Next Article in Journal
New Insights into the Relationship between Gut Microbiota and Radiotherapy for Cancer
Next Article in Special Issue
Role of Vitamin D Deficiency in the Pathogenesis of Cardiovascular and Cerebrovascular Diseases
Previous Article in Journal
Dietary Approaches to Stop Hypertension via Indo-Mediterranean Foods, May Be Superior to DASH Diet Intervention
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Nutrition Strategies Promoting Healthy Aging: From Improvement of Cardiovascular and Brain Health to Prevention of Age-Associated Diseases

1
Department of Public Health, Faculty of Medicine, Semmelweis University, 1089 Budapest, Hungary
2
Department of Haematology and Stem Cell Transplantation, National Institute for Haematology and Infectious Diseases, South Pest Central Hospital, 1097 Budapest, Hungary
3
Department of Neurosurgery, The University of Oklahoma Health Sciences Center, Oklahoma City, OK 73117, USA
4
Department of Health Promotion Sciences, College of Public Health, The University of Oklahoma Health Sciences Center, Oklahoma City, OK 73117, USA
5
Peggy and Charles Stephenson Oklahoma Cancer Center, Oklahoma City, OK 73104, USA
6
Department of Pulmonology, Semmelweis University, 1083 Budapest, Hungary
*
Authors to whom correspondence should be addressed.
Shared first authorship: These authors contributed equally to this work.
Nutrients 2023, 15(1), 47; https://doi.org/10.3390/nu15010047
Submission received: 1 December 2022 / Revised: 19 December 2022 / Accepted: 21 December 2022 / Published: 22 December 2022
(This article belongs to the Special Issue Dietary Interventions against Age-Related Vascular Pathologies)

Abstract

:
Background: An increasing number of studies suggest that diet plays an important role in regulating aging processes and modulates the development of the most important age-related diseases. Objective: The aim of this review is to provide an overview of the relationship between nutrition and critical age-associated diseases. Methods: A literature review was conducted to survey recent pre-clinical and clinical findings related to the role of nutritional factors in modulation of fundamental cellular and molecular mechanisms of aging and their role in prevention of the genesis of the diseases of aging. Results: Studies show that the development of cardiovascular and cerebrovascular diseases, neurodegenerative diseases, cognitive impairment and dementia can be slowed down or prevented by certain diets with anti-aging action. The protective effects of diets, at least in part, may be mediated by their beneficial macro- (protein, fat, carbohydrate) and micronutrient (vitamins, minerals) composition. Conclusions: Certain diets, such as the Mediterranean diet, may play a significant role in healthy aging by preventing the onset of certain diseases and by improving the aging process itself. This latter can be strengthened by incorporating fasting elements into the diet. As dietary recommendations change with age, this should be taken into consideration as well, when developing a diet tailored to the needs of elderly individuals. Future and ongoing clinical studies on complex anti-aging dietary interventions translating the results of preclinical investigations are expected to lead to novel nutritional guidelines for older adults in the near future.

1. Introduction

The aging of the population of the Western world represents one of the greatest challenges for sustainable development. At present, approximately 9% of the European population is older than 65. This is projected to increase to 25% by 2050. Morbidity and mortality continuously increase with age [1]. There are several conditions that affect an older adult’s life, such as sensory decline (decrease or loss of hearing, poor eyesight), cardiovascular diseases, diabetes mellitus, depression, dementia, sarcopenia [2,3,4,5,6,7,8,9], gait and balance abnormalities and falls [10,11,12,13,14,15,16,17,18,19,20], and chronic obstructive pulmonary disease [21]. These diseases share several risk factors. One of them is an unhealthy diet accompanied by diseases related to a poor diet [22], such as overweight, obesity [23] and metabolic syndrome, which may affect up to 75% of the population aged over 60 [24]. Preclinical studies provide ample evidence that several components of diet, such as protein, carbohydrate and fat content as well as intake of calories, play important roles in regulating aging processes and longevity and development of age-associated diseases, including cardiovascular and cerebrovascular diseases, cognitive impairment and dementia, and malignant diseases [19,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53]. Dietary regimens that favor reduced calorie intake were shown to delay aging and the genesis of age-associated diseases [31,47]. As the proportion of the elderly is on the rise, which is associated with an increased burden of disease and expenditure, evidence-based preventive policies and health promotion programs to promote healthy aging are needed to ensure the future functioning of social health care and pension systems. Healthy diets with anti-aging potential [54,55] are essential to prevent the development of chronic diseases and to promote healthy aging [56]. The aim of the present review is to summarize the role of diets in the development of common age-related chronic conditions (cardio- and cerebrovascular disease, neurodegeneration and neuropsychiatric disorders), and to provide an overview of their role in modulating the aging process itself.

2. Cellular and Molecular Mechanisms of Aging: Regulation by Nutrition and Diets

Geroscience research has identified basic biological mechanisms driving aging [57]. These mechanisms include pathways that drive organismal aging according to a biological timetable. These processes are regulated by changes in gene expression and epigenetic regulatory mechanisms [58,59,60] and are coordinated by endocrine pathways, such as the insulin/insulin-like growth factor-1 (IGF-1) signaling pathway [61,62]. There are multiple evolutionarily conserved nutrient sensing pathways, whose activity is regulated by nutrient levels and which regulate cellular metabolism and also influence aging and determine survival and lifespan (IGF-1, mammalian target of rapamycin (mTOR), sirtuins and AMP-activated protein kinase (AMPK) [30,42,63,64,65,66,67,68,69,70,71,72]. Additionally, certain aging phenotypes are also determined by stochastic processes, including macromolecular damage [73], which leads to the gradual “wearing and tearing” of cells and tissues [74]. Part of these age-related changes are driven by increased production of oxygen free radicals [75,76,77,78,79,80,81,82,83] and by the accumulation of cross-linked proteins [84,85] and DNA damage [86,87,88,89,90]. Some of this damage accumulates over time, while some is repaired by different repair mechanisms [91,92]. Importantly, both programmed mechanisms of aging and stochastic processes contributing to the genesis aging phenotypes are impacted by nutrition and diets (e.g., caloric restriction confers multifaceted anti-aging effects) [93,94]. There is strong evidence that unhealthy diets (e.g., a high fat diet, Western diets, and methionine-rich diets), because of accelerated cellular aging, exacerbate the development of age-related diseases and shorten lifespan in laboratory animals [95,96,97,98,99,100,101,102,103,104]. Development of aging phenotypes and the pathogenesis of age-related diseases are also influenced by a heightened state of low-grade sterile systemic inflammation (“inflamm-aging”) [105]. Recent findings highlight the association between inflammation and deleterious frailty status in older adults [106,107]. Inadequate diets (e.g., a high fat diet [95,96,97,98,99], diabetogenic diets, high methionine diets [108,109,110]) are an important contributing factor to this low-grade systemic inflammation, but it is also one of the easiest modifiable interventions for elderly individuals to intervene on the process of aging [105]. Preclinical studies have developed a wide range of dietary interventions to delay aging and prevent development of age-related diseases in laboratory animals. These include caloric restriction, methionine restriction, diets enriched with polyphenols, time-restricted feeding and intermittent fasting regimens. The results of preclinical research into the role of dietary factors in regulation of aging processes have been incorporated in the dietary recommendations developed for older adults. These findings, taken together with the results of epidemiological studies, led to the clinical development of various dietary patterns and interventions, including the Okinawan diet, Mediterranean diet, caloric restriction and intermittent fasting regimens, and high-carbohydrate-low-protein and ketogenic diets to combat aging in humans.

3. Dietary Factors Influencing Mortality

3.1. Nutritional Recommendations for the Elderly

Nearly 50% of premature deaths are linked to lifestyle, such as our diet [111]. A diet can be characterized quantitatively, by its calorie content, or qualitatively, for example, by its macro- and micronutrient content.
The amount of required daily calorie intake changes with age [112]. The average energy need of people over 60 depends on their anthropometric parameters and their daily physical activity [113]. A male with a sedentary lifestyle between 60 and 69 requires around 2000 calories a day, while above age 70 the required amount of 1900 calories is somewhat lower than that. In the case of an active lifestyle, these thresholds increase to 2600–2900 and 2500–2700 a day, respectively. In the case of an active life, females between 60 and 69 are advised to consume 2100–2300 calories a day, while females over 70 years of age should consume 2000 to 2200 calories daily [114].
Apart from the daily calorie intake, the macronutrient composition of the diet is also a key factor [115,116]. New evidence shows that the response to dietary fat intake may be based on individual circumstances and the rise in low-density lipoprotein (LDL) cholesterol caused by saturated fats may represent a normal rather than a pathologic response, with different factors, such as gut microbiota, mediating the response [117]. The reduction of dietary saturated fatty acids (SFAs) primarily lowers large LDL particles, less strongly associated with cardiovascular disease (CVD), while small atherogenic LDL particles, more strongly associated with CVD through their plasma residence time and enhanced oxidative susceptibility, are minimally affected by SFA content in the diet [118]. As protein utilization decreases over the age of 60, protein requirements increase with age (0.9–1.1 g/kg). This can be met by eating meat, especially poultry, fish, milk and dairy products. Approximately 55–60% of the energy should be covered by carbohydrates, preferably from sources such as whole grain products, vegetables, fruits and brown rice. Daily consumption of fruits and vegetables should exceed 400 g, which is also essential to reach the recommended daily intake of 30–40 g of fibers (Table 1). The intake of fats should not exceed 25–30% of the total daily energy intake [119], and vegetable oils should be favored as opposed to animal fats. Data reveal that a diet including vegetable fat rather than animal fat might be beneficial in type 2 diabetes prevention [119]. The weekly consumption of sea fish should be encouraged as well, as they help prevent cardiovascular diseases and certain mental conditions due to their high omega-3 fatty acid content [120,121].
Minerals and vitamins in our diet may also be essential in healthy aging. Both men and women should strive to consume 600 international units of vitamin D daily (e.g., from fish, egg yolk or supplements) [122]. Women aged over 50 and men aged over 70 are also recommended to consume 1200 mg calcium per day as a preventive measure for osteoporotic fractures. As people age, the amount of stomach acid decreases, which may reduce vitamin B12 intake and consequentially lead to symptoms such as depression and fatigue [119]. Supplements or fortified foods rich in vitamin B12 (e.g., orange juice, milk or yogurt) should also be considered to reach the daily recommendation of 2.4 micrograms of vitamin B12 [123]. Finally, consumption of zinc supplements is also advised in advanced age, as zinc helps the normal functioning of the immune system and has anti-inflammatory properties as well [124,125]. Recent studies are also evaluating the potential role of combining dietary interventions with repurposed drugs and supplements targeting aging mechanisms to optimize aging trajectories, including rescue of vascular function and prevention of age-related cognitive impairment.

3.2. Healthy Hydration for the Elderly

Appropriate fluid intake is necessary to the physiological ageing process [126]. It is known that dehydration is associated with increased hospitalization, morbidity and mortality. The elderly are vulnerable to hypohydration because of the physiological and cognitive changes occurring at advanced age [127,128].
The European Food Safety Authority (EFSA) published recommendations for age-specific fluid intake, and the current recommendation is 2.0 L/day for adult females and 2.5 L/day for adult males [129]. The European Society for Clinical Nutrition and Metabolism (ESPEN) guidelines [130] recommend slightly less daily fluid intake, as it also takes into account the fluid content of foods, hence: 1.6 L/day for females and 2.0 L/day for males [126].
The state of low body water (hypohydration) has deleterious effects on cardiovascular health. Observational studies in the United States have described that long-lasting low water intake increased the risk for adverse cardiovascular events. There is evidence that acute hypohydration induces endothelial dysfunction, increases sympathetic nervous system activity, and may worsen orthostatic tolerance. Thereby, hypohydration impairs vascular function and blood pressure regulation. This mechanism also plays a role in decreased mental and physical performance [128,131].
Hypohydration can lead to many diseases and health issues, e.g., fractures as a result of falls, urinary infections, constipation, pressure ulcers, kidney problems such as stones and functional impairment. Dehydration is also a risk for acute coronary events (OR 1.16, 95% CI 1.03–1.32), pneumonia (OR 1.23, 95% CI 1.13–1.34) and thromboembolism (OR 1.28, 95% CI 1.14–1.42) [132].
Appropriate fluid intake is important for preserved cognitive function [133]. The abnormalities of water homeostasis can lead to early expressions of neuronal dysfunction, chronic cerebral vasculopathy, brain atrophy and neurodegenerative disease [134]. Water consumption can also positively influence cognitive abilities and mood states [135,136,137].
Adequate hydration is also essential for prevention of chronic diseases (such as diabetes, heart disease, kidney stones or renal failure) or may delay their development [138]. Hypohydration seems to be the initial “cause” of many diseases: diabetes, kidney disease, obesity and cancer, leading to a shortened lifespan. Therefore, proper fluid intake might prevent older people from suffering acute health problems, improve health and reduce the risk of chronic diseases commonly occurring in the elderly [133].
In addition, proper hydration was a contributing factor to a faster, better recovery in a retrospective descriptive study conducted in patients after stroke in an intensive care unit [138].
The increase in the amount of consumed water together with the application of angiotensin antagonists can also reduce body weight and can improve metabolic status [133]. Increased fluid intake is recommended for type 2 diabetic patients as well, which can help prevent and can significantly reduce the occurrence of chronic diseases [133].

3.3. Experimental Diets Targeting Healthy Aging

Several types of diets have been developed to promote healthy aging [27]. Intermittent fasting significantly improved several aspects of the quality of life, decreased fatigue and significantly lowered IGF-1, which can act as an accelerator of tumor development and progression [27]. Longevity-promoting diets include caloric restriction, intermittent fasting (ideally with an 11–12 h daily eating period), restriction of methionine intake, and consumption of fruits and vegetables, legumes, whole grains and oilseeds rich in beneficial fats (e.g., nuts) [139]. The emphasis is on the consumption of fresh ingredients rather than semi-prepared or refined foods [139]. The diet is mostly plant-based with 45–60% of the daily calorie requirement provided by unrefined, complex carbohydrates, 10–15% by predominantly plant-based proteins and 25–35% by mostly plant-based fats [139]. The longevity diet combines the beneficial effects of an anti-inflammatory diet rich in vegetables and fruits with the positive aspects of fasting.
Franceschi et al. recently developed a new dietary recommendation for healthy aging, which is based on the Mediterranean diet [140] supplemented with vitamin D [141,142]. This diet does not only decrease the occurrence of age-related diseases, such as cardiovascular disease, cancer and osteoporosis [143], but may also be effective in reducing low-grade systemic inflammation [141,142]. The anti-aging effect of the diet is mediated by other factors apart from inflammation, such as its ability to prevent the shortening of chromosome telomeres and its ability to reduce lipid peroxide, hydrogen peroxide and tumor necrosis factor levels and increase nitric oxide levels [144].
Other dietary approaches for healthy aging have been proposed as well. Several studies have shown, for instance, that restricting calorie intake [101,145,146,147,148,149,150,151,152,153,154,155,156] is an effective way to combat many chronic diseases and metabolic disorders [155,157,158,159]. Other studies have shown that restriction from dietary methionine [160] can delay age-related pathophysiological manifestations. Fasting can mean abstaining from certain foods, for example, meat, or abstaining from foods altogether for a shorter or longer period of time [161] and has been shown to exert multiple beneficial effects in aging [162,163]. Fasting for 12 to 24 h triggers autophagy in cells, which does not only provide energy but also helps cells to renew and survive, and thus slows down the aging process [139]. Fasting methods may activate autophagy, especially during the late portion of the fasting period, and increase the levels of stem cells and regeneration in various tissues, especially during the re-feeding period [139]. Downstream consequences of these changes are improved metabolic function, reduced inflammation with delayed immunosenescence, reduced oxidative damage and improved proteostasis [139,164,165,166]. Because of this, fasting is being investigated as a potential augmentative therapy during cancer therapies [167]. Autophagy impairment has been linked to several diseases, such as Alzheimer’s and Parkinson’s diseases, diabetes, cancer, chronic inflammatory diseases, depression and chronic fatigue syndrome [168,169]. Increased enthusiasm about fasting and its associated benefits [170] has prompted interest from clinical researchers. A recent study found that short-term, time-restricted feeding is safe and feasible in non-obese healthy midlife and older adults [171].

4. Nutrition, Diets and Prevention of Cardiovascular Diseases

Unhealthy nutrition plays a central part in the development of cardiovascular diseases [172,173,174,175]. Unhealthy dietary patterns (e.g., low intake of vegetables and fruits, excessive intake of sodium, added sugars, etc.) may lead to conditions such as obesity, diabetes, hypertension and dyslipidemia that greatly burden the cardiovascular system [176,177]. A healthy diet can reduce the risk of cardiovascular disease by up to 30% [176,178]. The Mediterranean diet, as seen in PREDIMED (Prevención con Dieta Mediterránea) and PREDIMED-Plus studies, may have a positive effect on cardiovascular disease development [178,179]. This effect may be partially mediated by their weight-reducing effect [178,179]. By normalizing body weight, the unwanted consequences of obesity and overweight can be prevented (see Table 2). The cardioprotective elements of a balanced diet are summarized in Table 3, while the various nutritional recommendations to decrease cardiovascular risk are displayed in Table 4 [180,181,182].
In patients affected by cardiovascular diseases, the total fat intake needs to be reduced, and saturated fatty acids should be replaced with unsaturated fatty acids [183]. The risk of cardiovascular diseases can be reduced by 2–3% by replacing one energy percent of saturated fatty acid with polyunsaturated fatty acids. The intake of omega-3 fatty acids, including eicosapentaenoic acid and docosahexaenoic acid, are also highly recommended because of their cardioprotective properties. For the omega-3, the minimum healthy target intake is a combined 500 mg/day of eicosapentaenoic acid (EPA) and docosahexaenoic acid (DHA), and moderate intake (around 10g/day) of linoleic acid is recommended [184]. There are no universal guidelines for the recommended ratio of omega-6 to omega-3 fatty acids for the prevention of cardiovascular diseases, but some reports suggest that a proportion of 1:1 to 4:1 (omega-6: omega-3) is ideal [185]. The consumption of oil seeds and fish is also crucial for prevention, as 30 g of oilseeds daily decreases the risk of cardiovascular diseases by about 30% while eating fish twice a week reduces the risk of stroke by 6%. Conversely, the intake of trans fatty acids should be kept below 1 energy%, as trans fatty acids increase low-density lipoprotein levels and lower high-density lipoprotein levels, which are very strong risk factors for cardiovascular diseases [180,181,182,186]. Two recent studies concluded that there is inconsistent evidence on the relation of fatty acids to coronary heart disease and stroke risk, and that higher intakes of total and saturated fats were associated with lower likelihood of having hypertension, while higher intakes of short-chain saturated fatty acids (SCFAs) were inversely associated with dyslipidemia and diabetes [187,188]. SCFA metabolic remodeling was related to cognitive benefits, better antioxidant capacity, the attenuation of inflammation and longevity [189].
Sodium and potassium are also important dietary elements of cardiovascular disease prevention. Cutting down on salt intake by 1 g/day reduces blood pressure by 3.1 mmHg in patients with hypertension [190]. In healthy individuals, the decrease is less pronounced albeit still significant with an observed decrease of 1.6 mmHg [191]. In contrast, potassium intake should be encouraged as it has been described to have a positive effect on blood pressure (at least 4700 mg of potassium/day) [172,173].
Fiber may also play an important role in cardiovascular disease prevention. A 7 g/day increase in fiber intake reduces cardiovascular disease risk by 9%, while a consumption of 10 g/day fiber also lowers the risk of stroke by 16% and the risk of type 2 diabetes mellitus by 6%. A part of the beneficial effects of fiber could be mediated by their body weight lowering effects [192]. Several human experiments support that increased dietary fiber intake is accompanied by weight loss, prolonged mealtimes, increased satiety, improved glucose homeostasis and fatty acid metabolism, decreased calorie intake and increased microbiome diversity [193,194]. Fiber also mitigates the deposition of cholesterol [195] and lipoproteins [196,197] in the vessels by reducing the adsorption of bile acids as well as by modifying the microbiome [198].

Nutrition, Diets and Endothelial Protection

It is well-established that endothelium-derived nitric oxide (NO) is an important vasodilator gasotransmitter that regulates smooth muscle mediated vascular resistance and thereby organ blood flow. In addition to maintaining normal tissue perfusion, endothelium-derived NO is also involved in numerous vasoprotective signaling pathways. For instance, it inhibits platelet aggregation and inflammatory cell adhesion to endothelial cells. NO also regulates cell division and survival, disrupts pro-inflammatory cytokine-induced signaling pathways, preserves progenitor cell function, and regulates mitochondrial function and cellular energy metabolism.
Cocoa flavanols present in dark chocolate, soy isoflavones, beetroot juice, garlic, organ meat, leafy greens, citrus, pomegranate, nuts and seeds, and watermelon have been implicated in improvement of NO-mediated vasodilation [199]. By drinking beetroot juice, the high content of nitrates (NO3) is partially absorbed into blood through intestinal mucosa and converted to nitrite (NO2) through a non-enzymatic process. Recycled NO2 is reabsorbed and concentrated by the salivary glands and then secreted into saliva to be converted to NO by symbiotic bacteria located both in the oral cavity and stomach [200]. Due to the high dietary NO3 content, beet consumption is associated with numerous benefits, including improved cognitive function, enhanced athletic performance and lower blood pressure [201,202,203]. Garlic (Allium sativum) and similar plant species are rich sources of sulfur compounds. The sulfur constituents of garlic have been involved in the regulation of vascular homeostasis and the control of metabolic systems linked to nutrient metabolism. Recent evidence implicated one of these sulfur compounds, diallyl trisulfide (DATS), to alter the levels of gaseous signaling transmission of NO in mammalian tissues [204]. Leafy green vegetables (spinach, arugula, kale) and cruciferous vegetables are also very rich in NO3. The results of a clinical meta-analysis indicated a significant 15.8% reduced incidence of cardiovascular disease and slowed the progression of cognitive decline with the intake of green leafy vegetables [205]. Despite the fact that specific nutrients responsible for these effects are still under investigation, some of the hypothesized intervening variables found in leafy greens include folic acid, the antioxidants beta-carotene and vitamin E, soluble fiber, calcium and vitamin K [206].
Citrus fruits such as tangerines, oranges, lemons, grapefruits and kumquats are very rich in vitamin C, or ascorbate, which in preclinical studies has been shown to enhance endothelial nitric oxide synthase (eNOS) activity by stabilizing the eNOS cofactor tetrahydrobiopterin (BH4) [207]. In another study, vitamin C improved microvascular reactivity and peripheral tissue perfusion in septic shock patients [208]. More evidence from studies in patients with chronic endothelial dysfunction due to atherosclerosis, hypertension or diabetes have shown the beneficial effects of vitamin C on endothelial and nitric oxide dependent vasodilation. The rapid measured effects of ascorbate (only 1-h post-injection) may be mediated by increased NO availability, either through enhanced synthesis mediated by BH4 recycling, direct reduction of nitrite to NO, release of NO from nitrosothiols or by scavenging superoxide (O2) that would otherwise react with NO to form peroxynitrite (ONOO).
Pomegranate (Punica granatum L.) is another food that is both rich in NO3 and vitamins including ascorbate. Clinical investigation of the acute effects of pomegranate juice has attributed its vasculoprotective effects to the presence of hydrolysable tannins ellagitannins, ellagic acid, anthocyanins, and flavonoids have demonstrated in vivo that they reduce oxidative stress and platelet aggregation, diminish lipid uptake by macrophages, positively influence endothelial cell function and are involved in blood pressure regulation [209,210].
Berries are high in anthocyanin flavonoids, a polyphenol subtype with vascular beneficial properties including increased NO production and reducing oxidative stress and inflammation [211]. It has been previously shown that polyphenols from wine and grape berries’ extracts exert robust vasodilator effects, due to the increased eNOS expression and activity, consequently increasing NO production [212]. Moderate consumption of wine is recommended due to the deleterious effects of excessive alcohol consumption, which exerts pro-oxidant effects and depletes NO bioavailability via the endogenous NOS inhibitor asymmetric dimethylarginine, which mediates the decreased synthesis of NO [213].
Nuts and seeds are rich in L-arginine, the amino acid precursor for the production of NO. Watermelon contains L-citrulline, and recent evidence suggests that in cerebral arteries, L-citrulline is converted to L-arginine most likely via an argininosuccinate pathway to synthesize NO [214]. Multiple other studies have associated L-citrulline supplementation with increased NO synthesis, decreased blood pressure and increased peripheral blood flow [215].
Similarly, other dietary elements have been associated with decreased endothelium-deprive NO production, effectively worsening cerebrovascular function and cognitive outcomes. High table salt (NaCl) intake has been shown to reduce NO drastically; studies showed that it is recommended not to exceed 5 g/day, while the World Health Organization suggests consuming no more than 1 teaspoon of salt per day. Similarly, elevated blood glucose levels destabilize NO bioavailability, and recent studies suggest controlling glucose levels by decreasing consumption of refined sugars. High consumption of saturated fats increases low-density lipoprotein (LDL) cholesterol. Clinical studies have revealed that participants undergoing an LDL-lowering diet benefitted from improved endothelial function when compared with placebo participants. Interestingly, most of the deleterious nutrients are readily found in the standard American diet (SAD), which could in part explain why individuals in North America have higher rates of vascular inflammation, heart disease, brain strokes, coronary artery disease, and cognitive impairment in addition to being more likely to experience weight gain and constipation compared to other countries. In contrast, adherence to the Mediterranean diet was inversely associated with the prevalence of metabolic syndrome in the elderly [216]. Walnuts, part of the Mediterranean diet, significantly decreased triglyceride, total cholesterol and LDL cholesterol concentrations and presented no consequences on anthropometric and glycemic parameters [217]. Hydrosoluble micronutrients (such as polyphenols), lipophilic compounds (tocopherols and tocotrienols; n-3 PUFAs and n-6 PUFAs) and other plant molecules exert their antioxidant action through multiple mechanisms, including the activation of the Nrf2/ARE pathway and down-regulation of the NF-kB pathway, directly implicated in the inflammatory response [217].
The most consistently reconfirmed biological adaptations to the Mediterranean diet include strong antioxidant protection against oxidative stress and inflammation, a lipid-lowering effect, modulation of humoral factors involved in the pathogenesis of cancer, inhibition of energy-sensing pathways and better gut microbiota-mediated release of metabolites beneficial to metabolic health. Adequate nutrition is therefore a crucial component to maintain healthy cerebrovasculature to lead a healthy lifestyle, leading to improved cognitive outcomes in aging.

5. Nutrition, Diets and Healthy Brain Aging

5.1. Prevention of Cognitive Decline and Dementia

A healthy diet can also help prevent cognitive decline and maintain cognitive functioning in older individuals [218]. The occurrence of dementia increases with age, doubling every 5 years after the age of 65 [219]. Risk factors of dementia can be divided into two groups: modifiable and non-modifiable factors (shown in Table 5). Non-modifiable factors include age, sex and genetic factors, while modifiable factors include vascular and metabolic (hypertension, cholesterol, homocysteine, atherosclerosis, diabetes mellitus, etc.), lifestyle (diet, physical and mental activity, smoking, alcohol consumption, obesity, etc.), environmental, (trauma, pesticides, etc.), and disease-related factors (depression). Controlling modifiable risk factors can reduce the incidence of dementia by up to 35% [220]. Studies focusing specifically on the prevention of cognitive decline, such as the Finnish Geriatric Intervention Study to Prevent Cognitive Impairment and Disability Study (FINGER Study), also support these observations [221].
Modulation of metabolic pathways is one of the most promising strategies to promote resilience and delay aging [222]. Diet is one of the most important lifestyle factors associated with dementia [223], including Alzheimer’s disease [224]. The Mediterranean diet has a risk-reducing effect for dementia, which may be mediated by its cardioprotective effect [225]. The MIND diet (Mediterranean-DASH diet Intervention for Neurodegenerative Delay), which is based on the Mediterranean diet (see Table 6), has also been associated with better cognitive functioning [226,227,228].
The role of vitamin D and the vitamin B complex (mainly vitamin B12 and B3 and folic acid) has been confirmed as a protective factor for dementia [229,230,231]. While vitamin D is a potential anti-inflammatory and antioxidant vitamin, vitamin B may enhance the production of certain neurotransmitters and exert an antioxidant effect. The antioxidant properties of vitamin B12 are accomplished by different mechanisms, including the direct scavenging of reactive oxygen radicals, particularly superoxide in the cytosol and mitochondria, and indirectly by stimulating the scavenging of reactive oxygen radicals through the preservation of glutathione.
Mitochondria play a crucial role in mediating and amplifying the oxidative stress that contributes to the process of aging [232,233,234,235]. A growing number of studies have focused on elucidating the role of mitochondrial dysfunction [236], mitochondria-derived mitokines [237] and peptides [238] in the pathogenesis of age-related cognitive impairment. It has become very clear that significant pathophysiological disturbances are associated with mitochondrial aberrations in clinical studies [239] and preclinical models of aging. Mitochondria DNA copy number was found to be a predictor of physical performance in older men [240]. Possibly, the most important lifestyle factor determining adequate mitochondrial function is appropriate nutrition.
Important dietary compounds such as curcumin, astaxanthin, resveratrol, hydroxytyrosol, oleuropein, and spermidine (which are present in the Mediterranean diet) have been shown to exert powerful beneficial effects, enhancing mitophagy (the degradation of damaged mitochondria) by upregulating mitophagy mediators, and promoting new mitochondria biogenesis [241]. Due to this evidence, many laboratories interested in developing anti-aging interventions are targeting mitochondrial dysfunction [242,243,244], and investigating novel roles of mitochondria in longevity.
Caffeine has also sometimes been described to play a protective role in the development of certain forms of dementia, such as Alzheimer’s disease [245]. Caffeine is a non-selective antagonist of adenosine receptors, demonstrating a structural similarity to adenosine. The blocking of these receptors (A1, A2) at lower concentrations of caffeine (plasma concentration <250 µM) modulates glutamatergic, cholinergic, dopaminergic, serotoninergic and noradrenergic neurotransmission. Moreover, caffeine is an agonist of ryanodine receptors as well that play an important part in calcium release from the endoplasmic reticulum. Caffeine is also a non-selective competitive inhibitor of phosphodiesterase, the enzyme degrading cyclic adenosine monophosphate. However, these latter effects are only possible at higher doses (plasma concentration >250 µM). Moreover, caffeine also interferes with γ-aminobutyric acid type A receptors, protects against cell damage by reducing oxidative stress and may exert anti-inflammatory activity by decreasing proinflammatory (C-reactive protein, IL-1β, IL-6, IL-18, TNF-α) and increasing anti-inflammatory (IL-10, adiponectin) marker levels [245]. According to the European Food Safety Authority, 400 mg (approximately 5.7 mg/kg bodyweight) of caffeine a day does not raise any safety concerns (except in pregnancy) [246].
More than a dozen epidemiological studies have reported that reduced intake of omega-3 fatty acids or fish consumption is also associated with increased risk for age-related cognitive decline or dementia, such as Alzheimer’s disease [247,248,249]. Docosahexaenoic acid is neuroprotective via multiple mechanisms that include the reduction of arachidonic acid metabolites and the increase of trophic factors or downstream trophic signal transduction. Docosahexaenoic acid is specifically protective against Alzheimer’s disease via additional mechanisms as well. It limits the production and accumulation of the amyloid beta peptide that is widely believed to drive the disease; and it also suppresses several signal transduction pathways induced by amyloid beta peptides, including two major kinases that phosphorylate the microtubule-associated protein tau and promote neurofibrillary tangle pathology. Results to date suggest that docosahexaenoic acid may be more effective if it is supplemented early or used in conjunction with antioxidants [250].
Recent studies demonstrate that the microbiota is an important modifier of risk factors for age-related neurodegenerative diseases and frailty [251,252,253,254,255,256] and that changes in microbiome-derived metabolites modulate neuroinflammation and age-related cognitive dysfunction [257].

5.2. Nutritional Factors Contributing to Cerebrovascular Protection

The human brain is an exquisitely complex organ with high metabolic demands. For the average adult in a resting state, the brain consumes about 20 percent of the body’s energy expenditure despite it only comprising 2 percent of the body’s weight. Unlike muscles, which possess energy reserves, the brain needs to be constantly supplied with oxygen and nutrients to function adequately. This is ensured by tightly controlled regulation of cerebral blood flow (CBF). Changes in neuronal activity require prompt CBF adjustments to maintain cellular homeostasis [258,259], which is accomplished via neurovascular coupling (NVC; also known as functional hyperemia) [260,261]. Recent studies demonstrate that aging is associated with diminished NVC responses [262] and that impaired NVC is causally related to cognitive decline [263,264].
Preclinical and clinical evidence shows that the endothelium plays a critical role in mediating NVC responses [265,266,267,268]. Pathological conditions that promote endothelial dysfunction (including obesity, diabetes mellitus, hypertension, among others) [269] were shown to exacerbate age-related endothelial dysfunction and neurovascular impairment [270], contributing to cognitive decline. Cerebrovascular endothelium is also important to maintain the integrity of the blood–brain barrier (BBB). Disruption of the BBB also has been causally linked to cognitive impairment [271,272].
Cerebrovascular and endothelial health are importantly modulated by nutritional factors and dietary patterns [100,109,154,273,274]. High fat diets and diets containing high levels of methionine were shown to exert deleterious effects on endothelial cells, promoting vasodilator dysfunction, BBB disruption and neuroinflammation, microvascular rarefaction and neurovascular uncoupling [100,275,276]. Pro-inflammatory diets such as those high in processed foods and refined sugars also have a deleterious effect on cerebral blood vessels and brain function [277]. Additionally, many elderly individuals live with an array of systemic vascular comorbidities that exacerbate the diverse pathophysiological processes promoting cognitive dysfunction [278,279].
Diets enriched for polyphenols (e.g., resveratrol) confer multifaceted anti-aging vasoprotective effects [274,276,280]. There are aso studies showing that dietary intake of cocoa flavanols protects humans against vascular disease, as evidenced by improvements in peripheral endothelial function [281]. An interesting study tested the effects of cocoa consumption in 60 participants in which 17 volunteers presented with impaired NVC [282]. Participants with intact NVC responses showed no significant benefits from cocoa consumption. However, subjects with impaired NVC experienced a dramatic doubling in NVC responses after just a month of cocoa supplementation, while scores on standard cognitive tests increased by 30 percent [282].
Aging is associated with NAD+ depletion, a change that is proposed to be a major contributor to organ-specific aging processes [283,284]. NAD+ is a co-substrate for sirtuin enyzmes, which are regulators of important cellular processes of aging, mitochondrial function, stress resilience apoptosis and inflammation in the vasculature [285,286]. Emerging evidence shows that vascular and cerebrovascular aging is driven by NAD+ depletion [81,82]. Recent studies show that the restoration of NAD+ levels via administration of NAD+ precursor nicotinamide mononucleotide (NMN) in aged murine models leads to the rescue of cerebrovascular endothelial function and NVC [81,82]. Clinical trials utilized a different NAD+ precursor, nicotinamide riboside (NR), which successfully increased NAD+ levels is humans subjects [287,288], making it a possible candidate to restore NAD+ concentration in order to achieve anti-aging effects.

5.3. Prevention of Depression and Other Geriatric Psychological Disorders

Studies indicate that our diet may influence the pathogenesis of psychological disorders, such as depression [289,290,291,292,293]. A recent meta-analysis [289] showed that [5,19,33,294,295,296,297,298] a Mediterranean diet may reduce the risk for depression by up to 30% [289,290,291]. In contrast, consumption of a Western diet with its high intake of saturated fats, refined carbohydrates and processed foods, is associated with poorer mental health indicators [292]. Similar observations were made in the SMILES study [299], in which one group of participants was prescribed a special diet for 12 weeks, while the control group received only peer support during the same time. The 12-week program consisted of a diet made up of the following elements: whole grains (5–8 times per day); vegetables (6 times per day); fruits (3 times per day); leguminous vegetables (3–4 times per week); low-fat, sugar-free dairy products (2–3 times per day); raw and unsalted grains (once per day); fish (twice a week); lean red meats (3–4 times a week); chicken (2–3 times a week); eggs (no more than 6 times a week); olive oil (3 tablespoons a day); and a limited amount of extra sugary soft drinks, processed meats and fast foods (up to 3 times a week). During follow-up, participants in the 12-week diet program reported significantly fewer depressive symptoms than the control group. Similar conclusions were reached by other studies as well [300].
Fruits and vegetables seem to be especially important in the prevention of psychological disorders. People who eat more fruits and vegetables have fewer mental disorders including depression, stress and negative mood [301]. They also report higher rates of happiness, good mood and life satisfaction [302]. Moreover, the association between fruit and vegetable intake and mental health appears to be dose-dependent: the higher the intake, the better the mental health was reported by study participants [303,304]. The consumption of fruits and vegetables exerts its positive effect not only through its vitamin, antioxidant and fiber content, but also by affecting the gut microbiome [305].
Another frequently examined group of nutrients in terms of mental illnesses is fatty acids. Polyunsaturated fatty acids make up about half of the grey matter, of which one-third is made up of fatty acids from the omega-3 family [306]. As omega-3 is an essential fatty acid, it cannot be produced by the body and must be obtained through diet [307]. Results from epidemiological and neurobiological studies have suggested that a relative deficiency of omega-3 polyunsaturated fatty acids may predispose the individual to psychological disorders, including depression. Several other studies have also highlighted the association between lower omega-3 unsaturated fatty acid levels and depression [308,309]. Furthermore, several preliminary small case-control clinical trials also indicate that omega-3 fatty acid supplements may be useful in treating symptoms of depression after previous antidepressant treatment has failed [307,310]. The risk-reducing effect of omega-3 fatty acids may be conveyed by their anti-inflammatory properties [307,308,309,310].
Vitamins may also have an impact on our mental health. It has been long known that higher levels of homocysteine and lower levels of vitamin B9 (folic acid) increase the risk of depression in older ages. A randomized, double-blind, placebo-controlled clinical trial also showed that vitamin B6, B9 and B12 supplementation, along with lower homocysteine levels, enhances the effect of standard antidepressant treatment in old age [311]. Finally, both vitamin D [312] and vitamin B complex (B1/B6/B12: 180/180/1 mg/body weight kg, daily) [311] have been described to improve stress tolerance, anxiety and depressive symptoms as well. The association appears to be driven by the homeostatic, trophic and immunomodulatory effects of vitamin D. Furthermore, vitamin D also modulates the hypothalamic–pituitary–adrenal axis, which regulates the production of the monoamine neurotransmitters epinephrine, norepinephrine and dopamine in the adrenal cortex and also protects against the depletion of dopamine and serotonin [313]. The vitamin B complex, on the other hand, may exert its positive effect on mental health by improving the carbohydrate metabolism in nerve cells [311].

6. Conclusions

Nutrition and diets play a central part in the development of several age-related diseases, such as cardiovascular disease, neurodegenerative disease and dementia. A well-managed nutrition and a diet adapted to age is essential to maintain mental freshness and a good quality of life at old age. Diets also influence the process of aging itself. By adhering to certain diets, such as the Mediterranean diet, the onset of age-related diseases can be prevented or delayed. This is achieved directly by certain elements included in the diet, for instance, fruits, vegetables and omega-3 fatty acids, and indirectly by their positive effect on bodyweight management. Moreover, the introduction of fasting episodes into our diet may also contribute to healthier aging. Diet recommendations change with age, and this must be taken into consideration when developing a diet tailored to the needs of elderly patients. Future and ongoing clinical studies (e.g., ClinicalTrials.gov Identifier: NCT03702335 (Impact of Comprehensive Dietary Counseling on Dietary Quality, Mental Health, and Quality of Life in Older Adults); NCT05593939 (Slow Age: Interventions to Slow Aging in Humans); NCT02751866 (Early Intervention in Cognitive Aging) on complex anti-aging dietary interventions translating the results of promising preclinical investigations are expected to lead to novel nutritional guidelines for older adults in the near future.

Author Contributions

J.T.V., V.F.-P., T.C., M.F., Z.S. and A.F. designed the study, and wrote and published the manuscript. S.T., T.C. and J.F. supplemented and reviewed the manuscript. N.D., A.L., T.C., A.P. and J.T.V. prepared the manuscript for publication. All authors have read and agreed to the published version of the manuscript.

Funding

Project no. TKP2021-NKTA-47 has been implemented with the support provided by the Ministry of Innovation and Technology of Hungary from the National Research, Development and Innovation Fund, financed under the TKP2021-NKTA funding scheme. Funding for the project through the National Cardiovascular Laboratory Program (RRF-2.3.1-21-2022-00003) was provided by the Ministry of Innovation and Technology of Hungary from the National Research, Development and Innovation Fund. This work was also supported by the European University for Well-Being (EUniWell) program (grant agreement number: 101004093/EUniWell/EAC-A02-2019/EAC-A02-2019-1) and grants from the National Institute on Aging (NIA R03AG070479, NIA K01AG073614), the American Heart Association AHA CDA941290, the NIA-supported Geroscience Training Program in Oklahoma (T32AG052363), the NIA-supported Oklahoma Nathan Shock Center, and the NIGMS supported Center of Biomedical Research Excellence (CoBRE) (1P20GM125528-01A1).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Not applicable.

Conflicts of Interest

The authors declare that they have no competing interests.

Abbreviations

ADAlzheimer’s disease
ADLactivity of daily living
AMPKAMP-activated protein kinase
BBBblood-brain barrier
BPSDbehavioral and psychological symptoms of dementia
Cacalcium
CBFcerebral blood flow
CoQ10coenzyme Q10
CRcaloric restriction
CRPC-reactive protein
CVDcardiovascular disease
DATSdiallyl trisulfide
DASHDietary Approaches to Stop Hypertension
DHAdocosahexaenoic acid
DMdiabetes mellitus
DNAdeoxyribonucleic acid
E%energy%
EFSAEuropean Food Safety Authority
EPA eicosapentaenoic acid
eNOS endothelial nitric oxide synthase
EUEuropean Union
FINGER StudyFinnish Geriatric Intervention Study to Prevent Cognitive Impairment and Disability Study
HDLhigh-density lipoprotein
IGF-1insulin-like growth factor-1
IFintermittent fasting
LDLlow-density lipoprotein
MIND dietMediterranean-DASH diet Intervention for Neurodegenerative Delay
mTORmammalian target of rapamycin
NFBneurofibrillary tangles
NICENational Institute for Health and Care Excellence
NOnitric oxide
Pphosphorus
PREDIMED Prevención con Dieta Mediterránea
PUFApolyunsaturated fatty acid
RCTrandomized controlled trial
SFAsaturated fatty acid
tFAtrans fatty acid
WHOWorld Health Organization

References

  1. Mitchell, E.; Walker, R. Global ageing: Successes, challenges and opportunities. Br. J. Hosp. Med. 2020, 81, 1–9. [Google Scholar] [CrossRef]
  2. Alcazar, J.; Aagaard, P.; Haddock, B.; Kamper, R.S.; Hansen, S.K.; Prescott, E.; Alegre, L.M.; Frandsen, U.; Suetta, C. Age- and Sex-Specific Changes in Lower-Limb Muscle Power Throughout the Lifespan. J. Gerontol. A Biol. Sci. Med. Sci. 2020, 75, 1369–1378. [Google Scholar] [CrossRef]
  3. Cawthon, P.M.; Blackwell, T.; Cummings, S.R.; Orwoll, E.S.; Duchowny, K.A.; Kado, D.M.; Stone, K.L.; Ensrud, K.E.; Cauley, J.A.; Evans, W.J. Muscle Mass Assessed by the D3-Creatine Dilution Method and Incident Self-reported Disability and Mortality in a Prospective Observational Study of Community-Dwelling Older Men. J. Gerontol. A Biol. Sci. Med. Sci. 2021, 76, 123–130. [Google Scholar] [CrossRef]
  4. Cheung, A.S.; Gray, H.A.; Schache, A.G.; Hoermann, R.; Bicknell, J.; Joon, D.L.; Zajac, J.D.; Pandy, M.G.; Grossmann, M. Biomechanical Leg Muscle Function During Stair Ambulation in Men Receiving Androgen Deprivation Therapy. J. Gerontol. A Biol. Sci. Med. Sci. 2020, 75, 1715–1722. [Google Scholar] [CrossRef]
  5. Dorhout, B.G.; Doets, E.L.; van Dongen, E.J.I.; de Groot, L.; Haveman-Nies, A. In-Depth Analyses of the Effects of a Diet and Resistance Exercise Intervention in Older Adults: Who Benefits Most From ProMuscle in Practice? J. Gerontol. A Biol. Sci. Med. Sci. 2021, 76, 2204–2212. [Google Scholar] [CrossRef]
  6. Gemikonakli, G.; Mach, J.; Hilmer, S.N. Interactions Between the Aging Gut Microbiome and Common Geriatric Giants: Polypharmacy, Frailty, and Dementia. J. Gerontol. A Biol. Sci. Med. Sci. 2021, 76, 1019–1028. [Google Scholar] [CrossRef]
  7. Giuliani, H.K.; Shea, N.W.; Gerstner, G.R.; Mota, J.A.; Blackburn, J.T.; Ryan, E.D. The Influence of Age and Obesity-Altered Muscle Tissue Composition on Muscular Dimensional Changes: Impact on Strength and Function. J. Gerontol. A Biol. Sci. Med. Sci. 2020, 75, 2286–2294. [Google Scholar] [CrossRef] [PubMed]
  8. Le Couteur, D.G.; Handelsman, D.J.; Stanaway, F.; Waite, L.M.; Blyth, F.M.; Naganathan, V.; Cumming, R.G.; Hirani, V. Sarcopenic Obesity and Amino Acids: Concord Health and Ageing in Men Project. J. Gerontol. A Biol. Sci. Med. Sci. 2021, 76, 1000–1004. [Google Scholar] [CrossRef]
  9. Lu, Y.; Karagounis, L.G.; Ng, T.P.; Carre, C.; Narang, V.; Wong, G.; Tan, C.T.Y.; Zin Nyunt, M.S.; Gao, Q.; Abel, B.; et al. Systemic and Metabolic Signature of Sarcopenia in Community-Dwelling Older Adults. J. Gerontol. A Biol. Sci. Med. Sci. 2020, 75, 309–317. [Google Scholar] [CrossRef]
  10. Ackermans, T.; Francksen, N.; Lees, C.; Papatzika, F.; Arampatzis, A.; Baltzopoulos, V.; Lisboa, P.; Hollands, M.; O’Brien, T.; Maganaris, C. Prediction of Balance Perturbations and Falls on Stairs in Older People Using a Biomechanical Profiling Approach: A 12-Month Longitudinal Study. J. Gerontol. A Biol. Sci. Med. Sci. 2021, 76, 638–646. [Google Scholar] [CrossRef]
  11. Agrawal, Y.; Merfeld, D.M.; Horak, F.B.; Redfern, M.S.; Manor, B.; Westlake, K.P.; Holstein, G.R.; Smith, P.F.; Bhatt, T.; Bohnen, N.I.; et al. Aging, Vestibular Function, and Balance: Proceedings of a National Institute on Aging/National Institute on Deafness and Other Communication Disorders Workshop. J. Gerontol. A Biol. Sci. Med. Sci. 2020, 75, 2471–2480. [Google Scholar] [CrossRef] [PubMed]
  12. Battilana, F.; Steurer, S.; Rizzi, G.; Delgado, A.C.; Tan, K.R.; Handschin, C. Exercise-linked improvement in age-associated loss of balance is associated with increased vestibular input to motor neurons. Aging Cell 2020, 19, e13274. [Google Scholar] [CrossRef]
  13. Cai, Y.; Leveille, S.G.; Shi, L.; Chen, P.; You, T. Chronic Pain and Risk of Injurious Falls in Community-Dwelling Older Adults. J. Gerontol. A Biol. Sci. Med. Sci. 2021, 76, e179–e186. [Google Scholar] [CrossRef] [PubMed]
  14. Chen, T.Y.; Chan, A.; Andersen-Ranberg, K.; Herr, M.; Fors, S.; Jeune, B.; Herrmann, F.R.; Robine, J.M.; Gondo, Y.; Saito, Y. Prevalence and Correlates of Falls Among Centenarians: Results from the Five-Country Oldest Old Project (5-COOP). J. Gerontol. A Biol. Sci. Med. Sci. 2020, 75, 974–979. [Google Scholar] [CrossRef] [PubMed]
  15. De Souto Barreto, P.; Maltais, M.; Rosendahl, E.; Vellas, B.; Bourdel-Marchasson, I.; Lamb, S.E.; Pitkala, K.; Rolland, Y. Exercise Effects on Falls, Fractures, Hospitalizations, and Mortality in Older Adults With Dementia: An Individual-Level Patient Data Meta-analysis. J. Gerontol. A Biol. Sci. Med. Sci. 2021, 76, e203–e212. [Google Scholar] [CrossRef] [PubMed]
  16. Del Din, S.; Galna, B.; Lord, S.; Nieuwboer, A.; Bekkers, E.M.J.; Pelosin, E.; Avanzino, L.; Bloem, B.R.; Olde Rikkert, M.G.M.; Nieuwhof, F.; et al. Falls Risk in Relation to Activity Exposure in High-Risk Older Adults. J. Gerontol. A Biol. Sci. Med. Sci. 2020, 75, 1198–1205. [Google Scholar] [CrossRef] [PubMed]
  17. Donoghue, O.A.; Leahy, S.; Kenny, R.A. Longitudinal Associations Between Gait, Falls, and Disability in Community-Dwelling Older Adults With Type II Diabetes Mellitus: Findings From The Irish Longitudinal Study on Ageing (TILDA). J. Gerontol. A Biol. Sci. Med. Sci. 2021, 76, 906–913. [Google Scholar] [CrossRef]
  18. Ge, M.L.; Simonsick, E.M.; Dong, B.R.; Kasper, J.D.; Xue, Q.L. Frailty, With or Without Cognitive Impairment, Is a Strong Predictor of Recurrent Falls in a US Population-Representative Sample of Older Adults. J. Gerontol. A Biol. Sci. Med. Sci. 2021, 76, e354–e360. [Google Scholar] [CrossRef]
  19. Hruby, A.; Sahni, S.; Bolster, D.; Jacques, P.F. Protein Intake and Functional Integrity in Aging: The Framingham Heart Study Offspring. J. Gerontol. A Biol. Sci. Med. Sci. 2020, 75, 123–130. [Google Scholar] [CrossRef]
  20. S, G.R.N.; J, S.O.; A, B.D.; R, M.L.; Tiedemann, A. Does Obesity Increase the Risk and Severity of Falls in People Aged 60 Years and Older? A Systematic Review and Meta-analysis of Observational Studies. J. Gerontol. A Biol. Sci. Med. Sci. 2020, 75, 952–960. [Google Scholar] [CrossRef]
  21. Fekete, M.; Fazekas-Pongor, V.; Balazs, P.; Tarantini, S.; Nemeth, A.N.; Varga, J.T. Role of new digital technologies and telemedicine in pulmonary rehabilitation: Smart devices in the treatment of chronic respiratory diseases. Wien. Klin. Wochenschr. 2021, 133, 1201–1207. [Google Scholar] [CrossRef] [PubMed]
  22. Palliyaguru, D.L.; Rudderow, A.L.; Sossong, A.M.; Lewis, K.N.; Younts, C.; Pearson, K.J.; Bernier, M.; de Cabo, R. Perinatal diet influences health and survival in a mouse model of leukemia. Geroscience 2020, 42, 1147–1155. [Google Scholar] [CrossRef] [PubMed]
  23. Ayo-Martin, O.; Garcia-Garcia, J.; Hernandez-Fernandez, F.; Gomez-Hontanilla, M.; Gomez-Fernandez, I.; Andres-Fernandez, C.; Lamas, C.; Alfaro-Martinez, J.J.; Botella, F.; Segura, T. Cerebral hemodynamics in obesity: Relationship with sex, age, and adipokines in a cohort-based study. Geroscience 2021, 43, 1465–1479. [Google Scholar] [CrossRef]
  24. Kuzuya, M. Nutritional status related to poor health outcomes in older people: Which is better, obese or lean? Geriatr. Gerontol. Int. 2021, 21, 5–13. [Google Scholar] [CrossRef]
  25. Cole, J.H.; Marioni, R.E.; Harris, S.E.; Dear, I.J. Brain age and other bodily‘ages’: Implications for neuropsychiatry. Mol. Psychiatry 2019, 24, 266–281. [Google Scholar] [CrossRef] [Green Version]
  26. Allmeroth, K.; Hartman, M.D.; Purrio, M.; Mesaros, A.; Denzel, M.S. Hexosamine pathway activation improves memory but does not extend lifespan in mice. Aging Cell 2022, 21, e13711. [Google Scholar] [CrossRef] [PubMed]
  27. Anic, K.; Schmidt, M.W.; Furtado, L.; Weidenbach, L.; Battista, M.J.; Schmidt, M.; Schwab, R.; Brenner, W.; Ruckes, C.; Lotz, J.; et al. Intermittent Fasting-Short- and Long-Term Quality of Life, Fatigue, and Safety in Healthy Volunteers: A Prospective, Clinical Trial. Nutrients 2022, 14, 4216. [Google Scholar] [CrossRef]
  28. Bose, C.; Alves, I.; Singh, P.; Palade, P.T.; Carvalho, E.; Borsheim, E.; Jun, S.R.; Cheema, A.; Boerma, M.; Awasthi, S.; et al. Sulforaphane prevents age-associated cardiac and muscular dysfunction through Nrf2 signaling. Aging Cell 2020, 19, e13261. [Google Scholar] [CrossRef]
  29. Dodds, S.G.; Parihar, M.; Javors, M.; Nie, J.; Musi, N.; Dave Sharp, Z.; Hasty, P. Acarbose improved survival for Apc(+/Min) mice. Aging Cell 2020, 19, e13088. [Google Scholar] [CrossRef] [Green Version]
  30. Dumas, S.N.; Lamming, D.W. Next Generation Strategies for Geroprotection via mTORC1 Inhibition. J. Gerontol. A Biol. Sci. Med. Sci. 2020, 75, 14–23. [Google Scholar] [CrossRef]
  31. Duque, G.; Al Saedi, A.; Rivas, D.; Miard, S.; Ferland, G.; Picard, F.; Gaudreau, P. Differential Effects of Long-Term Caloric Restriction and Dietary Protein Source on Bone and Marrow Fat of the Aging Rat. J. Gerontol. A Biol. Sci. Med. Sci. 2020, 75, 2031–2036. [Google Scholar] [CrossRef] [PubMed]
  32. Fan, S.Z.; Lin, C.S.; Wei, Y.W.; Yeh, S.R.; Tsai, Y.H.; Lee, A.C.; Lin, W.S.; Wang, P.Y. Dietary citrate supplementation enhances longevity, metabolic health, and memory performance through promoting ketogenesis. Aging Cell 2021, 20, e13510. [Google Scholar] [CrossRef] [PubMed]
  33. Houston, D.K. The Role of Diet on Life and Health Span-Lessons Learned Over the Past 75 Years. J. Gerontol. A Biol. Sci. Med. Sci. 2020, 75, 1499–1501. [Google Scholar] [CrossRef] [PubMed]
  34. Juricic, P.; Gronke, S.; Partridge, L. Branched-Chain Amino Acids Have Equivalent Effects to Other Essential Amino Acids on Lifespan and Aging-Related Traits in Drosophila. J. Gerontol. A Biol. Sci. Med. Sci. 2020, 75, 24–31. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Kim, C.S.; Cha, L.; Sim, M.; Jung, S.; Chun, W.Y.; Baik, H.W.; Shin, D.M. Probiotic Supplementation Improves Cognitive Function and Mood with Changes in Gut Microbiota in Community-Dwelling Older Adults: A Randomized, Double-Blind, Placebo-Controlled, Multicenter Trial. J. Gerontol. A Biol. Sci. Med. Sci. 2021, 76, 32–40. [Google Scholar] [CrossRef] [PubMed]
  36. Lauretti, E.; Nenov, M.; Dincer, O.; Iuliano, L.; Pratico, D. Extra virgin olive oil improves synaptic activity, short-term plasticity, memory, and neuropathology in a tauopathy model. Aging Cell 2020, 19, e13076. [Google Scholar] [CrossRef] [Green Version]
  37. Li, Z.; Xu, K.; Guo, Y.; Ping, L.; Gao, Y.; Qiu, Y.; Ni, J.; Liu, Q.; Wang, Z. A high-fat diet reverses metabolic disorders and premature aging by modulating insulin and IGF1 signaling in SIRT6 knockout mice. Aging Cell 2020, 19, e13104. [Google Scholar] [CrossRef] [Green Version]
  38. Liao, C.Y.; Kummert, O.M.P.; Bair, A.M.; Alavi, N.; Alavi, J.; Miller, D.M.; Bagga, I.; Schempf, A.M.; Hsu, Y.M.; Woods, B.D.; et al. The Autophagy Inducer Spermidine Protects Against Metabolic Dysfunction During Overnutrition. J. Gerontol. A Biol. Sci. Med. Sci. 2021, 76, 1714–1725. [Google Scholar] [CrossRef]
  39. Mautz, B.S.; Lind, M.I.; Maklakov, A.A. Dietary Restriction Improves Fitness of Aging Parents But Reduces Fitness of Their Offspring in Nematodes. J. Gerontol. A Biol. Sci. Med. Sci. 2020, 75, 843–848. [Google Scholar] [CrossRef] [Green Version]
  40. Mazzei, G.; Ikegami, R.; Abolhassani, N.; Haruyama, N.; Sakumi, K.; Saito, T.; Saido, T.C.; Nakabeppu, Y. A high-fat diet exacerbates the Alzheimer’s disease pathology in the hippocampus of the App(NL-F/NL-F) knock-in mouse model. Aging Cell 2021, 20, e13429. [Google Scholar] [CrossRef]
  41. Pini, M.; Czibik, G.; Sawaki, D.; Mezdari, Z.; Braud, L.; Delmont, T.; Mercedes, R.; Martel, C.; Buron, N.; Marcelin, G.; et al. Adipose tissue senescence is mediated by increased ATP content after a short-term high-fat diet exposure. Aging Cell 2021, 20, e13421. [Google Scholar] [CrossRef] [PubMed]
  42. Plummer, J.D.; Johnson, J.E. Intermittent methionine restriction reduces IGF-1 levels and produces similar healthspan benefits to continuous methionine restriction. Aging Cell 2022, 21, e13629. [Google Scholar] [CrossRef] [PubMed]
  43. Ramirez-Tortosa, C.L.; Varela-Lopez, A.; Navarro-Hortal, M.D.; Ramos-Pleguezuelos, F.M.; Marquez-Lobo, B.; Ramirez-Tortosa, M.; Ochoa, J.J.; Battino, M.; Quiles, J.L. Longevity and Cause of Death in Male Wistar Rats Fed Lifelong Diets Based on Virgin Olive Oil, Sunflower Oil, or Fish Oil. J. Gerontol. A Biol. Sci. Med. Sci. 2020, 75, 442–451. [Google Scholar] [CrossRef] [PubMed]
  44. Tao, S.; Wang, Y.; Wu, J.; Zeng, T.; Cui, H.; Tao, Z.; Lei, L.; Yu, L.; Liu, A.; Wang, H.; et al. Long-term mid-onset dietary restriction rejuvenates hematopoietic stem cells and improves regeneration capacity of total bone marrow from aged mice. Aging Cell 2020, 19, e13241. [Google Scholar] [CrossRef] [PubMed]
  45. Trautman, M.E.; Richardson, N.E.; Lamming, D.W. Protein restriction and branched-chain amino acid restriction promote geroprotective shifts in metabolism. Aging Cell 2022, 21, e13626. [Google Scholar] [CrossRef]
  46. Unnikrishnan, A.; Matyi, S.; Garrett, K.; Ranjo-Bishop, M.; Allison, D.B.; Ejima, K.; Chen, X.; Dickinson, S.; Richardson, A. Reevaluation of the effect of dietary restriction on different recombinant inbred lines of male and female mice. Aging Cell 2021, 20, e13500. [Google Scholar] [CrossRef]
  47. Velingkaar, N.; Mezhnina, V.; Poe, A.; Makwana, K.; Tulsian, R.; Kondratov, R.V. Reduced caloric intake and periodic fasting independently contribute to metabolic effects of caloric restriction. Aging Cell 2020, 19, e13138. [Google Scholar] [CrossRef]
  48. Wahl, D.; Cavalier, A.N.; Smith, M.; Seals, D.R.; LaRocca, T.J. Healthy Aging Interventions Reduce Repetitive Element Transcripts. J. Gerontol. A Biol. Sci. Med. Sci. 2021, 76, 805–810. [Google Scholar] [CrossRef]
  49. Wallace, M.A.; Aguirre, N.W.; Marcotte, G.R.; Marshall, A.G.; Baehr, L.M.; Hughes, D.C.; Hamilton, K.L.; Roberts, M.N.; Lopez-Dominguez, J.A.; Miller, B.F.; et al. The ketogenic diet preserves skeletal muscle with aging in mice. Aging Cell 2021, 20, e13322. [Google Scholar] [CrossRef]
  50. Wang, H.; Zhu, S.; Wu, X.; Liu, Y.; Ge, J.; Wang, Q.; Gu, L. NAMPT reduction-induced NAD(+) insufficiency contributes to the compromised oocyte quality from obese mice. Aging Cell 2021, 20, e13496. [Google Scholar] [CrossRef]
  51. Wiesenborn, D.S.; Galvez, E.J.C.; Spinel, L.; Victoria, B.; Allen, B.; Schneider, A.; Gesing, A.; Al-Regaiey, K.A.; Strowig, T.; Schafer, K.H.; et al. The Role of Ames Dwarfism and Calorie Restriction on Gut Microbiota. J. Gerontol. A Biol. Sci. Med. Sci. 2020, 75, e1–e8. [Google Scholar] [CrossRef] [PubMed]
  52. Wu, Y.; Green, C.L.; Wang, G.; Yang, D.; Li, L.; Li, B.; Wang, L.; Li, M.; Li, J.; Xu, Y.; et al. Effects of dietary macronutrients on the hepatic transcriptome and serum metabolome in mice. Aging Cell 2022, 21, e13585. [Google Scholar] [CrossRef] [PubMed]
  53. Zhang, B.; Podolskiy, D.I.; Mariotti, M.; Seravalli, J.; Gladyshev, V.N. Systematic age-, organ-, and diet-associated ionome remodeling and the development of ionomic aging clocks. Aging Cell 2020, 19, e13119. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Ros, M.; Carrascosa, J.M. Current nutritional and pharmacological anti-aging interventions. Biochim. Biophys. Acta Mol. Basis Dis. 2020, 1866, 165612. [Google Scholar] [CrossRef] [PubMed]
  55. Oikawa, S.Y.; Brisbois, T.D.; van Loon, L.J.C.; Rollo, I. Eat like an athlete: Insights of sports nutrition science to support active aging in healthy older adults. Geroscience 2021, 43, 2485–2495. [Google Scholar] [CrossRef]
  56. Rudnicka, E.; Napierala, P.; Podfigurna, A.; Meczekalski, B.; Smolarczyk, R.; Grymowicz, M. The World Health Organization (WHO) approach to healthy ageing. Maturitas 2020, 139, 6–11. [Google Scholar] [CrossRef]
  57. Jin, K. Modern Biological Theories of Aging. Aging Dis. 2010, 1, 72–74. [Google Scholar]
  58. Huan, T.; Nguyen, S.; Colicino, E.; Ochoa-Rosales, C.; Hill, W.D.; Brody, J.A.; Soerensen, M.; Zhang, Y.; Baldassari, A.; Elhadad, M.A.; et al. Integrative analysis of clinical and epigenetic biomarkers of mortality. Aging Cell 2022, 21, e13608. [Google Scholar] [CrossRef]
  59. Robinson, O.; Chadeau Hyam, M.; Karaman, I.; Climaco Pinto, R.; Ala-Korpela, M.; Handakas, E.; Fiorito, G.; Gao, H.; Heard, A.; Jarvelin, M.R.; et al. Determinants of accelerated metabolomic and epigenetic aging in a UK cohort. Aging Cell 2020, 19, e13149. [Google Scholar] [CrossRef]
  60. Sandoval-Sierra, J.V.; Helbing, A.H.B.; Williams, E.G.; Ashbrook, D.G.; Roy, S.; Williams, R.W.; Mozhui, K. Body weight and high-fat diet are associated with epigenetic aging in female members of the BXD murine family. Aging Cell 2020, 19, e13207. [Google Scholar] [CrossRef]
  61. Fulop, T.; Witkowski, J.M.; Pawelec, G.; Alan, C.; Larbi, A. On the immunological theory of aging. Interdiscip. Top. Gerontol. 2014, 39, 163–176. [Google Scholar] [CrossRef] [PubMed]
  62. Farkas, A.; Szipocs, A.; Horvath, A.; Horvath, I.; Galffy, G.; Varga, J.; Galambos, K.; Kugler, S.; Nagy, A.; Szalai, Z. Establishment of relationships between native and inhalation device specific spirometric parameters as a step towards patient tailored inhalation device selection. Respir. Med. 2019, 154, 133–140. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Ballesteros-Alvarez, J.; Andersen, J.K. mTORC2: The other mTOR in autophagy regulation. Aging Cell 2021, 20, e13431. [Google Scholar] [CrossRef] [PubMed]
  64. Van Skike, C.E.; Lin, A.L.; Roberts Burbank, R.; Halloran, J.J.; Hernandez, S.F.; Cuvillier, J.; Soto, V.Y.; Hussong, S.A.; Jahrling, J.B.; Javors, M.A.; et al. mTOR drives cerebrovascular, synaptic, and cognitive dysfunction in normative aging. Aging Cell 2020, 19, e13057. [Google Scholar] [CrossRef] [PubMed]
  65. Fang, Y.; An, N.; Zhu, L.; Gu, Y.; Qian, J.; Jiang, G.; Zhao, R.; Wei, W.; Xu, L.; Zhang, G.; et al. Autophagy-Sirt3 axis decelerates hematopoietic aging. Aging Cell 2020, 19, e13232. [Google Scholar] [CrossRef] [PubMed]
  66. Iljas, J.D.; Wei, Z.; Homer, H.A. Sirt1 sustains female fertility by slowing age-related decline in oocyte quality required for post-fertilization embryo development. Aging Cell 2020, 19, e13204. [Google Scholar] [CrossRef]
  67. Li, Q.; Cheng, J.C.; Jiang, Q.; Lee, W.Y. Role of sirtuins in bone biology: Potential implications for novel therapeutic strategies for osteoporosis. Aging Cell 2021, 20, e13301. [Google Scholar] [CrossRef]
  68. Long, C.; Liu, H.; Zhan, W.; Chen, L.; Yu, Z.; Tian, S.; Xiang, Y.; Chen, S.; Tian, X.L. Chronological attenuation of NPRA/PKG/AMPK signaling promotes vascular aging and elevates blood pressure. Aging Cell 2022, 21, e13699. [Google Scholar] [CrossRef]
  69. Zhang, J.; He, Z.; Fedorova, J.; Logan, C.; Bates, L.; Davitt, K.; Le, V.; Murphy, J.; Li, M.; Wang, M.; et al. Alterations in mitochondrial dynamics with age-related Sirtuin1/Sirtuin3 deficiency impair cardiomyocyte contractility. Aging Cell 2021, 20, e13419. [Google Scholar] [CrossRef]
  70. Rahmani, J.; Montesanto, A.; Giovannucci, E.; Zand, H.; Barati, M.; Kopchick, J.J.; Mirisola, M.G.; Lagani, V.; Bawadi, H.; Vardavas, R.; et al. Association between IGF-1 levels ranges and all-cause mortality: A meta-analysis. Aging Cell 2022, 21, e13540. [Google Scholar] [CrossRef]
  71. Sherlala, R.A.; Kammerer, C.M.; Kuipers, A.L.; Wojczynski, M.K.; Ukraintseva, S.V.; Feitosa, M.F.; Mengel-From, J.; Zmuda, J.M.; Minster, R.L. Relationship Between Serum IGF-1 and BMI Differs by Age. J. Gerontol. A Biol. Sci. Med. Sci. 2021, 76, 1303–1308. [Google Scholar] [CrossRef]
  72. Zhang, W.B.; Ye, K.; Barzilai, N.; Milman, S. The antagonistic pleiotropy of insulin-like growth factor 1. Aging Cell 2021, 20, e13443. [Google Scholar] [CrossRef] [PubMed]
  73. Plagg, B.; Zerbe, S. How does the environment affect human ageing? An interdisciplinary review. J. Gerontol. Geriatr. 2020, 69, 53–67. [Google Scholar] [CrossRef]
  74. Sattaur, Z.; Lashley, L.K.; Golden, C.J. Wear and Tear Theory of Aging. Available online: https://www.assessmentpsychologyboard.org/edp/pdf/Wear_and_Tear_Theory_of_Aging.pdf (accessed on 5 May 2022).
  75. Tarantini, S.; Valcarcel-Ares, N.M.; Yabluchanskiy, A.; Springo, Z.; Fulop, G.A.; Ashpole, N.; Gautam, T.; Giles, C.B.; Wren, J.D.; Sonntag, W.E.; et al. Insulin-like growth factor 1 deficiency exacerbates hypertension-induced cerebral microhemorrhages in mice, mimicking the aging phenotype. Aging Cell 2017, 16, 469–479. [Google Scholar] [CrossRef] [PubMed]
  76. Tarantini, S.; Valcarcel-Ares, M.N.; Yabluchanskiy, A.; Tucsek, Z.; Hertelendy, P.; Kiss, T.; Gautam, T.; Zhang, X.A.; Sonntag, W.E.; de Cabo, R.; et al. Nrf2 deficiency exacerbates obesity-induced oxidative stress, neurovascular dysfunction, blood brain barrier disruption, neuroinflammation, amyloidogenic gene expression and cognitive decline in mice, mimicking the aging phenotype. J. Gerontol. A Biol. Sci. Med. Sci. 2018, 73, 853–863. [Google Scholar] [CrossRef] [PubMed]
  77. Tarantini, S.; Valcarcel-Ares, N.M.; Yabluchanskiy, A.; Fulop, G.A.; Hertelendy, P.; Gautam, T.; Farkas, E.; Perz, A.; Rabinovitch, P.S.; Sonntag, W.E.; et al. Treatment with the mitochondrial-targeted antioxidant peptide SS-31 rescues neurovascular coupling responses and cerebrovascular endothelial function and improves cognition in aged mice. Aging Cell 2018, 17, e12731. [Google Scholar] [CrossRef]
  78. Ungvari, Z.; Tarantini, S.; Donato, A.J.; Galvan, V.; Csiszar, A. Mechanisms of Vascular Aging. Circ. Res. 2018, 123, 849–867. [Google Scholar] [CrossRef]
  79. Csiszar, A.; Tarantini, S.; Yabluchanskiy, A.; Balasubramanian, P.; Kiss, T.; Farkas, E.; Baur, J.A.; Ungvari, Z.I. Role of endothelial NAD+ deficiency in age-related vascular dysfunction. Am. J. Physiol. Heart Circ. Physiol. 2019, 316, H1253–H1266. [Google Scholar] [CrossRef]
  80. Csiszar, A.; Yabluchanskiy, A.; Ungvari, A.; Ungvari, Z.; Tarantini, S. Overexpression of catalase targeted to mitochondria improves neurovascular coupling responses in aged mice. Geroscience 2019, 41, 609–617. [Google Scholar] [CrossRef]
  81. Kiss, T.; Balasubramanian, P.; Valcarcel-Ares, M.N.; Tarantini, S.; Yabluchanskiy, A.; Csipo, T.; Lipecz, A.; Reglodi, D.; Zhang, X.A.; Bari, F.; et al. Nicotinamide mononucleotide (NMN) treatment attenuates oxidative stress and rescues angiogenic capacity in aged cerebromicrovascular endothelial cells: A potential mechanism for the prevention of vascular cognitive impairment. Geroscience 2019, 41, 619–630. [Google Scholar] [CrossRef]
  82. Tarantini, S.; Valcarcel-Ares, M.N.; Toth, P.; Yabluchanskiy, A.; Tucsek, Z.; Kiss, T.; Hertelendy, P.; Kinter, M.; Ballabh, P.; Sule, Z.; et al. Nicotinamide mononucleotide (NMN) supplementation rescues cerebromicrovascular endothelial function and neurovascular coupling responses and improves cognitive function in aged mice. Redox Biol. 2019, 24, 101192. [Google Scholar] [CrossRef] [PubMed]
  83. Tarantini, S.; Yabluchanskiy, A.; Csipo, T.; Fulop, G.; Kiss, T.; Balasubramanian, P.; DelFavero, J.; Ahire, C.; Ungvari, A.; Nyul-Toth, A.; et al. Treatment with the poly(ADP-ribose) polymerase inhibitor PJ-34 improves cerebromicrovascular endothelial function, neurovascular coupling responses and cognitive performance in aged mice, supporting the NAD+ depletion hypothesis of neurovascular aging. Geroscience 2019, 41, 533–542. [Google Scholar] [CrossRef] [PubMed]
  84. Pathath, A. Theories of Aging. Int. J. Indian Psychol. 2017, 4, 15–22. [Google Scholar] [CrossRef]
  85. Fekete, M.; Fazekas-Pongor, V.; Szőllősi, G.; Varga, J.T. Metabolic consequences of chronic obstructive pulmonary disease. Orv. Hetil. 2021, 162, 185–191. [Google Scholar] [CrossRef] [PubMed]
  86. Copp, M.E.; Chubinskaya, S.; Bracey, D.N.; Shine, J.; Sessions, G.; Loeser, R.F.; Diekman, B.O. Comet assay for quantification of the increased DNA damage burden in primary human chondrocytes with aging and osteoarthritis. Aging Cell 2022, 21, e13698. [Google Scholar] [CrossRef] [PubMed]
  87. Tuttle, C.S.L.; Waaijer, M.E.C.; Slee-Valentijn, M.S.; Stijnen, T.; Westendorp, R.; Maier, A.B. Cellular senescence and chronological age in various human tissues: A systematic review and meta-analysis. Aging Cell 2020, 19, e13083. [Google Scholar] [CrossRef] [Green Version]
  88. Vougioukalaki, M.; Demmers, J.; Vermeij, W.P.; Baar, M.; Bruens, S.; Magaraki, A.; Kuijk, E.; Jager, M.; Merzouk, S.; Brandt, R.M.C.; et al. Different responses to DNA damage determine ageing differences between organs. Aging Cell 2022, 21, e13562. [Google Scholar] [CrossRef]
  89. Wang, L.; Lu, Z.; Zhao, J.; Schank, M.; Cao, D.; Dang, X.; Nguyen, L.N.; Nguyen, L.N.T.; Khanal, S.; Zhang, J.; et al. Selective oxidative stress induces dual damage to telomeres and mitochondria in human T cells. Aging Cell 2021, 20, e13513. [Google Scholar] [CrossRef]
  90. Wasserzug-Pash, P.; Rothman, R.; Reich, E.; Zecharyahu, L.; Schonberger, O.; Weiss, Y.; Srebnik, N.; Cohen-Hadad, Y.; Weintraub, A.; Ben-Ami, I.; et al. Loss of heterochromatin and retrotransposon silencing as determinants in oocyte aging. Aging Cell 2022, 21, e13568. [Google Scholar] [CrossRef]
  91. Stanic, S.M.; Matic, S.L. The biology and t heories of aging. Biol. Forum. 2019, 112, 79–89. [Google Scholar] [CrossRef]
  92. Vijg, J. From DNA damage to mutations: All roads lead to aging. Ageing Res. Rev. 2021, 68, 101316. [Google Scholar] [CrossRef] [PubMed]
  93. Okawa, T.; Nagai, M.; Hase, K. Dietary Intervention Impacts Immune Cell Functions and Dynamics by Inducing Metabolic Rewiring. Front. Immunol. 2020, 11, 623989. [Google Scholar] [CrossRef] [PubMed]
  94. Andrianova, N.V.; Buyan, M.I.; Bolikhova, A.K.; Zorov, D.B.; Plotnikov, E.Y. Dietary Restriction for Kidney Protection: Decline in Nephroprotective Mechanisms During Aging. Front. Physiol. 2021, 12, 699490. [Google Scholar] [CrossRef] [PubMed]
  95. Fekete, M.; Szollosi, G.; Tarantini, S.; Lehoczki, A.; Nemeth, A.N.; Bodola, C.; Varga, L.; Varga, J.T. Metabolic syndrome in patients with COPD: Causes and pathophysiological consequences. Physiol. Int. 2022. [Google Scholar] [CrossRef]
  96. Duan, Y.; Zeng, L.; Zheng, C.; Song, B.; Li, F.; Kong, X.; Xu, K. Inflammatory Links Between High Fat Diets and Diseases. Front. Immunol. 2018, 9, 2649. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Santos, L.S.; Cordeiro, G.S.; Matos, R.J.B.; Perez, G.S.; Silva, R.T.; Boaventura, G.T.; Barreto-Medeiros, J.M. High-fat diet promotes hypothalamic inflammation in animal models: A systematic review. Nutr. Rev. 2022, 80, 392–399. [Google Scholar] [CrossRef] [PubMed]
  98. Garman, T.S.; Setlow, B.; Orsini, C.A. Effects of a high-fat diet on impulsive choice in rats. Physiol. Behav. 2021, 229, 113260. [Google Scholar] [CrossRef]
  99. Bittencourt, A.; Brum, P.O.; Ribeiro, C.T.; Gasparotto, J.; Bortolin, R.C.; de Vargas, A.R.; Heimfarth, L.; de Almeida, R.F.; Moreira, J.C.F.; de Oliveira, J.; et al. High fat diet-induced obesity causes a reduction in brain tyrosine hydroxylase levels and non-motor features in rats through metabolic dysfunction, neuroinflammation and oxidative stress. Nutr. Neurosci. 2022, 25, 1026–1040. [Google Scholar] [CrossRef]
  100. Valcarcel-Ares, M.N.; Tucsek, Z.; Kiss, T.; Giles, C.B.; Tarantini, S.; Yabluchanskiy, A.; Balasubramanian, P.; Gautam, T.; Galvan, V.; Ballabh, P.; et al. Obesity in Aging Exacerbates Neuroinflammation, Dysregulating Synaptic Function-related Genes and Altering Eicosanoid Synthesis in the Mouse Hippocampus: Potential Role in Impaired Synaptic Plasticity and Cognitive Decline. J. Gerontol. A Biol. Sci. Med. Sci. 2019, 74, 290–298. [Google Scholar] [CrossRef]
  101. Pallauf, K.; Gunther, I.; Kuhn, G.; Chin, D.; de Pascual-Teresa, S.; Rimbach, G. The Potential of Resveratrol to Act as a Caloric Restriction Mimetic Appears to Be Limited: Insights from Studies in Mice. Adv. Nutr. 2021, 12, 995–1005. [Google Scholar] [CrossRef]
  102. Goedeke, L.; Murt, K.N.; Di Francesco, A.; Camporez, J.P.; Nasiri, A.R.; Wang, Y.; Zhang, X.M.; Cline, G.W.; de Cabo, R.; Shulman, G.I. Sex- and strain-specific effects of mitochondrial uncoupling on age-related metabolic diseases in high-fat diet-fed mice. Aging Cell 2022, 21, e13539. [Google Scholar] [CrossRef] [PubMed]
  103. Hyatt, J.K.; de Cabo, R.; Mattison, J.A. Resveratrol Blunts Mitochondrial Loss in Slow and Mixed Skeletal Muscle Phenotypes of Non-Human Primates following a Long-Term High Fat/Sugar Diet. J. Diet Suppl. 2022, 1–19. [Google Scholar] [CrossRef] [PubMed]
  104. Mishra, A.; Mirzaei, H.; Guidi, N.; Vinciguerra, M.; Mouton, A.; Linardic, M.; Rappa, F.; Barone, R.; Navarrete, G.; Wei, M.; et al. Fasting-mimicking diet prevents high-fat diet effect on cardiometabolic risk and lifespan. Nat. Metab. 2021, 3, 1342–1356. [Google Scholar] [CrossRef] [PubMed]
  105. Solana, C.; Tarazona, R.; Solana, R. Immunosenescence of Natural Killer Cells, Inflammation, and Alzheimer’s Disease. Int. J. Alzheimers Dis. 2018, 2018, 3128758. [Google Scholar] [CrossRef] [PubMed]
  106. Jung, S.J.; Lee, G.B.; Nishimi, K.; Chibnik, L.; Koenen, K.C.; Kim, H.C. Association between psychological resilience and cognitive function in older adults: Effect modification by inflammatory status. Geroscience 2021, 43, 2749–2760. [Google Scholar] [CrossRef] [PubMed]
  107. Marcos-Perez, D.; Sanchez-Flores, M.; Proietti, S.; Bonassi, S.; Costa, S.; Teixeira, J.P.; Fernandez-Tajes, J.; Pasaro, E.; Laffon, B.; Valdiglesias, V. Association of inflammatory mediators with frailty status in older adults: Results from a systematic review and meta-analysis. Geroscience 2020, 42, 1451–1473. [Google Scholar] [CrossRef] [PubMed]
  108. Wu, Q.; Gao, Z.J.; Yu, X.; Wang, P. Dietary regulation in health and disease. Signal Transduct. Target. Ther. 2022, 7, 252. [Google Scholar] [CrossRef]
  109. Kovalska, M.; Baranovicova, E.; Kalenska, D.; Tomascova, A.; Adamkov, M.; Kovalska, L.; Lehotsky, J. Methionine Diet Evoked Hyperhomocysteinemia Causes Hippocampal Alterations, Metabolomics Plasma Changes and Behavioral Pattern in Wild Type Rats. Int. J. Mol. Sci. 2021, 22, 4961. [Google Scholar] [CrossRef]
  110. Aissa, A.F.; Amaral, C.L.D.; Venancio, V.P.; Machado, C.D.S.; Hernandes, L.C.; Santos, P.; Curi, R.; Bianchi, M.L.P.; Antunes, L.M.G. Methionine-supplemented diet affects the expression of cardiovascular disease-related genes and increases inflammatory cytokines in mice heart and liver. J. Toxicol. Environ. Health A 2017, 80, 1116–1128. [Google Scholar] [CrossRef] [PubMed]
  111. Fekete, M.; Kerti, M.; Fazekas-Pongor, V.; Balazs, P.; Csizmadia, Z.; Nemeth, A.N.; Tarantini, S.; Varga, J.T. Effect of interval training with non-invasive ventilation in severe chronic obstructive pulmonary disease—A prospective cohort study with matched control group. Ann. Palliat. Med. 2021, 10, 5289–5298. [Google Scholar] [CrossRef]
  112. Institute of Medicine (US) Food Forum. Providing Healthy and Safe Foods As We Age: Workshop Summary; National Academies Press (US): Washington, DC, USA, 2010. [Google Scholar]
  113. Tur, J.A.; Bibiloni, M.D.M. Anthropometry, Body Composition and Resting Energy Expenditure in Human. Nutrients 2019, 11, 1891. [Google Scholar] [CrossRef] [PubMed]
  114. Sarkadi Nagy, E.; Bakacs, M.; Illés, É.; Nagy, B.; Varga, A.; Kis, O.; Schreiberné Molnár, E.; Martos, É. Hungarian Diet and Nutritional Status Survey—OTÁP2014. II. Energy and macronutrient intake of the Hungarian population. Orv. Hetil. 2017, 158, 587–597. [Google Scholar] [CrossRef] [PubMed]
  115. Szucs, B.; Petrekanits, M.; Varga, J. Effectiveness of a 4-week rehabilitation program on endothelial function, blood vessel elasticity in patients with chronic obstructive pulmonary disease. J. Thorac. Dis. 2018, 10, 6482–6490. [Google Scholar] [CrossRef] [PubMed]
  116. Schwingshackl, L.; Zahringer, J.; Beyerbach, J.; Werner, S.S.; Nagavci, B.; Heseker, H.; Koletzko, B.; Meerpohl, J.J. A Scoping Review of Current Guidelines on Dietary Fat and Fat Quality. Ann. Nutr. Metab. 2021, 77, 65–82. [Google Scholar] [CrossRef] [PubMed]
  117. Zinocker, M.K.; Svendsen, K.; Dankel, S.N. The homeoviscous adaptation to dietary lipids (HADL) model explains controversies over saturated fat, cholesterol, and cardiovascular disease risk. Am. J. Clin. Nutr. 2021, 113, 277–289. [Google Scholar] [CrossRef]
  118. Krauss, R.M.; Kris-Etherton, P.M. Public health guidelines should recommend reducing saturated fat consumption as much as possible: NO. Am. J. Clin. Nutr. 2020, 112, 19–24. [Google Scholar] [CrossRef]
  119. Neuenschwander, M.; Barbaresko, J.; Pischke, C.R.; Iser, N.; Beckhaus, J.; Schwingshackl, L.; Schlesinger, S. Intake of dietary fats and fatty acids and the incidence of type 2 diabetes: A systematic review and dose-response meta-analysis of prospective observational studies. PLoS Med. 2020, 17, e1003347. [Google Scholar] [CrossRef]
  120. Herforth, A.; Arimond, M.; Álvarez-Sánchez, C.; Coates, J.; Christianson, K.; Muehlhoff, E. A Global Review of Food-Based Dietary Guidelines. Adv. Nutr. 2019, 10, 590–605. [Google Scholar] [CrossRef] [Green Version]
  121. Cristina, N.M.; Lucia, D. Nutrition and Healthy Aging: Prevention and Treatment of Gastrointestinal Diseases. Nutrients 2021, 13, 4337. [Google Scholar] [CrossRef]
  122. Barber, T.M.; Kabisch, S.; Pfeiffer, A.F.H.; Weickert, M.O. The Health Benefits of Dietary Fibre. Nutrients 2020, 12, 3209. [Google Scholar] [CrossRef]
  123. Thompson, H.J. The Dietary Guidelines for Americans (2020-2025): Pulses, Dietary Fiber, and Chronic Disease Risk-A Call for Clarity and Action. Nutrients 2021, 13, 4034. [Google Scholar] [CrossRef] [PubMed]
  124. Wang, Y.; Jia, X.F.; Zhang, B.; Wang, Z.H.; Zhang, J.G.; Huang, F.F.; Su, C.; Ouyang, Y.F.; Zhao, J.; Du, W.W.; et al. Dietary Zinc Intake and Its Association with Metabolic Syndrome Indicators among Chinese Adults: An Analysis of the China Nutritional Transition Cohort Survey 2015. Nutrients 2018, 10, 572. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Olza, J.; Aranceta-Bartrina, J.; González-Gross, M.; Ortega, R.M.; Serra-Majem, L.; Varela-Moreiras, G.; Gil, Á. Reported Dietary Intake and Food Sources of Zinc, Selenium, and Vitamins A, E and C in the Spanish Population: Findings from the ANIBES Study. Nutrients 2017, 9, 697. [Google Scholar] [CrossRef]
  126. Masot, O.; Miranda, J.; Santamaria, A.L.; Paraiso Pueyo, E.; Pascual, A.; Botigue, T. Fluid Intake Recommendation Considering the Physiological Adaptations of Adults Over 65 Years: A Critical Review. Nutrients 2020, 12, 3383. [Google Scholar] [CrossRef] [PubMed]
  127. Lacey, J.; Corbett, J.; Forni, L.; Hooper, L.; Hughes, F.; Minto, G.; Moss, C.; Price, S.; Whyte, G.; Woodcock, T.; et al. A multidisciplinary consensus on dehydration: Definitions, diagnostic methods and clinical implications. Ann. Med. 2019, 51, 232–251. [Google Scholar] [CrossRef] [PubMed]
  128. Woodward, M. Guidelines to Effective Hydration in Aged Care Facilities. 2013. Available online: https://www.mcgill.ca/familymed/files/familymed/effective_hydration_in_elderly.pdf (accessed on 10 November 2022).
  129. European Food Safety Authority (EFSA). Scientific opinion on dietary reference values for water. EFSA J. 2010, 8, 1459. [Google Scholar]
  130. Volkert, D.; Beck, A.M.; Cederholm, T.; Cruz-Jentoft, A.; Goisser, S.; Hooper, L.; Kiesswetter, E.; Maggio, M.; Raynaud-Simon, A.; Sieber, C.C.; et al. ESPEN guideline on clinical nutrition and hydration in geriatrics. Clin. Nutr. 2019, 38, 10–47. [Google Scholar] [CrossRef] [Green Version]
  131. Watso, J.C.; Farquhar, W.B. Hydration Status and Cardiovascular Function. Nutrients 2019, 11, 1866. [Google Scholar] [CrossRef] [Green Version]
  132. Leung, A.A.; McAlister, F.A.; Finlayson, S.R.; Bates, D.W. Preoperative hypernatremia predicts increased perioperative morbidity and mortality. Am. J. Med. 2013, 126, 877–886. [Google Scholar] [CrossRef]
  133. Thornton, S.N. Angiotensin inhibition and longevity: A question of hydration. Pflug. Arch 2011, 461, 317–324. [Google Scholar] [CrossRef]
  134. Lauriola, M.; Mangiacotti, A.; D’Onofrio, G.; Cascavilla, L.; Paris, F.; Paroni, G.; Seripa, D.; Greco, A.; Sancarlo, D. Neurocognitive Disorders and Dehydration in Older Patients: Clinical Experience Supports the Hydromolecular Hypothesis of Dementia. Nutrients 2018, 10, 562. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Masento, N.A.; Golightly, M.; Field, D.T.; Butler, L.T.; van Reekum, C.M. Effects of hydration status on cognitive performance and mood. Br. J. Nutr. 2014, 111, 1841–1852. [Google Scholar] [CrossRef] [PubMed]
  136. Pross, N. Effects of Dehydration on Brain Functioning: A Life-Span Perspective. Ann. Nutr. Metab. 2017, 70 (Suppl. 1), 30–36. [Google Scholar] [CrossRef] [PubMed]
  137. Scherer, R.; Maroto-Sánchez, B.; Palacios, G.; González-Gross, M. Fluid intake and recommendations in older adults: More data are needed. Nutr. Bull. 2016, 41, 167–174. [Google Scholar] [CrossRef]
  138. Buoite Stella, A.; Gaio, M.; Furlanis, G.; Douglas, P.; Naccarato, M.; Manganotti, P. Fluid and energy intake in stroke patients during acute hospitalization in a stroke unit. J. Clin. Neurosci. 2019, 62, 27–32. [Google Scholar] [CrossRef]
  139. Longo, V.D.; Anderson, R.M. Nutrition, longevity and disease: From molecular mechanisms to interventions. Cell 2022, 185, 1455–1470. [Google Scholar] [CrossRef]
  140. Gensous, N.; Garagnani, P.; Santoro, A.; Giuliani, C.; Ostan, R.; Fabbri, C.; Milazzo, M.; Gentilini, D.; di Blasio, A.M.; Pietruszka, B.; et al. One-year Mediterranean diet promotes epigenetic rejuvenation with country- and sex-specific effects: A pilot study from the NU-AGE project. Geroscience 2020, 42, 687–701. [Google Scholar] [CrossRef] [Green Version]
  141. Berendsen, A.; Santoro, A.; Pini, E.; Cevenini, E.; Ostan, R.; Pietruszka, B.; Rolf, K.; Cano, N.; Caille, A.; Lyon-Belgy, N.; et al. A parallel randomized trial on the effect of a healthful diet on inflammageing and its consequences in European elderly people: Design of the NU-AGE dietary intervention study. Mech. Ageing Dev. 2013, 134, 523–530. [Google Scholar] [CrossRef]
  142. Clegg, M.E.; Williams, E.A. Optimizing nutrition in older people. Maturitas 2018, 112, 34–38. [Google Scholar] [CrossRef]
  143. Meehan, M.; Penckofer, S. The Role of Vitamin D in the Aging Adult. J. Aging Gerontol. 2014, 2, 60–71. [Google Scholar] [CrossRef] [Green Version]
  144. Capurso, C.; Bellanti, F.; Lo Buglio, A.; Vendemiale, G. The Mediterranean Diet Slows Down the Progression of Aging and Helps to Prevent the Onset of Frailty: A Narrative Review. Nutrients 2019, 12, 35. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Schubel, R.; Nattenmuller, J.; Sookthai, D.; Nonnenmacher, T.; Graf, M.E.; Riedl, L.; Schlett, C.L.; von Stackelberg, O.; Johnson, T.; Nabers, D.; et al. Effects of intermittent and continuous calorie restriction on body weight and metabolism over 50 wk: A randomized controlled trial. Am. J. Clin. Nutr. 2018, 108, 933–945. [Google Scholar] [CrossRef] [PubMed]
  146. Most, J.; Tosti, V.; Redman, L.M.; Fontana, L. Calorie restriction in humans: An update. Ageing Res. Rev. 2017, 39, 36–45. [Google Scholar] [CrossRef] [PubMed]
  147. Schubel, R.; Graf, M.E.; Nattenmuller, J.; Nabers, D.; Sookthai, D.; Gruner, L.F.; Johnson, T.; Schlett, C.L.; von Stackelberg, O.; Kirsten, R.; et al. The effects of intermittent calorie restriction on metabolic health: Rationale and study design of the HELENA Trial. Contemp. Clin. Trials 2016, 51, 28–33. [Google Scholar] [CrossRef]
  148. Most, J.; Redman, L.M. Impact of calorie restriction on energy metabolism in humans. Exp. Gerontol. 2020, 133, 110875. [Google Scholar] [CrossRef] [PubMed]
  149. Flanagan, E.W.; Most, J.; Mey, J.T.; Redman, L.M. Calorie Restriction and Aging in Humans. Annu. Rev. Nutr. 2020, 40, 105–133. [Google Scholar] [CrossRef]
  150. Dorling, J.L.; Martin, C.K.; Redman, L.M. Calorie restriction for enhanced longevity: The role of novel dietary strategies in the present obesogenic environment. Ageing Res. Rev. 2020, 64, 101038. [Google Scholar] [CrossRef]
  151. Caristia, S.; Vito, M.; Sarro, A.; Leone, A.; Pecere, A.; Zibetti, A.; Filigheddu, N.; Zeppegno, P.; Prodam, F.; Faggiano, F.; et al. Is Caloric Restriction Associated with Better Healthy Aging Outcomes? A Systematic Review and Meta-Analysis of Randomized Controlled Trials. Nutrients 2020, 12, 2290. [Google Scholar] [CrossRef]
  152. O’Leary, K. Health benefits of calorie restriction. Nat. Med. 2022. [Google Scholar] [CrossRef]
  153. Kokten, T.; Hansmannel, F.; Ndiaye, N.C.; Heba, A.C.; Quilliot, D.; Dreumont, N.; Arnone, D.; Peyrin-Biroulet, L. Calorie Restriction as a New Treatment of Inflammatory Diseases. Adv. Nutr. 2021, 12, 1558–1570. [Google Scholar] [CrossRef]
  154. Parikh, I.; Guo, J.; Chuang, K.H.; Zhong, Y.; Rempe, R.G.; Hoffman, J.D.; Armstrong, R.; Bauer, B.; Hartz, A.M.; Lin, A.L. Caloric restriction preserves memory and reduces anxiety of aging mice with early enhancement of neurovascular functions. Aging (Albany NY) 2016, 8, 2814–2826. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Miller, K.N.; Burhans, M.S.; Clark, J.P.; Howell, P.R.; Polewski, M.A.; DeMuth, T.M.; Eliceiri, K.W.; Lindstrom, M.J.; Ntambi, J.M.; Anderson, R.M. Aging and caloric restriction impact adipose tissue, adiponectin, and circulating lipids. Aging Cell 2017, 16, 497–507. [Google Scholar] [CrossRef] [PubMed]
  156. Wilkie, S.E.; Mulvey, L.; Sands, W.A.; Marcu, D.E.; Carter, R.N.; Morton, N.M.; Hine, C.; Mitchell, J.R.; Selman, C. Strain-specificity in the hydrogen sulphide signalling network following dietary restriction in recombinant inbred mice. Geroscience 2020, 42, 801–812. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Richardson, A. You Have Come A Long Way Baby: Five Decades of Research on the Biology of Aging From the Perspective of a Researcher Studying Aging. J. Gerontol. A Biol. Sci. Med. Sci. 2021, 76, 57–63. [Google Scholar] [CrossRef]
  158. Das, S.K.; Balasubramanian, P.; Weerasekara, Y.K. Nutrition modulation of human aging: The calorie restriction paradigm. Mol. Cell. Endocrinol. 2017, 455, 148–157. [Google Scholar] [CrossRef]
  159. Trivedi, S.; Pandey, R. 5’-Hydroxy-6, 7, 8, 3’, 4’-pentamethoxyflavone extends longevity mediated by DR-induced autophagy and oxidative stress resistance in C. elegans. Geroscience 2021, 43, 759–772. [Google Scholar] [CrossRef]
  160. Komninou, D.; Malloy, V.L.; Zimmerman, J.A.; Sinha, R.; Richie, J.P., Jr. Methionine restriction delays aging-related urogenital diseases in male Fischer 344 rats. Geroscience 2020, 42, 287–297. [Google Scholar] [CrossRef]
  161. Attina, A.; Leggeri, C.; Paroni, R.; Pivari, F.; Dei Cas, M.; Mingione, A.; Dri, M.; Marchetti, M.; Di Renzo, L. Fasting: How to Guide. Nutrients 2021, 13, 1570. [Google Scholar] [CrossRef]
  162. Duregon, E.; Pomatto-Watson, L.; Bernier, M.; Price, N.L.; de Cabo, R. Intermittent fasting: From calories to time restriction. Geroscience 2021, 43, 1083–1092. [Google Scholar] [CrossRef]
  163. Henderson, Y.O.; Bithi, N.; Link, C.; Yang, J.; Schugar, R.; Llarena, N.; Brown, J.M.; Hine, C. Late-life intermittent fasting decreases aging-related frailty and increases renal hydrogen sulfide production in a sexually dimorphic manner. Geroscience 2021, 43, 1527–1554. [Google Scholar] [CrossRef]
  164. De Cabo, R.; Mattson, M.P. Effects of Intermittent Fasting on Health, Aging, and Disease. N. Engl. J. Med. 2019, 381, 2541–2551. [Google Scholar] [CrossRef] [PubMed]
  165. Vidoni, C.; Ferraresi, A.; Esposito, A.; Maheshwari, C.; Dhanasekaran, D.N.; Mollace, V.; Isidoro, C. Calorie Restriction for Cancer Prevention and Therapy: Mechanisms, Expectations, and Efficacy. J. Cancer Prev. 2021, 26, 224–236. [Google Scholar] [CrossRef] [PubMed]
  166. Brandhorst, S.; Longo, V.D. Fasting and Caloric Restriction in Cancer Prevention and Treatment. Recent Results Cancer Res. 2016, 207, 241–266. [Google Scholar] [CrossRef] [PubMed]
  167. Brandhorst, S. Fasting and fasting-mimicking diets for chemotherapy augmentation. Geroscience 2021, 43, 1201–1216. [Google Scholar] [CrossRef]
  168. Kaya, A.; Phua, C.Z.J.; Lee, M.; Wang, L.; Tyshkovskiy, A.; Ma, S.; Barre, B.; Liu, W.; Harrison, B.R.; Zhao, X.; et al. Evolution of natural lifespan variation and molecular strategies of extended lifespan in yeast. eLife 2021, 10, e64860. [Google Scholar] [CrossRef]
  169. Fabrizio, P.; Pozza, F.; Pletcher, S.D.; Gendron, C.M.; Longo, V.D. Regulation of Longevity and Stress Resistance by Sch9 in Yeast. Science 2001, 292, 288–290. [Google Scholar] [CrossRef] [Green Version]
  170. Balasubramanian, P.; DelFavero, J.; Ungvari, A.; Papp, M.; Tarantini, A.; Price, N.; de Cabo, R.; Tarantini, S. Time-restricted feeding (TRF) for prevention of age-related vascular cognitive impairment and dementia. Ageing Res. Rev. 2020, 64, 101189. [Google Scholar] [CrossRef]
  171. Martens, C.R.; Rossman, M.J.; Mazzo, M.R.; Jankowski, L.R.; Nagy, E.E.; Denman, B.A.; Richey, J.J.; Johnson, S.A.; Ziemba, B.P.; Wang, Y.; et al. Short-term time-restricted feeding is safe and feasible in non-obese healthy midlife and older adults. Geroscience 2020, 42, 667–686. [Google Scholar] [CrossRef]
  172. Riley, L.; Guthold, R.; Cowan, M.; Savin, S.; Bhatti, L.; Armstrong, T.; Bonita, R. The World Health Organization STEPwise Approach to Noncommunicable Disease Risk-Factor Surveillance: Methods, Challenges, and Opportunities. Am. J. Public Health 2016, 106, 74–78. [Google Scholar] [CrossRef]
  173. Jayedi, A.; Soltani, S.; Abdolshahi, A.; Shab-Bidar, S. Healthy and unhealthy dietary patterns and the risk of chronic disease: An umbrella review of meta-analyses of prospective cohort studies. Br. J. Nutr. 2020, 124, 1133–1144. [Google Scholar] [CrossRef]
  174. Yu, E.; Malik, V.S.; Hu, F.B. Cardiovascular Disease Prevention by Diet Modification: JACC Health Promotion Series. J. Am. Coll. Cardiol. 2018, 72, 914–926. [Google Scholar] [CrossRef] [PubMed]
  175. Yumuk, V.; Tsigos, C.; Fried, M.; Schindler, K.; Busetto, L.; Micic, D.; Toplak, H.; Obesity Management Task Force of the European Association for the Study of Obesity. European Guidelines for Obesity Management in Adults. Obes. Facts 2015, 8, 402–424. [Google Scholar] [CrossRef] [PubMed]
  176. Péter, S.; Eggersdorfer, M.; van Asselt, D.; Buskens, E.; Detzel, P.; Freijer, K.; Koletzko, B.; Kraemer, K.; Kuipers, F.; Neufeld, L.; et al. Selected nutrients and their implications for health and disease across the lifespan: A roadmap. Nutrients 2014, 6, 6076–6094. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Valencia, A.P.; Nagaraj, N.; Osman, D.H.; Rabinovitch, P.S.; Marcinek, D.J. Are fat and sugar just as detrimental in old age? Geroscience 2021, 43, 1615–1625. [Google Scholar] [CrossRef]
  178. Estruch, R.; Ros, E.; Salas-Salvadó, J.; Covas, M.-I.; Corella, D.; Arós, F.; Gómez-Gracia, E.; Ruiz-Gutiérrez, V.; Fiol, M.; Lapetra, J.; et al. Primary Prevention of Cardiovascular Disease with a Mediterranean Diet Supplemented with Extra-Virgin Olive Oil or Nuts. N. Engl. J. Med. 2018, 378, e34. [Google Scholar] [CrossRef]
  179. Salas-Salvadó, J.; Díaz-López, A.; Ruiz-Canela, M.; Basora, J.; Fitó, M.; Corella, D.; Serra-Majem, L.; Wärnberg, J.; Romaguera, D.; Estruch, R.; et al. Effect of a Lifestyle Intervention Program With Energy-Restricted Mediterranean Diet and Exercise on Weight Loss and Cardiovascular Risk Factors: One-Year Results of the PREDIMED-Plus Trial. Diabetes Care 2019, 42, 777–788. [Google Scholar] [CrossRef] [Green Version]
  180. Piepoli, M.F.; Hoes, A.W.; Agewall, S.; Albus, C.; Brotons, C.; Catapano, A.L.; Cooney, M.T.; Corrà, U.; Cosyns, B.; Deaton, C.; et al. 2016 European Guidelines on cardiovascular disease prevention in clinical practice: The Sixth Joint Task Force of the European Society of Cardiology and Other Societies on Cardiovascular Disease Prevention in Clinical Practice (constituted by representatives of 10 societies and by invited experts)Developed with the special contribution of the European Association for Cardiovascular Prevention & Rehabilitation (EACPR). Eur. Heart J. 2016, 37, 2315–2381. [Google Scholar] [CrossRef]
  181. National Clinical Guideline Centre (UK). National Clinical Guideline Centre (UK). National Institute for Health and Clinical Excellence: Guidance. In Lipid Modification: Cardiovascular Risk Assessment and the Modification of Blood Lipids for the Primary and Secondary Prevention of Cardiovascular Disease; National Institute for Health and Care Excellence (UK): London, UK, 2014. [Google Scholar]
  182. Rabar, S.; Harker, M.; O’Flynn, N.; Wierzbicki, A.S. Lipid modification and cardiovascular risk assessment for the primary and secondary prevention of cardiovascular disease: Summary of updated NICE guidance. BMJ Br. Med. J. 2014, 349, g4356. [Google Scholar] [CrossRef]
  183. Hooper, L.; Martin, N.; Jimoh, O.F.; Kirk, C.; Foster, E.; Abdelhamid, A.S. Reduction in saturated fat intake for cardiovascular disease. Cochrane Database Syst. Rev. 2020, 5, CD011737. [Google Scholar] [CrossRef]
  184. Djuricic, I.; Calder, P.C. Beneficial Outcomes of Omega-6 and Omega-3 Polyunsaturated Fatty Acids on Human Health: An Update for 2021. Nutrients 2021, 13, 2421. [Google Scholar] [CrossRef]
  185. Sacks, F.M.; Lichtenstein, A.H.; Wu, J.H.Y.; Appel, L.J.; Creager, M.A.; Kris-Etherton, P.M.; Miller, M.; Rimm, E.B.; Rudel, L.L.; Robinson, J.G.; et al. Dietary Fats and Cardiovascular Disease: A Presidential Advisory From the American Heart Association. Circulation 2017, 136, e1–e23. [Google Scholar] [CrossRef]
  186. Islam, M.A.; Amin, M.N.; Siddiqui, S.A.; Hossain, M.P.; Sultana, F.; Kabir, M.R. Trans fatty acids and lipid profile: A serious risk factor to cardiovascular disease, cancer and diabetes. Diabetes Metab. Syndr. 2019, 13, 1643–1647. [Google Scholar] [CrossRef] [PubMed]
  187. Borges, M.C.; Schmidt, A.F.; Jefferis, B.; Wannamethee, S.G.; Lawlor, D.A.; Kivimaki, M.; Kumari, M.; Gaunt, T.R.; Ben-Shlomo, Y.; Tillin, T.; et al. Circulating Fatty Acids and Risk of Coronary Heart Disease and Stroke: Individual Participant Data Meta-Analysis in Up to 16 126 Participants. J. Am. Heart Assoc. 2020, 9, e013131. [Google Scholar] [CrossRef] [PubMed]
  188. Currenti, W.; Godos, J.; Alanazi, A.M.; Grosso, G.; Cincione, R.I.; La Vignera, S.; Buscemi, S.; Galvano, F. Dietary Fats and Cardio-Metabolic Outcomes in a Cohort of Italian Adults. Nutrients 2022, 14, 4294. [Google Scholar] [CrossRef] [PubMed]
  189. Cai, D.; Zhao, Z.; Zhao, L.; Dong, Y.; Wang, L.; Zhao, S.; Li, Q. The Age-Accompanied and Diet-Associated Remodeling of the Phospholipid, Amino Acid, and SCFA Metabolism of Healthy Centenarians from a Chinese Longevous Region: A Window into Exceptional Longevity. Nutrients 2022, 14, 4420. [Google Scholar] [CrossRef] [PubMed]
  190. Grillo, A.; Salvi, L.; Coruzzi, P.; Salvi, P.; Parati, G. Sodium Intake and Hypertension. Nutrients 2019, 11, 1970. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  191. Hyseni, L.; Elliot-Green, A.; Lloyd-Williams, F.; Kypridemos, C.; O’Flaherty, M.; McGill, R.; Orton, L.; Bromley, H.; Cappuccio, F.P.; Capewell, S. Systematic review of dietary salt reduction policies: Evidence for an effectiveness hierarchy? PLoS ONE 2017, 12, e0177535. [Google Scholar] [CrossRef] [Green Version]
  192. McRae, M.P. Dietary Fiber Is Beneficial for the Prevention of Cardiovascular Disease: An Umbrella Review of Meta-analyses. J. Chiropr. Med. 2017, 16, 289–299. [Google Scholar] [CrossRef]
  193. Mayengbam, S.; Lambert, J.E.; Parnell, J.A.; Tunnicliffe, J.M.; Nicolucci, A.C.; Han, J.; Sturzenegger, T.; Shearer, J.; Mickiewicz, B.; Vogel, H.J.; et al. Impact of dietary fiber supplementation on modulating microbiota-host-metabolic axes in obesity. J. Nutr. Biochem. 2019, 64, 228–236. [Google Scholar] [CrossRef]
  194. Solah, V.A.; Kerr, D.A.; Hunt, W.J.; Johnson, S.K.; Boushey, C.J.; Delp, E.J.; Meng, X.; Gahler, R.J.; James, A.P.; Mukhtar, A.S.; et al. Erratum: Effect of Fibre Supplementation on Body Weight and Composition, Frequency of Eating and Dietary Choice in Overweight Individuals Nutrients 2017, 9, 149. Nutrients 2017, 9, 409. [Google Scholar] [CrossRef] [Green Version]
  195. Soliman, G.A. Dietary Fiber, Atherosclerosis, and Cardiovascular Disease. Nutrients 2019, 11, 1155. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Jovanovski, E.; Mazhar, N.; Komishon, A.; Khayyat, R.; Li, D.; Blanco Mejia, S.; Khan, T.; A, L.J.; Smircic-Duvnjak, L.; J, L.S.; et al. Can dietary viscous fiber affect body weight independently of an energy-restrictive diet? A systematic review and meta-analysis of randomized controlled trials. Am. J. Clin. Nutr. 2020, 111, 471–485. [Google Scholar] [CrossRef]
  197. Guo, N.; Zhu, Y.; Tian, D.; Zhao, Y.; Zhang, C.; Mu, C.; Han, C.; Zhu, R.; Liu, X. Role of diet in stroke incidence: An umbrella review of meta-analyses of prospective observational studies. BMC Med. 2022, 20, 194. [Google Scholar] [CrossRef] [PubMed]
  198. Cronin, P.; Joyce, S.A.; O’Toole, P.W.; O’Connor, E.M. Dietary Fibre Modulates the Gut Microbiota. Nutrients 2021, 13, 1655. [Google Scholar] [CrossRef] [PubMed]
  199. Bjorklund, G.; Shanaida, M.; Lysiuk, R.; Butnariu, M.; Peana, M.; Sarac, I.; Strus, O.; Smetanina, K.; Chirumbolo, S. Natural Compounds and Products from an Anti-Aging Perspective. Molecules 2022, 27, 7084. [Google Scholar] [CrossRef] [PubMed]
  200. Ma, L.; Hu, L.; Feng, X.; Wang, S. Nitrate and Nitrite in Health and Disease. Aging Dis. 2018, 9, 938–945. [Google Scholar] [CrossRef] [Green Version]
  201. Gilchrist, M.; Winyard, P.G.; Fulford, J.; Anning, C.; Shore, A.C.; Benjamin, N. Dietary nitrate supplementation improves reaction time in type 2 diabetes: Development and application of a novel nitrate-depleted beetroot juice placebo. Nitric Oxide 2014, 40, 67–74. [Google Scholar] [CrossRef]
  202. Muggeridge, D.J.; Howe, C.C.; Spendiff, O.; Pedlar, C.; James, P.E.; Easton, C. A single dose of beetroot juice enhances cycling performance in simulated altitude. Med. Sci. Sport. Exerc. 2014, 46, 143–150. [Google Scholar] [CrossRef]
  203. Stanaway, L.; Rutherfurd-Markwick, K.; Page, R.; Wong, M.; Jirangrat, W.; Teh, K.H.; Ali, A. Acute Supplementation with Nitrate-Rich Beetroot Juice Causes a Greater Increase in Plasma Nitrite and Reduction in Blood Pressure of Older Compared to Younger Adults. Nutrients 2019, 11, 1683. [Google Scholar] [CrossRef] [Green Version]
  204. Rose, P.; Moore, P.K.; Zhu, Y.Z. Garlic and Gaseous Mediators. Trends Pharm. Sci. 2018, 39, 624–634. [Google Scholar] [CrossRef]
  205. Morris, M.C.; Wang, Y.; Barnes, L.L.; Bennett, D.A.; Dawson-Hughes, B.; Booth, S.L. Nutrients and bioactives in green leafy vegetables and cognitive decline: Prospective study. Neurology 2018, 90, e214–e222. [Google Scholar] [CrossRef] [PubMed]
  206. Pollock, R.L. The effect of green leafy and cruciferous vegetable intake on the incidence of cardiovascular disease: A meta-analysis. JRSM Cardiovasc. Dis. 2016, 5, 2048004016661435. [Google Scholar] [CrossRef] [PubMed]
  207. Ladurner, A.; Schmitt, C.A.; Schachner, D.; Atanasov, A.G.; Werner, E.R.; Dirsch, V.M.; Heiss, E.H. Ascorbate stimulates endothelial nitric oxide synthase enzyme activity by rapid modulation of its phosphorylation status. Free Radic. Biol. Med. 2012, 52, 2082–2090. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  208. Lavillegrand, J.R.; Raia, L.; Urbina, T.; Hariri, G.; Gabarre, P.; Bonny, V.; Bige, N.; Baudel, J.L.; Bruneel, A.; Dupre, T.; et al. Vitamin C improves microvascular reactivity and peripheral tissue perfusion in septic shock patients. Crit. Care 2022, 26, 25. [Google Scholar] [CrossRef]
  209. Wang, D.; Ozen, C.; Abu-Reidah, I.M.; Chigurupati, S.; Patra, J.K.; Horbanczuk, J.O.; Jozwik, A.; Tzvetkov, N.T.; Uhrin, P.; Atanasov, A.G. Vasculoprotective Effects of Pomegranate (Punica granatum L.). Front. Pharm. 2018, 9, 544. [Google Scholar] [CrossRef] [Green Version]
  210. Asgary, S.; Keshvari, M.; Sahebkar, A.; Hashemi, M.; Rafieian-Kopaei, M. Clinical investigation of the acute effects of pomegranate juice on blood pressure and endothelial function in hypertensive individuals. ARYA Atheroscler 2013, 9, 326–331. [Google Scholar]
  211. Oh, M.K.; Yoo, J.I.; Byun, H.; Chun, S.W.; Lim, S.K.; Jang, Y.J.; Lee, C.H. Efficacy of Combined Antigravity Treadmill and Conventional Rehabilitation After Hip Fracture in Patients With Sarcopenia. J. Gerontol. A Biol. Sci. Med. Sci. 2020, 75, e173–e181. [Google Scholar] [CrossRef]
  212. Festa, J.; Da Boit, M.; Hussain, A.; Singh, H. Potential Benefits of Berry Anthocyanins on Vascular Function. Mol. Nutr. Food Res. 2021, 65, e2100170. [Google Scholar] [CrossRef]
  213. Snopek, L.; Mlcek, J.; Sochorova, L.; Baron, M.; Hlavacova, I.; Jurikova, T.; Kizek, R.; Sedlackova, E.; Sochor, J. Contribution of Red Wine Consumption to Human Health Protection. Molecules 2018, 23, 1684. [Google Scholar] [CrossRef] [Green Version]
  214. Khalaf, D.; Kruger, M.; Wehland, M.; Infanger, M.; Grimm, D. The Effects of Oral l-Arginine and l-Citrulline Supplementation on Blood Pressure. Nutrients 2019, 11, 1679. [Google Scholar] [CrossRef] [Green Version]
  215. Figueroa, A.; Wong, A.; Jaime, S.J.; Gonzales, J.U. Influence of L-citrulline and watermelon supplementation on vascular function and exercise performance. Curr. Opin. Clin. Nutr. Metab. Care 2017, 20, 92–98. [Google Scholar] [CrossRef] [PubMed]
  216. Godos, J.; Zappala, G.; Bernardini, S.; Giambini, I.; Bes-Rastrollo, M.; Martinez-Gonzalez, M. Adherence to the Mediterranean diet is inversely associated with metabolic syndrome occurrence: A meta-analysis of observational studies. Int. J. Food Sci. Nutr. 2017, 68, 138–148. [Google Scholar] [CrossRef] [PubMed]
  217. Mates, L.; Popa, D.S.; Rusu, M.E.; Fizesan, I.; Leucuta, D. Walnut Intake Interventions Targeting Biomarkers of Metabolic Syndrome and Inflammation in Middle-Aged and Older Adults: A Systematic Review and Meta-Analysis of Randomized Controlled Trials. Antioxidants 2022, 11, 1412. [Google Scholar] [CrossRef] [PubMed]
  218. Melzer, T.M.; Manosso, L.M.; Yau, S.Y.; Gil-Mohapel, J.; Brocardo, P.S. In Pursuit of Healthy Aging: Effects of Nutrition on Brain Function. Int. J. Mol. Sci. 2021, 22, 5026. [Google Scholar] [CrossRef]
  219. Wolters, F.J.; Ikram, M.A. Epidemiology of Dementia: The Burden on Society, the Challenges for Research. Methods Mol. Biol. 2018, 1750, 3–14. [Google Scholar] [CrossRef]
  220. Rolandi, E.; Zaccaria, D.; Vaccaro, R.; Abbondanza, S.; Pettinato, L.; Davin, A.; Guaita, A. Estimating the potential for dementia prevention through modifiable risk factors elimination in the real-world setting: A population-based study. Alzheimers Res. 2020, 12, 94. [Google Scholar] [CrossRef]
  221. Kivipelto, M.; Solomon, A.; Ahtiluoto, S.; Ngandu, T.; Lehtisalo, J.; Antikainen, R.; Bäckman, L.; Hänninen, T.; Jula, A.; Laatikainen, T.; et al. The Finnish Geriatric Intervention Study to Prevent Cognitive Impairment and Disability (FINGER): Study design and progress. Alzheimers Dement. 2013, 9, 657–665. [Google Scholar] [CrossRef]
  222. Smith, W.K.; Ingram, D.K.; de Cabo, R.; Pasquina, P. Metabolic pathways and therapeutics to promote resilience, rehabilitation and delayed aging. Geroscience 2021, 43, 1069–1070. [Google Scholar] [CrossRef]
  223. Yusufov, M.; Weyandt, L.L.; Piryatinsky, I. Alzheimer’s disease and diet: A systematic review. Int. J. Neurosci. 2017, 127, 161–175. [Google Scholar] [CrossRef]
  224. Lauer, A.A.; Grimm, H.S.; Apel, B.; Golobrodska, N.; Kruse, L.; Ratanski, E.; Schulten, N.; Schwarze, L.; Slawik, T.; Sperlich, S.; et al. Mechanistic Link between Vitamin B12 and Alzheimer’s Disease. Biomolecules 2022, 12, 129. [Google Scholar] [CrossRef]
  225. Andreu-Reinon, M.E.; Chirlaque, M.D.; Gavrila, D.; Amiano, P.; Mar, J.; Tainta, M.; Ardanaz, E.; Larumbe, R.; Colorado-Yohar, S.M.; Navarro-Mateu, F.; et al. Mediterranean Diet and Risk of Dementia and Alzheimer’s Disease in the EPIC-Spain Dementia Cohort Study. Nutrients 2021, 13, 700. [Google Scholar] [CrossRef] [PubMed]
  226. Dhana, K.; James, B.D.; Agarwal, P.; Aggarwal, N.T.; Cherian, L.J.; Leurgans, S.E.; Barnes, L.L.; Bennett, D.A.; Schneider, J.A. MIND Diet, Common Brain Pathologies, and Cognition in Community-Dwelling Older Adults. J. Alzheimers Dis. 2021, 83, 683–692. [Google Scholar] [CrossRef] [PubMed]
  227. Di Fiore, N. Diet May Help Prevent Alzheimer’s: MIND Diet Rich in Vegetables, Berries, Whole Grains, Nuts. Rush University Medical Center Website. Available online: www.rush.edu/news/diet-may-help-prevent-alzheimers (accessed on 1 December 2022).
  228. Morris, M.C.; Tangney, C.C.; Wang, Y.; Sacks, F.M.; Barnes, L.L.; Bennett, D.A.; Aggarwal, N.T. MIND diet slows cognitive decline with aging. Alzheimer’s Dement. 2015, 11, 1015–1022. [Google Scholar] [CrossRef] [PubMed]
  229. Takács, I.; Dank, M.; Majnik, J.; Nagy, G.; Szabó, A.; Szabó, B.; Szekanecz, Z.; Sziller, I.; Toldy, E.; Tislér, A.; et al. Magyarországi konszenzusajánlás a D-vitamin szerepéről a betegségek megelőzésében és kezelésében. Orv. Hetil. 2022, 163, 575–584. [Google Scholar] [CrossRef]
  230. Hirani, V.; Cumming, R.G.; Blyth, F.; Naganathan, V.; Le Couteur, D.G.; Waite, L.M.; Handelsman, D.J.; Seibel, M.J. Cross-sectional and longitudinal associations between the active vitamin D metabolite (1,25 dihydroxyvitamin D) and haemoglobin levels in older Australian men: The Concord Health and Ageing in Men Project. AGE 2015, 37, 8. [Google Scholar] [CrossRef] [Green Version]
  231. D’Cunha, N.M.; Georgousopoulou, E.N.; Boyd, L.; Veysey, M.; Sturm, J.; O’Brien, B.; Lucock, M.; McKune, A.J.; Mellor, D.D.; Roach, P.D.; et al. Relationship Between B-Vitamin Biomarkers and Dietary Intake with Apolipoprotein E є4 in Alzheimer’s Disease. J. Nutr. Gerontol. Geriatr. 2019, 38, 173–195. [Google Scholar] [CrossRef]
  232. Berry, B.J.; Kaeberlein, M. An energetics perspective on geroscience: Mitochondrial protonmotive force and aging. Geroscience 2021, 43, 1591–1604. [Google Scholar] [CrossRef]
  233. Kaliszewska, A.; Allison, J.; Martini, M.; Arias, N. Improving Age-Related Cognitive Decline through Dietary Interventions Targeting Mitochondrial Dysfunction. Int. J. Mol. Sci. 2021, 22, 3574. [Google Scholar] [CrossRef]
  234. Rodriguez-Lopez, S.; Lopez-Bellon, S.; Gonzalez-Reyes, J.A.; Buron, M.I.; de Cabo, R.; Villalba, J.M. Mitochondrial adaptations in liver and skeletal muscle to pro-longevity nutritional and genetic interventions: The crosstalk between calorie restriction and CYB5R3 overexpression in transgenic mice. Geroscience 2020, 42, 977–994. [Google Scholar] [CrossRef]
  235. Zampino, M.; Brennan, N.A.; Kuo, P.L.; Spencer, R.G.; Fishbein, K.W.; Simonsick, E.M.; Ferrucci, L. Poor mitochondrial health and systemic inflammation? Test of a classic hypothesis in the Baltimore Longitudinal Study of Aging. Geroscience 2020, 42, 1175–1182. [Google Scholar] [CrossRef]
  236. Whitson, J.A.; Martin-Perez, M.; Zhang, T.; Gaffrey, M.J.; Merrihew, G.E.; Huang, E.; White, C.C.; Kavanagh, T.J.; Qian, W.J.; Campbell, M.D.; et al. Elamipretide (SS-31) treatment attenuates age-associated post-translational modifications of heart proteins. Geroscience 2021, 43, 2395–2412. [Google Scholar] [CrossRef] [PubMed]
  237. Conte, M.; Sabbatinelli, J.; Chiariello, A.; Martucci, M.; Santoro, A.; Monti, D.; Arcaro, M.; Galimberti, D.; Scarpini, E.; Bonfigli, A.R.; et al. Disease-specific plasma levels of mitokines FGF21, GDF15, and Humanin in type II diabetes and Alzheimer’s disease in comparison with healthy aging. Geroscience 2021, 43, 985–1001. [Google Scholar] [CrossRef] [PubMed]
  238. Kim, S.J.; Miller, B.; Kumagai, H.; Silverstein, A.R.; Flores, M.; Yen, K. Mitochondrial-derived peptides in aging and age-related diseases. Geroscience 2021, 43, 1113–1121. [Google Scholar] [CrossRef] [PubMed]
  239. Lagerwaard, B.; Nieuwenhuizen, A.G.; de Boer, V.C.J.; Keijer, J. In vivo assessment of mitochondrial capacity using NIRS in locomotor muscles of young and elderly males with similar physical activity levels. Geroscience 2020, 42, 299–310. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  240. Herbst, A.; Prior, S.J.; Lee, C.C.; Aiken, J.M.; McKenzie, D.; Hoang, A.; Liu, N.; Chen, X.; Xun, P.; Allison, D.B.; et al. Skeletal muscle mitochondrial DNA copy number and mitochondrial DNA deletion mutation frequency as predictors of physical performance in older men and women. Geroscience 2021, 43, 1253–1264. [Google Scholar] [CrossRef]
  241. Varghese, N.; Werner, S.; Grimm, A.; Eckert, A. Dietary Mitophagy Enhancer: A Strategy for Healthy Brain Aging? Antioxidants 2020, 9, 32. [Google Scholar] [CrossRef]
  242. Kiss, T.; Nyul-Toth, A.; Balasubramanian, P.; Tarantini, S.; Ahire, C.; Yabluchanskiy, A.; Csipo, T.; Farkas, E.; Wren, J.D.; Garman, L.; et al. Nicotinamide mononucleotide (NMN) supplementation promotes neurovascular rejuvenation in aged mice: Transcriptional footprint of SIRT1 activation, mitochondrial protection, anti-inflammatory, and anti-apoptotic effects. Geroscience 2020, 42, 527–546. [Google Scholar] [CrossRef]
  243. Kiss, T.; Tarantini, S.; Csipo, T.; Balasubramanian, P.; Nyul-Toth, A.; Yabluchanskiy, A.; Wren, J.D.; Garman, L.; Huffman, D.M.; Csiszar, A.; et al. Circulating anti-geronic factors from heterochonic parabionts promote vascular rejuvenation in aged mice: Transcriptional footprint of mitochondrial protection, attenuation of oxidative stress, and rescue of endothelial function by young blood. Geroscience 2020, 42, 727–748. [Google Scholar] [CrossRef]
  244. Saberi, M.; Zhang, X.; Mobasheri, A. Targeting mitochondrial dysfunction with small molecules in intervertebral disc aging and degeneration. Geroscience 2021, 43, 517–537. [Google Scholar] [CrossRef]
  245. Londzin, P.; Zamora, M.; Kakol, B.; Taborek, A.; Folwarczna, J. Potential of Caffeine in Alzheimer’s Disease-A Review of Experimental Studies. Nutrients 2021, 13, 537. [Google Scholar] [CrossRef]
  246. Temple, J.L.; Bernard, C.; Lipshultz, S.E.; Czachor, J.D.; Westphal, J.A.; Mestre, M.A. The Safety of Ingested Caffeine: A Comprehensive Review. Front. Psychiatry 2017, 8, 80. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  247. Cederholm, T. Fish consumption and omega-3 fatty acid supplementation for prevention or treatment of cognitive decline, dementia or Alzheimer’s disease in older adults—Any news? Curr. Opin. Clin. Nutr. Metab. Care 2017, 20, 104–109. [Google Scholar] [CrossRef] [PubMed]
  248. Marti Del Moral, A.; Fortique, F. Omega-3 fatty acids and cognitive decline: A systematic review. Nutr. Hosp. 2019, 36, 939–949. [Google Scholar] [CrossRef] [PubMed]
  249. Canhada, S.; Castro, K.; Perry, I.S.; Luft, V.C. Omega-3 fatty acids’ supplementation in Alzheimer’s disease: A systematic review. Nutr. Neurosci. 2018, 21, 529–538. [Google Scholar] [CrossRef]
  250. Cole, G.M.; Ma, Q.L.; Frautschy, S.A. Omega-3 fatty acids and dementia. Prostaglandins Leukot. Essent. Fat. Acids 2009, 81, 213–221. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  251. Brunt, V.E.; LaRocca, T.J.; Bazzoni, A.E.; Sapinsley, Z.J.; Miyamoto-Ditmon, J.; Gioscia-Ryan, R.A.; Neilson, A.P.; Link, C.D.; Seals, D.R. The gut microbiome-derived metabolite trimethylamine N-oxide modulates neuroinflammation and cognitive function with aging. Geroscience 2021, 43, 377–394. [Google Scholar] [CrossRef]
  252. Adriansjach, J.; Baum, S.T.; Lefkowitz, E.J.; Van Der Pol, W.J.; Buford, T.W.; Colman, R.J. Age-Related Differences in the Gut Microbiome of Rhesus Macaques. J. Gerontol. A Biol. Sci. Med. Sci. 2020, 75, 1293–1298. [Google Scholar] [CrossRef]
  253. Ahmadi, S.; Razazan, A.; Nagpal, R.; Jain, S.; Wang, B.; Mishra, S.P.; Wang, S.; Justice, J.; Ding, J.; McClain, D.A.; et al. Metformin Reduces Aging-Related Leaky Gut and Improves Cognitive Function by Beneficially Modulating Gut Microbiome/Goblet Cell/Mucin Axis. J. Gerontol. A Biol. Sci. Med. Sci. 2020, 75, e9–e21. [Google Scholar] [CrossRef]
  254. Ruiz-Ruiz, S.; Sanchez-Carrillo, S.; Ciordia, S.; Mena, M.C.; Mendez-Garcia, C.; Rojo, D.; Bargiela, R.; Zubeldia-Varela, E.; Martinez-Martinez, M.; Barbas, C.; et al. Functional microbiome deficits associated with ageing: Chronological age threshold. Aging Cell 2020, 19, e13063. [Google Scholar] [CrossRef]
  255. Sun, Y.; Baptista, L.C.; Roberts, L.M.; Jumbo-Lucioni, P.; McMahon, L.L.; Buford, T.W.; Carter, C.S. The Gut Microbiome as a Therapeutic Target for Cognitive Impairment. J. Gerontol. A Biol. Sci. Med. Sci. 2020, 75, 1242–1250. [Google Scholar] [CrossRef]
  256. Castro-Mejia, J.L.; Khakimov, B.; Krych, L.; Bulow, J.; Bechshoft, R.L.; Hojfeldt, G.; Mertz, K.H.; Garne, E.S.; Schacht, S.R.; Ahmad, H.F.; et al. Physical fitness in community-dwelling older adults is linked to dietary intake, gut microbiota, and metabolomic signatures. Aging Cell 2020, 19, e13105. [Google Scholar] [CrossRef] [Green Version]
  257. Feher, J.; Elo, A.; Istvan, L.; Nagy, Z.Z.; Radak, Z.; Scuderi, G.; Artico, M.; Kovacs, I. Microbiota mitochondria disorders as hubs for early age-related macular degeneration. Geroscience 2022, 44, 2623–2653. [Google Scholar] [CrossRef] [PubMed]
  258. Lee, H.K.; Kirkwood, A. Mechanisms of Homeostatic Synaptic Plasticity in vivo. Front. Cell. Neurosci. 2019, 13, 520. [Google Scholar] [CrossRef] [PubMed]
  259. Sinha, A.; Metzner, C.; Davey, N.; Adams, R.; Schmuker, M.; Steuber, V. Growth rules for the repair of Asynchronous Irregular neuronal networks after peripheral lesions. PLoS Comput. Biol. 2021, 17, e1008996. [Google Scholar] [CrossRef] [PubMed]
  260. Petzold, G.C.; Murthy, V.N. Role of astrocytes in neurovascular coupling. Neuron 2011, 71, 782–797. [Google Scholar] [CrossRef] [Green Version]
  261. Stobart, J.L.; Lu, L.; Anderson, H.D.; Mori, H.; Anderson, C.M. Astrocyte-induced cortical vasodilation is mediated by D-serine and endothelial nitric oxide synthase. Proc. Natl. Acad. Sci. USA 2013, 110, 3149–3154. [Google Scholar] [CrossRef] [Green Version]
  262. Lipecz, A.; Csipo, T.; Tarantini, S.; Hand, R.A.; Ngo, B.N.; Conley, S.; Nemeth, G.; Tsorbatzoglou, A.; Courtney, D.L.; Yabluchanska, V.; et al. Age-related impairment of neurovascular coupling responses: A dynamic vessel analysis (DVA)-based approach to measure decreased flicker light stimulus-induced retinal arteriolar dilation in healthy older adults. Geroscience 2019, 41, 341–349. [Google Scholar] [CrossRef]
  263. Tarantini, S.; Tran, C.H.T.; Gordon, G.R.; Ungvari, Z.; Csiszar, A. Impaired neurovascular coupling in aging and Alzheimer’s disease: Contribution of astrocyte dysfunction and endothelial impairment to cognitive decline. Exp. Gerontol. 2017, 94, 52–58. [Google Scholar] [CrossRef]
  264. Tarantini, S.; Hertelendy, P.; Tucsek, Z.; Valcarcel-Ares, M.N.; Smith, N.; Menyhart, A.; Farkas, E.; Hodges, E.L.; Towner, R.; Deak, F.; et al. Pharmacologically-induced neurovascular uncoupling is associated with cognitive impairment in mice. J. Cereb. Blood Flow Metab. 2015, 35, 1871–1881. [Google Scholar] [CrossRef] [Green Version]
  265. Balbi, M.; Ghosh, M.; Longden, T.A.; Jativa Vega, M.; Gesierich, B.; Hellal, F.; Lourbopoulos, A.; Nelson, M.T.; Plesnila, N. Dysfunction of mouse cerebral arteries during early aging. J. Cereb. Blood Flow Metab. 2015, 35, 1445–1453. [Google Scholar] [CrossRef]
  266. Fabiani, M.; Gordon, B.A.; Maclin, E.L.; Pearson, M.A.; Brumback-Peltz, C.R.; Low, K.A.; McAuley, E.; Sutton, B.P.; Kramer, A.F.; Gratton, G. Neurovascular coupling in normal aging: A combined optical, ERP and fMRI study. Neuroimage 2014, 85 Pt 1, 592–607. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  267. Barisano, G.; Montagne, A.; Kisler, K.; Schneider, J.A.; Wardlaw, J.M.; Zlokovic, B.V. Blood-brain barrier link to human cognitive impairment and Alzheimer’s Disease. Nat. Cardiovasc Res. 2022, 1, 108–115. [Google Scholar] [CrossRef] [PubMed]
  268. Toth, P.; Tarantini, S.; Tucsek, Z.; Ashpole, N.M.; Sosnowska, D.; Gautam, T.; Ballabh, P.; Koller, A.; Sonntag, W.E.; Csiszar, A.; et al. Resveratrol treatment rescues neurovascular coupling in aged mice:role of improved cerebromicrovascular endothelial function and down-regulation of NADPH oxidas. Am. J. Physiol. Heart Circ. Physiol. 2014, 306, H299–H308. [Google Scholar] [CrossRef] [PubMed]
  269. Gete, Y.G.; Koblan, L.W.; Mao, X.; Trappio, M.; Mahadik, B.; Fisher, J.P.; Liu, D.R.; Cao, K. Mechanisms of angiogenic incompetence in Hutchinson-Gilford progeria via downregulation of endothelial NOS. Aging Cell 2021, 20, e13388. [Google Scholar] [CrossRef] [PubMed]
  270. Rodriguez-Manas, L.; El-Assar, M.; Vallejo, S.; Lopez-Doriga, P.; Solis, J.; Petidier, R.; Montes, M.; Nevado, J.; Castro, M.; Gomez-Guerrero, C.; et al. Endothelial dysfunction in aged humans is related with oxidative stress and vascular inflammation. Aging Cell 2009, 8, 226–238. [Google Scholar] [CrossRef]
  271. Sweeney, M.D.; Sagare, A.P.; Zlokovic, B.V. Blood-brain barrier breakdown in Alzheimer disease and other neurodegenerative disorders. Nat. Rev. Neurol. 2018, 14, 133–150. [Google Scholar] [CrossRef]
  272. Sweeney, M.D.; Zhao, Z.; Montagne, A.; Nelson, A.R.; Zlokovic, B.V. Blood-Brain Barrier: From Physiology to Disease and Back. Physiol. Rev. 2019, 99, 21–78. [Google Scholar] [CrossRef]
  273. Graham, L.C.; Grabowska, W.A.; Chun, Y.; Risacher, S.L.; Philip, V.M.; Saykin, A.J.; Alzheimer’s Disease Neuroimaging, I.; Sukoff Rizzo, S.J.; Howell, G.R. Exercise prevents obesity-induced cognitive decline and white matter damage in mice. Neurobiol. Aging 2019, 80, 154–172. [Google Scholar] [CrossRef]
  274. Bernier, M.; Wahl, D.; Ali, A.; Allard, J.; Faulkner, S.; Wnorowski, A.; Sanghvi, M.; Moaddel, R.; Alfaras, I.; Mattison, J.A.; et al. Resveratrol supplementation confers neuroprotection in cortical brain tissue of nonhuman primates fed a high-fat/sucrose diet. Aging (Albany NY) 2016, 8, 899–916. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  275. Csipo, T.; Fulop, G.A.; Lipecz, A.; Tarantini, S.; Kiss, T.; Balasubramanian, P.; Csiszar, A.; Ungvari, Z.; Yabluchanskiy, A. Short-term weight loss reverses obesity-induced microvascular endothelial dysfunction. Geroscience 2018, 40, 337–346. [Google Scholar] [CrossRef]
  276. Jimoh, A.; Tanko, Y.; Ahmed, A.; Mohammed, A.; Ayo, J.O. Resveratrol prevents high-fat diet-induced obesity and oxidative stress in rabbits. Pathophysiology 2018, 25, 359–364. [Google Scholar] [CrossRef] [PubMed]
  277. Lesniewski, L.A.; Seals, D.R.; Walker, A.E.; Henson, G.D.; Blimline, M.W.; Trott, D.W.; Bosshardt, G.C.; LaRocca, T.J.; Lawson, B.R.; Zigler, M.C.; et al. Dietary rapamycin supplementation reverses age-related vascular dysfunction and oxidative stress, while modulating nutrient-sensing, cell cycle, and senescence pathways. Aging Cell 2017, 16, 17–26. [Google Scholar] [CrossRef] [PubMed]
  278. Gardner, A.W.; Montgomery, P.S.; Wang, M.; Shen, B.; Casanegra, A.I.; Silva-Palacios, F.; Ungvari, Z.; Yabluchanskiy, A.; Csiszar, A.; Waldstein, S.R. Cognitive decrement in older adults with symptomatic peripheral artery disease. Geroscience 2021, 43, 2455–2465. [Google Scholar] [CrossRef] [PubMed]
  279. Kerkhofs, D.; Wong, S.M.; Zhang, E.; Uiterwijk, R.; Hoff, E.I.; Jansen, J.F.A.; Staals, J.; Backes, W.H.; van Oostenbrugge, R.J. Blood-brain barrier leakage at baseline and cognitive decline in cerebral small vessel disease: A 2-year follow-up study. Geroscience 2021, 43, 1643–1652. [Google Scholar] [CrossRef]
  280. Labinskyy, N.; Csiszar, A.; Veress, G.; Stef, G.; Pacher, P.; Oroszi, G.; Wu, J.; Ungvari, Z. Vascular dysfunction in aging: Potential effects of resveratrol, an anti-inflammatory phytoestrogen. Curr. Med. Chem. 2006, 13, 989–996. [Google Scholar] [CrossRef] [PubMed]
  281. Gratton, G.; Weaver, S.R.; Burley, C.V.; Low, K.A.; Maclin, E.L.; Johns, P.W.; Pham, Q.S.; Lucas, S.J.E.; Fabiani, M.; Rendeiro, C. Dietary flavanols improve cerebral cortical oxygenation and cognition in healthy adults. Sci. Rep. 2020, 10, 19409. [Google Scholar] [CrossRef]
  282. Sorond, F.A.; Hurwitz, S.; Salat, D.H.; Greve, D.N.; Fisher, N.D. Neurovascular coupling, cerebral white matter integrity, and response to cocoa in older people. Neurology 2013, 81, 904–909. [Google Scholar] [CrossRef]
  283. Gomes, A.P.; Price, N.L.; Ling, A.J.; Moslehi, J.J.; Montgomery, M.K.; Rajman, L.; White, J.P.; Teodoro, J.S.; Wrann, C.D.; Hubbard, B.P.; et al. Declining NAD(+) induces a pseudohypoxic state disrupting nuclear-mitochondrial communication during aging. Cell 2013, 155, 1624–1638. [Google Scholar] [CrossRef] [Green Version]
  284. Massudi, H.; Grant, R.; Braidy, N.; Guest, J.; Farnsworth, B.; Guillemin, G.J. Age-associated changes in oxidative stress and NAD+ metabolism in human tissue. PLoS ONE 2012, 7, e42357. [Google Scholar] [CrossRef]
  285. Csiszar, A.; Labinskyy, N.; Podlutsky, A.; Kaminski, P.M.; Wolin, M.S.; Zhang, C.; Mukhopadhyay, P.; Pacher, P.; Hu, F.; de Cabo, R.; et al. Vasoprotective effects of resveratrol and SIRT1: Attenuation of cigarette smoke-induced oxidative stress and proinflammatory phenotypic alterations. Am. J. Physiol. Heart Circ. Physiol. 2008, 294, H2721–H2735. [Google Scholar] [CrossRef] [Green Version]
  286. Csiszar, A.; Labinskyy, N.; Jimenez, R.; Pinto, J.T.; Ballabh, P.; Losonczy, G.; Pearson, K.J.; de Cabo, R.; Ungvari, Z. Anti-oxidative and anti-inflammatory vasoprotective effects of caloric restriction in aging: Role of circulating factors and SIRT1. Mech. Ageing Dev. 2009, 130, 518–527. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  287. Martens, C.R.; Denman, B.A.; Mazzo, M.R.; Armstrong, M.L.; Reisdorph, N.; McQueen, M.B.; Chonchol, M.; Seals, D.R. Chronic nicotinamide riboside supplementation is well-tolerated and elevates NAD(+) in healthy middle-aged and older adults. Nat. Commun. 2018, 9, 1286. [Google Scholar] [CrossRef] [PubMed]
  288. Ito, T.K.; Sato, T.; Takanashi, Y.; Tamannaa, Z.; Kitamoto, T.; Odagiri, K.; Setou, M. A single oral supplementation of nicotinamide within the daily tolerable upper level increases blood NAD+ levels in healthy subjects. Transl. Med. Aging 2021, 5, 43–51. [Google Scholar] [CrossRef]
  289. Lai, J.S.; Hiles, S.; Bisquera, A.; Hure, A.J.; McEvoy, M.; Attia, J. A systematic review and meta-analysis of dietary patterns and depression in community-dwelling adults. Am. J. Clin. Nutr. 2014, 99, 181–197. [Google Scholar] [CrossRef] [Green Version]
  290. Lassale, C.; Batty, G.D.; Baghdadli, A.; Jacka, F.; Sanchez-Villegas, A.; Kivimaki, M.; Akbaraly, T. Healthy dietary indices and risk of depressive outcomes: A systematic review and meta-analysis of observational studies. Mol. Psychiatry 2019, 24, 965–986. [Google Scholar] [CrossRef] [Green Version]
  291. McGrory, S.; Cameron, J.R.; Pellegrini, E.; Warren, C.; Doubal, F.N.; Deary, I.J.; Dhillon, B.; Wardlaw, J.M.; Trucco, E.; MacGillivray, T.J. The application of retinal fundus camera imaging in dementia: A systematic review. Alzheimers Dement. (Amst.) 2017, 6, 91–107. [Google Scholar] [CrossRef] [Green Version]
  292. Lopez-Taboada, I.; Gonzalez-Pardo, H.; Conejo, N.M. Western Diet: Implications for Brain Function and Behavior. Front. Psychol. 2020, 11, 564413. [Google Scholar] [CrossRef]
  293. Cherian, L.; Wang, Y.; Holland, T.; Agarwal, P.; Aggarwal, N.; Morris, M.C. DASH and Mediterranean-Dash Intervention for Neurodegenerative Delay (MIND) Diets Are Associated With Fewer Depressive Symptoms Over Time. J. Gerontol. A Biol. Sci. Med. Sci. 2021, 76, 151–156. [Google Scholar] [CrossRef]
  294. Chopp-Hurley, J.N.; Wiebenga, E.G.; Keller, H.H.; Maly, M.R. Diet and Nutrition Risk Affect Mobility and General Health in Osteoarthritis: Data from the Canadian Longitudinal Study on Aging. J. Gerontol. A Biol. Sci. Med. Sci. 2020, 75, 2147–2155. [Google Scholar] [CrossRef]
  295. Das, A.; Cumming, R.G.; Naganathan, V.; Blyth, F.; Ribeiro, R.V.; Le Couteur, D.G.; Handelsman, D.J.; Waite, L.M.; Simpson, S.J.; Hirani, V. Prospective Associations Between Dietary Antioxidant Intake and Frailty in Older Australian Men: The Concord Health and Ageing in Men Project. J. Gerontol. A Biol. Sci. Med. Sci. 2020, 75, 348–356. [Google Scholar] [CrossRef]
  296. Duregon, E.; Bernier, M.; de Cabo, R. A Glance Back at the Journal of Gerontology-Coffee, Dietary Interventions and Life Span. J. Gerontol. A Biol. Sci. Med. Sci. 2020, 75, 2029–2030. [Google Scholar] [CrossRef] [PubMed]
  297. Huang, G.; Pencina, K.; Li, Z.; Apovian, C.M.; Travison, T.G.; Storer, T.W.; Gagliano-Juca, T.; Basaria, S.; Bhasin, S. Effect of Protein Intake on Visceral Abdominal Fat and Metabolic Biomarkers in Older Men With Functional Limitations: Results From a Randomized Clinical Trial. J. Gerontol. A Biol. Sci. Med. Sci. 2021, 76, 1084–1089. [Google Scholar] [CrossRef] [PubMed]
  298. Sheng, L.T.; Jiang, Y.W.; Pan, X.F.; Feng, L.; Yuan, J.M.; Pan, A.; Koh, W.P. Association Between Dietary Intakes of B Vitamins in Midlife and Cognitive Impairment in Late-Life: The Singapore Chinese Health Study. J. Gerontol. A Biol. Sci. Med. Sci. 2020, 75, 1222–1227. [Google Scholar] [CrossRef]
  299. Jacka, F.N.; O’Neil, A.; Opie, R.; Itsiopoulos, C.; Cotton, S.; Mohebbi, M.; Castle, D.; Dash, S.; Mihalopoulos, C.; Chatterton, M.L.; et al. A randomised controlled trial of dietary improvement for adults with major depression (the ‘SMILES’ trial). BMC Med. 2017, 15, 23. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  300. Parletta, N.; Zarnowiecki, D.; Cho, J.; Wilson, A.; Bogomolova, S.; Villani, A.; Itsiopoulos, C.; Niyonsenga, T.; Blunden, S.; Meyer, B.; et al. A Mediterranean-style dietary intervention supplemented with fish oil improves diet quality and mental health in people with depression: A randomized controlled trial (HELFIMED). Nutr. Neurosci. 2019, 22, 474–487. [Google Scholar] [CrossRef] [Green Version]
  301. Bishwajit, G.; O’Leary, D.P.; Ghosh, S.; Sanni, Y.; Shangfeng, T.; Zhanchun, F. Association between depression and fruit and vegetable consumption among adults in South Asia. BMC Psychiatry 2017, 17, 15. [Google Scholar] [CrossRef] [Green Version]
  302. Glabska, D.; Guzek, D.; Groele, B.; Gutkowska, K. Fruit and Vegetable Intake and Mental Health in Adults: A Systematic Review. Nutrients 2020, 12, 115. [Google Scholar] [CrossRef] [Green Version]
  303. Mujcic, R.; Oswald, A.J. Evolution of Well-Being and Happiness After Increases in Consumption of Fruit and Vegetables. Am. J. Public Health 2016, 106, 1504–1510. [Google Scholar] [CrossRef]
  304. Brookie, K.L.; Best, G.I.; Conner, T.S. Intake of Raw Fruits and Vegetables Is Associated With Better Mental Health Than Intake of Processed Fruits and Vegetables. Front. Psychol. 2018, 9, 487. [Google Scholar] [CrossRef]
  305. Opie, R.S.; Itsiopoulos, C.; Parletta, N.; Sanchez-Villegas, A.; Akbaraly, T.N.; Ruusunen, A.; Jacka, F.N. Dietary recommendations for the prevention of depression. Nutr. Neurosci. 2017, 20, 161–171. [Google Scholar] [CrossRef]
  306. McNamara, R.K.; Asch, R.H.; Lindquist, D.M.; Krikorian, R. Role of polyunsaturated fatty acids in human brain structure and function across the lifespan: An update on neuroimaging findings. Prostaglandins Leukot Essent Fat. Acids 2018, 136, 23–34. [Google Scholar] [CrossRef] [PubMed]
  307. Mocking, R.J.; Harmsen, I.; Assies, J.; Koeter, M.W.; Ruhé, H.G.; Schene, A.H. Meta-analysis and meta-regression of omega-3 polyunsaturated fatty acid supplementation for major depressive disorder. Transl. Psychiatry 2016, 6, e756. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  308. Park, S.J.; Lee, D.K.; Kim, B.; Na, K.S.; Lee, C.H.; Son, Y.D.; Lee, H.J. The Association between Omega-3 Fatty Acid Intake and Human Brain Connectivity in Middle-Aged Depressed Women. Nutrients 2020, 12, 2191. [Google Scholar] [CrossRef] [PubMed]
  309. Liao, Y.; Xie, B.; Zhang, H.; He, Q.; Guo, L.; Subramanieapillai, M.; Fan, B.; Lu, C.; McIntyre, R.S. Efficacy of omega-3 PUFAs in depression: A meta-analysis. Transl. Psychiatry 2019, 9, 190. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  310. Fekete, M.; Szőllősi, G.; Németh, A.N.; Varga, J.T. Clinical value of omega-3 polyunsaturated fatty acid supplementation in chronic obstructive pulmonary disease. Orv. Hetil. 2021, 162, 23–30. [Google Scholar] [CrossRef] [PubMed]
  311. Baltrusch, S. The Role of Neurotropic B Vitamins in Nerve Regeneration. Biomed Res. Int. 2021, 2021, 9968228. [Google Scholar] [CrossRef]
  312. Parker, G.B.; Brotchie, H.; Graham, R.K. Vitamin D and depression. J. Affect Disord 2017, 208, 56–61. [Google Scholar] [CrossRef]
  313. Menon, V.; Kar, S.K.; Suthar, N.; Nebhinani, N. Vitamin D and Depression: A Critical Appraisal of the Evidence and Future Directions. Indian J. Psychol. Med. 2020, 42, 11–21. [Google Scholar] [CrossRef]
Table 1. Summary of the health benefits of dietary fiber.
Table 1. Summary of the health benefits of dietary fiber.
EffectHealth Benefit
MetabolicImproved insulin sensitivity, reduced risk of developing T2D, improved glycaemic status and lipid profiles, reduced body weight and abdominal adiposity
Gut microfloraGut microbial viability and diversity, metabolites from gut microflora
CardiovascularChronic inflammation, cardiovascular risk, mortality
DepressionChronic inflammation, gut microbiota
Gastrointestinal localizedColonic health and integrity, colonic motility, colorectal carcinoma
Source: ref [122].
Table 2. Pathological consequences of obesity.
Table 2. Pathological consequences of obesity.
I. Metabolic consequencesDiabetes
Insulin resistance
Gout
Persistent inflammation
Metabolic
syndrome
Dyslipidemia
Hyperuricemia
II. Cardiovascular diseasesHypertension
Coronary heart disease
Venous thromboembolism
Stroke
Congestive heart failure
III. Respiratory
diseases
Asthma
Sleep apnea syndrome (OSAS)
Hypoxia
Hypoventilation syndrome
IV. TumorsEsophageal-, intestinal-, rectal-,
liver-, gall bladder-, pancreas-, kidney tumors
Leukemia, lymphoma, multiple myeloma
In women: endometrial, cervical, ovarian, breast cancer
Men: prostate cancer
VI. GastrointestinalGallbladder diseases
Non-alcoholic fatty liver disease
Gastroesophageal reflux
Abdominal and inguinal hernia
VII. Genitourinary system
and reproductive
organs
Urine loss
Irregular menstruation
Hirsutism
Hypertonia
Miscarriage
Esophageal abnormalities
Birth defects
Gestational diabetes
Infertility
Polycystic ovary
Pre-eclampsia
Large fetus
Fetal distress
Caesarean section
VIII. Psychological and
social disorders
Low self-esteem
Stigmatization
Anxiety, depression
Work and employment
problems
IX. Other pathologiesNephrosis syndrome
Complications of anesthesia
Idiopathic intracranial hypertension
Lymphoedema
Proteinuria
Periodontal diseases
Skin infections
Source: ref [175].
Table 3. Cardioprotective factors in nutrition based on the 2016 ESC guideline.
Table 3. Cardioprotective factors in nutrition based on the 2016 ESC guideline.
Saturated fatty acid <10 E%, achieved by replacing the excess with polyunsaturated fatty acid in the diet.
The amount of trans fatty acid should be reduced as much as possible by limiting the consumption of processed products and keeping the intake of natural trans fatty acid below <1 E%.
<5 g/day salt intake.
30–45 g/day dietary fiber, preferably whole grains.
≥200 g fruit (2–3 portions/day)
≥200 g vegetables (2–3 portions/day)
Fish 1–2 times/week, of which one should be fatty fish (high in fat).
30 g/day unsalted oilseeds.
Limit on alcoholic drinks: 2 glasses/day (20 g/day of alcohol) for men, 1 glass/day (10 g alcohol/day) for women.
Avoid sugary and alcoholic drinks.
Source: ref [180].
Table 4. Comparison of different recommendations to reduce cardiovascular risk.
Table 4. Comparison of different recommendations to reduce cardiovascular risk.
2016 European
Guideline
2014 NICE
Guideline
2020 NICE
Pathway
Vegetables≥2 dose/day2–3 dose/day2–3 dose/day
Fruit≥2 dose/day2–3 dose/day2–3 dose/day
Fish≥2 dose/week≥2 dose/day≥2 dose/day
Fat (saturated)<10 E%<30 E%30–35 E%
Added fats olive/rape and products made from these fats and oilsolive/rape and products made from these fats and oils
Fiber (g/day)30–45
Sodium (mg/day)<2500
Oil seeds 4–5 dose/week4–5 dose/week
Pulses vegetables 4–5 dose/week4–5 dose/week
Source: ref [180,181].
Table 5. Risk and protective factors for dementia.
Table 5. Risk and protective factors for dementia.
Risk FactorsProtective Factors
▪ Age▪ Genetic factors: some mutations in the Amyloid Precursor Protein gene, APOE ε2 allele
▪ Genetic factors: familial predisposition, APOE ε4 allele, other genes▪ Lifestyle factors: education, intellectual work, extensive social contacts, mental stimulation, physical activity
▪ Vascular and metabolic factors: arteriosclerosis, stroke, diabetes mellitus, hypertension, obesity, high cholesterol in middle age▪ Diet: low alcohol, Mediterranean diet, unsaturated fats, oils, vitamin B12, folic acid, vitamin D
▪ Lifestyle factors: smoking, inactivity, heavy alcohol consumption▪ Medications: antihypertensives, statins, hormone replacement therapy, NSAIDs
▪ Diet: saturated fats, hyperhomocysteinemia, vitamin deficiency
▪ Other factors: depression, trauma, toxic effects, infectious diseases
Table 6. The MIND (Mediterranean-DASH Diet Intervention for Neurodegenerative Delay) diet focuses on healthy foods.
Table 6. The MIND (Mediterranean-DASH Diet Intervention for Neurodegenerative Delay) diet focuses on healthy foods.
Include TheseLimit These
▪ Green leafy vegetables: every day▪ Red meats
▪ Other vegetables: at least once per day▪ Butter and stick margarine: less than 1 tablespoon per day
▪ Nuts: every day▪ Cheese: less than one serving per week
▪ Berries: at least twice per week▪ Pastries and sweets: limit
▪ Beans: every other day▪ Fried or fast food: less than one serving per week
▪ Whole grains: three times per day
▪ Fish: at least once per week
▪ Poultry: at least twice per week
▪ Olive oil
▪ Wine: one glass per day
Source: ref [228].
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Fekete, M.; Szarvas, Z.; Fazekas-Pongor, V.; Feher, A.; Csipo, T.; Forrai, J.; Dosa, N.; Peterfi, A.; Lehoczki, A.; Tarantini, S.; et al. Nutrition Strategies Promoting Healthy Aging: From Improvement of Cardiovascular and Brain Health to Prevention of Age-Associated Diseases. Nutrients 2023, 15, 47. https://doi.org/10.3390/nu15010047

AMA Style

Fekete M, Szarvas Z, Fazekas-Pongor V, Feher A, Csipo T, Forrai J, Dosa N, Peterfi A, Lehoczki A, Tarantini S, et al. Nutrition Strategies Promoting Healthy Aging: From Improvement of Cardiovascular and Brain Health to Prevention of Age-Associated Diseases. Nutrients. 2023; 15(1):47. https://doi.org/10.3390/nu15010047

Chicago/Turabian Style

Fekete, Monika, Zsofia Szarvas, Vince Fazekas-Pongor, Agnes Feher, Tamas Csipo, Judit Forrai, Norbert Dosa, Anna Peterfi, Andrea Lehoczki, Stefano Tarantini, and et al. 2023. "Nutrition Strategies Promoting Healthy Aging: From Improvement of Cardiovascular and Brain Health to Prevention of Age-Associated Diseases" Nutrients 15, no. 1: 47. https://doi.org/10.3390/nu15010047

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop