver X receptor, hepatocyte nuclear factor 4α, farnesoid X-activated receptor, and PPAR. Each of these receptors is, in turn, regulated by SREBP-1c( Reference Davidson 103 ). Both EPA and DHA reduce SREBP-1c, which is the main genetic switch controlling lipogenesis, and thereby reduces the amount of NEFA available for VLDL synthesis. EPA and DHA are highly unsaturated and are prone to peroxidation, which stimulates the degradation of ApoB, required for VLDL synthesis, and their presence in lipoproteins may also enhance postprandial chylomicron clearance by stimulating lipoprotein lipase activity( Reference Davidson 103 ).

Dietary carbohydrates

The amount and type of dietary carbohydrate has a significant impact on lipid profiles and the risk of developing AD risk factors such as CVD and type 2 diabetes. Long term consumption of high glycaemic load, refined carbohydrates and simple sugars can lead to IR and the metabolic syndrome( Reference Brand-Miller 54 , Reference Barclay, Brand-Miller and Mitchell 104 ).

In 1992, the US Department of Agriculture recommended that no more than 40 g of extra sugars should be added to a standard 8368 kJ (2000-calorie)-a-day diet. The liver rapidly absorbs and metabolises fructose, and exposure to large amounts of fructose leads to lipogenesis and TAG accumulation. Fructose is phosphorylated by fructokinase to fructose-1-phosphate, which is then metabolised to triose phosphates, glyceraldehydes and dihydroxyacetone phosphate. A key factor of fructose metabolism is that the entry of fructose via fructose-1-phosphate bypasses regulation by allosteric inhibition of phosphofructokinase by citrate and ATP, which is the main rate-controlling step in glycolysis( Reference Havel 105 ). In this way, fructose continually gets converted to fat.

Fructose conversion to glycerol-3-phosphate esterifies with NEFA to form TAG. The accumulation of TAG contributes to reduced insulin sensitivity, hepatic IR and impaired glucose intolerance( Reference Basciano, Federico and Adeli 106 ). As mentioned above fructose has been shown to up-regulate lipogenesis( Reference Kok, Roberfroid and Delzenne 107 ). Glucose and insulin directly regulate lipogenesis as insulin controls SREBP expression that regulates fatty acid and cholesterol synthesis by the activation of pathways involving enzymes such as 3-hydroxy-3-methyl-glutaryl-CoA reductase and fatty acid synthase.

Animal models have been used to investigate the relationship between chronic ingestion of simple sugars and neurogenesis( Reference van der Borght, Kohnke and Goransson 108 ). This study has shown that sugar (either sucrose or fructose) feeding to rats reduced the number of newly mature neurons in the dentate gyrus, the most prolific neurogenesis region of the hippocampus, with the number of apoptotic cells also being significantly increased. Interestingly, when the feeds contained either glucose or fructose, enhanced proliferation of new neurons was observed, probably resulting from an attempt to compensate for the increased apoptosis. The authors of this study have also suggested that the observed elevated TNF-α levels reduced the survival of new neurons by promoting apoptosis and impairing BBB function. Elevated inflammatory cytokines and TAG can both compromise the integrity of the BBB( Reference Banks 109 ). Studies have shown that the adipokines leptin and gut-derived ghrelin stimulate adult neurogenesis in the regions of the hippocampus, hypothalamus and brain stem that regulate feeding, and the vagus nerves connecting the brain and gut( Reference Kokoeva, Yin and Flier 110 Reference Pierce and Xu 113 ). A compromised BBB can also prevent the delivery of neuroprotective leptin and ghrelin to the hippocampus, and reduce neurogenesis( Reference van der Borght, Kohnke and Goransson 108 ). These elegant animal studies have shown that these effects appear to be related to fructose and sucrose ingestion and are not seen with glucose only. The down-regulated hippocampal neurogenesis also appears to be independent of glucose levels, insulin, insulin-like growth factor-1 and cortisol. Fructose is thought to be metabolised mainly by the liver and kidney; however, it now appears that fructose affects neuronal function and neurogenesis( Reference Funari, Crandall and Tolan 114 ). This may provide a plausible link between increased fructose consumption in sweetened foods and drinks and the escalating rates of obesity and metabolic disease, which are risk factors for the development of AD.

Key micronutrients

Specific nutrient deficiencies in the elderly may exacerbate existing pathology in the brain, particularly in the presence of other risk factors( Reference Dauncey 115 ). The formation and maintenance of neurons relies on an adequate supply of the building blocks and cofactors required for normal functioning of biochemical and neurotransmitter pathways, all of which must be obtained from the diet( Reference Kamphuis and Scheltens 116 ). In addition, mitochondrial decay due to oxidation is a feature of brain ageing and neurodegenerative disease and an adequate supply of nutrients that protect mitochondrial enzymes and mitochondrial membranes is crucial to support cellular energy generation and prevent neurological decline( Reference Liu and Ames 117 ).

Vitamins, minerals and other metabolites act as critical cofactors for the synthesis of mitochondrial enzymes and other pathways, and, therefore, diets that supply inadequate amounts of micronutrients can accelerate mitochondrial decay and neurodegeneration( Reference Pieczenik and Neustadt 118 ). Nutrients supporting mitochondrial function include B vitamins, vitamin C, cysteine, ubiquinone (co-enzyme Q10), α-lipoic acid, sulphur, Fe, Cu, Zn, Mn and Mg( Reference Pieczenik and Neustadt 118 ). Deficiencies of several nutrients have been linked to dementia, including vitamins A, B, D and E, Mg, Zn and essential fatty acids.

There have been several studies that have examined the status of various nutrients in AD and mild cognitive impairment, a condition which often but not always precedes AD( Reference Conquer, Tierney and Zecevic 119 Reference Baldeiras, Santana and Proenca 122 ). Several interventional clinical trials of nutrient supplementation have also been conducted to examine the effect on cognition( Reference Luchsinger and Mayeux 47 , Reference Dysken, Sano and Asthana 123 Reference Morris, Evans and Schneider 128 ). These studies have provided inconsistent and sometimes conflicting results, which may be explained by poor study design, heterogeneous populations in terms of stage of disease and the complexities of nutrient interventions in an elderly population who have significant neurodegeneration.

Various nutrients influence lipid metabolism and also have specific effects on lipid-related AD pathology. The importance of methylation pathways in phospholipid, sphingolipid and sterol metabolism is discussed below, as dietary intakes of vitamins B6, B12 and folate influence these pathways. Additionally, methylation is a crucial facilitator of epigenetic modifications controlling gene expression, and can be adversely affected. Abnormal methylation pathways can also elevate plasma homocysteine levels, a risk factor for developing AD.

Water-soluble vitamins

B vitamins

The B vitamins are interconnected and intimately involved with lipid, carbohydrate and protein metabolism via their role as co-enzymes in mitochondrial ATP generation. The metabolism of vitamin B6 (pyridoxine) depends on both vitamins B2 (riboflavin) and B3 (niacin). Synthesis of niacin from tryptophan requires the activated form of vitamin B6 (pyridoxal-5′-phosphate, PLP) to act as a cofactor for the enzyme kyureninase( Reference Cropper, Smith and Groff 44 , Reference Smith, Marks, Lieberman, Smith, Marks and Lieberman 129 ). In AD the activity of tyrosine hydoxylase, which catalyses the conversion of tyrosine to L-dopa (the precursor for dopamine), has been shown to be reduced in key regions of the brain( Reference Sawada, Hirata and Arai 130 ). In addition, in vitro studies have shown that the redox co-enzyme NADH can increase the activity of tyrosine hydroxylase and dopamine in cells by 6-fold( Reference Vrecko, Birkmayer and Krainz 131 ). Vitamin B5 (pantothenic acid) is the precursor to CoA, which forms the key intermediate acetyl-CoA, the molecule that links the metabolism of lipids, proteins and carbohydrates together and is the gateway to energy generation via ATP in the tricarboxylic acid/oxidative phosphorylation cycle( Reference Cropper, Smith and Groff 44 ). Due to the crucial role of the B vitamins in cell biochemistry and mitochondrial function, it is clear that any dietary or lifestyle factor that reduces their availability, including poor absorption, excess alcohol consumption and medication use, increases the risk of neurological damage.

Vitamin B1 (thiamin)

Subclinical thiamin deficiency is not uncommon, particularly in the elderly( Reference O'Keeffe 132 ) or those with high alcohol consumption as alcohol interferes with the active transport of thiamin out of the intestinal cells( Reference Cropper, Smith and Groff 44 ). Numerous thiamin-dependent processes have been shown to be significantly reduced in the AD brain( Reference Gibson and Blass 133 , Reference Gibson, Sheu and Blass 134 ), and deficiency has been associated with dementia( Reference Wyatt, Nelson and Hillman 135 ). The effects of thiamin deficiency on neurological function can be explained in part by considering the enzymes that use thiamin as a cofactor. The thiamin-dependent enzymes transketolase, pyruvate dehydrogenase and α-keto-glutarate dehydrogenase have crucial roles in glucose metabolism and energy generation, and reductions in enzyme activity can result in nerve cell damage( Reference Singleton and Martin 136 ). A reduction in transketolase also decreases the availability of reducing substances such as NADPH via the pentose phosphate pathway, which is required for lipid synthesis and the removal of reactive oxygen species. A reduction in pentoses, required for nucleic acids, co-enzymes and polysaccharides, is also observed, which can further compromise cellular function( Reference Singleton and Martin 136 ). Thiamin deficiency also leads to increased lactic acid in the brain, and is probably a result of reduced pyruvate entering the tricarboxylic acid cycle due to diminished pyruvate dehydrogenase activity. The reduction in α-keto-glutarate dehydrogenase activity may contribute to the decreased levels of several neurotransmitters such as γ-amino butyric acid, glutamate and aspartate( Reference Langlais and Zhang 137 ). This suggests that N-methyl d-aspartate receptor-mediated excitotoxicity may be involved in neuronal damage observed with thiamin deficiency, although this is yet to be established. Any cause of inadequate ATP production in neurons will result in an inability to pump the neurotransmitter glutamate out of the synaptic gap which will, therefore, continue to stimulate the neuron. ATP is required to control Na and Ca pumps, the balance of which is critical for nerve conduction. It has been shown that brain tissue of Alzheimer's patients contains reduced activities of thiamin-dependent enzymes, such as α-keto-glutarate dehydrogenase( Reference Gibson, Sheu and Blass 134 , Reference Heroux, Raghavendra Rao and Lavoie 138 Reference Mastrogiacoma, Lindsay and Bettendorff 140 ). Increased levels of cerebrospinal fluid phosphorylated tau, a biomarker of AD, have been observed in cases of thiamin deficiency( Reference Matsushita, Miyakawa and Maesato 141 ). It has also been shown in both cellular and animal models that thiamin deficiency induced oxidative stress promotes amyloidogenic processing of APP and Aβ accumulation by regulating BACE1 maturation( Reference Zhang, Yang and Li 142 , Reference Karuppagounder, Xu and Shi 143 ). However, the relationship between thiamin deficiency and AD in humans needs further clarification.

Vitamin B2 (riboflavin) and vitamin B3 (niacin)

Riboflavin is a precursor of the co-enzymes FMN and FAD, which are crucial redox cofactors in mitochondrial ATP generation and other biochemical pathways( Reference Liu and Ames 117 ). There is substantial evidence that oxidative stress is a contributing factor in both the development and progression of AD( Reference Markesbery 144 , Reference Perry, Cash and Smith 145 ). Riboflavin is vital in reducing oxidised glutathione (a major intracellular antioxidant) and restoring its antioxidant capacity( 146 ). FAD is also a co-enzyme for methylene tetrahydrofolate reductase (MTHFR) that converts homocysteine to methionine and xanthine oxidase and that produces uric acid( Reference Liu and Ames 117 ). Elevated homocysteine and low levels of reduced glutathione and uric acid are linked to both ageing and cognitive decline( Reference Seshadri, Beiser and Selhub 147 Reference Ames, Cathcart and Schwiers 149 ).

Niacin or nicotinic acid is required for the formation of the ubiquitous mitochondrial redox co-enzymes NAD and NADP, and severe deficiency of niacin or its precursor tryptophan causes pellagra, of which dementia is a main feature( Reference Cropper, Smith and Groff 44 ). Although the exact mechanism linking niacin deficiency and dementia has not been established, niacin has been shown to be important for DNA synthesis and repair, myelination and dendritic growth, cellular Ca signalling, and acts as a potent antioxidant in the brain mitochondria( Reference Hageman and Stierum 150 Reference Morris, Evans and Bienias 153 ). Niacin has been included in preparations used in trials, where improvements in cognitive test scores have been reported( Reference Battaglia, Bruni and Ardia 154 ), and a large prospective study using FFQ to estimate niacin intake concluded that dietary niacin may protect against AD and age-related cognitive decline( Reference Morris, Evans and Bienias 153 ). However, further research into the role of specific nutrients to improve cognition is required as studies using FFQ have limitations, particularly in the elderly.

Vitamin B6 (pyridoxine), vitamin B12 (cobalamin) and folate

Vitamin B6 depletion has been implicated in mood disorders, depression and decline in cognitive function, and may contribute to AD disease progression( Reference Malouf and Grimley Evans 155 ). It has been estimated that approximately 10 % of the US population consumes less than half the RDA of vitamin B6 ( Reference Wakimoto and Block 156 ) and poor vitamin B6 status is generally common in the elderly population( Reference Malouf and Grimley Evans 155 ). PLP is the active form and is involved in amino acid metabolism of neurotransmitters such as γ-amino butyric acid, serotonin, dopamine and noradrenalin. Taurine synthesis from cysteine also has a PLP-dependent step and taurine is a neuromodulatory compound( Reference Cropper, Smith and Groff 44 ). PLP is a co-enzyme for several enzymes in one-carbon metabolism and trans-sulphuration. These reactions are crucial in providing methyl groups for the remethylation of homocysteine, a process that also requires adequate vitamin B12 and folate.

As mentioned previously, methylation reactions in the brain are critical and, therefore, an adequate supply of methyl groups provided by folic acid, vitamin B12 and SAMe (S-adenosylmethionine) is vital. The elderly may be compromised by dietary deficiency, poor absorption and inadequate inter-conversion of folates as occurs with polymorphisms in the MTHFR gene. The importance of vitamin B12 has been described in relation to methylation reactions, which is also crucial for the production and maintenance of myelin; demyelination is thought to be one mechanism via which vitamin B12 deficiency affects CNS function( 146 ). Additionally, adenosylcobalamin, a metabolite of vitamin B12 acts as a mitochondrial cofactor for the production of succinyl-CoA. This is important for converting odd-chain fatty acids (not normally present in cell membranes) from propionate to succinate for oxidation in the tricarboxylic acid cycle. This conversion is inhibited with vitamin B12 deficiency and can result in odd-chain fatty acids being incorporated into myelin, therefore affecting nerve transmission( Reference Coker, de Klerk and Poll-The 157 ). Folate status also appears to be associated with the tissue levels of DHA, which are concentrated in neural tissue and affect receptor function and cell signalling( Reference Umhau, Dauphinais and Patel 158 ). Animal studies have shown that concentrations of DHA in platelets, erythrocytes and intestinal phospholipids are increased in rats fed supplemental folate( Reference Pita and Delgado 159 ). Furthermore, animal studies have demonstrated that dietary folate deficiency causes depletion of DHA in neural tissue( Reference Hirono and Wada 160 ). When there is folate sufficiency however, the increases in DHA observed are likely to be due to the efficient transfer of methyl groups from SAMe to phosphatidylethanolamine (PE) to form phosphatidylcholine (PC). Conversion of PE to PC that occurs in the liver is catalysed by phosphatidylethanolamine N-methyltransferase (PEMT), the conversion of which relies on methyl transfer and PC is critical for mobilisation of DHA from the liver into the plasma( Reference Umhau, Dauphinais and Patel 158 ).

Homocysteine, choline and methylation pathways

Homocysteine is a sulphur-containing amino acid that exists at a critical biochemical intersection in the ubiquitous methionine cycle, the function of which is to generate one-carbon methyl groups for transmethylation reactions and synthesise cysteine and taurine( 146 ) (see Fig. 2). Methionine is converted to SAMe, which is the most important methyl donor in the body and is critical for stabilising many macromolecules such as myelin and DNA. Methylation of DNA is an epigenetic modification that controls gene expression, including the highly methylated APP gene( Reference Rogaev, Lukiw and Lavrushina 161 ). Methylation via SAMe is involved in the synthesis of numerous compounds including PC, melatonin, serotonin, noradrenalin, co-enzyme Q10 and carnitine( Reference Miller 162 ). Additionally, SAMe is a crucial component of both phase I and phase II detoxification pathways in the liver, providing the methyl group directly. Furthermore, a balanced methionine cycle is required to supply taurine for bile acid synthesis and cysteine for both sulphur conjugation and for the formation of glutathione, which is both an antioxidant and a vital phase I and II component( Reference Miller 162 ). All these processes are crucial for normal neurological and metabolic function.

Fig. 2 Methylation pathway depicting interaction with phospholipids and sphingolipids. After donating the methyl group, S-adenosylmethionine (SAMe) is converted into homocysteine via S-adenosylhomocysteine (SAH). Homocysteine is then broken down by one of three pathways. First, it can be converted back to methionine by accepting a methyl group from methylcobalamin (vitamin B12), and second, it can be converted to methionine by accepting a methyl group from trimethylglycine (betaine), or third, it can be converted to cysteine and taurine via serine and activated vitamin B6 ( 146 ). The catabolism of homocysteine depends on an adequate supply of vitamin B6, folate and vitamin B12. The majority of the essential nutrient choline is present in phosphatidylcholine (PC) and sphingomyelin (SM), major components of all cell membranes. Additionally, PC and SM are precursors for the signalling molecules ceramide, platelet-activating factor and sphingophosphorylcholine( Reference Zeisel and Blusztajn 265 ). Choline is required for the synthesis of the neurotransmitter, acetylcholine, and as it is oxidised to trimethylglycine, plays a crucial role as a methyl donor in the methionine/homocysteine pathway. PE, phosphatidylethanolamine; PEMT, phosphatidylethanolamine methyltransferase; ACh, acetylcholine; B6, vitamin B6; B12, vitamin B12; THF, tetrahydrofolate. A colour version of this figure can be found online at http://www.journals.cambridge.org/bjn

The link between elevated homocysteine and AD risk has been firmly established in several studies( Reference Seshadri, Beiser and Selhub 147 , Reference Lehmann, Gottfries and Regland 163 , Reference Nilsson, Gustafson and Hultberg 164 ), and homocysteine levels are also correlated to vitamin B6, B12 and folate status( Reference Nilsson, Gustafson and Hultberg 165 Reference Kwok, Tang and Woo 167 ); however, the exact mechanisms underlying the elevated homocysteine, connected metabolites and AD, remain to be fully elucidated. It is probable that elevated homocysteine promotes AD by more than one mechanism, and while homocysteine serves as a surrogate marker for nutrient status (which when deficient can promote neurological damage in its own right), evidence suggests that homocysteine has direct actions on the brain. Cerebral microangiopathy has been demonstrated in stroke-associated hyperhomocysteinaemia( Reference Evers, Koch and Grotemeyer 168 ); endothelial dysfunction, oxidative damage and neuronal DNA damage are all reported consequences of elevated homocysteine( Reference Chambers, Ueland and Obeid 169 Reference Ho, Ortiz and Rogers 171 ). There also appears to be an enhancement of Aβ-mediated neurotoxicity and apoptosis when homocysteine levels are elevated( Reference White, Huang and Jobling 172 , Reference Ho, Collins and Dhitavat 173 ), and homocysteic acid, which is a metabolite of homocysteine, is possibly an N-methyl d-aspartate agonist itself, causing excitotoxicity and apoptosis( Reference Kruman, Culmsee and Chan 174 , Reference Olney, Price and Salles 175 ).

The conversion of S-adenosylhomocysteine (SAH) to homocysteine is a reversible reaction, and so elevated homocysteine can also lead to an increase in SAH which is a potent inhibitor of methyltransferase enzymes (including PEMT required for PC production), further reducing the capacity of SAMe to participate in methylation reactions. This adversely affects numerous biochemical processes( Reference James, Melnyk and Pogribna 176 ). The APP gene is highly methylated and decreased methylation may increase expression leading to increased Aβ production( Reference Rogaev, Lukiw and Lavrushina 161 , Reference West, Lee and Maroun 177 ). A link between elevated SAH and PUFA metabolism has also been established in AD( Reference Selley 178 ). The connection between homocysteine and lipid metabolism has been further highlighted as homocysteine-induced endoplasmic reticulum stress interferes with phospholipid metabolism. This has the effect of activating SREBP associated with increased expression of genes involved in cholesterol and TAG uptake and intracellular accumulation of cholesterol( Reference Werstuck, Lentz and Dayal 179 ). A regulatory feedback circuit has been identified, where SREBP-1 controls the production of SAMe and therefore PC production, which is dependent on methylation reactions, thereby giving merit to the notion that nutritional or genetic factors limiting the production of SAMe or PC, may activate SREBP-1 and contribute to metabolic dysfunction( Reference Walker, Jacobs and Watts 180 ). The inhibition of methyltransferase enzymes by SAH is very important in the brain as the alternate pathway for homocysteine catabolism via trimethylglycine (betaine) has no activity( Reference Obeid and Herrmann 42 ) (see Fig. 2). CNS cells cannot export SAH and therefore conversion to homocysteine and extracellular transport is the only way to remove SAH, causing a subsequent rise in plasma homocysteine. This raises the possibility that SAH could be the toxic metabolite, at least in some tissues( Reference James, Melnyk and Pogribna 176 ), and suggests why the question of elevated homocysteine and increased disease risk has been difficult to answer with certainty.

Dietary fat and cholesterol are transported to the liver via chylomicrons and then packaged into VLDL in the liver for delivery to other tissues. PC is an essential component of these lipoproteins, and deficiency of choline and PC results in fat accumulation in the liver. Requirement for choline depends on the status of other methyl group donors such as folate and SAMe. When dietary choline is inadequate, the liver has a back-up pathway to provide choline from PE via a three-step methylation pathway involving SAMe, a reaction catalysed by PEMT( Reference Vance, Walkey and Cui 181 ). Deficiencies of nutrients involved in this pathway could reduce the availability of choline. Interestingly, oestrogen induces endogenous synthesis of choline by up-regulating PEMT, which may put postmenopausal women at risk of choline deficiency, when dietary choline intakes are inadequate and oestrogen levels are declining( Reference Resseguie, Song and Niculescu 182 ). Additionally, disturbed choline transport is suggested to play a role in various neurological disorders, including AD( Reference Michel, Yuan and Ramsubir 183 ). Loss of cholinergic neurons is a feature of AD and various choline transporters have been investigated as potential targets for increased choline delivery to the neurons for acetylcholine synthesis( Reference Michel, Yuan and Ramsubir 183 Reference Slotkin, Nemeroff and Bissette 185 ). Studies have shown that one of the choline transport systems in erythrocytes is abnormal in patients with AD( Reference Miller, Jenden and Cummings 186 ).

Fat-soluble vitamins

Vitamin A (retinol, retinal and retinoic acid)

Vitamin A exerts hormone-like activity via binding of its metabolite retinoic acid to nuclear receptors. This process regulates cell differentiation, proliferation and apoptosis in adults, and influences binding of many other nuclear receptors involved in a diverse number of processes, including lipid metabolism( 146 , Reference Balmer and Blomhoff 187 ). Studies in rats have shown that vitamin A deficiency decreases the liver content of phospholipids( Reference Khanna and Reddy 188 ), probably as a result of lower PC synthesis and reduced availability of fatty acids. This may be explained by a low activation of the transcription factor PPARα by its coactivator retinoid X receptor, which together play pivotal roles in the regulation of genes involved in lipid metabolism( Reference Oliveros, Domeniconi and Vega 189 ). This process may adversely affect liver function itself in terms of lipid metabolism and bile production, and may also affect the supply of crucial phospholipids to other organs, including the brain. Vitamin A plays key roles in α-secretase production, acetylcholine transmission and in the regulation of excessive microglial activation( Reference Koryakina, Aeberhard and Kiefer 190 ), and vitamin A insufficiency has been shown to contribute to these processes in AD( Reference Goodman and Pardee 191 , Reference Shudo, Fukasawa and Nakagomi 192 ). There are two binding sites for the retinoic acid receptor just upstream from the α-secretase gene ADAM-10 (A disintegrin and metalloproteinase domain-containing protein 10), and retinoic acid has been shown to up-regulate the expression of ADAM-10( Reference Tippmann, Hundt and Schneider 193 ), thereby increasing non-amyloidogenic APP processing. In animal studies, vitamin A-deficient mice have been shown to have impaired ADAM-10 transcription and the addition of vitamin A up-regulated both ADAM-10 and APP, resulting in the reduced formation of Aβ and increased formation of the neuroprotective secreted β-APP ectodomain APPsα( Reference Corcoran, So and Maden 194 ). Additionally, dietary deficiency of vitamin A in adult rats leads to an increase in the deposition of Aβ in cerebral blood vessels( Reference Husson, Enderlin and Delacourte 195 ). Retinoic acid insufficiency has also been connected to reduced production of acetylcholine transferase, which inhibits the neurotransmitter function of acetylcholine, an additional feature of AD( Reference Shudo, Fukasawa and Nakagomi 192 ). Furthermore, as mentioned previously, inflammation is a key feature of AD, and retinoic acid is a powerful modulator of immune function by reducing Aβ-induced inflammation via suppression of IL-6 and inhibition of TNF-α, and increased production of anti-inflammatory cytokines such as IL-10. The net effect is to reduce microglial expression of inducible NO synthase and hence activation. These effects are thought to be mediated by inhibiting the translocation of NF-κB( Reference Shudo, Fukasawa and Nakagomi 192 ). Additionally, retinol has a crucial role in mitochondria as it is an essential cofactor for protein kinase C-delta, which acts as a nutritional sensor to regulate energy homeostasis. The protein kinase C-delta/retinol complex signals the pyruvate dehydrogenase complex to increase influx of pyruvate into the tricarboxylic acid cycle to produce more ATP( Reference Acin-Perez, Hoyos and Zhao 196 ). Therefore, vitamin A deficiency may contribute to the hypometabolism observed in the AD brain( Reference Craft 24 ). Vitamin A insufficiency can result from poor dietary intakes of vitamin A-rich foods, as well as a result of two recently identified SNP in the gene that converts the vitamin A precursor, β-carotene, to retinol (β-carotene 15,15′-monoxygenase). These SNP are present in 25–40 % of the population and result in reduced enzyme activity and conversion to retinol( Reference Leung, Hessel and Meplan 197 ). Individuals carrying these SNP may benefit from higher intakes of pre-formed retinol and not rely on carotenoid sources from plant foods.

Vitamin D

Vitamin D, a cholesterol metabolite, has been implicated as a factor in AD, and recently the presence of a ligand-mediated vitamin D receptor pathway in the CNS was confirmed( Reference Buell and Dawson-Hughes 198 ). Studies have shown lower serum vitamin D levels in AD( Reference Sato, Asoh and Oizumi 199 , Reference Scott, Peter and Tucker 200 ) and higher parathyroid hormone levels( Reference Ogihara, Miya and Morimoto 201 ). In addition, the comorbidities of osteoporosis (associated with vitamin D deficiency) and AD often exist( Reference Luckhaus, Mahabadi and Grass-Kapanke 202 , Reference Tysiewicz-Dudek, Pietraszkiewicz and Drozdzowska 203 ), both conditions being provoked by inflammatory processes and IR. Reduced mRNA for the vitamin D receptor has been shown in specific hippocampal regions in AD compared with normal controls( Reference Sutherland, Somerville and Yoong 204 ) and a higher frequency of vitamin D receptor polymorphisms in AD has also been reported( Reference Gezen-Ak, Dursun and Ertan 205 ). Vitamin D may protect the structure and integrity of neurons through detoxification pathways, and 1,25-dihydroxy vitamin D3 (the active vitamin D metabolite) inhibits inducible NO synthase, and, therefore, reduces inflammation and oxidation, preventing excessive microglial activation. Vitamin D up-regulates γ-glutamyl transpeptidase and increases glutathione synthesis, a critical intracellular antioxidant( Reference Baas, Prufer and Ittel 206 ). Furthermore, neurotrophins, such as the docosanoid, neuroprotectin D1, and glial-derived neurotrophic factors are proteins necessary for neuronal survival, and vitamin D up-regulates their synthesis( Reference Buell and Dawson-Hughes 198 ). In AD, there is loss of hippocampal cells, which has been attributed to elevated voltage-gated Ca channels, reduced Ca-buffering capacity and glucocorticoid neurotoxicity. Vitamin D is a major regulator of Ca homeostasis and can protect against excitotoxicity( Reference Buell and Dawson-Hughes 198 ). Insulin sensitivity and signalling is also linked to vitamin D( Reference Alvarez and Ashraf 207 ). A study of over 2000 men has shown a correlation between vitamin D levels and bioavailable testosterone( Reference Wehr, Pilz and Boehm 208 ), and low levels of testosterone in elderly men has been linked to mild cognitive impairment and AD( Reference Chu, Tam and Lee 209 , Reference Hogervorst, Bandelow and Combrinck 210 ). Moreover, animal studies have shown that vitamin D is crucial for gonad function and the production of sex steroids( Reference Kinuta, Tanaka and Moriwake 211 ).

Vitamin E

Vitamin E levels have been shown to be lower in AD( Reference Jimenez-Jimenez, de Bustos and Molina 212 ), and decreasing serum levels have been associated with poor memory in the elderly( Reference Perkins, Hendrie and Callahan 213 ). In addition to accumulating in circulating lipoproteins, vitamin E is also transported in plasma by phospholipid transfer protein. Studies have shown lower cerebrospinal fluid phospholipid transfer protein activity levels in AD compared with normal healthy controls, and this may result in reduced vitamin E transport to the brain and increased oxidative stress( Reference Vuletic, Peskind and Marcovina 214 , Reference Desrumaux, Risold and Schroeder 215 ). In neuronal cultures, vitamin E inhibits Aβ-induced lipid peroxidation and cell death( Reference Yatin, Varadarajan and Butterfield 216 , Reference Butterfield, Koppal and Subramaniam 217 ). Apart from its role as an antioxidant, there is increasing evidence that vitamin E may regulate gene activity as gene array studies have connected vitamin E deficiency with altered gene expression in the hippocampus of rats( Reference Rota, Rimbach and Minihane 218 ). This study showed down-regulation of 948 genes including those affecting growth hormone, thyroid hormones, insulin-like growth factor, neuronal growth factor, melatonin, dopaminergic neurotransmission and clearance of advanced glycation end products( Reference Rota, Rimbach and Minihane 218 ). Also, genes coding for proteins related to Aβ clearance were strongly down-regulated in the presence of vitamin E deficiency( Reference Rota, Rimbach and Minihane 218 ). In addition to protecting against lipid peroxidation and modifying gene expression, recent studies have shown that vitamin E can block intracellular accumulation of ceramides and cholesterol, providing a further role for this vitamin( Reference Cutler, Kelly and Storie 219 ). However, trials using supplemental vitamin E have produced conflicting results( Reference Kidd 7 , Reference Luchsinger and Mayeux 47 ), perhaps due to the varying forms of vitamin E used; naturally occurring vitamin E is a mixture of α- and γ-tocopherols, while the commonly used commercial form (d-α-tocopherol) contains just one isomer with activity. Additionally, a single isolated nutrient is less likely to produce a positive outcome due to the synergistic nature of nutrients and the fact that dietary sources contain so many other compounds. This, however, is likely to remain a limitation with respect to many such studies and should not be considered as unique to trials of vitamin E.

Zinc: a key mineral

Zn is a cofactor for numerous enzymes involved in DNA replication, repair and transcription, and is also a cofactor for the activity of the major intracellular antioxidant enzyme, superoxide dismutase( 146 ). Zn deficiency is a common feature in the elderly population( Reference Briefel, Bialostosky and Kennedy-Stephenson 220 ), and dietary intakes of Zn and subsequent status may be influenced by other dietary factors and nutrients, such as Cu. Brain and cerebrospinal fluid levels of Zn have been shown to be depleted in AD, and serum Zn levels are inversely correlated with senile plaque count( Reference Tully, Snowdon and Markesbery 221 ). However, a paradox exists with the role of Zn in terms of AD risk and pathology( Reference Cuajungco and Faget 222 ). Plasma levels may indicate Zn deficiency; however, Zn homeostasis in the brain is regulated in a way that is not reflected by peripheral levels, and levels in various regions of the brain may actually be increased( Reference Loef, von Stillfried and Walach 223 ). Despite this conflicting situation, subclinical Zn deficiency, as a risk factor for AD, is supported by several studies( Reference Stoltenberg, Bush and Bach 224 Reference Bhatnagar and Taneja 226 ). Redox metals, including Zn, Fe and Cu, are implicated in the pathophysiology of AD, by facilitating the neurotoxicity of Aβ( Reference Huang, Atwood and Hartshorn 227 , Reference Huang, Cuajungco and Atwood 228 ). The paradoxical role of Zn can be related to both its function in the mitochondria and in haem synthesis, which binds to Aβ to prevent aggregation( Reference Liu and Ames 117 , Reference Atamna 229 , Reference Atamna and Boyle 230 ). On the one hand, high levels as occurs with excessive Zn release (flooding) in response to oxidative stress may potentiate Aβ toxicity and activate apoptotic processes( Reference Cuajungco and Faget 222 , Reference Sensi and Jeng 231 ), yet Zn is also involved in mechanisms attempting to neutralise oxidative stress via metalloproteins, such as superoxide dismutase( Reference Huang, Cuajungco and Atwood 228 ). The Zn situation is complex in AD, but there is no doubt that abnormal cellular Zn mobilisation occurs, and the increase in Zn concentrations observed in affected areas may reflect an increase in the amount of Zn/Cu superoxide dismutase, when, in fact, tissue Zn in unaffected areas is depleted( Reference Furuta, Price and Pardo 232 ). Furthermore, insulin-degrading enzyme, which is a Zn metallopeptidase, breaks down both Aβ and plasma insulin levels, and an isoform is also found in the mitochondria( Reference Leissring, Farris and Wu 233 ). Zn depletion could, therefore, reduce the activity of this enzyme and hinder the clearance of Aβ.

Zn is intimately connected to lipid and fatty acid metabolism. The animals that are raised with n-3 PUFA-deficient diets demonstrate an increased expression of the Zn transporter protein (ZnT3, which loads Zn into synaptic vesicles), with an associated increased Zn level in the hippocampus and a decreased plasma Zn level( Reference Jayasooriya, Ackland and Mathai 234 ). Studies using neuroblastoma cell lines exposed to DHA have shown a decrease in Zn uptake and in the expression of ZnT3, and an associated reduction in apoptosis when compared with cells grown in a DHA-enriched medium( Reference Suphioglu, De Mel and Kumar 235 ). Therefore, it could be speculated that in DHA-depleted cells, there is an increase in desaturase activity (required for DHA production), for which Zn is a cofactor, in an attempt to generate additional long-chain PUFA. The interaction of Zn, DHA and vitamin D (which enhances intestinal absorption of Zn) may also provide an explanation for the beneficial effect of fish consumption as fatty fish are rich sources of all three nutrients( Reference Potocnik, van Rensburg and Hon 236 ). The notion of such micronutrient synergy is supported by the systematic review of Loef et al. ( Reference Loef, von Stillfried and Walach 223 ), who concluded that while it is not currently possible to determine whether Zn supplementation confers benefit in the context of AD prevention or treatment, animal studies have suggested that the effect of dietary Zn on cognition is contingent upon the presence of additional nutrients.

Dietary phytochemicals

Components of food are now recognised as having a profound influence on signalling pathways involving energy metabolism and synaptic plasticity, thereby influencing cognitive function. It appears that at a molecular level, these phytochemicals activate adaptive cellular stress responses, a phenomenon known as hormesis. These hormetic pathways involve transcription factors and kinases that influence the expression of genes encoding antioxidant enzymes, hepatic detoxification enzymes, protein chaperones, anti-inflammatory proteins and neurotrophic factors( Reference Mattson 67 ). Examples of these hormetic pathways include SIRT1, NF-κB and nuclear factor-erythroid-2-related factor 2–antioxidant response element. The NAD-dependent deacetylase SIRT1 has been shown to attenuate amyloidogenic processing of APP, resulting in reduced production of Aβ in both cell culture and transgenic AD mouse models( Reference Wang, Fivecoat and Ho 237 , Reference Donmez, Wang and Cohen 238 ). This effect is mediated by an up-regulation of the transcription of α-secretase, by deacetylating and thereby activating the retinoic acid receptor that binds to the promotor of the ADAM-10 gene. Additionally, ADAM-10 also enhances the Notch pathway that up-regulates the transcription of genes involved in neurogenesis including BDNF( Reference Costa, Drew and Silva 239 ). Activation of SIRT1 increases the gene expression of antioxidant response proteins (forkhead box O proteins), reduces inflammatory proteins (NF-κB) and increases mitochondrial biogenesis via PPARγ coactivator-1α( Reference Bonda, Lee and Camins 240 ). Collectively, these effects result in reduced Aβ, reduced oxidative stress, reduced inflammation and enhanced resistance to Aβ-mediated toxicity and apoptosis.

Energy restriction( Reference Civitarese, Carling and Heilbronn 241 ) and n-3 fatty acids( Reference Xue, Yang and Wang 242 ) are known to activate SIRT1. It has also been reported that dietary supplementation with n-3 fatty acids was effective in reversing the reduced SIRT1 expression observed in rats with traumatic brain injury (notably, in humans, traumatic brain injury is proposed to increase cerebral Aβ levels and increase AD risk in later life)( Reference Wu, Ying and Gomez-Pinilla 243 Reference Nemetz, Leibson and Naessens 246 ). In addition, the diet provides polyphenolic plant compounds that activate sirtuins, and may, therefore, be protective against some of the pathological processes contributing to AD. These compounds may protect against AD pathology by both SIRT1-dependent and SIRT1-independent mechanisms. These polyphenols include resveratrol( Reference Borra, Smith and Denu 247 ) and its metabolite piceatannol( Reference Allard, Perez and Zou 248 ) found in grapes, wine, peanuts, blueberries, bilberries and cranberries. The polyphenol fisetin found in foods such as strawberries, apples and grapes has been shown to stimulate signalling pathways that enhance long-term memory( Reference Maher, Akaishi and Abe 249 ). Quercetin, a polyphenol found in many foods including apples, onions, citrus fruits, green vegetables and berries, has significant anti-inflammatory and antioxidant properties and in cell cultures has been shown to attenuate Aβ production( Reference Ansari, Abdul and Joshi 250 ). Furthermore, lutein and zeaxanthin are two carotenoids that can accumulate in retinal and brain tissue( Reference Craft, Haitema and Garnett 251 ), and lower brain levels, as assessed by retinal pigment, have been observed in individuals with mild cognitive impairment compared with cognitively normal subjects( Reference Johnson 252 ).

Several other plant phytochemicals such as curcumin (the active constituent of turmeric), catechins (green tea) and anthocyanins (berries and pomegranate) have been shown to attenuate amyloid and tau pathology in addition to their general antioxidant and anti-inflammatory properties( Reference Son, Camandola and Mattson 64 , Reference GM, Lim and Yang 253 ). Curcumin has been shown to target multiple sites in the AD cascade. Redox metals such as Fe have been implicated in the aggregation and toxicity of Aβ. The structure of curcumin confers metal-chelating activity and inhibits metal-catalysed lipid peroxidation( Reference Venkatesan and Rao 254 ). Furthermore, by the inhibition of c-Jun N-terminal kinase/activator protein-1 and NF-κB gene transcription, curcumin reduces the expression of inflammatory cytokines such as TNF-α, cyclo-oxygenase-2 and inducible NO synthase; elevations of which have been implicated in AD pathology( Reference Aggarwal, Kumar and Bharti 255 , Reference Lim, Chu and Yang 256 ). In turn, the reduction in inflammatory cytokines may limit the induction of BACE, thereby reducing the amyloidogenic processing of APP( Reference GM, Lim and Yang 253 ). However, despite the multimodal activities of curcumin, randomised placebo-controlled clinical trials of curcumin therapy conducted in early to moderate AD patients have, to date, failed, with no differences in the measures of cognition, blood Aβ, or cerebrospinal fluid Aβ and tau species observed( Reference Baum, Lam and Cheung 257 , Reference Ringman, Frautschy and Teng 258 ). Advanced cerebral pathology and limited curcumin bioavailability were cited as possible reasons for the lack of observed effects.

Studies comparing metal-associated (Cu and Zn) Aβ species demonstrated that the green tea extract epigallocatechin-3-gallate (EGCG) mitigated the toxicity of both free Aβ and metal-associated toxicity in cultured cells. In addition, aggregation of EGCG-bound Aβ species was attenuated, demonstrating anti-amyloidogenic properties of EGCG( Reference Hyung, DeToma and Brender 259 ). Whether EGCG can demonstrate such anti-amyloid capacity in humans, however, remains to be determined, and while oral dosing in humans has been shown to yield appreciable plasma EGCG concentrations( Reference Lee, Maliakal and Chen 260 ), its ability to cross the BBB and enter the CNS has not been proven.

Studies in animals and cell-culture models have also shown that anthocyanins have direct effects on APP processing and reducing Aβ toxicity( Reference Vepsalainen, Koivisto and Pekkarinen 261 Reference Ye, Meng and Yan 263 ). Notably, anthocyanin treatment has been shown to prevent memory impairment in a rodent model of AD, with the regulation of cholinergic neurotransmission implicated as a potential mechanism of action( Reference Gutierres, Carvalho and Schetinger 264 ). These promising biochemical and behavioural results suggest that additional in vivo and human studies are warranted.

Conclusion

The numerous integrated processes leading to the development and proliferation of AD are complex and involve genetics, epigenetics, metabolic dysfunction, hormonal circuits and toxic injury. Nutrition, by its influence on all of these contributing factors, is likely to have a profound impact on AD pathology. Therefore, understanding the effect of macro and micronutrients on neuronal biochemistry and AD pathology will provide opportunities for dietary manipulation to promote neuronal resistance to insults and reduce brain injury. The present review has highlighted some of the key nutrients involved both directly in neuronal biochemistry and indirectly by influencing peripheral metabolism, which, in turn, can promote AD pathology. Deficiencies or excesses of nutrients are unlikely to cause AD in their own right, but when combined with genetic factors and metabolic disease may accelerate existing pathology. AD pathology develops over many years before clinical symptoms appear, providing the opportunity to develop interventions, including nutritional approaches, which could slow or stop disease progression well before any clinical manifestations are apparent. Such interventions could include the use of intermittent energy restriction regimens, dietary phytochemicals and n-3 fatty acids to activate sirtuins, and help ameliorate poor metabolic health. Diets composed of energy-dense, nutrient-poor foods and containing high saturated fats, high simple sugars and low fibre will continue to drive metabolic disease and associated morbidities, including AD.

Micronutrient insufficiencies are often present with metabolic disease due to inadequate intakes and increased requirement as part of altered metabolism and medication use. In an ageing individual, intakes and assimilation of nutrients can be compromised, and may partly explain the conflicting results of clinical studies using nutrient interventions. In addition, many of these studies are poorly designed in terms of the stage of disease and extent of neuronal injury, and further trials need to be conducted that consider intervention before significant neurodegeneration is present. The profound impact of nutrition on the processes contributing to AD pathology cannot be denied; however, the lack of consistency in study data highlights the complexity of interpreting results from nutritional intervention trials. Future trials should account for the bioavailability of supplements used and the synergistic nature of nutrients in general: isolated nutrients are not present in food. Strategies need to be employed to reduce the predicted explosion in the number of AD cases, and as nutrition is part of the complex AD puzzle, one strategy might be the assignment of a specialist nutritionist as part of ongoing healthcare of an ageing individual, particularly in those with increased risk of developing this devastating disease.

Acknowledgements

The authors thank Curtin University and Edith Cowan University for facilitating this collaboration.

The present review received no specific grant from any funding agency, commercial or not-for-profit sectors.

The authors' contributions are as follows: R. C., W. H., A. Mc. and S. R. R.-S. wrote the article. R. C. had primary responsibility for the final content. All authors read and approved the final manuscript.

There are no conflicts of interest.

References

1 Blennow, K, de Leon, MJ & Zetterberg, H (2006) Alzheimer's disease. Lancet 368, 387403.Google Scholar
2 Roses, A, Alberts, M & Strittmatter, W (1992) Alzheimer's disease – reassessing the data. Curr Biol 2, 79.Google Scholar
3 Brookmeyer, R, Johnson, E, Ziegler-Graham, K, et al. (2007) Forecasting the global burden of Alzheimer's disease. Alzheimers Dement 3, 186191.Google Scholar
4 Ferri, CP, Prince, M, Brayne, C, et al. (2005) Global prevalence of dementia: a Delphi consensus study. Lancet 366, 21122117.Google Scholar
5 Dementia across Australia 2011–2050 (2011) Deloitte access economics for Alzheimer's Australia.Google Scholar
6 Access, delaying onset of Alzheimer's disease: predictions and issues (internet) 2004–2009 report no. 30.Google Scholar
7 Kidd, PM (2008) Alzheimer's disease, amnestic mild cognitive impairment, and age-associated memory impairment: current understanding and progress toward integrative prevention. Altern Med Rev 13, 85115.Google ScholarPubMed
8 Verdile, G, Fuller, S, Atwood, CS, et al. (2004) The role of beta amyloid in Alzheimer's disease: still a cause of everything or the only one who got caught? Pharmacol Res 50, 397409.Google Scholar
9 Borroni, B, Akkawi, N, Martini, G, et al. (2002) Microvascular damage and platelet abnormalities in early Alzheimer's disease. J Neurol Sci 203–204, 189193.Google Scholar
10 Krishnaswamy, S, Verdile, G, Groth, D, et al. (2009) The structure and function of Alzheimer's gamma secretase enzyme complex. Crit Rev Clin Lab Sci 46, 282301.CrossRefGoogle ScholarPubMed
11 De Strooper, B (2010) Proteases and proteolysis in Alzheimer disease: a multifactorial view on the disease process. Physiol Rev 90, 465494.Google Scholar
12 Castellani, RJ, Lee, HG, Siedlak, SL, et al. (2009) Reexamining Alzheimer's disease: evidence for a protective role for amyloid-beta protein precursor and amyloid-beta. J Alzheimers Dis 18, 447452.Google Scholar
13 Ballatore, C, Lee, VM & Trojanowski, JQ (2007) Tau-mediated neurodegeneration in Alzheimer's disease and related disorders. Nat Rev Neurosci 8, 663672.CrossRefGoogle ScholarPubMed
14 Small, SA & Duff, K (2008) Linking Abeta and tau in late-onset Alzheimer's disease: a dual pathway hypothesis. Neuron 60, 534542.Google Scholar
15 Liang, Z, Liu, F, Iqbal, K, et al. (2009) Dysregulation of tau phosphorylation in mouse brain during excitotoxic damage. J Alzheimers Dis 17, 531539.Google Scholar
16 Iqbal, K, Liu, F, Gong, CX, et al. (2009) Mechanisms of tau-induced neurodegeneration. Acta Neuropathol 118, 5369.CrossRefGoogle ScholarPubMed
17 Corder, EH, Saunders, AM, Strittmatter, WJ, et al. (1993) Gene dose of apolipoprotein E type 4 allele and the risk of Alzheimer's disease in late onset families. Science 261, 921923.CrossRefGoogle ScholarPubMed
18 Di Paolo, G & Kim, TW (2011) Linking lipids to Alzheimer's disease: cholesterol and beyond. Nat Rev Neurosci 12, 284296.CrossRefGoogle ScholarPubMed
19 Han, X (2010) Multi-dimensional mass spectrometry-based shotgun lipidomics and the altered lipids at the mild cognitive impairment stage of Alzheimer's disease. Biochim Biophys Acta 1801, 774783.Google Scholar
20 Han, X & Gross, RW (2005) Shotgun lipidomics: electrospray ionization mass spectrometric analysis and quantitation of cellular lipidomes directly from crude extracts of biological samples. Mass Spectrom Rev 24, 367412.Google Scholar
21 Frisardi, V, Panza, F, Seripa, D, et al. (2011) Glycerophospholipids and glycerophospholipid-derived lipid mediators: a complex meshwork in Alzheimer's disease pathology. Prog Lipid Res 50, 313330.CrossRefGoogle ScholarPubMed
22 Farooqui, AA, Farooqui, T, Panza, F, et al. (2012) Metabolic syndrome as a risk factor for neurological disorders. Cell Mol Life Sci 69, 741762.Google Scholar
23 Frisardi, V & Imbimbo, BP (2012) Metabolic-cognitive syndrome: metabolic approach for the management of Alzheimer's disease risk. J Alzheimers Dis 30, Suppl. 2, S1S4.Google Scholar
24 Craft, S (2009) The role of metabolic disorders in Alzheimer disease and vascular dementia: two roads converged. Arch Neurol 66, 300305.CrossRefGoogle ScholarPubMed
25 Merlo, S, Spampinato, S, Canonico, PL, et al. (2010) Alzheimer's disease: brain expression of a metabolic disorder? Trends Endocrinol Metab 21, 537544.Google Scholar
26 Luchsinger, JA (2012) Type 2 diabetes and cognitive impairment: linking mechanisms. J Alzheimers Dis 30, Suppl. 2, S185S198.CrossRefGoogle ScholarPubMed
27 Moreira, PI (2012) Alzheimer's disease and diabetes: an integrative view of the role of mitochondria, oxidative stress, and insulin. J Alzheimers Dis 30, Suppl. 2, S199S215.Google Scholar
28 Kim, JK, Fillmore, JJ, Chen, Y, et al. (2001) Tissue-specific overexpression of lipoprotein lipase causes tissue-specific insulin resistance. Proc Natl Acad Sci U S A 98, 75227527.CrossRefGoogle ScholarPubMed
29 Shulman, GI (2000) Cellular mechanisms of insulin resistance. J Clin Invest 106, 171176.CrossRefGoogle ScholarPubMed
30 Yu, C, Chen, Y, Cline, GW, et al. (2002) Mechanism by which fatty acids inhibit insulin activation of insulin receptor substrate-1 (IRS-1)-associated phosphatidylinositol 3-kinase activity in muscle. J Biol Chem 277, 5023050236.Google Scholar
31 Borst, SE (2004) The role of TNF-alpha in insulin resistance. Endocrine 23, 177182.Google Scholar
32 Ruan, H & Lodish, HF (2003) Insulin resistance in adipose tissue: direct and indirect effects of tumor necrosis factor-alpha. Cytokine Growth Factor Rev 14, 447455.Google Scholar
33 Arbo, I, Halle, C, Malik, D, et al. (2011) Insulin induces fatty acid desaturase expression in human monocytes. Scand J Clin Lab Invest 71, 330339.Google Scholar
34 Lopez-Ramirez, MA, Wu, D, Pryce, G, et al. (2014) MicroRNA-155 negatively affects blood–brain barrier function during neuroinflammation. FASEB J 28, 25512565.Google Scholar
35 Craft, S (2005) Insulin resistance syndrome and Alzheimer's disease: age- and obesity-related effects on memory, amyloid, and inflammation. Neurobiol Aging 26, Suppl. 1, 6569.Google Scholar
36 Chiu, SL, Chen, CM & Cline, HT (2008) Insulin receptor signaling regulates synapse number, dendritic plasticity, and circuit function in vivo . Neuron 58, 708719.Google Scholar
37 Shoelson, SE, Lee, J & Goldfine, AB (2006) Inflammation and insulin resistance. J Clin Invest 116, 17931801.CrossRefGoogle ScholarPubMed
38 Chavez, JA & Summers, SA (2012) A ceramide-centric view of insulin resistance. Cell Metab 15, 585594.Google Scholar
39 Summers, SA (2006) Ceramides in insulin resistance and lipotoxicity. Prog Lipid Res 45, 4272.CrossRefGoogle ScholarPubMed
40 van Asselt, DZ, de Groot, LC, van Staveren, WA, et al. (1998) Role of cobalamin intake and atrophic gastritis in mild cobalamin deficiency in older Dutch subjects. Am J Clin Nutr 68, 328334.Google Scholar
41 Krasinski, SD, Russell, RM, Samloff, IM, et al. (1986) Fundic atrophic gastritis in an elderly population. Effect on hemoglobin and several serum nutritional indicators. J Am Geriatr Soc 34, 800806.Google Scholar
42 Obeid, R & Herrmann, W (2006) Mechanisms of homocysteine neurotoxicity in neurodegenerative diseases with special reference to dementia. FEBS Lett 580, 29943005.Google Scholar
43 Tucker, KL, Qiao, N, Scott, T, et al. (2005) High homocysteine and low B vitamins predict cognitive decline in aging men: the Veterans Affairs Normative Aging Study. Am J Clin Nutr 82, 627635.Google Scholar
44 Cropper, SS, Smith, JL & Groff, JL (2005) Advanced Nutrition and Human Metabolism, 4th ed. Belmont, CA: Thomson Wadsworth.Google Scholar
45 Polidori, MC, Pientka, L & Mecocci, P (2012) A review of the major vascular risk factors related to Alzheimer's disease. J Alzheimers Dis 32, 521530.Google Scholar
46 Patterson, C, Feightner, J, Garcia, A, et al. (2007) General risk factors for dementia: a systematic evidence review. Alzheimers Dement 3, 341347.CrossRefGoogle ScholarPubMed
47 Luchsinger, JA & Mayeux, R (2004) Dietary factors and Alzheimer's disease. Lancet Neurol 3, 579587.CrossRefGoogle ScholarPubMed
48 Dosunmu, R, Wu, J, Basha, MR, et al. (2007) Environmental and dietary risk factors in Alzheimer's disease. Expert Rev Neurother 7, 887900.CrossRefGoogle ScholarPubMed
49 Schiepers, OJ, de Groot, RH, Jolles, J, et al. (2010) Fish consumption, not fatty acid status, is related to quality of life in a healthy population. Prostaglandins Leukot Essent Fatty Acids 83, 3135.Google Scholar
50 Brayne, C, Gao, L, Matthews, F, et al. (2005) Challenges in the epidemiological investigation of the relationships between physical activity, obesity, diabetes, dementia and depression. Neurobiol Aging 26, Suppl. 1, 610.Google Scholar
51 Convit, A (2005) Links between cognitive impairment in insulin resistance: an explanatory model. Neurobiol Aging 26, Suppl. 1, 3135.Google Scholar
52 Greenwood, CE & Winocur, G (2005) High-fat diets, insulin resistance and declining cognitive function. Neurobiol Aging 26, Suppl. 1, 4245.Google Scholar
53 Brand-Miller, J, Hayne, S, Petocz, P, et al. (2003) Low-glycemic index diets in the management of diabetes: a meta-analysis of randomized controlled trials. Diabetes Care 26, 22612267.CrossRefGoogle ScholarPubMed
54 Brand-Miller, JC (2003) Glycemic load and chronic disease. Nutr Rev 61, S49S55.Google Scholar
55 Molteni, R, Barnard, RJ, Ying, Z, et al. (2002) A high-fat, refined sugar diet reduces hippocampal brain-derived neurotrophic factor, neuronal plasticity, and learning. Neuroscience 112, 803814.Google Scholar
56 Stangl, D & Thuret, S (2009) Impact of diet on adult hippocampal neurogenesis. Genes Nutr 4, 271282.Google Scholar
57 Kanoski, SE & Davidson, TL (2011) Western diet consumption and cognitive impairment: links to hippocampal dysfunction and obesity. Physiol Behav 103, 5968.CrossRefGoogle ScholarPubMed
58 Poo, MM (2001) Neurotrophins as synaptic modulators. Nat Rev Neurosci 2, 2432.Google Scholar
59 Castren, E, Berninger, B, Leingartner, A, et al. (1998) Regulation of brain-derived neurotrophic factor mRNA levels in hippocampus by neuronal activity. Prog Brain Res 117, 5764.CrossRefGoogle ScholarPubMed
60 Scarmeas, N, Stern, Y, Tang, MX, et al. (2006) Mediterranean diet and risk for Alzheimer's disease. Ann Neurol 59, 912921.Google Scholar
61 Gu, Y, Nieves, JW, Stern, Y, et al. (2010) Food combination and Alzheimer disease risk: a protective diet. Arch Neurol 67, 699706.Google Scholar
62 Solfrizzi, V, Panza, F & Capurso, A (2003) The role of diet in cognitive decline. J Neural Transm 110, 95110.Google Scholar
63 Morris, MC, Evans, DA, Bienias, JL, et al. (2002) Dietary intake of antioxidant nutrients and the risk of incident Alzheimer disease in a biracial community study. JAMA 287, 32303237.Google Scholar
64 Son, TG, Camandola, S & Mattson, MP (2008) Hormetic dietary phytochemicals. Neuromolecular Med 10, 236246.Google Scholar
65 Gomez-Pinilla, F (2008) Brain foods: the effects of nutrients on brain function. Nat Rev Neurosci 9, 568578.Google Scholar
66 Gomez-Pinilla, F (2008) The influences of diet and exercise on mental health through hormesis. Ageing Res Rev 7, 4962.CrossRefGoogle ScholarPubMed
67 Mattson, MP (2008) Hormesis and disease resistance: activation of cellular stress response pathways. Hum Exp Toxicol 27, 155162.Google Scholar
68 Mattson, MP (2008) Dietary factors, hormesis and health. Ageing Res Rev 7, 4348.CrossRefGoogle ScholarPubMed
69 Sofi, F, Cesari, F, Abbate, R, et al. (2008) Adherence to Mediterranean diet and health status: meta-analysis. BMJ 337, a1344.Google Scholar
70 Panagiotakos, DB, Dimakopoulou, K, Katsouyanni, K, et al. (2009) Mediterranean diet and inflammatory response in myocardial infarction survivors. Int J Epidemiol 38, 856866.Google Scholar
71 Dietschy, JM (1998) Dietary fatty acids and the regulation of plasma low density lipoprotein cholesterol concentrations. J Nutr 128, Suppl. 2, 444S448S.Google Scholar
72 Woollett, LA, Spady, DK & Dietschy, JM (1992) Regulatory effects of the saturated fatty acids 6:0 through 18:0 on hepatic low density lipoprotein receptor activity in the hamster. J Clin Invest 89, 11331141.Google Scholar
73 Woollett, LA, Spady, DK & Dietschy, JM (1992) Saturated and unsaturated fatty acids independently regulate low density lipoprotein receptor activity and production rate. J Lipid Res 33, 7788.Google Scholar
74 Caggiula, AW & Mustad, VA (1997) Effects of dietary fat and fatty acids on coronary artery disease risk and total and lipoprotein cholesterol concentrations: epidemiologic studies. Am J Clin Nutr 65, Suppl. 5, 1597S1610S.Google Scholar
75 Kris-Etherton, PM & Yu, S (1997) Individual fatty acid effects on plasma lipids and lipoproteins: human studies. Am J Clin Nutr 65, Suppl. 5, 1628S1644S.CrossRefGoogle ScholarPubMed
76 Kris-Etherton, PM, Yu, S, Etherton, TD, et al. (1997) Fatty acids and progression of coronary artery disease. Am J Clin Nutr 65, 10881090.Google Scholar
77 Gill, JM & Sattar, N (2009) Ceramides: a new player in the inflammation-insulin resistance paradigm? Diabetologia 52, 24752477.Google Scholar
78 Han, X (2005) Lipid alterations in the earliest clinically recognizable stage of Alzheimer's disease: implication of the role of lipids in the pathogenesis of Alzheimer's disease. Curr Alzheimer Res 2, 6577.Google Scholar
79 Han, X, Holtzman, DM, McKeel, DW Jr, et al. (2002) Substantial sulfatide deficiency and ceramide elevation in very early Alzheimer's disease: potential role in disease pathogenesis. J Neurochem 82, 809818.Google Scholar
80 Matsuzaki, K (2010) Ganglioside cluster-mediated aggregation and cytotoxicity of amyloid beta-peptide: molecular mechanism and inhibition. Yakugaku Zasshi 130, 511515 (in Japanese).Google Scholar
81 He, X, Huang, Y, Li, B, et al. (2010) Deregulation of sphingolipid metabolism in Alzheimer's disease. Neurobiol Aging 31, 398408.Google Scholar
82 Grimm, MO, Haupenthal, VJ, Rothhaar, TL, et al. (2013) Effect of different phospholipids on alpha-secretase activity in the non-amyloidogenic pathway of Alzheimer's disease. Int J Mol Sci 14, 58795898.CrossRefGoogle ScholarPubMed
83 Ascherio, A, Katan, MB, Zock, PL, et al. (1999) Trans fatty acids and coronary heart disease. N Engl J Med 340, 19941998.Google Scholar
84 Morris, MC, Evans, DA, Bienias, JL, et al. (2003) Dietary fats and the risk of incident Alzheimer disease. Arch Neurol 60, 194200.Google Scholar
85 Bowman, GL, Silbert, LC, Howieson, D, et al. (2012) Nutrient biomarker patterns, cognitive function, and MRI measures of brain aging. Neurology 78, 241249.Google Scholar
86 Khan, SA & Vanden Heuvel, JP (2003) Role of nuclear receptors in the regulation of gene expression by dietary fatty acids (review). J Nutr Biochem 14, 554567.CrossRefGoogle ScholarPubMed
87 Vanden Heuvel, JP (2009) Cardiovascular disease-related genes and regulation by diet. Curr Atheroscler Rep 11, 448455.Google Scholar
88 Vanden Heuvel, JP (2004) Diet, fatty acids, and regulation of genes important for heart disease. Curr Atheroscler Rep 6, 432440.Google Scholar
89 Horrocks, LA & Farooqui, AA (2004) Docosahexaenoic acid in the diet: its importance in maintenance and restoration of neural membrane function. Prostaglandins Leukot Essent Fatty Acids 70, 361372.Google Scholar
90 Sastry, PS (1985) Lipids of nervous tissue: composition and metabolism. Prog Lipid Res 24, 69176.Google Scholar
91 Bordoni, A, Hrelia, S, Lorenzini, A, et al. (1998) Dual influence of aging and vitamin B6 deficiency on delta-6-desaturation of essential fatty acids in rat liver microsomes. Prostaglandins Leukot Essent Fatty Acids 58, 417420.Google Scholar
92 Das, UN (2010) A defect in Δ6 and Δ5 desaturases may be a factor in the initiation and progression of insulin resistance, the metabolic syndrome and ischemic heart disease in South Asians. Lipids Health Dis 9, 130.Google Scholar
93 Horrobin, DF (1981) Loss of delta-6-desaturase activity as a key factor in aging. Med Hypotheses 7, 12111220.Google Scholar
94 Montuschi, P, Barnes, P & Roberts, LJ 2nd (2007) Insights into oxidative stress: the isoprostanes. Curr Med Chem 14, 703717.CrossRefGoogle ScholarPubMed
95 Oster, T & Pillot, T (2010) Docosahexaenoic acid and synaptic protection in Alzheimer's disease mice. Biochim Biophys Acta 1801, 791798.Google Scholar
96 Simopoulos, AP (2008) The importance of the omega-6/omega-3 fatty acid ratio in cardiovascular disease and other chronic diseases. Exp Biol Med 233, 674688.Google Scholar
97 Lord, RS & Bralley, JA (2008) Laboratory Evaluations for Integrative and Functional Medicine, 2nd ed. Georgia: Metametrix Institute.Google Scholar
98 Infante, JP & Huszagh, VA (1997) On the molecular etiology of decreased arachidonic (20:4n-6), docosapentaenoic (22:5n-6) and docosahexaenoic (22:6n-3) acids in Zellweger syndrome and other peroxisomal disorders. Mol Cell Biochem 168, 101115.Google Scholar
99 Astarita, G, Jung, KM, Berchtold, NC, et al. (2010) Deficient liver biosynthesis of docosahexaenoic acid correlates with cognitive impairment in Alzheimer's disease. PLoS One 5, e12538.Google Scholar
100 Kou, J, Kovacs, GG, Hoftberger, R, et al. (2011) Peroxisomal alterations in Alzheimer's disease. Acta Neuropathol 122, 271283.Google Scholar
101 Cunnane, SC (1988) Evidence that adverse effects of zinc deficiency on essential fatty acid composition in rats are independent of food intake. Br J Nutr 59, 273278.Google Scholar
102 Cunnane, SC (1988) Role of zinc in lipid and fatty acid metabolism and in membranes. Prog Food Nutr Sci 12, 151188.Google Scholar
103 Davidson, MH (2006) Mechanisms for the hypotriglyceridemic effect of marine omega-3 fatty acids. Am J Cardiol 98, 27i33i.Google Scholar
104 Barclay, AW, Brand-Miller, JC & Mitchell, P (2003) Glycemic index, glycemic load and diabetes in a sample of older Australians. Asia Pac J Clin Nutr Suppl. 12, S11.Google Scholar
105 Havel, PJ (2005) Dietary fructose: implications for dysregulation of energy homeostasis and lipid/carbohydrate metabolism. Nutr Rev 63, 133157.Google Scholar
106 Basciano, H, Federico, L & Adeli, K (2005) Fructose, insulin resistance, and metabolic dyslipidemia. Nutr Metab 2, 5.Google Scholar
107 Kok, N, Roberfroid, M & Delzenne, N (1996) Dietary oligofructose modifies the impact of fructose on hepatic triacylglycerol metabolism. Metabolism 45, 15471550.Google Scholar
108 van der Borght, K, Kohnke, R, Goransson, N, et al. (2011) Reduced neurogenesis in the rat hippocampus following high fructose consumption. Regul Pept 167, 2630.Google Scholar
109 Banks, WA (2008) The blood–brain barrier as a cause of obesity. Curr Pharm Des 14, 16061614.Google Scholar
110 Kokoeva, MV, Yin, H & Flier, JS (2005) Neurogenesis in the hypothalamus of adult mice: potential role in energy balance. Science 310, 679683.Google Scholar
111 McNay, DE, Briancon, N, Kokoeva, MV, et al. (2012) Remodeling of the arcuate nucleus energy-balance circuit is inhibited in obese mice. J Clin Invest 122, 142152.Google Scholar
112 Migaud, M, Batailler, M, Segura, S, et al. (2010) Emerging new sites for adult neurogenesis in the mammalian brain: a comparative study between the hypothalamus and the classical neurogenic zones. Eur J Neurosci 32, 20422052.Google Scholar
113 Pierce, AA & Xu, AW (2010) De novo neurogenesis in adult hypothalamus as a compensatory mechanism to regulate energy balance. J Neurosci 30, 723730.Google Scholar
114 Funari, VA, Crandall, JE & Tolan, DR (2007) Fructose metabolism in the cerebellum. Cerebellum 6, 130140.Google Scholar
115 Dauncey, MJ (2009) New insights into nutrition and cognitive neuroscience. Proc Nutr Soc 68, 408415.Google Scholar
116 Kamphuis, PJ & Scheltens, P (2010) Can nutrients prevent or delay onset of Alzheimer's disease? J Alzheimers Dis 20, 765775.Google Scholar
117 Liu, J & Ames, BN (2005) Reducing mitochondrial decay with mitochondrial nutrients to delay and treat cognitive dysfunction, Alzheimer's disease, and Parkinson's disease. Nutr Neurosci 8, 6789.Google Scholar
118 Pieczenik, SR & Neustadt, J (2007) Mitochondrial dysfunction and molecular pathways of disease. Exp Mol Pathol 83, 8492.Google Scholar
119 Conquer, JA, Tierney, MC, Zecevic, J, et al. (2000) Fatty acid analysis of blood plasma of patients with Alzheimer's disease, other types of dementia, and cognitive impairment. Lipids 35, 13051312.CrossRefGoogle ScholarPubMed
120 Rinaldi, P, Polidori, MC, Metastasio, A, et al. (2003) Plasma antioxidants are similarly depleted in mild cognitive impairment and in Alzheimer's disease. Neurobiol Aging 24, 915919.Google Scholar
121 Quadri, P, Fragiacomo, C, Pezzati, R, et al. (2004) Homocysteine, folate, and vitamin B-12 in mild cognitive impairment, Alzheimer disease, and vascular dementia. Am J Clin Nutr 80, 114122.Google Scholar
122 Baldeiras, I, Santana, I, Proenca, MT, et al. (2008) Peripheral oxidative damage in mild cognitive impairment and mild Alzheimer's disease. J Alzheimers Dis 15, 117128.Google Scholar
123 Dysken, MW, Sano, M, Asthana, S, et al. (2014) Effect of vitamin E and memantine on functional decline in Alzheimer disease: the TEAM-AD VA cooperative randomized trial. JAMA 311, 3344.Google Scholar
124 Littlejohns, TJ, Henley, WE, Lang, IA, et al. (2014) Vitamin D and the risk of dementia and Alzheimer disease. Neurology 83, 920928.Google Scholar
125 Douaud, G, Refsum, H, de Jager, CA, et al. (2013) Preventing Alzheimer's disease-related gray matter atrophy by B-vitamin treatment. Proc Natl Acad Sci U S A 110, 95239528.Google Scholar
126 Balk, E, Chung, M, Raman, G, et al. (2006) B vitamins and berries and age-related neurodegenerative disorders. Evid Rep Technol Assess 1161.Google Scholar
127 Aisen, PS, Schneider, LS, Sano, M, et al. (2008) High-dose B vitamin supplementation and cognitive decline in Alzheimer disease: a randomized controlled trial. JAMA 300, 17741783.Google Scholar
128 Morris, MC, Evans, DA, Schneider, JA, et al. (2006) Dietary folate and vitamins B-12 and B-6 not associated with incident Alzheimer's disease. J Alzheimer's Dis 9, 435443.Google Scholar
129 Smith, C, Marks, AD & Lieberman, M (2005) Protein digestion and amino acid absorption. In Marks' Basic Medical Biochemistry: A Clinical Approach, p. 695 [Smith, C, Marks, AD and Lieberman, M, editors]. Philadelphia: Lippincott Williams & Wilkins.Google Scholar
130 Sawada, M, Hirata, Y, Arai, H, et al. (1987) Tyrosine hydroxylase, tryptophan hydroxylase, biopterin, and neopterin in the brains of normal controls and patients with senile dementia of Alzheimer type. J Neurochem 48, 760764.Google Scholar
131 Vrecko, K, Birkmayer, JG & Krainz, J (1993) Stimulation of dopamine biosynthesis in cultured PC 12 phaeochromocytoma cells by the coenzyme nicotinamide adeninedinucleotide (NADH). J Neural Transm Park Dis Dement Sect 5, 147156.Google Scholar
132 O'Keeffe, ST (2000) Thiamine deficiency in elderly people. Age Ageing 29, 99101.Google Scholar
133 Gibson, GE & Blass, JP (2007) Thiamine-dependent processes and treatment strategies in neurodegeneration. Antioxid Redox Signal 9, 16051619.Google Scholar
134 Gibson, GE, Sheu, KF, Blass, JP, et al. (1988) Reduced activities of thiamine-dependent enzymes in the brains and peripheral tissues of patients with Alzheimer's disease. Arch Neurol 45, 836840.Google Scholar
135 Wyatt, DT, Nelson, D & Hillman, RE (1991) Age-dependent changes in thiamin concentrations in whole blood and cerebrospinal fluid in infants and children. Am J Clin Nutr 53, 530536.Google Scholar
136 Singleton, CK & Martin, PR (2001) Molecular mechanisms of thiamine utilization. Curr Mol Med 1, 197207.CrossRefGoogle ScholarPubMed
137 Langlais, PJ & Zhang, SX (1993) Extracellular glutamate is increased in thalamus during thiamine deficiency-induced lesions and is blocked by MK-801. J Neurochem 61, 21752182.Google Scholar
138 Heroux, M, Raghavendra Rao, VL, Lavoie, J, et al. (1996) Alterations of thiamine phosphorylation and of thiamine-dependent enzymes in Alzheimer's disease. Metab Brain Dis 11, 8188.Google Scholar
139 Mastrogiacoma, F, Bettendorff, L, Grisar, T, et al. (1996) Brain thiamine, its phosphate esters, and its metabolizing enzymes in Alzheimer's disease. Ann Neurol 39, 585591.Google Scholar
140 Mastrogiacoma, F, Lindsay, JG, Bettendorff, L, et al. (1996) Brain protein and alpha-ketoglutarate dehydrogenase complex activity in Alzheimer's disease. Ann Neurol 39, 592598.Google Scholar
141 Matsushita, S, Miyakawa, T, Maesato, H, et al. (2008) Elevated cerebrospinal fluid tau protein levels in Wernicke's encephalopathy. Alcohol Clin Exp Res 32, 10911095.Google Scholar
142 Zhang, Q, Yang, G, Li, W, et al. (2011) Thiamine deficiency increases beta-secretase activity and accumulation of beta-amyloid peptides. Neurobiol Aging 32, 4253.Google Scholar
143 Karuppagounder, SS, Xu, H, Shi, Q, et al. (2009) Thiamine deficiency induces oxidative stress and exacerbates the plaque pathology in Alzheimer's mouse model. Neurobiol Aging 30, 15871600.Google Scholar
144 Markesbery, WR (1997) Oxidative stress hypothesis in Alzheimer's disease. Free Radic Biol Med 23, 134147.Google Scholar
145 Perry, G, Cash, AD & Smith, MA (2002) Alzheimer disease and oxidative stress. J Biomed Biotechnol 2, 120123.Google Scholar
146 Micronutrient Information Centre, Linus Pauling Institute, Micronutrient Research For Optimum Health, Oregon State University (2014) Vitamin B12 . http://lpi.oregonstate.edu/infocenter/vitamins/vitaminB12/#alzheimer.Google Scholar
147 Seshadri, S, Beiser, A, Selhub, J, et al. (2002) Plasma homocysteine as a risk factor for dementia and Alzheimer's disease. N Engl J Med 346, 476483.Google Scholar
148 Joseph, JA, Denisova, N, Fisher, D, et al. (1998) Membrane and receptor modifications of oxidative stress vulnerability in aging. Nutritional considerations. Ann N Y Acad Sci 854, 268276.Google Scholar
149 Ames, BN, Cathcart, R, Schwiers, E, et al. (1981) Uric acid provides an antioxidant defense in humans against oxidant- and radical-caused aging and cancer: a hypothesis. Proc Natl Acad Sci U S A 78, 68586862.Google Scholar
150 Hageman, GJ & Stierum, RH (2001) Niacin, poly(ADP-ribose) polymerase-1 and genomic stability. Mutat Res 475, 4556.Google Scholar
151 Nakashima, Y & Suzue, R (1984) Influence of nicotinic acid on cerebroside synthesis in the brain of developing rats. J Nutr Sci Vitaminol 30, 525534.Google Scholar
152 Melo, SS, Meirelles, MS, Jordao Junior, AA, et al. (2000) Lipid peroxidation in nicotinamide-deficient and nicotinamide-supplemented rats. Int J Vitam Nutr Res 70, 321323.Google Scholar
153 Morris, MC, Evans, DA, Bienias, JL, et al. (2004) Dietary niacin and the risk of incident Alzheimer's disease and of cognitive decline. J Neurol Neurosurg Psychiatry 75, 10931099.Google Scholar
154 Battaglia, A, Bruni, G, Ardia, A, et al. (1989) Nicergoline in mild to moderate dementia. A multicenter, double-blind, placebo-controlled study. J Am Geriatr Soc 37, 295302.Google Scholar
155 Malouf, R & Grimley Evans, J (2003) The effect of vitamin B6 on cognition. The Cochrane Database of Systematic Reviews 2003, issue 4 , CD004393.Google Scholar
156 Wakimoto, P & Block, G (2001) Dietary intake, dietary patterns, and changes with age: an epidemiological perspective. J Gerontol A Biol Sci Med Sci 56, 6580.Google Scholar
157 Coker, M, de Klerk, JB, Poll-The, BT, et al. (1996) Plasma total odd-chain fatty acids in the monitoring of disorders of propionate, methylmalonate and biotin metabolism. J Inherit Metab Dis 19, 743751.Google Scholar
158 Umhau, JC, Dauphinais, KM, Patel, SH, et al. (2006) The relationship between folate and docosahexaenoic acid in men. Eur J Clin Nutr 60, 352357.Google Scholar
159 Pita, ML & Delgado, MJ (2000) Folate administration increases n-3 polyunsaturated fatty acids in rat plasma and tissue lipids. Thromb Haemost 84, 420423.Google Scholar
160 Hirono, H & Wada, Y (1978) Effects of dietary folate deficiency on developmental increase of myelin lipids in rat brain. J Nutr 108, 766772.Google Scholar
161 Rogaev, EI, Lukiw, WJ, Lavrushina, O, et al. (1994) The upstream promoter of the beta-amyloid precursor protein gene (APP) shows differential patterns of methylation in human brain. Genomics 22, 340347.Google Scholar
162 Miller, AL (2003) The methionine–homocysteine cycle and its effects on cognitive diseases. Altern Med Rev 8, 719.Google Scholar
163 Lehmann, M, Gottfries, CG & Regland, B (1999) Identification of cognitive impairment in the elderly: homocysteine is an early marker. Dement Geriatr Cogn Disord 10, 1220.Google Scholar
164 Nilsson, K, Gustafson, L & Hultberg, B (2002) Relation between plasma homocysteine and Alzheimer's disease. Dement Geriatr Cogn Disord 14, 712.Google Scholar
165 Nilsson, K, Gustafson, L & Hultberg, B (2001) Improvement of cognitive functions after cobalamin/folate supplementation in elderly patients with dementia and elevated plasma homocysteine. Int J Geriatr Psychiatry 16, 609614.Google Scholar
166 Snowdon, DA, Tully, CL, Smith, CD, et al. (2000) Serum folate and the severity of atrophy of the neocortex in Alzheimer disease: findings from the Nun study. Am J Clin Nutr 71, 993998.Google Scholar
167 Kwok, T, Tang, C, Woo, J, et al. (1998) Randomized trial of the effect of supplementation on the cognitive function of older people with subnormal cobalamin levels. Int J Geriatr Psychiatry 13, 611616.Google Scholar
168 Evers, S, Koch, HG, Grotemeyer, KH, et al. (1997) Features, symptoms, and neurophysiological findings in stroke associated with hyperhomocysteinemia. Arch Neurol 54, 12761282.Google Scholar
169 Chambers, JC, Ueland, PM, Obeid, OA, et al. (2000) Improved vascular endothelial function after oral B vitamins: an effect mediated through reduced concentrations of free plasma homocysteine. Circulation 102, 24792483.Google Scholar
170 Christen, Y (2000) Oxidative stress and Alzheimer disease. Am J Clin Nutr 71, 621S629S.Google Scholar
171 Ho, PI, Ortiz, D, Rogers, E, et al. (2002) Multiple aspects of homocysteine neurotoxicity: glutamate excitotoxicity, kinase hyperactivation and DNA damage. J Neurosci Res 70, 694702.Google Scholar
172 White, AR, Huang, X, Jobling, MF, et al. (2001) Homocysteine potentiates copper- and amyloid beta peptide-mediated toxicity in primary neuronal cultures: possible risk factors in the Alzheimer's-type neurodegenerative pathways. J Neurochem 76, 15091520.Google Scholar
173 Ho, PI, Collins, SC, Dhitavat, S, et al. (2001) Homocysteine potentiates beta-amyloid neurotoxicity: role of oxidative stress. J Neurochem 78, 249253.Google Scholar
174 Kruman, II, Culmsee, C, Chan, SL, et al. (2000) Homocysteine elicits a DNA damage response in neurons that promotes apoptosis and hypersensitivity to excitotoxicity. J Neurosci 20, 69206926.Google Scholar
175 Olney, JW, Price, MT, Salles, KS, et al. (1987) l-Homocysteic acid: an endogenous excitotoxic ligand of the NMDA receptor. Brain Res Bull 19, 597602.Google Scholar
176 James, SJ, Melnyk, S, Pogribna, M, et al. (2002) Elevation in S-adenosylhomocysteine and DNA hypomethylation: potential epigenetic mechanism for homocysteine-related pathology. J Nutr 132, Suppl. 8, 2361S2366S.Google Scholar
177 West, RL, Lee, JM & Maroun, LE (1995) Hypomethylation of the amyloid precursor protein gene in the brain of an Alzheimer's disease patient. J Mol Neurosci 6, 141146.Google Scholar
178 Selley, ML (2007) A metabolic link between S-adenosylhomocysteine and polyunsaturated fatty acid metabolism in Alzheimer's disease. Neurobiol Aging 28, 18341839.Google Scholar
179 Werstuck, GH, Lentz, SR, Dayal, S, et al. (2001) Homocysteine-induced endoplasmic reticulum stress causes dysregulation of the cholesterol and triglyceride biosynthetic pathways. J Clin Invest 107, 12631273.Google Scholar
180 Walker, AK, Jacobs, RL, Watts, JL, et al. (2011) A conserved SREBP-1/phosphatidylcholine feedback circuit regulates lipogenesis in metazoans. Cell 147, 840852.Google Scholar
181 Vance, DE, Walkey, CJ & Cui, Z (1997) Phosphatidylethanolamine N-methyltransferase from liver. Biochim Biophys Acta 1348, 142150.Google Scholar
182 Resseguie, M, Song, J, Niculescu, MD, et al. (2007) Phosphatidylethanolamine N-methyltransferase (PEMT) gene expression is induced by estrogen in human and mouse primary hepatocytes. FASEB J 21, 26222632.Google Scholar
183 Michel, V, Yuan, Z, Ramsubir, S, et al. (2006) Choline transport for phospholipid synthesis. Exp Biol Med 231, 490504.Google Scholar
184 Wecker, L (1990) Dietary choline: a limiting factor for the synthesis of acetylcholine by the brain. Adv Neurol 51, 139145.Google Scholar
185 Slotkin, TA, Nemeroff, CB, Bissette, G, et al. (1994) Overexpression of the high affinity choline transporter in cortical regions affected by Alzheimer's disease. Evidence from rapid autopsy studies. J Clin Invest 94, 696702.Google Scholar
186 Miller, BL, Jenden, DJ, Cummings, JL, et al. (1986) Abnormal erythrocyte choline and influx in Alzheimer's disease. Life Sci 38, 485490.Google Scholar
187 Balmer, JE & Blomhoff, R (2002) Gene expression regulation by retinoic acid. J Lipid Res 43, 17731808.Google Scholar
188 Khanna, A & Reddy, TS (1983) Effect of undernutrition and vitamin A deficiency on the phospholipid composition of rat tissues at 21 days of age. I. Liver, spleen and kidney. Int J Vitam Nutr Res 53, 38.Google Scholar
189 Oliveros, LB, Domeniconi, MA, Vega, VA, et al. (2007) Vitamin A deficiency modifies lipid metabolism in rat liver. Br J Nutr 97, 263272.Google Scholar
190 Koryakina, A, Aeberhard, J, Kiefer, S, et al. (2009) Regulation of secretases by all-trans-retinoic acid. FEBS J 276, 26452655.Google Scholar
191 Goodman, AB & Pardee, AB (2003) Evidence for defective retinoid transport and function in late onset Alzheimer's disease. Proc Natl Acad Sci U S A 100, 29012905.Google Scholar
192 Shudo, K, Fukasawa, H, Nakagomi, M, et al. (2009) Towards retinoid therapy for Alzheimer's disease. Curr Alzheimer Res 6, 302311.Google Scholar
193 Tippmann, F, Hundt, J, Schneider, A, et al. (2009) Up-regulation of the alpha-secretase ADAM10 by retinoic acid receptors and acitretin. FASEB J 23, 16431654.Google Scholar
194 Corcoran, JP, So, PL & Maden, M (2004) Disruption of the retinoid signalling pathway causes a deposition of amyloid beta in the adult rat brain. Eur J Neurosci 20, 896902.Google Scholar
195 Husson, M, Enderlin, V, Delacourte, A, et al. (2006) Retinoic acid normalizes nuclear receptor mediated hypo-expression of proteins involved in beta-amyloid deposits in the cerebral cortex of vitamin A deprived rats. Neurobiol Dis 23, 110.Google Scholar
196 Acin-Perez, R, Hoyos, B, Zhao, F, et al. (2010) Control of oxidative phosphorylation by vitamin A illuminates a fundamental role in mitochondrial energy homoeostasis. FASEB J 24, 627636.Google Scholar
197 Leung, WC, Hessel, S, Meplan, C, et al. (2009) Two common single nucleotide polymorphisms in the gene encoding beta-carotene 15,15′-monoxygenase alter beta-carotene metabolism in female volunteers. FASEB J 23, 10411053.CrossRefGoogle ScholarPubMed
198 Buell, JS & Dawson-Hughes, B (2008) Vitamin D and neurocognitive dysfunction: preventing “D”ecline? Mol Aspects Med 29, 415422.Google Scholar
199 Sato, Y, Asoh, T & Oizumi, K (1998) High prevalence of vitamin D deficiency and reduced bone mass in elderly women with Alzheimer's disease. Bone 23, 555557.Google Scholar
200 Scott, TM, Peter, I, Tucker, KL, et al. (2006) The Nutrition, Aging, and Memory in Elders (NAME) study: design and methods for a study of micronutrients and cognitive function in a homebound elderly population. Int J Geriatr Psychiatry 21, 519528.Google Scholar
201 Ogihara, T, Miya, K & Morimoto, S (1990) Possible participation of calcium-regulating factors in senile dementia in elderly female subjects. Gerontology 36, Suppl. 1, 2530.Google Scholar
202 Luckhaus, C, Mahabadi, B, Grass-Kapanke, B, et al. (2009) Blood biomarkers of osteoporosis in mild cognitive impairment and Alzheimer's disease. J Neural Transm 116, 905911.Google Scholar
203 Tysiewicz-Dudek, M, Pietraszkiewicz, F & Drozdzowska, B (2008) Alzheimer's disease and osteoporosis: common risk factors or one condition predisposing to the other? Ortop Traumatol Rehabil 10, 315323.Google Scholar
204 Sutherland, MK, Somerville, MJ, Yoong, LK, et al. (1992) Reduction of vitamin D hormone receptor mRNA levels in Alzheimer as compared to Huntington hippocampus: correlation with calbindin-28k mRNA levels. Brain Res Mol Brain Res 13, 239250.Google Scholar
205 Gezen-Ak, D, Dursun, E, Ertan, T, et al. (2007) Association between vitamin D receptor gene polymorphism and Alzheimer's disease. Tohoku J Exp Med 212, 275282.Google Scholar
206 Baas, D, Prufer, K, Ittel, ME, et al. (2000) Rat oligodendrocytes express the vitamin D(3) receptor and respond to 1,25-dihydroxyvitamin D(3). Glia 31, 5968.Google Scholar
207 Alvarez, JA & Ashraf, A (2010) Role of vitamin D in insulin secretion and insulin sensitivity for glucose homeostasis. Int J Endocrinol 2010, 351385.Google Scholar
208 Wehr, E, Pilz, S, Boehm, BO, et al. (2010) Association of vitamin D status with serum androgen levels in men. Clin Endocrinol (Oxf) 73, 243248.Google Scholar
209 Chu, LW, Tam, S, Lee, PW, et al. (2008) Bioavailable testosterone is associated with a reduced risk of amnestic mild cognitive impairment in older men. Clin Endocrinol (Oxf) 68, 589598.Google Scholar
210 Hogervorst, E, Bandelow, S, Combrinck, M, et al. (2004) Low free testosterone is an independent risk factor for Alzheimer's disease. Exp Gerontol 39, 16331639.Google Scholar
211 Kinuta, K, Tanaka, H, Moriwake, T, et al. (2000) Vitamin D is an important factor in estrogen biosynthesis of both female and male gonads. Endocrinology 141, 13171324.Google Scholar
212 Jimenez-Jimenez, FJ, de Bustos, F, Molina, JA, et al. (1997) Cerebrospinal fluid levels of alpha-tocopherol (vitamin E) in Alzheimer's disease. J Neural Transm 104, 703710.Google Scholar
213 Perkins, AJ, Hendrie, HC, Callahan, CM, et al. (1999) Association of antioxidants with memory in a multiethnic elderly sample using the Third National Health and Nutrition Examination Survey. Am J Epidemiol 150, 3744.Google Scholar
214 Vuletic, S, Peskind, ER, Marcovina, SM, et al. (2005) Reduced CSF PLTP activity in Alzheimer's disease and other neurologic diseases; PLTP induces ApoE secretion in primary human astrocytes in vitro . J Neurosci Res 80, 406413.Google Scholar
215 Desrumaux, C, Risold, PY, Schroeder, H, et al. (2005) Phospholipid transfer protein (PLTP) deficiency reduces brain vitamin E content and increases anxiety in mice. FASEB J 19, 296297.Google Scholar
216 Yatin, SM, Varadarajan, S & Butterfield, DA (2000) Vitamin E prevents Alzheimer's amyloid beta-peptide (1–42)-induced neuronal protein oxidation and reactive oxygen species production. J Alzheimers Dis 2, 123131.Google Scholar
217 Butterfield, DA, Koppal, T, Subramaniam, R, et al. (1999) Vitamin E as an antioxidant/free radical scavenger against amyloid beta-peptide-induced oxidative stress in neocortical synaptosomal membranes and hippocampal neurons in culture: insights into Alzheimer's disease. Rev Neurosci 10, 141149.Google Scholar
218 Rota, C, Rimbach, G, Minihane, AM, et al. (2005) Dietary vitamin E modulates differential gene expression in the rat hippocampus: potential implications for its neuroprotective properties. Nutr Neurosci 8, 2129.Google Scholar
219 Cutler, RG, Kelly, J, Storie, K, et al. (2004) Involvement of oxidative stress-induced abnormalities in ceramide and cholesterol metabolism in brain aging and Alzheimer's disease. Proc Natl Acad Sci U S A 101, 20702075.Google Scholar
220 Briefel, RR, Bialostosky, K, Kennedy-Stephenson, J, et al. (2000) Zinc intake of the U.S. population: findings from the third National Health and Nutrition Examination Survey, 1988–1994. J Nutr 130, Suppl. 5S, 1367S1373S.Google Scholar
221 Tully, CL, Snowdon, DA & Markesbery, WR (1995) Serum zinc, senile plaques, and neurofibrillary tangles: findings from the Nun Study. Neuroreport 6, 21052108.Google Scholar
222 Cuajungco, MP & Faget, KY (2003) Zinc takes the center stage: its paradoxical role in Alzheimer's disease. Brain Res Brain Res Rev 41, 4456.Google Scholar
223 Loef, M, von Stillfried, N & Walach, H (2012) Zinc diet and Alzheimer's disease: a systematic review. Nutr Neurosci 15, 212.Google Scholar
224 Stoltenberg, M, Bush, AI, Bach, G, et al. (2007) Amyloid plaques arise from zinc-enriched cortical layers in APP/PS1 transgenic mice and are paradoxically enlarged with dietary zinc deficiency. Neuroscience 150, 357369.Google Scholar
225 Black, MM (2003) Micronutrient deficiencies and cognitive functioning. J Nutr 133, Suppl. 2, 3927S3931S.Google Scholar
226 Bhatnagar, S & Taneja, S (2001) Zinc and cognitive development. Br J Nutr 85, Suppl. 2, S139S145.Google Scholar
227 Huang, X, Atwood, CS, Hartshorn, MA, et al. (1999) The A beta peptide of Alzheimer's disease directly produces hydrogen peroxide through metal ion reduction. Biochemistry 38, 76097616.Google Scholar
228 Huang, X, Cuajungco, MP, Atwood, CS, et al. (2000) Alzheimer's disease, beta-amyloid protein and zinc. J Nutr 130, Suppl. 5S, 1488S1492S.Google Scholar
229 Atamna, H (2006) Heme binding to amyloid-beta peptide: mechanistic role in Alzheimer's disease. J Alzheimers Dis 10, 255266.Google Scholar
230 Atamna, H & Boyle, K (2006) Amyloid-beta peptide binds with heme to form a peroxidase: relationship to the cytopathologies of Alzheimer's disease. Proc Natl Acad Sci U S A 103, 33813386.Google Scholar
231 Sensi, SL & Jeng, JM (2004) Rethinking the excitotoxic ionic milieu: the emerging role of Zn(2+) in ischemic neuronal injury. Curr Mol Med 4, 87111.Google Scholar
232 Furuta, A, Price, DL, Pardo, CA, et al. (1995) Localization of superoxide dismutases in Alzheimer's disease and Down's syndrome neocortex and hippocampus. Am J Pathol 146, 357367.Google Scholar
233 Leissring, MA, Farris, W, Wu, X, et al. (2004) Alternative translation initiation generates a novel isoform of insulin-degrading enzyme targeted to mitochondria. Biochem J 383, 439446.Google Scholar
234 Jayasooriya, AP, Ackland, ML, Mathai, ML, et al. (2005) Perinatal omega-3 polyunsaturated fatty acid supply modifies brain zinc homeostasis during adulthood. Proc Natl Acad Sci U S A 102, 71337138.Google Scholar
235 Suphioglu, C, De Mel, D, Kumar, L, et al. (2010) The omega-3 fatty acid, DHA, decreases neuronal cell death in association with altered zinc transport. FEBS Lett 584, 612618.Google Scholar
236 Potocnik, FC, van Rensburg, SJ, Hon, D, et al. (2006) Oral zinc augmentation with vitamins A and D increases plasma zinc concentration: implications for burden of disease. Metab Brain Dis 21, 139147.Google Scholar
237 Wang, J, Fivecoat, H, Ho, L, et al. (2010) The role of Sirt1: at the crossroad between promotion of longevity and protection against Alzheimer's disease neuropathology. Biochim Biophys Acta 1804, 16901694.Google Scholar
238 Donmez, G, Wang, D, Cohen, DE, et al. (2010) SIRT1 suppresses beta-amyloid production by activating the alpha-secretase gene ADAM10 . Cell 142, 320332.Google Scholar
239 Costa, RM, Drew, C & Silva, AJ (2005) Notch to remember. Trends Neurosci 28, 429435.Google Scholar
240 Bonda, DJ, Lee, HG, Camins, A, et al. (2011) The sirtuin pathway in ageing and Alzheimer disease: mechanistic and therapeutic considerations. Lancet Neurol 10, 275279.Google Scholar
241 Civitarese, AE, Carling, S, Heilbronn, LK, et al. (2007) Calorie restriction increases muscle mitochondrial biogenesis in healthy humans. PLoS Med 4, e76.Google Scholar
242 Xue, B, Yang, Z, Wang, X, et al. (2012) Omega-3 polyunsaturated fatty acids antagonize macrophage inflammation via activation of AMPK/SIRT1 pathway. PLOS ONE 7, e45990.Google Scholar
243 Wu, A, Ying, Z & Gomez-Pinilla, F (2007) Omega-3 fatty acids supplementation restores mechanisms that maintain brain homeostasis in traumatic brain injury. J Neurotrauma 24, 15871595.Google Scholar
244 Hong, YT, Veenith, T, Dewar, D, et al. (2014) Amyloid imaging with carbon 11-labeled Pittsburgh compound B for traumatic brain injury. JAMA Neurol 71, 2331.Google Scholar
245 Lye, TC & Shores, EA (2000) Traumatic brain injury as a risk factor for Alzheimer's disease: a review. Neuropsychol Rev 10, 115129.Google Scholar
246 Nemetz, PN, Leibson, C, Naessens, JM, et al. (1999) Traumatic brain injury and time to onset of Alzheimer's disease: a population-based study. Am J Epidemiol 149, 3240.Google Scholar
247 Borra, MT, Smith, BC & Denu, JM (2005) Mechanism of human SIRT1 activation by resveratrol. J Biol Chem 280, 1718717195.Google Scholar
248 Allard, JS, Perez, E, Zou, S, et al. (2009) Dietary activators of Sirt1. Mol Cell Endocrinol 299, 5863.Google Scholar
249 Maher, P, Akaishi, T & Abe, K (2006) Flavonoid fisetin promotes ERK-dependent long-term potentiation and enhances memory. Proc Natl Acad Sci U S A 103, 1656816573.Google Scholar
250 Ansari, MA, Abdul, HM, Joshi, G, et al. (2009) Protective effect of quercetin in primary neurons against Abeta(1–42): relevance to Alzheimer's disease. J Nutr Biochem 20, 269275.Google Scholar
251 Craft, NE, Haitema, TB, Garnett, KM, et al. (2004) Carotenoid, tocopherol, and retinol concentrations in elderly human brain. J Nutr Health Aging 8, 156162.Google Scholar
252 Johnson, EJ (2012) A possible role for lutein and zeaxanthin in cognitive function in the elderly. Am J Clin Nutr 96, 1161S1165S.Google Scholar
253 GM, Cole, Lim, GP, Yang, F, et al. (2005) Prevention of Alzheimer's disease: omega-3 fatty acid and phenolic anti-oxidant interventions. Neurobiol Aging 26, Suppl. 1, 133136.Google Scholar
254 Venkatesan, P & Rao, MN (2000) Structure–activity relationships for the inhibition of lipid peroxidation and the scavenging of free radicals by synthetic symmetrical curcumin analogues. J Pharm Pharmacol 52, 11231128.Google Scholar
255 Aggarwal, BB, Kumar, A & Bharti, AC (2003) Anticancer potential of curcumin: preclinical and clinical studies. Anticancer Res 23, 363398.Google Scholar
256 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
257 Baum, L, Lam, CW, Cheung, SK, et al. (2008) Six-month randomized, placebo-controlled, double-blind, pilot clinical trial of curcumin in patients with Alzheimer disease. J Clin Psychopharmacol 28, 110113.Google Scholar
258 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. Alzheimers Res Ther 4, 43.Google Scholar
259 Hyung, SJ, DeToma, AS, Brender, JR, et al. (2013) Insights into antiamyloidogenic properties of the green tea extract ( − )-epigallocatechin-3-gallate toward metal-associated amyloid-beta species. Proc Natl Acad Sci U S A 110, 37433748.Google Scholar
260 Lee, MJ, Maliakal, P, Chen, L, et al. (2002) Pharmacokinetics of tea catechins after ingestion of green tea and ( − )-epigallocatechin-3-gallate by humans: formation of different metabolites and individual variability. Cancer Epidemiol Biomarkers Prev 11, 10251032.Google Scholar
261 Vepsalainen, S, Koivisto, H, Pekkarinen, E, et al. (2013) Anthocyanin-enriched bilberry and blackcurrant extracts modulate amyloid precursor protein processing and alleviate behavioral abnormalities in the APP/PS1 mouse model of Alzheimer's disease. J Nutr Biochem 24, 360370.Google Scholar
262 Hartman, RE, Shah, A, Fagan, AM, et al. (2006) Pomegranate juice decreases amyloid load and improves behavior in a mouse model of Alzheimer's disease. Neurobiol Dis 24, 506515.Google Scholar
263 Ye, J, Meng, X, Yan, C, et al. (2010) Effect of purple sweet potato anthocyanins on beta-amyloid-mediated PC-12 cells death by inhibition of oxidative stress. Neurochem Res 35, 357365.Google Scholar
264 Gutierres, JM, Carvalho, FB & Schetinger, MR (2014) Anthocyanins restore behavioral and biochemical changes caused by streptozotocin-induced sporadic dementia of Alzheimer's type. Life Sci 96, 717.Google Scholar
265 Zeisel, SH & Blusztajn, JK (1994) Choline and human nutrition. Ann Rev Nutr 14, 269296.Google Scholar
Figure 0

Fig. 1 Ceramides – the toxic intermediate linking metabolic dysfunction, inflammatory cytokines and insulin resistance? When adipose tissue exceeds its storage capacity, adipokines increase inflammation that increases ceramides. This inhibits insulin signalling, further increasing lipolysis and increasing the release of fatty acids for ceramide synthesis. Ceramide promotes apoptosis and elevated SFA inhibit the B-cell lymphoma 2 (Bcl2) anti-apoptotic protein family of anti-apoptotic proteins. iNOS, inducible nitric oxide synthase; IRS-1, insulin receptor substrate-1; PI3-K, phosphatidylinositol-3 kinase; Akt/PKB, Akt also known as protein kinase B, a serine/threonine-specific protein kinase. A colour version of this figure can be found online at http://www.journals.cambridge.org/bjn

Figure 1

Fig. 2 Methylation pathway depicting interaction with phospholipids and sphingolipids. After donating the methyl group, S-adenosylmethionine (SAMe) is converted into homocysteine via S-adenosylhomocysteine (SAH). Homocysteine is then broken down by one of three pathways. First, it can be converted back to methionine by accepting a methyl group from methylcobalamin (vitamin B12), and second, it can be converted to methionine by accepting a methyl group from trimethylglycine (betaine), or third, it can be converted to cysteine and taurine via serine and activated vitamin B6(146). The catabolism of homocysteine depends on an adequate supply of vitamin B6, folate and vitamin B12. The majority of the essential nutrient choline is present in phosphatidylcholine (PC) and sphingomyelin (SM), major components of all cell membranes. Additionally, PC and SM are precursors for the signalling molecules ceramide, platelet-activating factor and sphingophosphorylcholine(265). Choline is required for the synthesis of the neurotransmitter, acetylcholine, and as it is oxidised to trimethylglycine, plays a crucial role as a methyl donor in the methionine/homocysteine pathway. PE, phosphatidylethanolamine; PEMT, phosphatidylethanolamine methyltransferase; ACh, acetylcholine; B6, vitamin B6; B12, vitamin B12; THF, tetrahydrofolate. A colour version of this figure can be found online at http://www.journals.cambridge.org/bjn