Next Article in Journal
Association between Dietary Acid Load and Hypertension in Chinese Adults: Analysis of the China Health and Nutrition Survey (2009)
Next Article in Special Issue
Association between Vitamin D and Short-Term Functional Outcomes in Acute Ischemic Stroke
Previous Article in Journal
The Effect of Probiotic Supplements on Metabolic Parameters of People with Type 2 Diabetes in Greece—A Randomized, Double-Blind, Placebo-Controlled Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Role of Methionine-Rich Diet in Unhealthy Cerebrovascular and Brain Aging: Mechanisms and Implications for Cognitive Impairment

by
Anna Ungvari
1,*,†,
Rafal Gulej
2,3,†,
Boglarka Csik
2,3,4,
Peter Mukli
2,3,4,
Sharon Negri
2,3,
Stefano Tarantini
2,3,4,5,6,
Andriy Yabluchanskiy
2,3,4,5,6,
Zoltan Benyo
7,8,
Anna Csiszar
2,3,5,6,9 and
Zoltan Ungvari
2,3,4,5,6
1
Department of Public Health, Semmelweis University, 1089 Budapest, Hungary
2
Vascular Cognitive Impairment, Neurodegeneration and Healthy Brain Aging Program, Department of Neurosurgery, University of Oklahoma Health Sciences Center, Oklahoma City, OK 73104, USA
3
Oklahoma Center for Geroscience and Healthy Brain Aging, University of Oklahoma Health Sciences Center, Oklahoma City, OK 73104, USA
4
International Training Program in Geroscience, Department of Public Health, Doctoral School of Basic and Translational Medicine, Semmelweis University, 1089 Budapest, Hungary
5
Stephenson Cancer Center, University of Oklahoma, Oklahoma City, OK 73104, USA
6
Department of Health Promotion Sciences, College of Public Health, University of Oklahoma Health Sciences Center, Oklahoma City, OK 73104, USA
7
Institute of Translational Medicine, Semmelweis University, 1094 Budapest, Hungary
8
Cerebrovascular and Neurocognitive Disorders Research Group, Eötvös Loránd Research Network, Semmelweis University, 1094 Budapest, Hungary
9
International Training Program in Geroscience, Department of Translational Medicine, Doctoral School of Basic and Translational Medicine, Semmelweis University, 1089 Budapest, Hungary
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Nutrients 2023, 15(21), 4662; https://doi.org/10.3390/nu15214662
Submission received: 6 October 2023 / Revised: 28 October 2023 / Accepted: 31 October 2023 / Published: 3 November 2023

Abstract

:
As aging societies in the western world face a growing prevalence of vascular cognitive impairment and Alzheimer’s disease (AD), understanding their underlying causes and associated risk factors becomes increasingly critical. A salient concern in the western dietary context is the high consumption of methionine-rich foods such as red meat. The present review delves into the impact of this methionine-heavy diet and the resultant hyperhomocysteinemia on accelerated cerebrovascular and brain aging, emphasizing their potential roles in cognitive impairment. Through a comprehensive exploration of existing evidence, a link between high methionine intake and hyperhomocysteinemia and oxidative stress, mitochondrial dysfunction, inflammation, and accelerated epigenetic aging is drawn. Moreover, the microvascular determinants of cognitive deterioration, including endothelial dysfunction, reduced cerebral blood flow, microvascular rarefaction, impaired neurovascular coupling, and blood–brain barrier (BBB) disruption, are explored. The mechanisms by which excessive methionine consumption and hyperhomocysteinemia might drive cerebromicrovascular and brain aging processes are elucidated. By presenting an intricate understanding of the relationships among methionine-rich diets, hyperhomocysteinemia, cerebrovascular and brain aging, and cognitive impairment, avenues for future research and potential therapeutic interventions are suggested.

Graphical Abstract

1. Introduction

The Western world is undergoing a profound demographic shift, with an increasingly aging population [1,2,3]. This trend presents significant challenges, particularly regarding age-related cognitive decline [4,5]. Cognitive impairment and dementia have emerged as critical public health concerns, as they not only affect the well-being of individuals but also strain healthcare systems and social support networks [4,6,7,8,9]. With the prevalence of cognitive decline on the rise, gaining a comprehensive understanding of the underlying mechanisms contributing to this phenomenon and the lifestyle factors that influence it becomes of paramount importance.
Epidemiological studies have shed light on the prevalence and impact of cognitive decline and dementia across the aging populations of the developed world [10,11,12,13]. Alzheimer’s disease and other neurodegenerative disorders have received considerable attention due to their devastating effects on memory, cognition, and daily functioning [14,15]. However, emerging evidence suggests that vascular cognitive impairment and dementia (VCID) may be among the most significant contributors to age-related cognitive decline [16,17,18]. VCID is associated with cerebrovascular pathologies and functional impairment of the cerebral microcirculation, such as small vessel disease, dysregulation of cerebral blood flow (CBF), blood–brain barrier (BBB) disruption, cerebral microhemorrhages, and cerebral infarcts, which contribute to cognitive impairments independent of neurodegenerative changes [19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34].
Significant progress has been made in elucidating the cellular and molecular mechanisms underlying cerebrovascular and brain aging, as well as age-related pathologies affecting the central nervous system [28,29,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60]. Accumulating evidence suggests that dietary factors play a crucial role in modulating the aging process [61,62,63,64,65] and the development of age-related diseases [66,67,68,69,70,71,72,73,74]. Importantly, dietary factors have also emerged as key determinants in shaping the trajectories of age-related cognitive decline and cerebrovascular and brain health [75]. Unhealthy dietary patterns, characterized by excessive consumption of high-fat, high-cholesterol, high-sugar, ultra-processed, and calorie-dense foods, are associated with detrimental effects on the brain, accelerating aging processes and increasing the risk of cognitive impairments [75].
Among dietary factors of interest, high methionine intake has been implicated as a potential contributor to unhealthy cerebrovascular and brain aging [76,77,78,79,80,81,82]. A methionine-rich diet refers to a dietary pattern that is high in methionine, an essential sulfur-containing amino acid. Methionine is an essential building block for protein synthesis and is involved in various biochemical processes. Beyond its role in protein synthesis, methionine also serves as a crucial methyl donor in cellular metabolism. Methionine is involved in the synthesis of S-adenosylmethionine (SAM), a universal methyl donor used in numerous methylation reactions. An excess intake of methionine, found abundantly in certain foods, can lead to increased oxidative stress, mitochondrial dysfunction, and inflammation when consumed in excess. Moreover, methionine metabolism can generate homocysteine, which has been associated with vascular dysfunction and cognitive impairments. Epidemiological studies have consistently highlighted the potential detrimental effects of high methionine consumption and hyperhomocysteinemia (a condition characterized by elevated levels of homocysteine, a metabolite of methionine, in the blood) on health outcomes, particularly in relation to age-related cardiovascular and cerebrovascular diseases and neurodegenerative disorders. Hyperhomocysteinemia is characterized by elevated levels of total homocysteine in the blood, exceeding the standard plasma homocysteine (Hcy) range of 5 to 15 μmol/L. For adults, the recommended daily allowance (RDA) for methionine is around 19 mg per kilogram of body weight, which translates to about 1.3 g for an average adult. A particular study highlighted a prevalence of 35.4% for this condition, with a notable gender disparity—45.4% in men compared to 28.5% in women [83]. Importantly, recent evidence links high methionine consumption and dysregulation of DNA methylation in relation to aging. The concept is emerging that dietary methionine intake plays an important role in modulation of epigenetic modification of DNA that regulates gene expression patterns, cellular function, and the aging process itself [84,85,86].
Preclinical studies utilizing laboratory animals have provided valuable insights into the effects of high-methionine diets and hyperhomocysteinemia on cardiovascular, cerebrovascular, and brain aging [77,87,88,89,90,91,92,93,94,95]. Animal models fed a methionine-rich diet have demonstrated accelerated aging-related changes, including increased vascular oxidative stress, mitochondrial dysfunction, inflammation, and impaired cognitive performance [77,87,88,89,90,91,92,93,94,95]. These studies have revealed the detrimental effects of high methionine intake on multiple facets of vascular and brain health and provide a foundation for further investigation into the underlying mechanisms. In contrast, animals on low methionine diets (“methionine restriction”) exhibit a youthful phenotype and increased health span, including attenuated oxidative stress, inflammation, and improved mitochondrial function [76,79,88,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110,111,112]. These findings underscore the importance of understanding the impact of high and low methionine consumption on cerebrovascular health and its potential implications for cognitive impairment.
In light of these observations, this review aims to comprehensively examine the available evidence regarding the role of a methionine-rich diet in unhealthy cerebrovascular and brain aging, with a specific focus on its implications for cognitive impairment. By evaluating epidemiological data, preclinical studies, and the molecular mechanisms involved, we aim to provide a deeper understanding of the impact of high methionine consumption on cerebrovascular health, brain aging, and cognitive function. By examining the impact of a methionine-rich diet on cognitive decline and age-related pathologies, we aim to shed light on the importance of healthy dietary choices in supporting optimal brain health and mitigating the risk of cognitive impairment and dementia in aging populations. Understanding the impact of high methionine consumption on cerebrovascular and brain aging is essential for developing targeted dietary interventions to promote healthy aging and maintain cognitive function throughout the aging process.

2. Methionine-Rich Diet, Hyperhomocysteinemia, and Aging

2.1. Dietary Sources of Methionine

Methionine can be obtained through dietary sources. Certain foods are particularly high in methionine and can contribute to a methionine-rich diet. Foods that are known to be rich sources of methionine include red meat (beef, lamb, pork), fish (such as salmon, tuna), and dairy products (milk, cheese, yogurt), and some plant-based foods (e.g., legumes).
Red meat consumption in particular has been a topic of increasing interest in the context of aging and age-related diseases [113]. Access to fresh vegetables as compared to meat can vary based on factors like location, socioeconomic status, and individual circumstances. For instance, in developed countries like the U.S., meat constitutes a significant part of the diet, contributing to daily energy, protein, and fat intake due to its availability and preservation ease. On the other hand, fresh vegetables, crucial for a balanced diet, might not be as easily accessible in certain areas, which can be exacerbated by factors like neighborhood disparities and income levels. The accessibility of healthier food options, like fresh vegetables, is known to facilitate healthier diets and potentially lower the risk of overweight and obesity issues. Strategies to improve access to fresh vegetables, especially in comparison to meat and fast food outlets, are being explored to promote better nutrition and health outcomes across different communities [114]. Epidemiological studies have indicated that regular and high intake of red meat may be associated with accelerated aging processes and an increased risk of age-related diseases [115,116,117,118,119,120,121,122,123,124,125,126,127]. Of particular concern is the relationship between red meat consumption and cardiovascular and cerebrovascular diseases [120,121,123,127,128,129,130,131,132,133,134,135,136,137]. The potential adverse effects on health of red meat consumption have been attributed to an increased intake of cholesterol, saturated fat, and methionine [128,129,130,131,132,133,134,135,136,137]. Considering these factors, reducing red meat consumption and opting for healthier protein sources may have beneficial implications for the prevention of age-related cardiovascular and cerebrovascular diseases.

2.2. Methionine Metabolism

Methionine metabolism is a complex and highly regulated process that begins when methionine is taken up from food sources during digestion. Once ingested, methionine is absorbed into the bloodstream and transported to various tissues throughout the body. In the liver, the major site of methionine metabolism, methionine can undergo two main pathways. Through the transmethylation pathway, methionine is converted into SAM through a reaction catalyzed by the enzyme methionine adenosyltransferase. SAM is a universal methyl donor involved in various methylation reactions. It donates methyl groups to molecules such as DNA, histones, proteins, and neurotransmitters, playing a crucial role in epigenetic regulation, protein synthesis, and neurotransmitter metabolism. Excess SAM can be converted back to methionine through the methionine cycle.
Under normal conditions, a smaller portion of methionine is metabolized through the transsulfuration pathway. Here, methionine is converted into homocysteine, a process that requires the enzyme cystathionine beta-synthase (CBS). Subsequently, homocysteine is converted into cysteine, another important sulfur-containing amino acid. Normal methionine-homocysteine metabolism involves a tightly regulated series of biochemical reactions, including the remethylation pathway, that maintains the balance of these two sulfur-containing amino acids in the body. Folate and vitamin B12 are essential cofactors for the remethylation of homocysteine back to methionine. In this process, homocysteine is converted to methionine through the actions of methionine synthase, a vitamin B12-dependent enzyme. Folate donates a methyl group to homocysteine, which is transferred by methionine synthase, along with a methyl group from 5-methyltetrahydrofolate (the active form of folate), to produce methionine. Vitamin B6 is required for the transsulfuration pathway, where homocysteine is converted into cysteine through a series of reactions involving the enzyme cystathionine gamma-lyase (CGL). This well-coordinated system ensures that the levels of methionine and homocysteine are tightly regulated. Proper methionine-homocysteine metabolism is vital for various cellular processes, including protein synthesis, epigenetic regulation, neurotransmitter metabolism, and maintaining redox balance. Disruptions in this finely balanced metabolism, arising from deficiencies in the necessary vitamins, genetic mutations impacting the enzymes involved, or an excess intake of methionine, can lead to a condition known as hyperhomocysteinemia. This condition is associated with an elevated risk of various health issues, including cardiovascular and cerebrovascular diseases, neurodegenerative disorders, and other related health complications (as detailed below).

2.3. Hyperhomocysteinemia

Hyperhomocysteinemia is a condition characterized by elevated levels of homocysteine, a sulfur-containing amino acid derived from methionine metabolism [77,138,139]. Homocysteine is normally metabolized through a process that requires specific vitamins, including folate, vitamin B12, and vitamin B6. The increased dietary intake of methionine, especially if combined with impairment of this metabolic pathway due to deficiencies in these vitamins or genetic variations affecting the enzymes involved in homocysteine metabolism, leads to hyperhomocysteinemia [138,139].
In addition to high methionine intake, hyperhomocysteinemia itself can affect various organ systems. Elevated homocysteine levels have been associated with an increased risk of cardiovascular diseases, including atherosclerosis, arterial thrombosis, and stroke [138,139,140,141,142,143,144,145,146]. Hyperhomocysteinemia has also been linked to cognitive impairment, neurodegenerative diseases, such as Alzheimer’s disease and Parkinson’s disease, and an increased risk of age-related macular degeneration [138,139]. Managing hyperhomocysteinemia involves addressing the underlying causes and optimizing vitamin intake. Supplementation with folate, vitamin B12, and vitamin B6 can help lower homocysteine levels in individuals with deficiencies or impaired metabolism.

2.4. Effects of Methionine-Rich Diet and Hyperhomocysteinemia on Cellular Mechanisms of Aging

The mechanisms of aging represent a complex interplay of cellular processes that lead to the gradual decline of physiological functions and an increased susceptibility to age-related diseases. In recent years, research has shed light on the potential impact of dietary factors in general and methionine intake in particular on the aging process [76,79,88,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110,111,112]. High methionine intake has been associated with detrimental effects on cellular homeostasis and overall health, promoting various cellular and molecular mechanisms of aging and the genesis of age-related diseases [76,79,88,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110,111,112].
Conversely, evidence delineates a pleiotropic relationship between methionine intake and health outcomes. Methionine restriction—though potentially beneficial—can engender adverse effects including the onset of hair greying [147], atherosclerosis precipitated by improper homocysteine conversion [148], and, under severe restriction, the manifestation of fatty liver and anemia as demonstrated in rodent studies [149]. Furthermore, the utilization of methionine restriction as a therapeutic strategy in oncology is deemed impracticable due to the concomitant risks of muscle degradation and sarcopenia, thereby potentially hastening patient deterioration.
In this section, we explore some of the key mechanisms through which high methionine consumption can accelerate the aging process. From its role in increased oxidative stress, DNA damage, and induction of mitochondrial dysfunction to epigenetic modifications and altered protein homeostasis, we delve into the intricate pathways that link high methionine intake to the accelerated aging phenotype. Understanding these mechanisms is crucial for devising strategies to mitigate the adverse effects of a methionine-rich diet and potentially promote healthier aging outcomes.

2.4.1. Oxidative Stress and Mitochondrial Dysfunction

Oxidative stress is a hallmark of aging and is implicated in the pathogenesis of various age-related diseases. Importantly, high methionine intake has been associated with increased cellular oxidative stress [93,150,151,152,153]. Conversely, dietary methionine restriction was shown to be associated with attenuated oxidative stress [88,107,108,109,110]. Methionine metabolism also produces toxic byproducts, such as homocysteine, leading to hyperhomocysteinemia, which also contributes to oxidative damage. The mechanisms by which high methionine intake and hyperhomocysteinemia exacerbate oxidative stress in aging likely include an imbalance between the increased production of reactive oxygen species (ROS) and the impaired ability of antioxidant systems to neutralize them [57,154,155,156,157].
There is strong evidence that high methionine intake and hyperhomocysteinemia increase mitochondrial ROS production, whereas dietary methionine restriction attenuates mitochondrial oxidative stress [88,110]. Mitochondria are crucial for energy production and play a vital role in cellular function. High methionine intake can disrupt mitochondrial function by increasing oxidative stress, impairing mitochondrial DNA integrity, and affecting mitochondrial dynamics. These disruptions can lead to mitochondrial dysfunction, compromised energy production, and bioenergetic impairment. Mitochondrial oxidative stress and mitochondrial dysfunction are hallmarks of aging and can contribute to various age-related diseases, including cardiovascular and cerebrovascular diseases and neurodegenerative disorders [44,45,48,51,54,55,60,158,159,160,161,162,163,164].
High methionine intake and/or hyperhomocysteinemia have also been shown to up-regulate major cellular oxidant systems, including NAD(P)H oxidases [90,91,92,93,94,95]. NAD(P)H oxidases, also known as NOX enzymes, are a family of enzymes contributing to age-related increases in ROS production in cells of the cardiovascular and immune systems as well as other organs [93,165,166]. In the context of high methionine intake and hyperhomocysteinemia, increased NAD(P)H oxidase activity can further exacerbate age-related oxidative stress [93,165]. Elevated homocysteine levels, in particular, have been associated with increased ROS production and oxidative damage in humans [167].
Both mitochondrial ROS generation and the activity and expression of cellular NAD(P)H oxidases are regulated by humoral factors [35,40,41,44,59,168,169,170] and proinflammatory cytokines, such as TNFα [93,171,172]. Previous studies have shown that in high-methionine-fed rodent models there is significant up-regulation of TNFα [93,173]. It is likely that increased inflammatory status associated with high methionine intake contributes to increased oxidative stress both in the cardiovascular system and the central nervous system [93,174,175,176,177,178]. High methionine intake and/or hyperhomocysteinemia were also reported to up-regulate xanthine oxidase [90] and cyclooxygenases [91,95]. Additionally, high methionine intake can also lead to the generation of ROS via the auto-oxidation of the sulfur atom in the SH-containing metabolites of methionine.

2.4.2. Inflammation

Chronic low-grade inflammation, often referred to as inflammaging, is associated with accelerated aging and age-related cardiovascular, cerebrovascular, and brain diseases [36,37,54,179,180,181,182,183,184,185,186,187,188,189,190,191,192,193,194]. Methionine-rich diets have been shown to induce inflammation through multiple mechanisms, including the activation of pro-inflammatory signaling pathways and the generation of inflammatory mediators [78,99,108,153,195]. Chronic inflammation can contribute to accelerated cerebrovascular and brain aging by promoting increased oxidative stress and endothelial dysfunction, compromising mitochondrial energetics and increasing the susceptibility to microvascular damage and cerebrovascular senescence. A heightened state of inflammation also contributes to the pathogenesis of neurodegenerative disorders [181,191,196,197,198,199,200,201,202,203,204,205].

2.4.3. Epigenetic Regulation of Aging Processes

Epigenetic modifications, including DNA methylation, histone modifications, and non-coding RNA expression, play a crucial role in regulating gene expression patterns, cellular function, genomic stability, cellular identity, and organismal aging [206,207,208,209,210,211,212,213,214,215,216,217]. DNA methylation, one of the major epigenetic mechanisms, involves the addition of a methyl group to DNA molecules, typically occurring at cytosine residues within CpG dinucleotides [208,211,212,218]. Methionine serves as a methyl donor in the synthesis of SAM, the principal supplier of methyl groups for DNA methylation reactions. Dysregulation of DNA methylation patterns is associated with accelerated aging [212,219] and is directly linked to the premature development of cognitive impairment and dementia [220,221,222]. High methionine intake can influence DNA methylation patterns, potentially contributing to the accelerated epigenetic aging [84,85,86,223,224,225]. Diet-induced changes in DNA methylation patterns can affect gene expression profiles, disrupt cellular functions, and increase susceptibility to age-related diseases.

2.4.4. DNA Damage and Repair Mechanisms

Increased oxidative stress leads to DNA damage accumulation in aged cells [226]. High methionine intake may contribute to increased DNA damage and impairments in DNA repair mechanisms [225]. Methionine metabolism generates reactive oxygen species (ROS) and reactive nitrogen species (RNS), which can lead to DNA damage [100,227]. Moreover, high methionine levels can disrupt the balance of one-carbon metabolism, affecting the availability of key molecules involved in DNA repair processes. Folate, a crucial B-vitamin, plays a pivotal role in DNA maintenance and repair, and its deficiency can contribute to DNA damage, especially when combined with high methionine intake [228]. Accumulated DNA damage and compromised repair mechanisms can accelerate the aging process and increase the risk of age-related diseases. DNA repair mechanisms play a pivotal role in maintaining genomic stability and preventing age-related phenotypic alteration (including induction of DNA-damage-mediated cellular senescence [191,215,229,230,231,232,233,234]) and functional decline. However, high methionine intake and hyperhomocysteinemia can interfere with these repair processes, contributing to DNA damage accumulation. Understanding the interplay between high methionine intake, hyperhomocysteinemia, and DNA damage/repair mechanisms is crucial for unraveling the mechanisms underpinning unhealthy cerebrovascular and brain aging.

2.4.5. Cellular Senescence

Cellular senescence is a cellular stress response, characterized by a state of irreversible growth arrest in which cells lose their ability to divide and proliferate, even in the presence of stimuli that would typically promote cell division. This state is often accompanied by distinct changes in cell morphology and function. Cellular senescence can be triggered by a variety of stressors, including DNA damage, telomere shortening, and exposure to oxidative stress. Senescent cells remain metabolically active but play a role in aging biology by secreting pro-inflammatory molecules and other factors that can influence tissue function and contribute to the pathogenesis of age-related diseases [46,57,157,235,236,237,238,239,240,241,242,243].
Importantly, cellular senescence has emerged as a pivotal player in the biology of cerebrovascular and brain aging, with profound implications for brain health [37,42,46,244]. Senescent cells accumulate over time in the brain and the cerebral microcirculation, and secrete a complex array of pro-inflammatory factors, collectively known as the senescence-associated secretory phenotype (SASP) [245]. This SASP can promote chronic inflammation, oxidative stress, and tissue dysfunction, all of which are central drivers of brain aging and the genesis of cognitive decline and neurodegenerative diseases.
In the context of cerebrovascular and neurovascular aging, senescence of endothelial cells, pericytes, astrocytes, and paravascular microglia contributes to endothelial dysfunction, neurovascular dysfunction, blood brain barrier dysfunction, and genesis of cerebral microhemorrhages. These changes can lead to diminished cerebral blood flow, impairing nutrient and oxygen delivery to the aging brain, disruption of neuronal networks, and neuroinflammation. Furthermore, senescent cells within the cerebral circulation can disrupt neurovascular coupling, affecting the brain’s ability to respond to changing metabolic demands [42]. This, in turn, can contribute to cognitive decline and increased vulnerability to neurodegenerative diseases. The elimination of senescent cells [246] has recently been reported to improve lifespan and health span in rodents [247,248,249,250,251], consistent with the notion that senescent cells drive organismal aging. Importantly, clearance of senescent cells was shown to rejuvenate cerebromicrovascular function in mouse models of aging and accelerated brain aging [42,252,253].
Remarkably, a high-methionine diet has been implicated in the induction of cellular senescence [254,255,256]. Excessive methionine intake can lead to the generation of ROS, increased oxidative stress, and DNA damage, a known inducer of senescence. Moreover, methionine-derived homocysteine can promote endothelial dysfunction and inflammation, further exacerbating the senescent phenotype within the cerebral vasculature. Consequently, understanding the role of cellular senescence in vascular aging and its relationship with high methionine intake is of paramount importance. Efforts to detect senescent cells in animal models fed a high-methionine diet could involve profiling their transcriptomic signatures, utilizing techniques such as single-cell RNA sequencing [46], flow cytometry [253], or immunohistology. Furthermore, it is advisable to investigate the impact of senolytic treatments on these animal models, as this could yield valuable insights into the potential therapeutic effects of senolytics in the context of accelerated cerebrovascular and brain aging.

2.4.6. Altered Protein Homeostasis and Proteostasis Network

Protein homeostasis, or proteostasis, is a critical cellular process responsible for maintaining the proper folding, function, and degradation of proteins, which is essential for maintaining cellular function and preventing protein aggregation, a hallmark of aging and neurodegenerative diseases [257,258,259,260,261,262,263,264]. High methionine intake may disrupt protein homeostasis by increasing oxidative stress and impairing the function of protein quality control mechanisms, including the ubiquitin-proteasome system and autophagy [265,266]. This disruption can lead to the accumulation of misfolded or damaged proteins, further contributing to cellular dysfunction and accelerated aging processes. Hyperhomocysteinemia can further exacerbate proteostatic disturbances [267]. Elevated homocysteine levels are associated with increased oxidative stress, which can promote protein oxidation and misfolding [268]. Homocysteine-induced oxidative modifications can lead to protein aggregates and impair the function of proteasomal and lysosomal degradation pathways, further compromising proteostasis.

3. Methionine-Rich Diet and Vascular Contributions to Cognitive Impairment

The brain’s elevated metabolic requirements are sustained by an intricate microcirculatory network, estimated to encompass approximately 600 km in length within the human body. This cerebral microcirculation serves as a vital conduit, ensuring the precise distribution of essential resources such as oxygen, glucose, and other nutrients to neural tissues. Additionally, it actively participates in the removal of metabolic waste products, maintains the ionic balance crucial for neuronal function, orchestrates the formation and integrity of the blood–brain barrier (BBB), and governs the transport of diverse substances across this barrier. Thus, the health of the microvasculature assumes a pivotal role in preserving normal cognitive and neuronal functions [20,21,45,47,49,50,51,269,270,271,272,273,274,275,276]. It is increasingly evident that dysfunction and damage within the cerebral microcirculation constitute significant contributors to age-related cognitive decline [17,29,273,274,275,277,278,279]. Clinical investigations have furnished evidence suggesting that the consumption of a methionine-rich diet, coupled with hyperhomocysteinemia, can incite dysregulation in cerebral blood flow, directly impacting cognition [280,281,282,283]. Experimental studies have extended these clinical observations, providing valuable mechanistic insights into the synergistic interplay between the aging process and diet-induced accelerated cellular aging, particularly concerning cerebromicrovascular function [173,178,284]. This section offers a comprehensive exploration of the distinct pathogenic roles played by endothelial dysfunction, neurovascular impairment, microvascular rarefaction, and blood–brain barrier disruption in the genesis of vascular cognitive impairment (VCI). Furthermore, we scrutinize the potential involvement of a methionine-rich diet and hyperhomocysteinemia in exacerbating these pathogenic processes, shedding light on the intricate interconnections between diet, vascular health, and cognitive outcomes.

3.1. Atherosclerosis and Stroke

Both hyperhomocysteinemia and increased methionine intake are associated with atherosclerotic vascular diseases, including stroke [140,141,142,143,144,145,146,285,286,287]. Previous studies also confirmed significant genetic associations between premature ischaemic stroke and haplotypes in various genes involved in methionine metabolism [285,288,289]. Preclinical studies distinguished between the effects of excess dietary methionine from those of genetic forms of hyperhomocysteinemia and provides evidence that consumption of a methionine-rich diet per se promotes atherogenesis [87,290,291,292,293,294,295].

3.2. Endothelial Dysfunction and Dysregulation of Cerebral Blood Flow

Endothelial cells play a crucial role in maintaining vascular homeostasis. Endothelium-dependent, NO-mediated dilation of cerebral microvessels plays a crucial role both in maintaining cerebral blood flow [279,296] and mediating neurovascular coupling responses [40,47,48,55,154,252,297,298,299,300,301]. Neurovascular coupling refers to the tight regulation of cerebral blood flow in response to neuronal activity. This mechanism ensures that an adequate supply of oxygen and nutrients is delivered to active brain regions.
With aging, endothelial dysfunction occurs, characterized by impaired endothelium-dependent vasodilation, increased oxidative stress, and inflammation [24,42,44,47,48,51,52,55,57,59,162,168,171,183,269,270,271,301,302,303,304,305,306]. This dysfunction contributes to cognitive decline by compromising cerebral blood flow regulation, impairing nutrient and oxygen delivery to the brain, and disrupting the blood–brain barrier (BBB) [20,21,27,30,43,47,48,55,269,272,273,274,279,307,308,309]. Clinical studies have shown associations between endothelial dysfunction and cognitive impairment in aging populations [24,302]. Age-related impairments in neurovascular coupling, due to endothelial dysfunction, have been observed and are associated with cognitive dysfunction [21,35,42,47,48,51,55,270].
Earlier research has indicated that consumption of a methionine-rich diet and an increase in plasma homocysteine levels can impair endothelium-mediated vasodilation in humans [150,310,311,312,313,314,315,316], mimicking the aging phenotype. Additionally, preclinical studies have demonstrated that high-methionine diets or hyperhomocysteinemia can exacerbate endothelial dysfunction [90,91,92,93,94,95,317]. This occurs through a reduction in NO bioavailability due to increased vascular oxidative stress [90,91,92,93,94]. Preliminary preclinical evidence suggests that high-methionine diets and hyperhomocysteinemia may also disrupt the synthesis of vasoactive arachidonic acid metabolites within the vasculature [91,95,306]. Endothelial dysfunction induced by high-methionine diets or hyperhomocysteinemia can further disrupt neurovascular coupling by impairing the dilation of cerebral arterioles in response to neuronal activation [281,318]. Clinical and preclinical studies have reported associations between high methionine intake or hyperhomocysteinemia, dysregulation of cerebral blood flow, and cognitive dysfunction [173,178,280,282,283,284,319].

3.3. Microvascular Rarefaction

Microvascular rarefaction refers to a reduction in the density and branching of small blood vessels, impairing microcirculation and nutrient exchange in the brain. Aging is associated with microvascular rarefaction, which can compromise cerebral perfusion and contribute to cognitive decline [26,43,58,279,301,320,321]. Crucially, there exists a direct correlation between the degree of age-related capillary rarefaction observed in the hippocampus and the severity of cognitive decline. This correlation serves as supplementary evidence affirming the intimate link between the disruption of cerebral blood flow and impaired neuronal function. High-methionine diets or hyperhomocysteinemia have been shown to accelerate microvascular rarefaction, promoting the loss of capillaries and impairing the integrity of the neurovascular unit [77]. The mechanisms that potentially contribute to cerebromicrovascular rarefaction in the context of aging may encompass several factors, including diminished endothelial senescence, endothelial NO availability [322,323,324,325], pericyte depletion [301], heightened endothelial cell apoptosis [326,327], reduced levels of pro-angiogenic factors like VEGF [328] and IGF-1 [44,58,329,330,331], and compromised endothelial angiogenic processes [58,159,301,332,333,334,335]. Further studies are warranted to determine how consumption of a diet rich in methionine impacts these synergistic mechanisms in the cerebral microcirculation.

3.4. Blood–Brain Barrier Disruption and Neuroinflammation

The blood–brain barrier (BBB) regulates the exchange of substances between the blood and the brain, maintaining brain homeostasis [29,273,274]. Aging is associated with BBB dysfunction, characterized by increased permeability and compromised barrier integrity [43,204,309,336,337,338,339]. BBB disruption can lead to neuroinflammation, neuronal damage, and cognitive impairment [29,30,273,274,309]. Preclinical evidence suggests that high-methionine diets or hyperhomocysteinemia may contribute to BBB disruption, allowing the entry of harmful substances into the brain and exacerbating neuroinflammation and cognitive decline [78,284,318,340,341,342,343]. Understanding the potential mechanisms underlying BBB disruption induced by high methionine intake and hyperhomocysteinemia and its role in accelerated brain aging is crucial for identifying potential therapeutic interventions that target cerebrovascular health and mitigate the detrimental effects of high methionine consumption on cognitive function. The mechanisms contributing to increased BBB permeability are likely to be multifaceted. Firstly, there may be alterations in the expression levels of tight junction and adherens junction proteins potentially compromising BBB integrity [344,345]. Moreover, high methionine intake and hyperhomocysteinemia are expected to induce post-translational modifications of tight junction proteins. These modifications could impact the stability and appropriate cellular localization of these proteins. Pericytes, crucial structural components of the BBB, play a pivotal role in maintaining its integrity. In this context, it’s noteworthy that animal models of hyperhomocysteinemia exhibit pericyte damage [344,345]. There is also increasing evidence that cellular senescence promotes BBB disruption [244,252,346,347]. Additionally, the cells that constitute the BBB exhibit a high metabolic rate, consistent with their elevated energy demands to support the activities of ATP-dependent transporters. Both high methionine intake and hyperhomocysteinemia were shown to compromise mitochondrial function in the vasculature [100,174,175].
A significant outcome of BBB disruption is the leakage of plasma components, including IgG, thrombin, and fibrinogen, into the brain parenchyma [348]. This heightened infiltration of plasma proteins across the BBB serves as a catalyst for neuroinflammation, primarily driven by the activation of resident immune cells, particularly microglia [348]. Notably, the increased presence of activated microglia in the hippocampi is closely linked to the aggravated impairment of long-term potentiation (LTP) in excitatory synaptic transmission. LTP is a fundamental cellular mechanism associated with learning and memory [349].

4. Methionine-Rich Diet and Synaptic Function/Neuronal Health

4.1. Effects of High Methionine Intake and Hyperhomocysteinemia on Synaptic Plasticity and Neurotransmitter Systems

Synaptic plasticity, the ability of synapses to strengthen or weaken over time in response to changes in their activity, is fundamental to learning and memory processes. High methionine intake and hyperhomocysteinemia have been shown to impact synaptic plasticity by disrupting key molecular signaling pathways involved in synaptic function [223,350,351,352,353,354]. Preclinical studies have demonstrated that a methionine-rich diet can impair long-term potentiation (LTP), a cellular mechanism underlying synaptic plasticity, leading to deficits in learning and memory [351,354,355]. Elevated homocysteine levels, stemming from high methionine intake, have been linked to reduced levels of brain-derived neurotrophic factor (BDNF) [356], a molecule essential for synaptic plasticity. Reduced BDNF impairs LTP. Furthermore, homocysteine can induce oxidative stress, further inhibiting LTP and leading to synaptic dysfunction. High methionine intake and hyperhomocysteinemia can also influence various neurotransmitter systems [357]. Elevated homocysteine levels have been associated with disruptions in the dopaminergic, serotonergic, and cholinergic neurotransmitter systems. These disturbances can lead to mood disorders, cognitive deficits, and other neurological issues.

4.2. Neuronal Damage and Neurodegenerative Processes Induced by High Methionine Intake and Hyperhomocysteinemia

Neurodegenerative diseases, characterized by the progressive loss of structure and function of neurons, represent a burgeoning global health concern [6,358]. Conditions such as Alzheimer’s disease and Parkinson’s disease have seen a notable increase in prevalence, primarily due to the aging global population. Alzheimer’s disease, in particular, stands out as the most common form of dementia, affecting millions worldwide, with its incidence expected to triple by 2050. These diseases not only inflict immense emotional and physical burdens on patients and caregivers but also pose significant socioeconomic challenges. The rising prevalence underscores the urgency of understanding their pathophysiology, developing effective preventive strategies, including dietary interventions.
Preclinical studies have shown that a methionine-rich diet and hyperhomocysteinemia can induce neurodegenerative disease-like changes, including protein aggregation and neuroinflammation. Clinical evidence has also suggested associations between high methionine intake or hyperhomocysteinemia, imaging biomarkers of accelerated aging of the central nervous system [359,360], and an increased risk of neurodegenerative diseases, such as Alzheimer’s disease and Parkinson’s disease. A sustained high-methionine diet and hyperhomocysteinemia have been implicated in direct and indirect neuronal damage [77,81,176,352,361,362]. Oxidative stress is one mechanism by which high methionine intake and hyperhomocysteinemia induces cellular damage. Elevated levels of homocysteine increase the production of reactive oxygen species, leading to oxidative stress and subsequent damage to neuronal DNA, proteins, and lipids. Furthermore, a high-methionine diet and hyperhomocysteinemia has been linked to mitochondrial dysfunction in neurons. Mitochondrial dysfunctions can trigger a cascade of events, including the release of pro-apoptotic proteins, leading to neuronal cell death. Moreover, homocysteine can activate microglia, leading to neuroinflammation, another mechanism contributing to neurodegeneration [342,363]. Besides, there are indications that high methionine levels might influence the formation and aggregation of beta-amyloid plaques, a hallmark of Alzheimer’s disease, further suggesting a link between methionine-rich diets and neurodegenerative processes [363,364,365,366,367,368,369,370,371,372,373,374,375]. Elevated homocysteine levels have also been associated with increased tau hyperphosphorylation, a hallmark of Alzheimer’s disease [176,370]. Additionally, studies have shown a correlation between hyperhomocysteinemia and enhanced production of amyloid-beta, the primary component of amyloid plaques in Alzheimer’s disease [89,373].

4.3. Methionine-Rich Diet, Hyperhomocysteinemia, and Altered Functional Connectivity

Functional connectivity refers to the temporal correlation of neural activity between different brain regions, which reflects the coordination and integration of neural networks. Disruptions in functional connectivity with aging [376] have been implicated in various cognitive impairments and neurological disorders [377,378,379,380,381,382,383,384,385,386,387,388]. Brain connectivity, essential for efficient cognitive function, can be influenced by dietary factors [389,390,391]. In particular, hyperhomocysteinemia has been shown to influence functional connectivity patterns in the brain [82,392]. These studies suggest that hyperhomocysteinemia can alter the default mode network, a key brain network active during rest and implicated in memory processes. These alterations in functional connectivity may contribute to cognitive impairments observed in individuals consuming a methionine-rich diet. The exact mechanisms underlying the effects of high methionine intake on functional connectivity are still being explored. However, it is believed that oxidative stress, inflammation, and mitochondrial dysfunction induced by high methionine consumption can cause white matter damage [368,393] and disrupt the normal communication and coordination between brain regions, leading to altered functional connectivity. Understanding the impact of a methionine-rich diet on functional connectivity provides further insights into the potential mechanisms through which high methionine intake may contribute to cognitive impairments.

5. Conclusions and Perspectives

In summary, the relationship between elevated methionine intake, hyperhomocysteinemia, and the aging processes affecting cerebrovascular and brain health stands as a pivotal area of investigation in understanding age-related cognitive decline. This review has shed light on the multifaceted pathways through which these dietary and metabolic factors accelerate aging mechanisms within the cerebrovascular system and brain, ultimately contributing to cognitive impairment. From the initiation of oxidative stress and mitochondrial dysfunction to the induction of chronic inflammation and accelerated epigenetic aging, high methionine consumption and hyperhomocysteinemia set forth a cascade of detrimental effects.
The body of evidence available underscores the pivotal role played by microvascular health in the context of age-related cognitive impairment, offering illuminating insights into how factors such as endothelial dysfunction, reduced cerebral blood flow, microvascular rarefaction, compromised neurovascular coupling, and blood–brain barrier disruption collectively contribute to the complex landscape of cognitive decline. Importantly, it should be noted that these factors are modifiable through dietary interventions and lifestyle adjustments, providing promising avenues for the preservation of cognitive function and the promotion of healthy brain aging. The profound impact on synaptic function and neuronal health further underscores the extensive consequences of these dietary factors.
Looking forward, further research should delve deeper into the intricate molecular mechanisms underpinning the relationship between methionine-rich diets, hyperhomocysteinemia, and cerebrovascular and brain aging. It is imperative to investigate the dose-response relationship between methionine intake and cerebrovascular/brain aging to establish optimal intake thresholds. Additionally, efforts should focus on elucidating the molecular mechanisms through which high methionine intake affects microvascular dysfunction, synaptic function, and neuronal health. Exploring the interactions between methionine metabolism and other dietary factors in modulating cerebrovascular health and brain aging will be crucial in gaining a comprehensive understanding of these processes. Investigations into potential interventions to mitigate the detrimental effects of high methionine intake on cerebrovascular health and brain aging are warranted. Strategies targeting these mechanisms, such as dietary modifications, supplementation, or pharmacological interventions, should be further explored in preclinical and clinical settings.
To optimize cerebrovascular and brain health and counteract the adverse effects of a methionine-rich diet, several dietary recommendations and potential therapeutic strategies can be considered [394,395,396,397,398,399]. These include optimizing methionine intake by adopting a balanced diet rich in nutrient-dense whole foods, favoring plant-based diets [400] like the mediterranean diet [401] or vegetarian/vegan diets. Such diets naturally limit methionine intake while providing essential amino acids. Increasing folate intake, found in foods like leafy greens and legumes, can aid in homocysteine metabolism. An antioxidant-rich diet comprising fruits, vegetables, nuts, and seeds can help combat oxidative stress linked to high methionine intake. Regular physical exercise has demonstrated benefits in improving cerebrovascular function, reducing oxidative stress, and enhancing cognitive performance [402]. Moreover, pharmacological interventions, such as antioxidants, anti-inflammatory agents, or compounds targeting mitochondrial function, may hold promise for mitigating the detrimental effects of high methionine intake on cerebrovascular and brain aging. Nevertheless, further research is essential to validate the effectiveness and safety of these interventions in clinical contexts.

Author Contributions

Conceptualization, A.U., R.G., B.C., P.M., S.N., S.T., A.Y., Z.B., A.C. and Z.U.; writing—original draft preparation, A.U., R.G., B.C., P.M., S.N., S.T., A.Y., Z.B., A.C. and Z.U.; writing—review and editing, A.U., R.G., B.C., P.M., S.N., S.T., A.Y., Z.B., A.C. and Z.U.; funding acquisition, S.T., A.Y., Z.B., A.C. and Z.U. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by grants from the American Heart Association (ST: AHA CDA941290, RG: AHA 916225), the Oklahoma Center for the Advancement of Science and Technology, the National Institute on Aging (RF1AG072295, R01AG055395, R01AG068295; R01AG070915, K01AG073614, K01AG073613, R03AG070479), the National Institute of Neurological Disorders and Stroke (R01NS100782), the National Cancer Institute (R01CA255840), the Oklahoma Shared Clinical and Translational Resources (U54GM104938) with an Institutional Development Award (IDeA) from NIGMS, the Presbyterian Health Foundation, the Reynolds Foundation, the Oklahoma Nathan Shock Center (P30AG050911), and the Cellular and Molecular GeroScience CoBRE (P20GM125528), the NCI Cancer Center Support Grant (P30 CA225520) and the Oklahoma Tobacco Settlement Endowment Trust. AU was supported by Project no. TKP2021-NKTA-47, 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; by funding through the National Cardiovascular Laboratory Program (RRF-2.3.1-21-2022-00003) provided by the Ministry of Innovation and Technology of Hungary from the National Research, Development and Innovation Fund; Project no. 135784 implemented with the support provided from the National Research, Development and Innovation Fund of Hungary, financed under the K_20 funding scheme and the European University for Well-Being (EUniWell) program (grant agreement number: 101004093/EUniWell/EAC-A02-2019/EAC-A02-2019-1). The funding sources had no role in the study design; in the collection, analysis and interpretation of data; in the writing of the report; and in the decision to submit the article for publication. The content is solely the responsibility of the authors and does not necessarily represent the official views of the National Institutes of Health, the American Heart Association, or the Presbyterian Health Foundation. The 3.5 version of ChatGPT, developed by OpenAI, was used as a language tool to refine our writing, enhancing the clarity of our work.

Conflicts of Interest

The authors declare no competing financial interest.

References

  1. Ungvari, Z.; Adany, R. The future of healthy aging: Translation of geroscience discoveries to public health practice. Eur. J. Public Health 2021, 31, 455–456. [Google Scholar] [CrossRef] [PubMed]
  2. Ortman, J.M.; Velkoff, V.A.; Hogan, H. An Aging Nation: The Older Population in the United States, Current Population Reports; U.S. Census Bureau: Washington, DC, USA, 2014; pp. 25–1140.
  3. Malhotra, R.; Bautista, M.A.C.; Muller, A.M.; Aw, S.; Koh, G.C.H.; Theng, Y.L.; Hoskins, S.J.; Wong, C.H.; Miao, C.; Lim, W.S.; et al. The Aging of a Young Nation: Population Aging in Singapore. Gerontologist 2019, 59, 401–410. [Google Scholar] [CrossRef] [PubMed]
  4. 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] [PubMed]
  5. Qian, Y.; Chen, X.; Tang, D.; Kelley, A.S.; Li, J. Prevalence of Memory-Related Diagnoses among U.S. Older Adults with Early Symptoms of Cognitive Impairment. J. Gerontol. A Biol. Sci. Med. Sci. 2021, 76, 1846–1853. [Google Scholar] [CrossRef] [PubMed]
  6. Fang, X.; Zhang, J.; Roman, R.J.; Fan, F. From 1901 to 2022, how far are we from truly understanding the pathogenesis of age-related dementia? Geroscience 2022, 44, 1879–1883. [Google Scholar] [CrossRef]
  7. Enroth, L.; Raitanen, J.; Halonen, P.; Tiainen, K.; Jylha, M. Trends of Physical Functioning, Morbidity, and Disability-Free Life Expectancy among the Oldest Old: Six Repeated Cross-Sectional Surveys Between 2001 and 2018 in the Vitality 90+ Study. J. Gerontol. A Biol. Sci. Med. Sci. 2021, 76, 1227–1233. [Google Scholar] [CrossRef]
  8. Ullrich, P.; Werner, C.; Bongartz, M.; Eckert, T.; Abel, B.; Schonstein, A.; Kiss, R.; Hauer, K. Increasing Life-Space Mobility in Community-Dwelling Older Persons with Cognitive Impairment Following Rehabilitation: A Randomized Controlled Trial. J. Gerontol. A Biol. Sci. Med. Sci. 2021, 76, 1988–1996. [Google Scholar] [CrossRef]
  9. Wadley, V.G.; Bull, T.P.; Zhang, Y.; Barba, C.; Bryan, R.N.; Crowe, M.; Desiderio, L.; Deutsch, G.; Erus, G.; Geldmacher, D.S.; et al. Cognitive Processing Speed Is Strongly Related to Driving Skills, Financial Abilities, and Other Instrumental Activities of Daily Living in Persons with Mild Cognitive Impairment and Mild Dementia. J. Gerontol. A Biol. Sci. Med. Sci. 2021, 76, 1829–1838. [Google Scholar] [CrossRef]
  10. Akushevich, I.; Yashkin, A.; Kovtun, M.; Kravchenko, J.; Arbeev, K.; Yashin, A.I. Forecasting prevalence and mortality of Alzheimer’s disease using the partitioning models. Exp. Gerontol. 2023, 174, 112133. [Google Scholar] [CrossRef]
  11. Barnes, D.E.; Yaffe, K. The projected effect of risk factor reduction on Alzheimer’s disease prevalence. Lancet Neurol. 2011, 10, 819–828. [Google Scholar] [CrossRef]
  12. Breitner, J.C.; Welsh, K.A.; Gau, B.A.; McDonald, W.M.; Steffens, D.C.; Saunders, A.M.; Magruder, K.M.; Helms, M.J.; Plassman, B.L.; Folstein, M.F.; et al. Alzheimer’s disease in the National Academy of Sciences-National Research Council Registry of Aging Twin Veterans. III. Detection of cases, longitudinal results, and observations on twin concordance. Arch. Neurol. 1995, 52, 763–771. [Google Scholar] [CrossRef] [PubMed]
  13. Gillis, C.; Montenigro, P.; Nejati, M.; Maserejian, N. Estimating prevalence of early Alzheimer’s disease in the United States, accounting for racial and ethnic diversity. Alzheimers Dement. 2023, 19, 1841–1848. [Google Scholar] [CrossRef] [PubMed]
  14. Bendayan, R.; Zhu, Y.; Federman, A.D.; Dobson, R.J.B. Multimorbidity Patterns and Memory Trajectories in Older Adults: Evidence From the English Longitudinal Study of Aging. J. Gerontol. A Biol. Sci. Med. Sci. 2021, 76, 867–875. [Google Scholar] [CrossRef]
  15. Fassier, P.; Kang, J.H.; Lee, I.M.; Grodstein, F.; Vercambre, M.N. Vigorous Physical Activity and Cognitive Trajectory Later in Life: Prospective Association and Interaction by Apolipoprotein E e4 in the Nurses’ Health Study. J. Gerontol. A Biol. Sci. Med. Sci. 2022, 77, 817–825. [Google Scholar] [CrossRef] [PubMed]
  16. Kirkpatrick, A.C.; Stoner, J.A.; Donna-Ferreira, F.; Malatinszky, G.C.; Guthery, L.D.; Scott, J.; Prodan, C.I. High rates of undiagnosed vascular cognitive impairment among American Indian veterans. Geroscience 2019, 41, 69–76. [Google Scholar] [CrossRef] [PubMed]
  17. Iadecola, C.; Duering, M.; Hachinski, V.; Joutel, A.; Pendlebury, S.T.; Schneider, J.A.; Dichgans, M. Vascular Cognitive Impairment and Dementia: JACC Scientific Expert Panel. J. Am. Coll. Cardiol. 2019, 73, 3326–3344. [Google Scholar] [CrossRef]
  18. van der Flier, W.M.; Skoog, I.; Schneider, J.A.; Pantoni, L.; Mok, V.; Chen, C.L.H.; Scheltens, P. Vascular cognitive impairment. Nat. Rev. Dis. Primers 2018, 4, 18003. [Google Scholar] [CrossRef]
  19. Balasubramanian, P.; Kiss, T.; Tarantini, S.; Nyul-Toth, A.; Ahire, C.; Yabluchanskiy, A.; Csipo, T.; Lipecz, A.; Tabak, A.; Institoris, A.; et al. Obesity-induced cognitive impairment in older adults: A microvascular perspective. Am. J. Physiol. Heart Circ. Physiol. 2021, 320, H740–H761. [Google Scholar] [CrossRef]
  20. Csipo, T.; Lipecz, A.; Fulop, G.A.; Hand, R.A.; Ngo, B.N.; Dzialendzik, M.; Tarantini, S.; Balasubramanian, P.; Kiss, T.; Yabluchanska, V.; et al. Age-related decline in peripheral vascular health predicts cognitive impairment. Geroscience 2019, 41, 125–136. [Google Scholar] [CrossRef]
  21. Csipo, T.; Mukli, P.; Lipecz, A.; Tarantini, S.; Bahadli, D.; Abdulhussein, O.; Owens, C.; Kiss, T.; Balasubramanian, P.; Nyul-Toth, A.; et al. Assessment of age-related decline of neurovascular coupling responses by functional near-infrared spectroscopy (fNIRS) in humans. Geroscience 2019, 41, 495–509. [Google Scholar] [CrossRef]
  22. 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]
  23. Nyul-Toth, A.; Fulop, G.A.; Tarantini, S.; Kiss, T.; Ahire, C.; Faakye, J.A.; Ungvari, A.; Toth, P.; Toth, A.; Csiszar, A.; et al. Cerebral venous congestion exacerbates cerebral microhemorrhages in mice. Geroscience 2022, 44, 805–816. [Google Scholar] [CrossRef] [PubMed]
  24. Owens, C.D.; Mukli, P.; Csipo, T.; Lipecz, A.; Silva-Palacios, F.; Dasari, T.W.; Tarantini, S.; Gardner, A.W.; Montgomery, P.S.; Waldstein, S.R.; et al. Microvascular dysfunction and neurovascular uncoupling are exacerbated in peripheral artery disease, increasing the risk of cognitive decline in older adults. Am. J. Physiol. Heart Circ. Physiol. 2022, 322, H924–H935. [Google Scholar] [CrossRef] [PubMed]
  25. Ungvari, Z.; Tarantini, S.; Kirkpatrick, A.C.; Csiszar, A.; Prodan, C.I. Cerebral microhemorrhages: Mechanisms, consequences, and prevention. Am. J. Physiol. Heart Circ. Physiol. 2017, 312, H1128–H1143. [Google Scholar] [CrossRef]
  26. Ungvari, Z.; Toth, P.; Tarantini, S.; Prodan, C.I.; Sorond, F.; Merkely, B.; Csiszar, A. Hypertension-induced cognitive impairment: From pathophysiology to public health. Nat. Rev. Nephrol. 2021, 17, 639–654. [Google Scholar] [CrossRef]
  27. Montagne, A.; Barnes, S.R.; Nation, D.A.; Kisler, K.; Toga, A.W.; Zlokovic, B.V. Imaging subtle leaks in the blood-brain barrier in the aging human brain: Potential pitfalls, challenges, and possible solutions. Geroscience 2022, 44, 1339–1351. [Google Scholar] [CrossRef]
  28. Zlokovic, B.V.; Gottesman, R.F.; Bernstein, K.E.; Seshadri, S.; McKee, A.; Snyder, H.; Greenberg, S.M.; Yaffe, K.; Schaffer, C.B.; Yuan, C.; et al. Vascular contributions to cognitive impairment and dementia (VCID): A report from the 2018 National Heart, Lung, and Blood Institute and National Institute of Neurological Disorders and Stroke Workshop. Alzheimers Dement. 2020, 16, 1714–1733. [Google Scholar] [CrossRef]
  29. 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]
  30. Nation, D.A.; Sweeney, M.D.; Montagne, A.; Sagare, A.P.; D’Orazio, L.M.; Pachicano, M.; Sepehrband, F.; Nelson, A.R.; Buennagel, D.P.; Harrington, M.G.; et al. Blood-brain barrier breakdown is an early biomarker of human cognitive dysfunction. Nat. Med. 2019, 25, 270–276. [Google Scholar] [CrossRef]
  31. Fitzgibbon-Collins, L.K.; Heckman, G.A.; Bains, I.; Noguchi, M.; McIlroy, W.E.; Hughson, R.L. Older Adults’ Drop in Cerebral Oxygenation on Standing Correlates with Postural Instability and May Improve with Sitting Prior to Standing. J. Gerontol. A Biol. Sci. Med. Sci. 2021, 76, 1124–1133. [Google Scholar] [CrossRef]
  32. Beason-Held, L.L.; Fournier, D.; Shafer, A.T.; Fabbri, E.; An, Y.; Huang, C.W.; Bilgel, M.; Wong, D.F.; Ferrucci, L.; Resnick, S.M. Disease Burden Affects Aging Brain Function. J. Gerontol. A Biol. Sci. Med. Sci. 2022, 77, 1810–1818. [Google Scholar] [CrossRef] [PubMed]
  33. Cai, M.; Jacob, M.A.; Norris, D.G.; de Leeuw, F.E.; Tuladhar, A.M. Longitudinal Relation Between Structural Network Efficiency, Cognition, and Gait in Cerebral Small Vessel Disease. J. Gerontol. A Biol. Sci. Med. Sci. 2022, 77, 554–560. [Google Scholar] [CrossRef] [PubMed]
  34. Siejka, T.P.; Srikanth, V.K.; Hubbard, R.E.; Moran, C.; Beare, R.; Wood, A.G.; Collyer, T.A.; Gujjari, S.; Phan, T.G.; Callisaya, M.L. Frailty Is Associated with Cognitive Decline Independent of Cerebral Small Vessel Disease and Brain Atrophy. J. Gerontol. A Biol. Sci. Med. Sci. 2022, 77, 1819–1826. [Google Scholar] [CrossRef] [PubMed]
  35. Toth, L.; Czigler, A.; Hegedus, E.; Komaromy, H.; Amrein, K.; Czeiter, E.; Yabluchanskiy, A.; Koller, A.; Orsi, G.; Perlaki, G.; et al. Age-related decline in circulating IGF-1 associates with impaired neurovascular coupling responses in older adults. Geroscience 2022, 44, 2771–2783. [Google Scholar] [CrossRef]
  36. Kiss, T.; Nyúl-Tóth, Á.; Gulej, R.; Tarantini, S.; Csipo, T.; Mukli, P.; Ungvari, A.; Balasubramanian, P.; Yabluchanskiy, A.; Benyo, Z.; et al. Old blood from heterochronic parabionts accelerates vascular aging in young mice: Transcriptomic signature of pathologic smooth muscle remodeling. Geroscience 2022, 44, 953–981. [Google Scholar] [CrossRef]
  37. Kiss, T.; Nyul-Toth, A.; DelFavero, J.; Balasubramanian, P.; Tarantini, S.; Faakye, J.; Gulej, R.; Ahire, C.; Ungvari, A.; Yabluchanskiy, A.; et al. Spatial transcriptomic analysis reveals inflammatory foci defined by senescent cells in the white matter, hippocampi and cortical grey matter in the aged mouse brain. Geroscience 2022, 44, 661–681. [Google Scholar] [CrossRef]
  38. Gosalia, J.; Montgomery, P.S.; Zhang, S.; Pomilla, W.A.; Wang, M.; Liang, M.; Csiszar, A.; Ungvari, Z.; Yabluchanskiy, A.; Proctor, D.N.; et al. Increased pulse wave velocity is related to impaired working memory and executive function in older adults with metabolic syndrome. Geroscience 2022, 44, 2831–2844. [Google Scholar] [CrossRef]
  39. Whitehead, S.N.; Bruno, A.; Burns, J.M.; Carmichael, S.T.; Csiszar, A.; Edwards, J.D.; Elahi, F.M.; Faraco, G.; Gould, D.B.; Gustafson, D.R.; et al. Expanding the horizon of research into the pathogenesis of the white matter diseases: Proceedings of the 2021 Annual Workshop of the Albert Research Institute for White Matter and Cognition. Geroscience 2021, 16, 1714–1733. [Google Scholar] [CrossRef]
  40. Tarantini, S.; Nyul-Toth, A.; Yabluchanskiy, A.; Csipo, T.; Mukli, P.; Balasubramanian, P.; Ungvari, A.; Toth, P.; Benyo, Z.; Sonntag, W.E.; et al. Endothelial deficiency of insulin-like growth factor-1 receptor (IGF1R) impairs neurovascular coupling responses in mice, mimicking aspects of the brain aging phenotype. Geroscience 2021, 43, 2387–2394. [Google Scholar] [CrossRef]
  41. Tarantini, S.; Balasubramanian, P.; Yabluchanskiy, A.; Ashpole, N.M.; Logan, S.; Kiss, T.; Ungvari, A.; Nyul-Toth, A.; Schwartzman, M.L.; Benyo, Z.; et al. IGF1R signaling regulates astrocyte-mediated neurovascular coupling in mice: Implications for brain aging. Geroscience 2021, 43, 901–911. [Google Scholar] [CrossRef]
  42. Tarantini, S.; Balasubramanian, P.; Delfavero, J.; Csipo, T.; Yabluchanskiy, A.; Kiss, T.; Nyul-Toth, A.; Mukli, P.; Toth, P.; Ahire, C.; et al. Treatment with the BCL-2/BCL-xL inhibitor senolytic drug ABT263/Navitoclax improves functional hyperemia in aged mice. Geroscience 2021, 43, 2427–2440. [Google Scholar] [CrossRef] [PubMed]
  43. Nyul-Toth, A.; Tarantini, S.; DelFavero, J.; Yan, F.; Balasubramanian, P.; Yabluchanskiy, A.; Ahire, C.; Kiss, T.; Csipo, T.; Lipecz, A.; et al. Demonstration of age-related blood-brain barrier disruption and cerebromicrovascular rarefaction in mice by longitudinal intravital two-photon microscopy and optical coherence tomography. Am. J. Physiol. Heart Circ. Physiol. 2021, 320, H1370–H1392. [Google Scholar] [CrossRef] [PubMed]
  44. 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] [PubMed]
  45. 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]
  46. Kiss, T.; Nyul-Toth, A.; Balasubramanian, P.; Tarantini, S.; Ahire, C.; DelFavero, J.; Yabluchanskiy, A.; Csipo, T.; Farkas, E.; Wiley, G.; et al. Single-cell RNA sequencing identifies senescent cerebromicrovascular endothelial cells in the aged mouse brain. Geroscience 2020, 42, 429–444. [Google Scholar] [CrossRef]
  47. 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]
  48. 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]
  49. Kiss, T.; Giles, C.B.; Tarantini, S.; Yabluchanskiy, A.; Balasubramanian, P.; Gautam, T.; Csipo, T.; Nyul-Toth, A.; Lipecz, A.; Szabo, C.; et al. Nicotinamide mononucleotide (NMN) supplementation promotes anti-aging miRNA expression profile in the aorta of aged mice, predicting epigenetic rejuvenation and anti-atherogenic effects. Geroscience 2019, 41, 419–439. [Google Scholar] [CrossRef]
  50. Farias Quipildor, G.E.; Mao, K.; Hu, Z.; Novaj, A.; Cui, M.H.; Gulinello, M.; Branch, C.A.; Gubbi, S.; Patel, K.; Moellering, D.R.; et al. Central IGF-1 protects against features of cognitive and sensorimotor decline with aging in male mice. Geroscience 2019, 41, 185–208. [Google Scholar] [CrossRef]
  51. 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]
  52. 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] [PubMed]
  53. Van Skike, C.E.; Jahrling, J.B.; Olson, A.B.; Sayre, N.L.; Hussong, S.A.; Ungvari, Z.; Lechleiter, J.D.; Galvan, V. Inhibition of mTOR protects the blood-brain barrier in models of Alzheimer’s disease and vascular cognitive impairment. Am. J. Physiol. Heart Circ. Physiol. 2018, 314, H693–H703. [Google Scholar] [CrossRef] [PubMed]
  54. Ungvari, Z.; Tarantini, S.; Donato, A.J.; Galvan, V.; Csiszar, A. Mechanisms of Vascular Aging. Circ. Res. 2018, 123, 849–867. [Google Scholar] [CrossRef] [PubMed]
  55. 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] [PubMed]
  56. Fulop, G.A.; Ramirez-Perez, F.I.; Kiss, T.; Tarantini, S.; Valcarcel Ares, M.N.; Toth, P.; Yabluchanskiy, A.; Conley, S.M.; Ballabh, P.; Martinez-Lemus, L.A.; et al. IGF-1 deficiency Promotes Pathological Remodeling of Cerebral Arteries: A Potential Mechanism Contributing to the Pathogenesis of Intracerebral Hemorrhages in Aging. J. Gerontol. A Biol. Sci. Med. Sci. 2018, 74, 446–454. [Google Scholar] [CrossRef]
  57. Fulop, G.A.; Kiss, T.; Tarantini, S.; Balasubramanian, P.; Yabluchanskiy, A.; Farkas, E.; Bari, F.; Ungvari, Z.; Csiszar, A. Nrf2 deficiency in aged mice exacerbates cellular senescence promoting cerebrovascular inflammation. Geroscience 2018, 40, 513–521. [Google Scholar] [CrossRef]
  58. Tarantini, S.; Tucsek, Z.; Valcarcel-Ares, M.N.; Toth, P.; Gautam, T.; Giles, C.B.; Ballabh, P.; Wei, J.Y.; Wren, J.D.; Ashpole, N.M.; et al. Circulating IGF-1 deficiency exacerbates hypertension-induced microvascular rarefaction in the mouse hippocampus and retrosplenial cortex: Implications for cerebromicrovascular and brain aging. Age 2016, 38, 273–289. [Google Scholar] [CrossRef]
  59. Toth, P.; Tarantini, S.; Ashpole, N.M.; Tucsek, Z.; Milne, G.L.; Valcarcel-Ares, N.M.; Menyhart, A.; Farkas, E.; Sonntag, W.E.; Csiszar, A.; et al. IGF-1 deficiency impairs neurovascular coupling in mice: Implications for cerebromicrovascular aging. Aging Cell 2015, 14, 1034–1044. [Google Scholar] [CrossRef]
  60. Springo, Z.; Tarantini, S.; Toth, P.; Tucsek, Z.; Koller, A.; Sonntag, W.E.; Csiszar, A.; Ungvari, Z. Aging Exacerbates Pressure-Induced Mitochondrial Oxidative Stress in Mouse Cerebral Arteries. J. Gerontol. A Biol. Sci. Med. Sci. 2015, 70, 1355–1359. [Google Scholar] [CrossRef]
  61. Le Couteur, D.G.; Solon-Biet, S.; Cogger, V.C.; Mitchell, S.J.; Senior, A.; de Cabo, R.; Raubenheimer, D.; Simpson, S.J. The impact of low-protein high-carbohydrate diets on aging and lifespan. Cell. Mol. Life Sci. 2016, 73, 1237–1252. [Google Scholar] [CrossRef]
  62. Minor, R.K.; Allard, J.S.; Younts, C.M.; Ward, T.M.; de Cabo, R. Dietary interventions to extend life span and health span based on calorie restriction. J. Gerontol. A Biol. Sci. Med. Sci. 2010, 65, 695–703. [Google Scholar] [CrossRef] [PubMed]
  63. 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]
  64. 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] [PubMed]
  65. Montegut, L.; de Cabo, R.; Zitvogel, L.; Kroemer, G. Science-Driven Nutritional Interventions for the Prevention and Treatment of Cancer. Cancer Discov. 2022, 12, 2258–2279. [Google Scholar] [CrossRef]
  66. Guo, J.; Schupf, N.; Cruz, E.; Stern, Y.; Mayeux, R.P.; Gu, Y. Association Between Mediterranean Diet and Functional Status in Older Adults: A Longitudinal Study Based on the Washington Heights-Inwood Columbia Aging Project. J. Gerontol. A Biol. Sci. Med. Sci. 2022, 77, 1873–1881. [Google Scholar] [CrossRef]
  67. Guralnik, J.M.; Sternberg, A.L.; Mitchell, C.M.; Blackford, A.L.; Schrack, J.; Wanigatunga, A.A.; Michos, E.; Juraschek, S.P.; Szanton, S.; Kalyani, R.; et al. Effects of Vitamin D on Physical Function: Results from the STURDY Trial. J. Gerontol. A Biol. Sci. Med. Sci. 2022, 77, 1585–1592. [Google Scholar] [CrossRef]
  68. Maroto-Rodriguez, J.; Delgado-Velandia, M.; Ortola, R.; Garcia-Esquinas, E.; Martinez-Gomez, D.; Struijk, E.A.; Lopez-Garcia, E.; Rodriguez-Artalejo, F.; Sotos-Prieto, M. A Mediterranean Lifestyle and Frailty Incidence in Older Adults: The Seniors-ENRICA-1 Cohort. J. Gerontol. A Biol. Sci. Med. Sci. 2022, 77, 1845–1852. [Google Scholar] [CrossRef]
  69. Promislow, D.E.L. A New Concept in Diet Restriction Is Cleaning Up! J. Gerontol. A Biol. Sci. Med. Sci. 2021, 76, 599–600. [Google Scholar] [CrossRef]
  70. Ramaker, M.E.; Corcoran, D.L.; Apsley, A.T.; Kobor, M.S.; Kraus, V.B.; Kraus, W.E.; Lin, D.T.S.; Orenduff, M.C.; Pieper, C.F.; Waziry, R.; et al. Epigenome-wide Association Study Analysis of Calorie Restriction in Humans, CALERIETM Trial Analysis. J. Gerontol. A Biol. Sci. Med. Sci. 2022, 77, 2395–2401. [Google Scholar] [CrossRef]
  71. Sotos-Prieto, M.; Ortola, R.; Lopez-Garcia, E.; Rodriguez-Artalejo, F.; Garcia-Esquinas, E. Adherence to the Mediterranean Diet and Physical Resilience in Older Adults: The Seniors-ENRICA Cohort. J. Gerontol. A Biol. Sci. Med. Sci. 2021, 76, 505–512. [Google Scholar] [CrossRef]
  72. Stephen, R.; Ngandu, T.; Liu, Y.; Peltonen, M.; Antikainen, R.; Kemppainen, N.; Laatikainen, T.; Lotjonen, J.; Rinne, J.; Strandberg, T.; et al. Change in CAIDE Dementia Risk Score and Neuroimaging Biomarkers During a 2-Year Multidomain Lifestyle Randomized Controlled Trial: Results of a Post-Hoc Subgroup Analysis. J. Gerontol. A Biol. Sci. Med. Sci. 2021, 76, 1407–1414. [Google Scholar] [CrossRef] [PubMed]
  73. Talegawkar, S.A.; Jin, Y.; Xue, Q.L.; Tanaka, T.; Simonsick, E.M.; Tucker, K.L.; Ferrucci, L. Dietary Pattern Trajectories in Middle Age and Physical Function in Older Age. J. Gerontol. A Biol. Sci. Med. Sci. 2021, 76, 513–519. [Google Scholar] [CrossRef] [PubMed]
  74. Tessier, A.J.; Presse, N.; Rahme, E.; Ferland, G.; Bherer, L.; Chevalier, S. Milk, Yogurt, and Cheese Intake Is Positively Associated with Cognitive Executive Functions in Older Adults of the Canadian Longitudinal Study on Aging. J. Gerontol. A Biol. Sci. Med. Sci. 2021, 76, 2223–2231. [Google Scholar] [CrossRef] [PubMed]
  75. Power, R.; Prado-Cabrero, A.; Mulcahy, R.; Howard, A.; Nolan, J.M. The Role of Nutrition for the Aging Population: Implications for Cognition and Alzheimer’s Disease. Annu. Rev. Food Sci. Technol. 2019, 10, 619–639. [Google Scholar] [CrossRef]
  76. Xi, Y.; Zhang, Y.; Zhou, Y.; Liu, Q.; Chen, X.; Liu, X.; Grune, T.; Shi, L.; Hou, M.; Liu, Z. Effects of methionine intake on cognitive function in mild cognitive impairment patients and APP/PS1 Alzheimer’s Disease model mice: Role of the cystathionine-beta-synthase/H(2)S pathway. Redox Biol. 2023, 59, 102595. [Google Scholar] [CrossRef]
  77. Troen, A.M.; Shea-Budgell, M.; Shukitt-Hale, B.; Smith, D.E.; Selhub, J.; Rosenberg, I.H. B-vitamin deficiency causes hyperhomocysteinemia and vascular cognitive impairment in mice. Proc. Natl. Acad. Sci. USA 2008, 105, 12474–12479. [Google Scholar] [CrossRef]
  78. Nuru, M.; Muradashvili, N.; Kalani, A.; Lominadze, D.; Tyagi, N. High methionine, low folate and low vitamin B6/B12 (HM-LF-LV) diet causes neurodegeneration and subsequent short-term memory loss. Metab. Brain Dis. 2018, 33, 1923–1934. [Google Scholar] [CrossRef]
  79. Feng, C.; Jiang, Y.; Li, S.; Ge, Y.; Shi, Y.; Tang, X.; Le, G. Methionine Restriction Improves Cognitive Ability by Alleviating Hippocampal Neuronal Apoptosis through H19 in Middle-Aged Insulin-Resistant Mice. Nutrients 2022, 14, 4503. [Google Scholar] [CrossRef]
  80. Fu, J.; Zhu, Y.; Sun, Y.; Liu, Q.; Duan, H.; Huang, L.; Zhou, D.; Wang, Z.; Zhao, J.; Li, Z.; et al. Circulating Amyloid-beta and Methionine-Related Metabolites to Predict the Risk of Mild Cognitive Impairment: A Nested Case-Control Study. J. Alzheimers Dis. 2022, 90, 389–404. [Google Scholar] [CrossRef]
  81. Ostrakhovitch, E.A.; Tabibzadeh, S. Homocysteine and age-associated disorders. Ageing Res. Rev. 2019, 49, 144–164. [Google Scholar] [CrossRef]
  82. Silberstein, R.B.; Pipingas, A.; Scholey, A.B. Homocysteine Modulates Brain Functional Connectivity in a Memory Retrieval Task. J. Alzheimers Dis. 2022, 90, 199–209. [Google Scholar] [CrossRef] [PubMed]
  83. Yang, Y.; Zeng, Y.; Yuan, S.; Xie, M.; Dong, Y.; Li, J.; He, Q.; Ye, X.; Lv, Y.; Hocher, C.F.; et al. Prevalence and risk factors for hyperhomocysteinemia: A population-based cross-sectional study from Hunan, China. BMJ Open 2021, 11, e048575. [Google Scholar] [CrossRef] [PubMed]
  84. Aissa, A.F.; Tryndyak, V.P.; de Conti, A.; Rita Thomazela Machado, A.; Tuttis, K.; da Silva Machado, C.; Hernandes, L.C.; Wellington da Silva Santos, P.; Mara Serpeloni, J.; Pogribny, I.P.; et al. Epigenetic changes induced in mice liver by methionine-supplemented and methionine-deficient diets. Food Chem. Toxicol. 2022, 163, 112938. [Google Scholar] [CrossRef] [PubMed]
  85. Navik, U.; Sheth, V.G.; Kabeer, S.W.; Tikoo, K. Dietary Supplementation of Methyl Donor l-Methionine Alters Epigenetic Modification in Type 2 Diabetes. Mol. Nutr. Food Res. 2019, 63, e1801401. [Google Scholar] [CrossRef]
  86. Singh, M.; George, A.K.; Eyob, W.; Homme, R.P.; Stansic, D.; Tyagi, S.C. High-methionine diet in skeletal muscle remodeling: Epigenetic mechanism of homocysteine-mediated growth retardation. Can. J. Physiol. Pharmacol. 2021, 99, 56–63. [Google Scholar] [CrossRef]
  87. Troen, A.M.; Lutgens, E.; Smith, D.E.; Rosenberg, I.H.; Selhub, J. The atherogenic effect of excess methionine intake. Proc. Natl. Acad. Sci. USA 2003, 100, 15089–15094. [Google Scholar] [CrossRef]
  88. Caro, P.; Gomez, J.; Sanchez, I.; Naudi, A.; Ayala, V.; Lopez-Torres, M.; Pamplona, R.; Barja, G. Forty percent methionine restriction decreases mitochondrial oxygen radical production and leak at complex I during forward electron flow and lowers oxidative damage to proteins and mitochondrial DNA in rat kidney and brain mitochondria. Rejuvenation Res. 2009, 12, 421–434. [Google Scholar] [CrossRef]
  89. McCampbell, A.; Wessner, K.; Marlatt, M.W.; Wolffe, C.; Toolan, D.; Podtelezhnikov, A.; Yeh, S.; Zhang, R.; Szczerba, P.; Tanis, K.Q.; et al. Induction of Alzheimer’s-like changes in brain of mice expressing mutant APP fed excess methionine. J. Neurochem. 2011, 116, 82–92. [Google Scholar] [CrossRef]
  90. Bagi, Z.; Ungvari, Z.; Koller, A. Xanthine oxidase-derived reactive oxygen species convert flow-induced arteriolar dilation to constriction in hyperhomocysteinemia: Possible role of peroxynitrite. Arterioscler. Thromb. Vasc. Biol. 2002, 22, 28–33. [Google Scholar] [CrossRef]
  91. Bagi, Z.; Ungvari, Z.; Szollar, L.; Koller, A. Flow-Induced Constriction in Arterioles of Hyperhomocysteinemic Rats Is Due to Impaired Nitric Oxide and Enhanced Thromboxane A(2) Mediation. Arterioscler. Thromb. Vasc. Biol. 2001, 21, 233–237. [Google Scholar] [CrossRef]
  92. Ungvari, Z.; Csiszar, A.; Bagi, Z.; Koller, A. Impaired nitric oxide-mediated flow-induced coronary dilation in hyperhomocysteinemia: Morphological and functional evidence for increased peroxynitrite formation. Am. J. Pathol. 2002, 161, 145–153. [Google Scholar] [CrossRef] [PubMed]
  93. Ungvari, Z.; Csiszar, A.; Edwards, J.G.; Kaminski, P.M.; Wolin, M.S.; Kaley, G.; Koller, A. Increased superoxide production in coronary arteries in hyperhomocysteinemia: Role of tumor necrosis factor-alpha, NAD(P)H oxidase, and inducible nitric oxide synthase. Arterioscler. Thromb. Vasc. Biol. 2003, 23, 418–424. [Google Scholar] [CrossRef] [PubMed]
  94. Ungvari, Z.; Pacher, P.; Rischak, K.; Szollar, L.; Koller, A. Dysfunction of nitric oxide mediation in isolated rat arterioles with methionine diet-induced hyperhomocysteinemia. Arterioscler. Thromb. Vasc. Biol. 1999, 19, 1899–1904. [Google Scholar] [CrossRef] [PubMed]
  95. Ungvari, Z.; Sarkadi-Nagy, E.; Bagi, Z.; Szollar, L.; Koller, A. Simultaneously Increased TxA2 Activity in Isolated Arterioles and Platelets of Rats with Hyperhomocysteinemia. Arterioscler. Thromb. Vasc. Biol. 2000, 20, 1203–1208. [Google Scholar] [CrossRef]
  96. Ables, G.P.; Ouattara, A.; Hampton, T.G.; Cooke, D.; Perodin, F.; Augie, I.; Orentreich, D.S. Dietary methionine restriction in mice elicits an adaptive cardiovascular response to hyperhomocysteinemia. Sci. Rep. 2015, 5, 8886. [Google Scholar] [CrossRef]
  97. Ables, G.P.; Perrone, C.E.; Orentreich, D.; Orentreich, N. Methionine-restricted C57BL/6J mice are resistant to diet-induced obesity and insulin resistance but have low bone density. PLoS ONE 2012, 7, e51357. [Google Scholar] [CrossRef]
  98. Barcena, C.; Quiros, P.M.; Durand, S.; Mayoral, P.; Rodriguez, F.; Caravia, X.M.; Marino, G.; Garabaya, C.; Fernandez-Garcia, M.T.; Kroemer, G.; et al. Methionine Restriction Extends Lifespan in Progeroid Mice and Alters Lipid and Bile Acid Metabolism. Cell Rep. 2018, 24, 2392–2403. [Google Scholar] [CrossRef]
  99. Ghosh, S.; Wanders, D.; Stone, K.P.; Van, N.T.; Cortez, C.C.; Gettys, T.W. A systems biology analysis of the unique and overlapping transcriptional responses to caloric restriction and dietary methionine restriction in rats. FASEB J. 2014, 28, 2577–2590. [Google Scholar] [CrossRef]
  100. Gomez, J.; Caro, P.; Sanchez, I.; Naudi, A.; Jove, M.; Portero-Otin, M.; Lopez-Torres, M.; Pamplona, R.; Barja, G. Effect of methionine dietary supplementation on mitochondrial oxygen radical generation and oxidative DNA damage in rat liver and heart. J. Bioenerg. Biomembr. 2009, 41, 309–321. [Google Scholar] [CrossRef]
  101. Hine, C.; Mitchell, J.R. Calorie restriction and methionine restriction in control of endogenous hydrogen sulfide production by the transsulfuration pathway. Exp. Gerontol. 2015, 68, 26–32. [Google Scholar] [CrossRef]
  102. Hipkiss, A.R. On methionine restriction, suppression of mitochondrial dysfunction and aging. Rejuvenation Res. 2008, 11, 685–688. [Google Scholar] [CrossRef] [PubMed]
  103. Johnson, J.E.; Johnson, F.B. Methionine restriction activates the retrograde response and confers both stress tolerance and lifespan extension to yeast, mouse and human cells. PLoS ONE 2014, 9, e97729. [Google Scholar] [CrossRef] [PubMed]
  104. Jove, M.; Ayala, V.; Ramirez-Nunez, O.; Naudi, A.; Cabre, R.; Spickett, C.M.; Portero-Otin, M.; Pamplona, R. Specific lipidome signatures in central nervous system from methionine-restricted mice. J. Proteome Res. 2013, 12, 2679–2689. [Google Scholar] [CrossRef] [PubMed]
  105. 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 2019, 42, 287–297. [Google Scholar] [CrossRef] [PubMed]
  106. Lees, E.K.; Krol, E.; Grant, L.; Shearer, K.; Wyse, C.; Moncur, E.; Bykowska, A.S.; Mody, N.; Gettys, T.W.; Delibegovic, M. Methionine restriction restores a younger metabolic phenotype in adult mice with alterations in fibroblast growth factor 21. Aging Cell 2014, 13, 817–827. [Google Scholar] [CrossRef] [PubMed]
  107. Maddineni, S.; Nichenametla, S.; Sinha, R.; Wilson, R.P.; Richie, J.P., Jr. Methionine restriction affects oxidative stress and glutathione-related redox pathways in the rat. Exp. Biol. Med. 2013, 238, 392–399. [Google Scholar] [CrossRef]
  108. Perrone, C.E.; Malloy, V.L.; Orentreich, D.S.; Orentreich, N. Metabolic adaptations to methionine restriction that benefit health and lifespan in rodents. Exp. Gerontol. 2013, 48, 654–660. [Google Scholar] [CrossRef]
  109. Sanchez-Roman, I.; Barja, G. Regulation of longevity and oxidative stress by nutritional interventions: Role of methionine restriction. Exp. Gerontol. 2013, 48, 1030–1042. [Google Scholar] [CrossRef]
  110. Sanz, A.; Caro, P.; Ayala, V.; Portero-Otin, M.; Pamplona, R.; Barja, G. Methionine restriction decreases mitochondrial oxygen radical generation and leak as well as oxidative damage to mitochondrial DNA and proteins. FASEB J. 2006, 20, 1064–1073. [Google Scholar] [CrossRef]
  111. Stone, K.P.; Wanders, D.; Orgeron, M.; Cortez, C.C.; Gettys, T.W. Mechanisms of increased in vivo insulin sensitivity by dietary methionine restriction in mice. Diabetes 2014, 63, 3721–3733. [Google Scholar] [CrossRef]
  112. Sun, L.; Sadighi Akha, A.A.; Miller, R.A.; Harper, J.M. Life-span extension in mice by preweaning food restriction and by methionine restriction in middle age. J. Gerontol. A Biol. Sci. Med. Sci. 2009, 64, 711–722. [Google Scholar] [CrossRef] [PubMed]
  113. Kouvari, M.; Tyrovolas, S.; Panagiotakos, D.B. Red meat consumption and healthy ageing: A review. Maturitas 2016, 84, 17–24. [Google Scholar] [CrossRef] [PubMed]
  114. Daniel, C.R.; Cross, A.J.; Koebnick, C.; Sinha, R. Trends in meat consumption in the USA. Public Health Nutr. 2011, 14, 575–583. [Google Scholar] [CrossRef] [PubMed]
  115. Al-Shaar, L.; Satija, A.; Wang, D.D.; Rimm, E.B.; Smith-Warner, S.A.; Stampfer, M.J.; Hu, F.B.; Willett, W.C. Red meat intake and risk of coronary heart disease among US men: Prospective cohort study. BMJ 2020, 371, m4141. [Google Scholar] [CrossRef] [PubMed]
  116. Dong, X.; Zhuang, Z.; Zhao, Y.; Song, Z.; Xiao, W.; Wang, W.; Li, Y.; Huang, N.; Jia, J.; Liu, Z.; et al. Unprocessed Red Meat and Processed Meat Consumption, Plasma Metabolome, and Risk of Ischemic Heart Disease: A Prospective Cohort Study of UK Biobank. J. Am. Heart Assoc. 2023, 12, e027934. [Google Scholar] [CrossRef]
  117. Fan, B.; Huang, X.; Zhao, J.V. Exploration of Metabolic Biomarkers Linking Red Meat Consumption to Ischemic Heart Disease Mortality in the UK Biobank. Nutrients 2023, 15, 1865. [Google Scholar] [CrossRef]
  118. Johnston, B.C.; Zeraatkar, D.; Han, M.A.; Vernooij, R.W.M.; Valli, C.; El Dib, R.; Marshall, C.; Stover, P.J.; Fairweather-Taitt, S.; Wojcik, G.; et al. Unprocessed Red Meat and Processed Meat Consumption: Dietary Guideline Recommendations From the Nutritional Recommendations (NutriRECS) Consortium. Ann. Intern. Med. 2019, 171, 756–764. [Google Scholar] [CrossRef]
  119. Mota, J.O.; Guillou, S.; Pierre, F.; Membre, J.M. Public health risk-benefit assessment of red meat in France: Current consumption and alternative scenarios. Food Chem. Toxicol. 2021, 149, 111994. [Google Scholar] [CrossRef]
  120. Shi, W.; Huang, X.; Schooling, C.M.; Zhao, J.V. Red meat consumption, cardiovascular diseases, and diabetes: A systematic review and meta-analysis. Eur. Heart J. 2023, 44, 2626–2635. [Google Scholar] [CrossRef]
  121. Singh, B.; Khan, A.A.; Anamika, F.; Munjal, R.; Munjal, J.; Jain, R. Red Meat Consumption and its Relationship with Cardiovascular Health: A Review of Pathophysiology and Literature. Cardiol. Rev. 2023. [Google Scholar] [CrossRef]
  122. Sun, L.; Yuan, J.L.; Chen, Q.C.; Xiao, W.K.; Ma, G.P.; Liang, J.H.; Chen, X.K.; Wang, S.; Zhou, X.X.; Wu, H.; et al. Red meat consumption and risk for dyslipidaemia and inflammation: A systematic review and meta-analysis. Front. Cardiovasc. Med. 2022, 9, 996467. [Google Scholar] [CrossRef] [PubMed]
  123. Wang, M.; Ma, H.; Song, Q.; Zhou, T.; Hu, Y.; Heianza, Y.; Manson, J.E.; Qi, L. Red meat consumption and all-cause and cardiovascular mortality: Results from the UK Biobank study. Eur. J. Nutr. 2022, 61, 2543–2553. [Google Scholar] [CrossRef] [PubMed]
  124. Zeraatkar, D.; Guyatt, G.H.; Alonso-Coello, P.; Bala, M.M.; Rabassa, M.; Han, M.A.; Vernooij, R.W.M.; Valli, C.; El Dib, R.; Johnston, B.C. Red and Processed Meat Consumption and Risk for All-Cause Mortality and Cardiometabolic Outcomes. Ann. Intern. Med. 2020, 172, 511–512. [Google Scholar] [CrossRef] [PubMed]
  125. Zeraatkar, D.; Johnston, B.C.; Bartoszko, J.; Cheung, K.; Bala, M.M.; Valli, C.; Rabassa, M.; Sit, D.; Milio, K.; Sadeghirad, B.; et al. Effect of Lower Versus Higher Red Meat Intake on Cardiometabolic and Cancer Outcomes: A Systematic Review of Randomized Trials. Ann. Intern. Med. 2019, 171, 721–731. [Google Scholar] [CrossRef] [PubMed]
  126. Zhang, J.; Hayden, K.; Jackson, R.; Schutte, R. Association of red and processed meat consumption with cardiovascular morbidity and mortality in participants with and without obesity: A prospective cohort study. Clin. Nutr. 2021, 40, 3643–3649. [Google Scholar] [CrossRef] [PubMed]
  127. Saneei, P.; Saadatnia, M.; Shakeri, F.; Beykverdi, M.; Keshteli, A.H.; Esmaillzadeh, A. A case-control study on red meat consumption and risk of stroke among a group of Iranian adults. Public Health Nutr. 2015, 18, 1084–1090. [Google Scholar] [CrossRef]
  128. Amiano, P.; Chamosa, S.; Etxezarreta, N.; Arriola, L.; Sanchez, M.J.; Ardanaz, E.; Molina-Montes, E.; Chirlaque, M.D.; Moreno-Iribas, C.; Huerta, J.M.; et al. Unprocessed red meat and processed meat consumption and risk of stroke in the Spanish cohort of the European Prospective Investigation into Cancer and Nutrition (EPIC). Eur. J. Clin. Nutr. 2016, 70, 313–319. [Google Scholar] [CrossRef]
  129. Chen, F.; Hu, W.; Chen, S.; Si, A.; Zhang, Y.; Ma, J. Stroke mortality attributable to high red meat intake in China and South Korea: An age-period-cohort and joinpoint analysis. Front. Nutr. 2022, 9, 921592. [Google Scholar] [CrossRef]
  130. Chen, G.C.; Lv, D.B.; Pang, Z.; Liu, Q.F. Red and processed meat consumption and risk of stroke: A meta-analysis of prospective cohort studies. Eur. J. Clin. Nutr. 2013, 67, 91–95. [Google Scholar] [CrossRef]
  131. Kaluza, J.; Wolk, A.; Larsson, S.C. Red meat consumption and risk of stroke: A meta-analysis of prospective studies. Stroke 2012, 43, 2556–2560. [Google Scholar] [CrossRef]
  132. Kim, K.; Hyeon, J.; Lee, S.A.; Kwon, S.O.; Lee, H.; Keum, N.; Lee, J.K.; Park, S.M. Role of Total, Red, Processed, and White Meat Consumption in Stroke Incidence and Mortality: A Systematic Review and Meta-Analysis of Prospective Cohort Studies. J. Am. Heart Assoc. 2017, 6, e005983. [Google Scholar] [CrossRef] [PubMed]
  133. Larsson, S.C.; Virtamo, J.; Wolk, A. Red meat consumption and risk of stroke in Swedish men. Am. J. Clin. Nutr. 2011, 94, 417–421. [Google Scholar] [CrossRef] [PubMed]
  134. Larsson, S.C.; Virtamo, J.; Wolk, A. Red meat consumption and risk of stroke in Swedish women. Stroke 2011, 42, 324–329. [Google Scholar] [CrossRef] [PubMed]
  135. Micha, R.; Wallace, S.K.; Mozaffarian, D. Red and processed meat consumption and risk of incident coronary heart disease, stroke, and diabetes mellitus: A systematic review and meta-analysis. Circulation 2010, 121, 2271–2283. [Google Scholar] [CrossRef]
  136. Veno, S.K.; Bork, C.S.; Jakobsen, M.U.; Lundbye-Christensen, S.; Bach, F.W.; McLennan, P.L.; Tjonneland, A.; Schmidt, E.B.; Overvad, K. Substitution of Fish for Red Meat or Poultry and Risk of Ischemic Stroke. Nutrients 2018, 10, 1648. [Google Scholar] [CrossRef]
  137. Yang, C.; Pan, L.; Sun, C.; Xi, Y.; Wang, L.; Li, D. Red Meat Consumption and the Risk of Stroke: A Dose-Response Meta-analysis of Prospective Cohort Studies. J. Stroke Cerebrovasc. Dis. 2016, 25, 1177–1186. [Google Scholar] [CrossRef]
  138. Morris, M.S.; Jacques, P.F.; Rosenberg, I.H.; Selhub, J.; National, H.; Nutrition Examination, S. Hyperhomocysteinemia associated with poor recall in the third National Health and Nutrition Examination Survey. Am. J. Clin. Nutr. 2001, 73, 927–933. [Google Scholar] [CrossRef]
  139. Selhub, J. The many facets of hyperhomocysteinemia: Studies from the Framingham cohorts. J. Nutr. 2006, 136, 1726S–1730S. [Google Scholar] [CrossRef]
  140. Holmen, M.; Hvas, A.M.; Arendt, J.F.H. Hyperhomocysteinemia and Ischemic Stroke: A Potential Dose-Response Association-A Systematic Review and Meta-analysis. TH Open 2021, 5, e420–e437. [Google Scholar] [CrossRef]
  141. Kumral, E.; Saruhan, G.; Aktert, D.; Orman, M. Association of Hyperhomocysteinemia with Stroke Recurrence after Initial Stroke. J. Stroke Cerebrovasc. Dis. 2016, 25, 2047–2054. [Google Scholar] [CrossRef]
  142. Lu, Z.H.; Li, J.; Li, X.L.; Ding, M.; Mao, C.J.; Zhu, X.Y.; Liu, C.F. Hypertension with Hyperhomocysteinemia Increases the Risk of Early Cognitive Impairment after First-Ever Ischemic Stroke. Eur. Neurol. 2019, 82, 75–85. [Google Scholar] [CrossRef] [PubMed]
  143. Mizuno, T.; Hoshino, T.; Ishizuka, K.; Toi, S.; Takahashi, S.; Wako, S.; Arai, S.; Kitagawa, K. Hyperhomocysteinemia Increases Vascular Risk in Stroke Patients with Chronic Kidney Disease. J. Atheroscler. Thromb. 2022, 30, 1198–1209. [Google Scholar] [CrossRef] [PubMed]
  144. Pang, H.; Fu, Q.; Cao, Q.; Hao, L.; Zong, Z. Sex differences in risk factors for stroke in patients with hypertension and hyperhomocysteinemia. Sci. Rep. 2019, 9, 14313. [Google Scholar] [CrossRef] [PubMed]
  145. Poddar, R. Hyperhomocysteinemia is an emerging comorbidity in ischemic stroke. Exp. Neurol. 2021, 336, 113541. [Google Scholar] [CrossRef] [PubMed]
  146. Rawashdeh, S.I.; Al-Mistarehi, A.H.; Yassin, A.; Rabab’ah, W.; Skaff, H.; Ibdah, R. A Concurrent Ischemic Stroke, Myocardial Infarction, and Aortic Thrombi in a Young Patient with Hyperhomocysteinemia: A Case Report. Int. Med. Case Rep. J. 2020, 13, 581–590. [Google Scholar] [CrossRef]
  147. Wood, J.M.; Decker, H.; Hartmann, H.; Chavan, B.; Rokos, H.; Spencer, J.D.; Hasse, S.; Thornton, M.J.; Shalbaf, M.; Paus, R.; et al. Senile hair graying: H2O2-mediated oxidative stress affects human hair color by blunting methionine sulfoxide repair. FASEB J. 2009, 23, 2065–2075. [Google Scholar] [CrossRef]
  148. Refsum, H.; Ueland, P.M.; Nygard, O.; Vollset, S.E. Homocysteine and cardiovascular disease. Annu. Rev. Med. 1998, 49, 31–62. [Google Scholar] [CrossRef]
  149. Oz, H.S.; Chen, T.S.; Neuman, M. Methionine deficiency and hepatic injury in a dietary steatohepatitis model. Dig. Dis. Sci. 2008, 53, 767–776. [Google Scholar] [CrossRef]
  150. Chao, C.L.; Kuo, T.L.; Lee, Y.T. Effects of methionine-induced hyperhomocysteinemia on endothelium- dependent vasodilation and oxidative status in healthy adults. Circulation 2000, 101, 485–490. [Google Scholar] [CrossRef]
  151. Di Minno, M.N.; Pezzullo, S.; Palmieri, V.; Coppola, A.; D’Angelo, A.; Sampietro, F.; Cavalca, V.; Tremoli, E.; Di Minno, G. Genotype-independent in vivo oxidative stress following a methionine loading test: Maximal platelet activation in subjects with early-onset thrombosis. Thromb. Res. 2011, 128, e43–e48. [Google Scholar] [CrossRef]
  152. Yalcinkaya-Demirsoz, S.; Depboylu, B.; Dogru-Abbasoglu, S.; Unlucerci, Y.; Uysal, M. Effects of high methionine diet on oxidative stress in serum, apo-B containing lipoproteins, heart, and aorta in rabbits. Ann. Clin. Lab. Sci. 2009, 39, 386–391. [Google Scholar] [PubMed]
  153. Zhang, R.; Ma, J.; Xia, M.; Zhu, H.; Ling, W. Mild hyperhomocysteinemia induced by feeding rats diets rich in methionine or deficient in folate promotes early atherosclerotic inflammatory processes. J. Nutr. 2004, 134, 825–830. [Google Scholar] [CrossRef] [PubMed]
  154. 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, in press. [Google Scholar] [CrossRef] [PubMed]
  155. Ungvari, Z.; Bailey-Downs, L.; Gautam, T.; Sosnowska, D.; Wang, M.; Monticone, R.E.; Telljohann, R.; Pinto, J.T.; de Cabo, R.; Sonntag, W.E.; et al. Age-associated vascular oxidative stress, Nrf2 dysfunction and NF-kB activation in the non-human primate Macaca mulatta. J. Gerontol. A Biol. Sci. Med. Sci. 2011, 66, 866–875. [Google Scholar] [CrossRef] [PubMed]
  156. Ungvari, Z.; Bailey-Downs, L.; Sosnowska, D.; Gautam, T.; Koncz, P.; Losonczy, G.; Ballabh, P.; de Cabo, R.; Sonntag, W.E.; Csiszar, A. Vascular oxidative stress in aging: A homeostatic failure due to dysregulation of Nrf2-mediated antioxidant response. Am. J. Physiol. Heart Circ. Physiol. 2011, 301, H363–H372. [Google Scholar] [CrossRef]
  157. Ungvari, Z.; Tarantini, S.; Nyul-Toth, A.; Kiss, T.; Yabluchanskiy, A.; Csipo, T.; Balasubramanian, P.; Lipecz, A.; Benyo, Z.; Csiszar, A. Nrf2 dysfunction and impaired cellular resilience to oxidative stressors in the aged vasculature: From increased cellular senescence to the pathogenesis of age-related vascular diseases. Geroscience 2019, 41, 727–738. [Google Scholar] [CrossRef]
  158. Addabbo, F.; Ratliff, B.; Park, H.C.; Kuo, M.C.; Ungvari, Z.; Csiszar, A.; Krasnikov, B.; Sodhi, K.; Zhang, F.; Nasjletti, A.; et al. The Krebs cycle and mitochondrial mass are early victims of endothelial dysfunction: Proteomic approach. Am. J. Pathol. 2009, 174, 34–43. [Google Scholar] [CrossRef]
  159. Csiszar, A.; Gautam, T.; Sosnowska, D.; Tarantini, S.; Banki, E.; Tucsek, Z.; Toth, P.; Losonczy, G.; Koller, A.; Reglodi, D.; et al. Caloric restriction confers persistent anti-oxidative, pro-angiogenic, and anti-inflammatory effects and promotes anti-aging miRNA expression profile in cerebromicrovascular endothelial cells of aged rats. Am. J. Physiol. Heart Circ. Physiol. 2014, 307, H292–H306. [Google Scholar] [CrossRef]
  160. Csiszar, A.; Labinskyy, N.; Orosz, Z.; Ungvari, Z. Altered mitochondrial energy metabolism may play a role in vascular aging. Med. Hypotheses 2006, 67, 904–908. [Google Scholar] [CrossRef]
  161. Csiszar, A.; Sosnowska, D.; Wang, M.; Lakatta, E.G.; Sonntag, W.E.; Ungvari, Z. Age-associated proinflammatory secretory phenotype in vascular smooth muscle cells from the non-human primate Macaca mulatta: Reversal by resveratrol treatment. J. Gerontol. A Biol. Sci. Med. Sci. 2012, 67, 811–820. [Google Scholar] [CrossRef]
  162. Csiszar, A.; Wang, M.; Lakatta, E.G.; Ungvari, Z.I. Inflammation and endothelial dysfunction during aging: Role of NF-κB. J. Appl. Physiol. 2008, 105, 1333–1341. [Google Scholar] [CrossRef] [PubMed]
  163. Dai, D.F.; Rabinovitch, P.S.; Ungvari, Z. Mitochondria and cardiovascular aging. Circ. Res. 2012, 110, 1109–1124. [Google Scholar] [CrossRef] [PubMed]
  164. Ungvari, Z.; Orosz, Z.; Labinskyy, N.; Rivera, A.; Xiangmin, Z.; Smith, K.; Csiszar, A. Increased mitochondrial H2O2 production promotes endothelial NF-kappaB activation in aged rat arteries. Am. J. Physiol. Heart Circ. Physiol. 2007, 293, H37–H47. [Google Scholar] [CrossRef]
  165. Dayal, S.; Arning, E.; Bottiglieri, T.; Boger, R.H.; Sigmund, C.D.; Faraci, F.M.; Lentz, S.R. Cerebral vascular dysfunction mediated by superoxide in hyperhomocysteinemic mice. Stroke 2004, 35, 1957–1962. [Google Scholar] [CrossRef] [PubMed]
  166. Song, S.; Kertowidjojo, E.; Ojaimi, C.; Martin-Fernandez, B.; Kandhi, S.; Wolin, M.; Hintze, T.H. Long-term methionine-diet induced mild hyperhomocysteinemia associated cardiac metabolic dysfunction in multiparous rats. Physiol. Rep. 2015, 3, e12292. [Google Scholar] [CrossRef] [PubMed]
  167. Ventura, E.; Durant, R.; Jaussent, A.; Picot, M.C.; Morena, M.; Badiou, S.; Dupuy, A.M.; Jeandel, C.; Cristol, J.P. Homocysteine and inflammation as main determinants of oxidative stress in the elderly. Free. Radic. Biol. Med. 2009, 46, 737–744. [Google Scholar] [CrossRef] [PubMed]
  168. 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]
  169. 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]
  170. Bailey-Downs, L.C.; Mitschelen, M.; Sosnowska, D.; Toth, P.; Pinto, J.T.; Ballabh, P.; Valcarcel-Ares, M.N.; Farley, J.; Koller, A.; Henthorn, J.C.; et al. Liver-specific knockdown of IGF-1 decreases vascular oxidative stress resistance by impairing the Nrf2-dependent antioxidant response: A novel model of vascular aging. J. Gerontol. Biol. Med. Sci. 2012, 67, 313–329. [Google Scholar] [CrossRef]
  171. Csiszar, A.; Labinskyy, N.; Smith, K.; Rivera, A.; Orosz, Z.; Ungvari, Z. Vasculoprotective effects of anti-TNFalfa treatment in aging. Am. J. Pathol. 2007, 170, 388–698. [Google Scholar] [CrossRef]
  172. Zhang, H.; Zhang, J.; Ungvari, Z.; Zhang, C. Resveratrol improves endothelial function: Role of TNF{alpha} and vascular oxidative stress. Arterioscler. Thromb. Vasc. Biol. 2009, 29, 1164–1171. [Google Scholar] [CrossRef] [PubMed]
  173. Sudduth, T.L.; Powell, D.K.; Smith, C.D.; Greenstein, A.; Wilcock, D.M. Induction of hyperhomocysteinemia models vascular dementia by induction of cerebral microhemorrhages and neuroinflammation. J. Cereb. Blood Flow Metab. 2013, 33, 708–715. [Google Scholar] [CrossRef] [PubMed]
  174. Chen, S.; Dong, Z.; Zhao, Y.; Sai, N.; Wang, X.; Liu, H.; Huang, G.; Zhang, X. Homocysteine induces mitochondrial dysfunction involving the crosstalk between oxidative stress and mitochondrial pSTAT3 in rat ischemic brain. Sci. Rep. 2017, 7, 6932. [Google Scholar] [CrossRef]
  175. Kaplan, P.; Tatarkova, Z.; Sivonova, M.K.; Racay, P.; Lehotsky, J. Homocysteine and Mitochondria in Cardiovascular and Cerebrovascular Systems. Int. J. Mol. Sci. 2020, 21, 7698. [Google Scholar] [CrossRef] [PubMed]
  176. Mahaman, Y.A.R.; Huang, F.; Wu, M.; Wang, Y.; Wei, Z.; Bao, J.; Salissou, M.T.M.; Ke, D.; Wang, Q.; Liu, R.; et al. Moringa Oleifera Alleviates Homocysteine-Induced Alzheimer’s Disease-Like Pathology and Cognitive Impairments. J. Alzheimers Dis. 2018, 63, 1141–1159. [Google Scholar] [CrossRef] [PubMed]
  177. Moretti, R.; Giuffre, M.; Caruso, P.; Gazzin, S.; Tiribelli, C. Homocysteine in Neurology: A Possible Contributing Factor to Small Vessel Disease. Int. J. Mol. Sci. 2021, 22, 2051. [Google Scholar] [CrossRef] [PubMed]
  178. Yin, Y.L.; Chen, Y.; Ren, F.; Wang, L.; Zhu, M.L.; Lu, J.X.; Wang, Q.Q.; Lu, C.B.; Liu, C.; Bai, Y.P.; et al. Nitrosative stress induced by homocysteine thiolactone drives vascular cognitive impairments via GTP cyclohydrolase 1 S-nitrosylation in vivo. Redox Biol. 2022, 58, 102540. [Google Scholar] [CrossRef]
  179. Cuervo, A.M.; Huffman, D.M.; Vijg, J.; Milman, S.; Singh, R.; Barzilai, N. Einstein-Nathan Shock Center: Translating the hallmarks of aging to extend human health span. Geroscience 2021, 43, 2167–2182. [Google Scholar] [CrossRef]
  180. Franceschi, C.; Campisi, J. Chronic inflammation (inflammaging) and its potential contribution to age-associated diseases. J. Gerontol. A Biol. Sci. Med. Sci. 2014, 69 (Suppl. S1), S4–S9. [Google Scholar] [CrossRef]
  181. Meszaros, A.; Molnar, K.; Nogradi, B.; Hernadi, Z.; Nyul-Toth, A.; Wilhelm, I.; Krizbai, I.A. Neurovascular Inflammaging in Health and Disease. Cells 2020, 9, 1614. [Google Scholar] [CrossRef]
  182. Royce, G.H.; Brown-Borg, H.M.; Deepa, S.S. The potential role of necroptosis in inflammaging and aging. Geroscience 2019, 41, 795–811. [Google Scholar] [CrossRef] [PubMed]
  183. Ungvari, Z.; Tarantini, S.; Sorond, F.; Merkely, B.; Csiszar, A. Mechanisms of Vascular Aging, A Geroscience Perspective: JACC Focus Seminar. J. Am. Coll. Cardiol. 2020, 75, 931–941. [Google Scholar] [CrossRef] [PubMed]
  184. Gomez, C.R. Role of heat shock proteins in aging and chronic inflammatory diseases. Geroscience 2021, 43, 2515–2532. [Google Scholar] [CrossRef] [PubMed]
  185. Trial, J.; Diaz Lankenau, R.; Angelini, A.; Tovar Perez, J.E.; Taffet, G.E.; Entman, M.L.; Cieslik, K.A. Treatment with a DC-SIGN ligand reduces macrophage polarization and diastolic dysfunction in the aging female but not male mouse hearts. Geroscience 2021, 43, 881–899. [Google Scholar] [CrossRef]
  186. Mahalakshmi, A.M.; Ray, B.; Tuladhar, S.; Bhat, A.; Bishir, M.; Bolla, S.R.; Yang, J.; Essa, M.M.; Chidambaram, S.B.; Guillemin, G.J.; et al. Sleep, brain vascular health and ageing. Geroscience 2020, 42, 1257–1283. [Google Scholar] [CrossRef] [PubMed]
  187. Huffman, D.M.; Csiszar, A.; Ungvari, Z. Heterochronic blood exchange attenuates age-related neuroinflammation and confers cognitive benefits: Do microvascular protective effects play a role? Geroscience 2021, 43, 111–113. [Google Scholar] [CrossRef]
  188. Mehdipour, M.; Mehdipour, T.; Skinner, C.M.; Wong, N.; Liu, C.; Chen, C.C.; Jeon, O.H.; Zuo, Y.; Conboy, M.J.; Conboy, I.M. Plasma dilution improves cognition and attenuates neuroinflammation in old mice. Geroscience 2021, 43, 1–18. [Google Scholar] [CrossRef]
  189. Thadathil, N.; Nicklas, E.H.; Mohammed, S.; Lewis, T.L., Jr.; Richardson, A.; Deepa, S.S. Necroptosis increases with age in the brain and contributes to age-related neuroinflammation. Geroscience 2021, 43, 2345–2361. [Google Scholar] [CrossRef]
  190. Towner, R.A.; Gulej, R.; Zalles, M.; Saunders, D.; Smith, N.; Lerner, M.; Morton, K.A.; Richardson, A. Rapamycin restores brain vasculature, metabolism, and blood-brain barrier in an inflammaging model. Geroscience 2021, 43, 563–578. [Google Scholar] [CrossRef]
  191. Dorigatti, A.O.; Riordan, R.; Yu, Z.; Ross, G.; Wang, R.; Reynolds-Lallement, N.; Magnusson, K.; Galvan, V.; Perez, V.I. Brain cellular senescence in mouse models of Alzheimer’s disease. Geroscience 2022, 44, 1157–1168. [Google Scholar] [CrossRef]
  192. Ritzel, R.M.; Li, Y.; Lei, Z.; Carter, J.; He, J.; Choi, H.M.C.; Khan, N.; Li, H.; Allen, S.; Lipinski, M.M.; et al. Functional and transcriptional profiling of microglial activation during the chronic phase of TBI identifies an age-related driver of poor outcome in old mice. Geroscience 2022, 44, 1407–1440. [Google Scholar] [CrossRef]
  193. Cribb, L.; Hodge, A.M.; Yu, C.; Li, S.X.; English, D.R.; Makalic, E.; Southey, M.C.; Milne, R.L.; Giles, G.G.; Dugue, P.A. Inflammation and Epigenetic Aging Are Largely Independent Markers of Biological Aging and Mortality. J. Gerontol. A Biol. Sci. Med. Sci. 2022, 77, 2378–2386. [Google Scholar] [CrossRef] [PubMed]
  194. Dugue, P.A.; Hodge, A.M.; Ulvik, A.; Ueland, P.M.; Midttun, O.; Rinaldi, S.; MacInnis, R.J.; Li, S.X.; Meyer, K.; Navionis, A.S.; et al. Association of Markers of Inflammation, the Kynurenine Pathway and B Vitamins with Age and Mortality, and a Signature of Inflammaging. J. Gerontol. A Biol. Sci. Med. Sci. 2022, 77, 826–836. [Google Scholar] [CrossRef] [PubMed]
  195. Xu, Y.; Tian, Y.; Wei, H.J.; Dong, J.F.; Zhang, J.N. Methionine diet-induced hyperhomocysteinemia accelerates cerebral aneurysm formation in rats. Neurosci. Lett. 2011, 494, 139–144. [Google Scholar] [CrossRef] [PubMed]
  196. Bagi, Z.; Kroenke, C.D.; Fopiano, K.A.; Tian, Y.; Filosa, J.A.; Sherman, L.S.; Larson, E.B.; Keene, C.D.; Degener O’Brien, K.; Adeniyi, P.A.; et al. Association of cerebral microvascular dysfunction and white matter injury in Alzheimer’s disease. Geroscience 2022, 44, 1–14. [Google Scholar] [CrossRef]
  197. Custodero, C.; Ciavarella, A.; Panza, F.; Gnocchi, D.; Lenato, G.M.; Lee, J.; Mazzocca, A.; Sabba, C.; Solfrizzi, V. Role of inflammatory markers in the diagnosis of vascular contributions to cognitive impairment and dementia: A systematic review and meta-analysis. Geroscience 2022, 44, 1373–1392. [Google Scholar] [CrossRef]
  198. Sanfilippo, C.; Castrogiovanni, P.; Vinciguerra, M.; Imbesi, R.; Ulivieri, M.; Fazio, F.; Blennow, K.; Zetterberg, H.; Di Rosa, M. A sex-stratified analysis of neuroimmune gene expression signatures in Alzheimer’s disease brains. Geroscience 2023, 45, 523–541. [Google Scholar] [CrossRef]
  199. Wan, Y.W.; Al-Ouran, R.; Mangleburg, C.G.; Perumal, T.M.; Lee, T.V.; Allison, K.; Swarup, V.; Funk, C.C.; Gaiteri, C.; Allen, M.; et al. Meta-Analysis of the Alzheimer’s Disease Human Brain Transcriptome and Functional Dissection in Mouse Models. Cell Rep. 2020, 32, 107908. [Google Scholar] [CrossRef]
  200. Zhang, P.; Kishimoto, Y.; Grammatikakis, I.; Gottimukkala, K.; Cutler, R.G.; Zhang, S.; Abdelmohsen, K.; Bohr, V.A.; Misra Sen, J.; Gorospe, M.; et al. Senolytic therapy alleviates Abeta-associated oligodendrocyte progenitor cell senescence and cognitive deficits in an Alzheimer’s disease model. Nat. Neurosci. 2019, 22, 719–728. [Google Scholar] [CrossRef]
  201. Ising, C.; Venegas, C.; Zhang, S.; Scheiblich, H.; Schmidt, S.V.; Vieira-Saecker, A.; Schwartz, S.; Albasset, S.; McManus, R.M.; Tejera, D.; et al. NLRP3 inflammasome activation drives tau pathology. Nature 2019, 575, 669–673. [Google Scholar] [CrossRef]
  202. Ahmad, M.H.; Fatima, M.; Mondal, A.C. Influence of microglia and astrocyte activation in the neuroinflammatory pathogenesis of Alzheimer’s disease: Rational insights for the therapeutic approaches. J. Clin. Neurosci. 2019, 59, 6–11. [Google Scholar] [CrossRef] [PubMed]
  203. Rajendran, L.; Paolicelli, R.C. Microglia-Mediated Synapse Loss in Alzheimer’s Disease. J. Neurosci. 2018, 38, 2911–2919. [Google Scholar] [CrossRef] [PubMed]
  204. Bowman, G.L.; Dayon, L.; Kirkland, R.; Wojcik, J.; Peyratout, G.; Severin, I.C.; Henry, H.; Oikonomidi, A.; Migliavacca, E.; Bacher, M.; et al. Blood-brain barrier breakdown, neuroinflammation, and cognitive decline in older adults. Alzheimers Dement. 2018, 14, 1640–1650. [Google Scholar] [CrossRef] [PubMed]
  205. Sivandzade, F.; Prasad, S.; Bhalerao, A.; Cucullo, L. NRF2 and NF-B interplay in cerebrovascular and neurodegenerative disorders: Molecular mechanisms and possible therapeutic approaches. Redox Biol. 2019, 21, 101059. [Google Scholar] [CrossRef] [PubMed]
  206. 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]
  207. de Lima Camillo, L.P.; Quinlan, R.B.A. A ride through the epigenetic landscape: Aging reversal by reprogramming. Geroscience 2021, 43, 463–485. [Google Scholar] [CrossRef]
  208. Horvath, S.; Zoller, J.A.; Haghani, A.; Jasinska, A.J.; Raj, K.; Breeze, C.E.; Ernst, J.; Vaughan, K.L.; Mattison, J.A. Epigenetic clock and methylation studies in the rhesus macaque. Geroscience 2021, 43, 2441–2453. [Google Scholar] [CrossRef]
  209. Horvath, S.; Zoller, J.A.; Haghani, A.; Lu, A.T.; Raj, K.; Jasinska, A.J.; Mattison, J.A.; Salmon, A.B. DNA methylation age analysis of rapamycin in common marmosets. Geroscience 2021, 43, 2413–2425. [Google Scholar] [CrossRef]
  210. Mendenhall, A.R.; Martin, G.M.; Kaeberlein, M.; Anderson, R.M. Cell-to-cell variation in gene expression and the aging process. Geroscience 2021, 43, 181–196. [Google Scholar] [CrossRef]
  211. Schachtschneider, K.M.; Schook, L.B.; Meudt, J.J.; Shanmuganayagam, D.; Zoller, J.A.; Haghani, A.; Li, C.Z.; Zhang, J.; Yang, A.; Raj, K.; et al. Epigenetic clock and DNA methylation analysis of porcine models of aging and obesity. Geroscience 2021, 43, 2467–2483. [Google Scholar] [CrossRef]
  212. Fraszczyk, E.; Thio, C.H.L.; Wackers, P.; Dolle, M.E.T.; Bloks, V.W.; Hodemaekers, H.; Picavet, H.S.; Stynenbosch, M.; Verschuren, W.M.M.; Snieder, H.; et al. DNA methylation trajectories and accelerated epigenetic aging in incident type 2 diabetes. Geroscience 2022, 44, 2671–2684. [Google Scholar] [CrossRef] [PubMed]
  213. Milicic, L.; Vacher, M.; Porter, T.; Dore, V.; Burnham, S.C.; Bourgeat, P.; Shishegar, R.; Doecke, J.; Armstrong, N.J.; Tankard, R.; et al. Comprehensive analysis of epigenetic clocks reveals associations between disproportionate biological ageing and hippocampal volume. Geroscience 2022, 44, 1807–1823. [Google Scholar] [CrossRef] [PubMed]
  214. Ng, G.Y.; Sheng, D.; Bae, H.G.; Kang, S.W.; Fann, D.Y.; Park, J.; Kim, J.; Alli-Shaik, A.; Lee, J.; Kim, E.; et al. Integrative epigenomic and transcriptomic analyses reveal metabolic switching by intermittent fasting in brain. Geroscience 2022, 44, 2171–2194. [Google Scholar] [CrossRef] [PubMed]
  215. Pearce, E.E.; Alsaggaf, R.; Katta, S.; Dagnall, C.; Aubert, G.; Hicks, B.D.; Spellman, S.R.; Savage, S.A.; Horvath, S.; Gadalla, S.M. Telomere length and epigenetic clocks as markers of cellular aging: A comparative study. Geroscience 2022, 44, 1861–1869. [Google Scholar] [CrossRef] [PubMed]
  216. Vetter, V.M.; Sommerer, Y.; Kalies, C.H.; Spira, D.; Bertram, L.; Demuth, I. Vitamin D supplementation is associated with slower epigenetic aging. Geroscience 2022, 44, 1847–1859. [Google Scholar] [CrossRef]
  217. Yusipov, I.; Kondakova, E.; Kalyakulina, A.; Krivonosov, M.; Lobanova, N.; Bacalini, M.G.; Franceschi, C.; Vedunova, M.; Ivanchenko, M. Accelerated epigenetic aging and inflammatory/immunological profile (ipAGE) in patients with chronic kidney disease. Geroscience 2022, 44, 817–834. [Google Scholar] [CrossRef]
  218. Horvath, S.; Lin, D.T.S.; Kobor, M.S.; Zoller, J.A.; Said, J.W.; Morgello, S.; Singer, E.; Yong, W.H.; Jamieson, B.D.; Levine, A.J. HIV, pathology and epigenetic age acceleration in different human tissues. Geroscience 2022, 44, 1609–1620. [Google Scholar] [CrossRef]
  219. Nwanaji-Enwerem, J.C.; Colicino, E.; Gao, X.; Wang, C.; Vokonas, P.; Boyer, E.W.; Baccarelli, A.A.; Schwartz, J. Associations of Plasma Folate and Vitamin B6 with Blood DNA Methylation Age: An Analysis of One-Carbon Metabolites in the VA Normative Aging Study. J. Gerontol. A Biol. Sci. Med. Sci. 2021, 76, 760–769. [Google Scholar] [CrossRef]
  220. Shadyab, A.H.; McEvoy, L.K.; Horvath, S.; Whitsel, E.A.; Rapp, S.R.; Espeland, M.A.; Resnick, S.M.; Manson, J.E.; Chen, J.C.; Chen, B.H.; et al. Association of Epigenetic Age Acceleration with Incident Mild Cognitive Impairment and Dementia among Older Women. J. Gerontol. A Biol. Sci. Med. Sci. 2022, 77, 1239–1244. [Google Scholar] [CrossRef]
  221. Liu, H.; Lutz, M.; Luo, S.; Alzheimer’s Disease Neuroimaging, I. Genetic Association Between Epigenetic Aging-Acceleration and the Progression of Mild Cognitive Impairment to Alzheimer’s Disease. J. Gerontol. A Biol. Sci. Med. Sci. 2022, 77, 1734–1742. [Google Scholar] [CrossRef]
  222. Vaccarino, V.; Huang, M.; Wang, Z.; Hui, Q.; Shah, A.J.; Goldberg, J.; Smith, N.; Kaseer, B.; Murrah, N.; Levantsevych, O.M.; et al. Epigenetic Age Acceleration and Cognitive Decline: A Twin Study. J. Gerontol. A Biol. Sci. Med. Sci. 2021, 76, 1854–1863. [Google Scholar] [CrossRef] [PubMed]
  223. Kalani, A.; Chaturvedi, P.; Kalani, K.; Kamat, P.K.; Chaturvedi, P. A high methionine, low folate and vitamin B(6)/B(12) containing diet can be associated with memory loss by epigenetic silencing of netrin-1. Neural Regen. Res. 2019, 14, 1247–1254. [Google Scholar] [CrossRef] [PubMed]
  224. Waterland, R.A. Assessing the effects of high methionine intake on DNA methylation. J. Nutr. 2006, 136, 1706S–1710S. [Google Scholar] [CrossRef] [PubMed]
  225. Amaral, C.L.; Bueno Rde, B.; Burim, R.V.; Queiroz, R.H.; Bianchi Mde, L.; Antunes, L.M. The effects of dietary supplementation of methionine on genomic stability and p53 gene promoter methylation in rats. Mutat. Res. 2011, 722, 78–83. [Google Scholar] [CrossRef]
  226. Grasselli, C.; Bombelli, S.; Eriani, S.; Domenici, G.; Galluccio, R.; Tropeano, C.; De Marco, S.; Bolognesi, M.M.; Torsello, B.; Bianchi, C.; et al. DNA Damage in Circulating Hematopoietic Progenitor Stem Cells as Promising Biological Sensor of Frailty. J. Gerontol. A Biol. Sci. Med. Sci. 2022, 77, 1279–1286. [Google Scholar] [CrossRef]
  227. Ognik, K.; Konieczka, P.; Mikulski, D.; Jankowski, J. The effect of different dietary ratios of lysine and arginine in diets with high or low methionine levels on oxidative and epigenetic DNA damage, the gene expression of tight junction proteins and selected metabolic parameters in Clostridium perfringens-challenged turkeys. Vet. Res. 2020, 51, 50. [Google Scholar] [CrossRef] [PubMed]
  228. Fenech, M. Folate, DNA damage and the aging brain. Mech. Ageing Dev. 2010, 131, 236–241. [Google Scholar] [CrossRef]
  229. Hense, J.D.; Garcia, D.N.; Isola, J.V.; Alvarado-Rincon, J.A.; Zanini, B.M.; Prosczek, J.B.; Stout, M.B.; Mason, J.B.; Walsh, P.T.; Brieno-Enriquez, M.A.; et al. Senolytic treatment reverses obesity-mediated senescent cell accumulation in the ovary. Geroscience 2022, 44, 1747–1759. [Google Scholar] [CrossRef]
  230. Fielding, R.A.; Atkinson, E.J.; Aversa, Z.; White, T.A.; Heeren, A.A.; Achenbach, S.J.; Mielke, M.M.; Cummings, S.R.; Pahor, M.; Leeuwenburgh, C.; et al. Associations between biomarkers of cellular senescence and physical function in humans: Observations from the lifestyle interventions for elders (LIFE) study. Geroscience 2022, 44, 2757–2770. [Google Scholar] [CrossRef]
  231. Dungan, C.M.; Figueiredo, V.C.; Wen, Y.; VonLehmden, G.L.; Zdunek, C.J.; Thomas, N.T.; Mobley, C.B.; Murach, K.A.; Brightwell, C.R.; Long, D.E.; et al. Senolytic treatment rescues blunted muscle hypertrophy in old mice. Geroscience 2022, 44, 1925–1940. [Google Scholar] [CrossRef]
  232. Bloom, S.I.; Tucker, J.R.; Lim, J.; Thomas, T.G.; Stoddard, G.J.; Lesniewski, L.A.; Donato, A.J. Aging results in DNA damage and telomere dysfunction that is greater in endothelial versus vascular smooth muscle cells and is exacerbated in atheroprone regions. Geroscience 2022, 44, 2741–2755. [Google Scholar] [CrossRef] [PubMed]
  233. Karin, O.; Alon, U. Senescent cell accumulation mechanisms inferred from parabiosis. Geroscience 2021, 43, 329–341. [Google Scholar] [CrossRef] [PubMed]
  234. Yousefzadeh, M.J.; Wilkinson, J.E.; Hughes, B.; Gadela, N.; Ladiges, W.C.; Vo, N.; Niedernhofer, L.J.; Huffman, D.M.; Robbins, P.D. Heterochronic parabiosis regulates the extent of cellular senescence in multiple tissues. Geroscience 2020, 42, 951–961. [Google Scholar] [CrossRef] [PubMed]
  235. Baker, D.J.; Petersen, R.C. Cellular senescence in brain aging and neurodegenerative diseases: Evidence and perspectives. J. Clin. Investig. 2018, 128, 1208–1216. [Google Scholar] [CrossRef]
  236. Csipo, T.; Lipecz, A.; Ashpole, N.M.; Balasubramanian, P.; Tarantini, S. Astrocyte senescence contributes to cognitive decline. Geroscience 2019, 42, 51–55. [Google Scholar] [CrossRef]
  237. Lawrence, I.; Bene, M.; Nacarelli, T.; Azar, A.; Cohen, J.Z.; Torres, C.; Johannes, G.; Sell, C. Correlations between age, functional status, and the senescence-associated proteins HMGB2 and p16(INK4a). Geroscience 2018, 40, 193–199. [Google Scholar] [CrossRef]
  238. Campisi, J. Aging, cellular senescence, and cancer. Annu. Rev. Physiol. 2013, 75, 685–705. [Google Scholar] [CrossRef]
  239. Childs, B.G.; Baker, D.J.; Wijshake, T.; Conover, C.A.; Campisi, J.; van Deursen, J.M. Senescent intimal foam cells are deleterious at all stages of atherosclerosis. Science 2016, 354, 472–477. [Google Scholar] [CrossRef]
  240. Chinta, S.J.; Woods, G.; Rane, A.; Demaria, M.; Campisi, J.; Andersen, J.K. Cellular senescence and the aging brain. Exp. Gerontol. 2014, 68, 3–7. [Google Scholar] [CrossRef]
  241. Baker, D.J.; Childs, B.G.; Durik, M.; Wijers, M.E.; Sieben, C.J.; Zhong, J.; Saltness, R.A.; Jeganathan, K.B.; Verzosa, G.C.; Pezeshki, A.; et al. Naturally occurring p16(Ink4a)-positive cells shorten healthy lifespan. Nature 2016, 530, 184–189. [Google Scholar] [CrossRef]
  242. Bussian, T.J.; Aziz, A.; Meyer, C.F.; Swenson, B.L.; van Deursen, J.M.; Baker, D.J. Clearance of senescent glial cells prevents tau-dependent pathology and cognitive decline. Nature 2018, 562, 578–582. [Google Scholar] [CrossRef] [PubMed]
  243. Wang, J.; Uryga, A.K.; Reinhold, J.; Figg, N.; Baker, L.; Finigan, A.; Gray, K.; Kumar, S.; Clarke, M.; Bennett, M. Vascular Smooth Muscle Cell Senescence Promotes Atherosclerosis and Features of Plaque Vulnerability. Circulation 2015, 132, 1909–1919. [Google Scholar] [CrossRef] [PubMed]
  244. Yamazaki, Y.; Baker, D.J.; Tachibana, M.; Liu, C.C.; van Deursen, J.M.; Brott, T.G.; Bu, G.; Kanekiyo, T. Vascular Cell Senescence Contributes to Blood-Brain Barrier Breakdown. Stroke 2016, 47, 1068–1077. [Google Scholar] [CrossRef] [PubMed]
  245. Han, X.; Lei, Q.; Xie, J.; Liu, H.; Li, J.; Zhang, X.; Zhang, T.; Gou, X. Potential Regulators of the Senescence-Associated Secretory Phenotype During Senescence and Aging. J. Gerontol. A Biol. Sci. Med. Sci. 2022, 77, 2207–2218. [Google Scholar] [CrossRef]
  246. Romashkan, S.; Chang, H.; Hadley, E.C. National Institute on Aging Workshop: Repurposing Drugs or Dietary Supplements for Their Senolytic or Senomorphic Effects: Considerations for Clinical Trials. J. Gerontol. A Biol. Sci. Med. Sci. 2021, 76, 1144–1152. [Google Scholar] [CrossRef]
  247. Baar, M.P.; Brandt, R.M.C.; Putavet, D.A.; Klein, J.D.D.; Derks, K.W.J.; Bourgeois, B.R.M.; Stryeck, S.; Rijksen, Y.; van Willigenburg, H.; Feijtel, D.A.; et al. Targeted Apoptosis of Senescent Cells Restores Tissue Homeostasis in Response to Chemotoxicity and Aging. Cell 2017, 169, 132–147.e116. [Google Scholar] [CrossRef]
  248. Farr, J.N.; Xu, M.; Weivoda, M.M.; Monroe, D.G.; Fraser, D.G.; Onken, J.L.; Negley, B.A.; Sfeir, J.G.; Ogrodnik, M.B.; Hachfeld, C.M.; et al. Targeting cellular senescence prevents age-related bone loss in mice. Nat. Med. 2017, 23, 1072–1079. [Google Scholar] [CrossRef]
  249. Xu, M.; Pirtskhalava, T.; Farr, J.N.; Weigand, B.M.; Palmer, A.K.; Weivoda, M.M.; Inman, C.L.; Ogrodnik, M.B.; Hachfeld, C.M.; Fraser, D.G.; et al. Senolytics improve physical function and increase lifespan in old age. Nat. Med. 2018, 24, 1246–1256. [Google Scholar] [CrossRef]
  250. Patil, P.; Dong, Q.; Wang, D.; Chang, J.; Wiley, C.; Demaria, M.; Lee, J.; Kang, J.; Niedernhofer, L.J.; Robbins, P.D.; et al. Systemic clearance of p16. Aging Cell 2019, 18, e12927. [Google Scholar] [CrossRef]
  251. Baker, D.J.; Wijshake, T.; Tchkonia, T.; LeBrasseur, N.K.; Childs, B.G.; van de Sluis, B.; Kirkland, J.L.; van Deursen, J.M. Clearance of p16Ink4a-positive senescent cells delays ageing-associated disorders. Nature 2011, 479, 232–236. [Google Scholar] [CrossRef]
  252. Ahire, C.; Nyul-Toth, A.; DelFavero, J.; Gulej, R.; Faakye, J.A.; Tarantini, S.; Kiss, T.; Kuan-Celarier, A.; Balasubramanian, P.; Ungvari, A.; et al. Accelerated cerebromicrovascular senescence contributes to cognitive decline in a mouse model of paclitaxel (Taxol)-induced chemobrain. Aging Cell 2023, 22, e13832. [Google Scholar] [CrossRef] [PubMed]
  253. Yabluchanksiy, A.; Tarantini, S.; Balasubramaniam, P.; Kiss, T.; Csipo, T.; Fulop, G.A.; Lipecz, A.; delFavero, J.; Nyul-Toth, A.; Sonntag, W.E.; et al. Pharmacological or genetic depletion of senescent astrocytes prevents whole brain irradiation-induced impairment of neurovascular coupling responses protecting cognitive function in mice. Geroscience 2020, in press. [Google Scholar] [CrossRef] [PubMed]
  254. Xing, S.S.; Li, J.; Chen, L.; Yang, Y.F.; He, P.L.; Li, J.; Yang, J. Salidroside attenuates endothelial cellular senescence via decreasing the expression of inflammatory cytokines and increasing the expression of SIRT3. Mech. Ageing Dev. 2018, 175, 1–6. [Google Scholar] [CrossRef]
  255. Albertini, E.; Koziel, R.; Durr, A.; Neuhaus, M.; Jansen-Durr, P. Cystathionine beta synthase modulates senescence of human endothelial cells. Aging 2012, 4, 664–673. [Google Scholar] [CrossRef]
  256. Koziel, R.; Ruckenstuhl, C.; Albertini, E.; Neuhaus, M.; Netzberger, C.; Bust, M.; Madeo, F.; Wiesner, R.J.; Jansen-Durr, P. Methionine restriction slows down senescence in human diploid fibroblasts. Aging Cell 2014, 13, 1038–1048. [Google Scholar] [CrossRef] [PubMed]
  257. Auzmendi-Iriarte, J.; Matheu, A. Impact of Chaperone-Mediated Autophagy in Brain Aging: Neurodegenerative Diseases and Glioblastoma. Front. Aging Neurosci. 2020, 12, 630743. [Google Scholar] [CrossRef] [PubMed]
  258. Chocron, E.S.; Munkacsy, E.; Kim, H.S.; Karpowicz, P.; Jiang, N.; Van Skike, C.E.; DeRosa, N.; Banh, A.Q.; Palavicini, J.P.; Wityk, P.; et al. Genetic and pharmacologic proteasome augmentation ameliorates Alzheimer’s-like pathology in mouse and fly APP overexpression models. Sci. Adv. 2022, 8, eabk2252. [Google Scholar] [CrossRef]
  259. Hou, Y.; Dan, X.; Babbar, M.; Wei, Y.; Hasselbalch, S.G.; Croteau, D.L.; Bohr, V.A. Ageing as a risk factor for neurodegenerative disease. Nat. Rev. Neurol. 2019, 15, 565–581. [Google Scholar] [CrossRef]
  260. Taylor, R.C.; Hetz, C. Mastering organismal aging through the endoplasmic reticulum proteostasis network. Aging Cell 2020, 19, e13265. [Google Scholar] [CrossRef]
  261. Cozachenco, D.; Ribeiro, F.C.; Ferreira, S.T. Defective proteostasis in Alzheimer’s disease. Ageing Res. Rev. 2023, 85, 101862. [Google Scholar] [CrossRef]
  262. Kulkarni, A.; Preeti, K.; Tryphena, K.P.; Srivastava, S.; Singh, S.B.; Khatri, D.K. Proteostasis in Parkinson’s disease: Recent development and possible implication in diagnosis and therapeutics. Ageing Res. Rev. 2023, 84, 101816. [Google Scholar] [CrossRef] [PubMed]
  263. Mukherjee, S.; Mishra, A.K.; Peer, G.D.G.; Bagabir, S.A.; Haque, S.; Pandey, R.P.; Raj, V.S.; Jain, N.; Pandey, A.; Kar, S.K. The Interplay of the Unfolded Protein Response in Neurodegenerative Diseases: A Therapeutic Role of Curcumin. Front. Aging Neurosci. 2021, 13, 767493. [Google Scholar] [CrossRef] [PubMed]
  264. Weng, F.L.; He, L. Disrupted ubiquitin proteasome system underlying tau accumulation in Alzheimer’s disease. Neurobiol. Aging 2021, 99, 79–85. [Google Scholar] [CrossRef] [PubMed]
  265. Nichenametla, S.N.; Mattocks, D.A.L.; Malloy, V.L.; Pinto, J.T. Sulfur amino acid restriction-induced changes in redox-sensitive proteins are associated with slow protein synthesis rates. Ann. N. Y. Acad. Sci. 2018, 1418, 80–94. [Google Scholar] [CrossRef]
  266. Yang, Y.; Zhang, J.; Wu, G.; Sun, J.; Wang, Y.; Guo, H.; Shi, Y.; Cheng, X.; Tang, X.; Le, G. Dietary methionine restriction regulated energy and protein homeostasis by improving thyroid function in high fat diet mice. Food Funct. 2018, 9, 3718–3731. [Google Scholar] [CrossRef]
  267. Jakubowski, H. Proteomic exploration of cystathionine beta-synthase deficiency: Implications for the clinic. Expert Rev. Proteom. 2020, 17, 751–765. [Google Scholar] [CrossRef]
  268. Reddy, V.S.; Trinath, J.; Reddy, G.B. Implication of homocysteine in protein quality control processes. Biochimie 2019, 165, 19–31. [Google Scholar] [CrossRef]
  269. Wiedenhoeft, T.; Tarantini, S.; Nyul-Toth, A.; Yabluchanskiy, A.; Csipo, T.; Balasubramanian, P.; Lipecz, A.; Kiss, T.; Csiszar, A.; Csiszar, A.; et al. Fusogenic liposomes effectively deliver resveratrol to the cerebral microcirculation and improve endothelium-dependent neurovascular coupling responses in aged mice. Geroscience 2019, 41, 711–725. [Google Scholar] [CrossRef]
  270. 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]
  271. 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 prevention of vascular cognitive impairment. Geroscience 2019, 41, 619–630. [Google Scholar] [CrossRef]
  272. Fulop, G.A.; Ahire, C.; Csipo, T.; Tarantini, S.; Kiss, T.; Balasubramanian, P.; Yabluchanskiy, A.; Farkas, E.; Toth, A.; Nyul-Toth, A.; et al. Cerebral venous congestion promotes blood-brain barrier disruption and neuroinflammation, impairing cognitive function in mice. Geroscience 2019, 41, 575–589. [Google Scholar] [CrossRef] [PubMed]
  273. 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] [PubMed]
  274. Sweeney, M.D.; Kisler, K.; Montagne, A.; Toga, A.W.; Zlokovic, B.V. The role of brain vasculature in neurodegenerative disorders. Nat. Neurosci. 2018, 21, 1318–1331. [Google Scholar] [CrossRef] [PubMed]
  275. Iadecola, C.; Gottesman, R.F. Cerebrovascular Alterations in Alzheimer Disease. Circ. Res. 2018, 123, 406–408. [Google Scholar] [CrossRef]
  276. Iadecola, C. The Neurovascular Unit Coming of Age: A Journey through Neurovascular Coupling in Health and Disease. Neuron 2017, 96, 17–42. [Google Scholar] [CrossRef]
  277. Iadecola, C.; Gottesman, R.F. Neurovascular and Cognitive Dysfunction in Hypertension. Circ. Res. 2019, 124, 1025–1044. [Google Scholar] [CrossRef]
  278. Iadecola, C.; Park, L.; Capone, C. Threats to the mind: Aging, amyloid, and hypertension. Stroke 2009, 40, S40–S44. [Google Scholar] [CrossRef]
  279. Toth, P.; Tarantini, S.; Csiszar, A.; Ungvari, Z. Functional vascular contributions to cognitive impairment and dementia: Mechanisms and consequences of cerebral autoregulatory dysfunction, endothelial impairment, and neurovascular uncoupling in aging. Am. J. Physiol. Heart Circ. Physiol. 2017, 312, H1–H20. [Google Scholar] [CrossRef]
  280. Li, P.; Ma, S.; Ma, X.; Ding, D.; Zhu, X.; Zhang, H.; Liu, J.; Mu, J.; Zhang, M. Reversal of neurovascular decoupling and cognitive impairment in patients with end-stage renal disease during a hemodialysis session: Evidence from a comprehensive fMRI analysis. Hum. Brain Mapp. 2023, 44, 989–1001. [Google Scholar] [CrossRef]
  281. Rosengarten, B.; Osthaus, S.; Auch, D.; Kaps, M. Effects of acute hyperhomocysteinemia on the neurovascular coupling mechanism in healthy young adults. Stroke 2003, 34, 446–451. [Google Scholar] [CrossRef]
  282. Toda, N.; Okamura, T. Hyperhomocysteinemia impairs regional blood flow: Involvements of endothelial and neuronal nitric oxide. Pflugers Arch. 2016, 468, 1517–1525. [Google Scholar] [CrossRef] [PubMed]
  283. van Laar, P.J.; van der Graaf, Y.; Mali, W.P.; van der Grond, J.; Hendrikse, J.; Group, S.S. Effect of cerebrovascular risk factors on regional cerebral blood flow. Radiology 2008, 246, 198–204. [Google Scholar] [CrossRef] [PubMed]
  284. Braun, D.J.; Abner, E.; Bakshi, V.; Goulding, D.S.; Grau, E.M.; Lin, A.L.; Norris, C.M.; Sudduth, T.L.; Webster, S.J.; Wilcock, D.M.; et al. Blood Flow Deficits and Cerebrovascular Changes in a Dietary Model of Hyperhomocysteinemia. ASN Neuro 2019, 11, 1759091419865788. [Google Scholar] [CrossRef] [PubMed]
  285. Giusti, B.; Saracini, C.; Bolli, P.; Magi, A.; Martinelli, I.; Peyvandi, F.; Rasura, M.; Volpe, M.; Lotta, L.A.; Rubattu, S.; et al. Early-onset ischaemic stroke: Analysis of 58 polymorphisms in 17 genes involved in methionine metabolism. Thromb. Haemost. 2010, 104, 231–242. [Google Scholar] [CrossRef]
  286. Wu, G.H.; Kong, F.Z.; Dong, X.F.; Wu, D.F.; Guo, Q.Z.; Shen, A.R.; Cheng, Q.Z.; Luo, W.F. Association between hyperhomocysteinemia and stroke with atherosclerosis and small artery occlusion depends on homocysteine metabolism-related vitamin levels in Chinese patients with normal renal function. Metab. Brain Dis. 2017, 32, 859–865. [Google Scholar] [CrossRef]
  287. Larsson, S.C.; Mannisto, S.; Virtanen, M.J.; Kontto, J.; Albanes, D.; Virtamo, J. Folate, vitamin B6, vitamin B12, and methionine intakes and risk of stroke subtypes in male smokers. Am. J. Epidemiol. 2008, 167, 954–961. [Google Scholar] [CrossRef]
  288. Williams, S.R.; Yang, Q.; Chen, F.; Liu, X.; Keene, K.L.; Jacques, P.; Chen, W.M.; Weinstein, G.; Hsu, F.C.; Beiser, A.; et al. Genome-wide meta-analysis of homocysteine and methionine metabolism identifies five one carbon metabolism loci and a novel association of ALDH1L1 with ischemic stroke. PLoS Genet. 2014, 10, e1004214. [Google Scholar] [CrossRef]
  289. Kim, O.J.; Hong, S.P.; Ahn, J.Y.; Hong, S.H.; Hwang, T.S.; Kim, S.O.; Yoo, W.; Oh, D.; Kim, N.K. Influence of combined methionine synthase (MTR 2756A > G) and methylenetetrahydrofolate reductase (MTHFR 677C > T) polymorphisms to plasma homocysteine levels in Korean patients with ischemic stroke. Yonsei Med. J. 2007, 48, 201–209. [Google Scholar] [CrossRef]
  290. Selhub, J.; Troen, A.M. Sulfur amino acids and atherosclerosis: A role for excess dietary methionine. Ann. N. Y. Acad. Sci. 2016, 1363, 18–25. [Google Scholar] [CrossRef]
  291. Toborek, M.; Kopieczna-Grzebieniak, E.; Drozdz, M.; Wieczorek, M. Increased lipid peroxidation as a mechanism of methionine-induced atherosclerosis in rabbits. Atherosclerosis 1995, 115, 217–224. [Google Scholar] [CrossRef]
  292. Yang, A.N.; Zhang, H.P.; Sun, Y.; Yang, X.L.; Wang, N.; Zhu, G.; Zhang, H.; Xu, H.; Ma, S.C.; Zhang, Y.; et al. High-methionine diets accelerate atherosclerosis by HHcy-mediated FABP4 gene demethylation pathway via DNMT1 in ApoE(−/−) mice. FEBS Lett. 2015, 589, 3998–4009. [Google Scholar] [CrossRef] [PubMed]
  293. Zhou, J.; Moller, J.; Danielsen, C.C.; Bentzon, J.; Ravn, H.B.; Austin, R.C.; Falk, E. Dietary supplementation with methionine and homocysteine promotes early atherosclerosis but not plaque rupture in ApoE-deficient mice. Arterioscler. Thromb. Vasc. Biol. 2001, 21, 1470–1476. [Google Scholar] [CrossRef] [PubMed]
  294. Zulli, A.; Hare, D.L. High dietary methionine plus cholesterol stimulates early atherosclerosis and late fibrous cap development which is associated with a decrease in GRP78 positive plaque cells. Int. J. Exp. Pathol. 2009, 90, 311–320. [Google Scholar] [CrossRef] [PubMed]
  295. Zulli, A.; Hare, D.L.; Buxton, B.F.; Black, M.J. High dietary methionine plus cholesterol exacerbates atherosclerosis formation in the left main coronary artery of rabbits. Atherosclerosis 2004, 176, 83–89. [Google Scholar] [CrossRef] [PubMed]
  296. Demchenko, I.T.; Luchakov, Y.I.; Moskvin, A.N.; Gutsaeva, D.R.; Allen, B.W.; Thalmann, E.D.; Piantadosi, C.A. Cerebral blood flow and brain oxygenation in rats breathing oxygen under pressure. J. Cereb. Blood Flow Metab. 2005, 25, 1288–1300. [Google Scholar] [CrossRef]
  297. Tarantini, S.; Fulop, G.A.; Kiss, T.; Farkas, E.; Zolei-Szenasi, D.; Galvan, V.; Toth, P.; Csiszar, A.; Ungvari, Z.; Yabluchanskiy, A. Demonstration of impaired neurovascular coupling responses in TG2576 mouse model of Alzheimer’s disease using functional laser speckle contrast imaging. Geroscience 2017, 39, 465–473. [Google Scholar] [CrossRef]
  298. Tarantini, S.; Yabluchanksiy, A.; Fulop, G.A.; Hertelendy, P.; Valcarcel-Ares, M.N.; Kiss, T.; Bagwell, J.M.; O’Connor, D.; Farkas, E.; Sorond, F.; et al. Pharmacologically induced impairment of neurovascular coupling responses alters gait coordination in mice. Geroscience 2017, 39, 601–614. [Google Scholar] [CrossRef]
  299. Toth, P.; Tarantini, S.; Davila, A.; Valcarcel-Ares, M.N.; Tucsek, Z.; Varamini, B.; Ballabh, P.; Sonntag, W.E.; Baur, J.A.; Csiszar, A.; et al. Purinergic glio-endothelial coupling during neuronal activity: Role of P2Y1 receptors and eNOS in functional hyperemia in the mouse somatosensory cortex. Am. J. Physiol. Heart Circ. Physiol. 2015, 309, H1837–H1845. [Google Scholar] [CrossRef]
  300. 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 downregulation of NADPH oxidase. Am. J. Physiol. Heart Circ. Physiol. 2014, 306, H299–H308. [Google Scholar] [CrossRef]
  301. Tucsek, Z.; Toth, P.; Tarantini, S.; Sosnowska, D.; Gautam, T.; Warrington, J.P.; Giles, C.B.; Wren, J.D.; Koller, A.; Ballabh, P.; et al. Aging exacerbates obesity-induced cerebromicrovascular rarefaction, neurovascular uncoupling, and cognitive decline in mice. J. Gerontol. A Biol. Sci. Med. Sci. 2014, 69, 1339–1352. [Google Scholar] [CrossRef]
  302. Vendemiale, G.; Romano, A.D.; Dagostino, M.; de Matthaeis, A.; Serviddio, G. Endothelial dysfunction associated with mild cognitive impairment in elderly population. Aging Clin. Exp. Res. 2013, 25, 247–255. [Google Scholar] [CrossRef] [PubMed]
  303. Bhayadia, R.; Schmidt, B.M.; Melk, A.; Homme, M. Senescence-Induced Oxidative Stress Causes Endothelial Dysfunction. J. Gerontol. A Biol. Sci. Med. Sci. 2016, 71, 161–169. [Google Scholar] [CrossRef] [PubMed]
  304. de Picciotto, N.E.; Gano, L.B.; Johnson, L.C.; Martens, C.R.; Sindler, A.L.; Mills, K.F.; Imai, S.; Seals, D.R. Nicotinamide mononucleotide supplementation reverses vascular dysfunction and oxidative stress with aging in mice. Aging Cell 2016, 15, 522–530. [Google Scholar] [CrossRef] [PubMed]
  305. Seals, D.R.; Nagy, E.E.; Moreau, K.L. Aerobic exercise training and vascular function with ageing in healthy men and women. J. Physiol. 2019, 597, 4901–4914. [Google Scholar] [CrossRef] [PubMed]
  306. Csiszar, A.; Ungvari, Z.; Edwards, J.G.; Kaminski, P.M.; Wolin, M.S.; Koller, A.; Kaley, G. Aging-induced phenotypic changes and oxidative stress impair coronary arteriolar function. Circ. Res. 2002, 90, 1159–1166. [Google Scholar] [CrossRef]
  307. Istvan, L.; Czako, C.; Elo, A.; Mihaly, Z.; Sotonyi, P.; Varga, A.; Ungvari, Z.; Csiszar, A.; Yabluchanskiy, A.; Conley, S.; et al. Imaging retinal microvascular manifestations of carotid artery disease in older adults: From diagnosis of ocular complications to understanding microvascular contributions to cognitive impairment. Geroscience 2021, 43, 1703–1723. [Google Scholar] [CrossRef]
  308. Tarantini, S.; Yabluchanskiy, A.; Lindsey, M.L.; Csiszar, A.; Ungvari, Z. Effect of genetic depletion of MMP-9 on neurological manifestations of hypertension-induced intracerebral hemorrhages in aged mice. Geroscience 2021, 43, 2611–2619. [Google Scholar] [CrossRef]
  309. Montagne, A.; Barnes, S.R.; Sweeney, M.D.; Halliday, M.R.; Sagare, A.P.; Zhao, Z.; Toga, A.W.; Jacobs, R.E.; Liu, C.Y.; Amezcua, L.; et al. Blood-brain barrier breakdown in the aging human hippocampus. Neuron 2015, 85, 296–302. [Google Scholar] [CrossRef]
  310. Bellamy, M.F.; McDowell, I.F.; Ramsey, M.W.; Brownlee, M.; Bones, C.; Newcombe, R.G.; Lewis, M.J. Hyperhomocysteinemia after an oral methionine load acutely impairs endothelial function in healthy adults. Circulation 1998, 98, 1848–1852. [Google Scholar] [CrossRef]
  311. Bellamy, M.F.; McDowell, I.F.; Ramsey, M.W.; Brownlee, M.; Newcombe, R.G.; Lewis, M.J. Oral folate enhances endothelial function in hyperhomocysteinaemic subjects. Eur. J. Clin. Investig. 1999, 29, 659–662. [Google Scholar] [CrossRef]
  312. Chambers, J.C.; McGregor, A.; Jean-Marie, J.; Obeid, O.A.; Kooner, J.S. Demonstration of rapid onset vascular endothelial dysfunction after hyperhomocysteinemia: An effect reversible with vitamin C therapy. Circulation 1999, 99, 1156–1160. [Google Scholar] [CrossRef] [PubMed]
  313. Chambers, J.C.; Obeid, O.A.; Kooner, J.S. Physiological increments in plasma homocysteine induce vascular endothelial dysfunction in normal human subjects. Arterioscler. Thromb. Vasc. Biol. 1999, 19, 2922–2927. [Google Scholar] [CrossRef] [PubMed]
  314. Chambers, J.C.; Ueland, P.M.; Obeid, O.A.; Wrigley, J.; Refsum, H.; Kooner, J.S. Improved vascular endothelial function after oral B vitamins: An effect mediated through reduced concentrations of free plasma homocysteine. Circulation 2000, 102, 2479–2483. [Google Scholar] [CrossRef] [PubMed]
  315. Woo, K.S.; Chook, P.; Lolin, Y.I.; Cheung, A.S.; Chan, L.T.; Sun, Y.Y.; Sanderson, J.E.; Metreweli, C.; Celermajer, D.S. Hyperhomocyst(e)inemia is a risk factor for arterial endothelial dysfunction in humans. Circulation 1997, 96, 2542–2544. [Google Scholar] [CrossRef] [PubMed]
  316. Woo, K.S.; Chook, P.; Lolin, Y.I.; Sanderson, J.E.; Metreweli, C.; Celermajer, D.S. Folic acid improves arterial endothelial function in adults with hyperhomocystinemia. J. Am. Coll. Cardiol. 1999, 34, 2002–2006. [Google Scholar] [CrossRef]
  317. Weiss, N.; Zhang, Y.Y.; Heydrick, S.; Bierl, C.; Loscalzo, J. Overexpression of cellular glutathione peroxidase rescues homocyst(e)ine-induced endothelial dysfunction. Proc. Natl. Acad. Sci. USA 2001, 98, 12503–12508. [Google Scholar] [CrossRef]
  318. Kamat, P.K.; Vacek, J.C.; Kalani, A.; Tyagi, N. Homocysteine Induced Cerebrovascular Dysfunction: A Link to Alzheimer’s Disease Etiology. Open Neurol. J. 2015, 9, 9–14. [Google Scholar] [CrossRef]
  319. Carey, A.; Fossati, S. Hypertension and hyperhomocysteinemia as modifiable risk factors for Alzheimer’s disease and dementia: New evidence, potential therapeutic strategies, and biomarkers. Alzheimers Dement. 2023, 19, 671–695. [Google Scholar] [CrossRef]
  320. Faber, J.E.; Zhang, H.; Lassance-Soares, R.M.; Prabhakar, P.; Najafi, A.H.; Burnett, M.S.; Epstein, S.E. Aging causes collateral rarefaction and increased severity of ischemic injury in multiple tissues. Arterioscler. Thromb. Vasc. Biol. 2011, 31, 1748–1756. [Google Scholar] [CrossRef]
  321. Riddle, D.R.; Sonntag, W.E.; Lichtenwalner, R.J. Microvascular plasticity in aging. Ageing Res. Rev. 2003, 2, 149–168. [Google Scholar] [CrossRef]
  322. Chantler, P.D.; Shrader, C.D.; Tabone, L.E.; d’Audiffret, A.C.; Huseynova, K.; Brooks, S.D.; Branyan, K.W.; Grogg, K.A.; Frisbee, J.C. Cerebral Cortical Microvascular Rarefaction in Metabolic Syndrome is Dependent on Insulin Resistance and Loss of Nitric Oxide Bioavailability. Microcirculation 2015, 22, 435–445. [Google Scholar] [CrossRef] [PubMed]
  323. Frisbee, J.C. Remodeling of the skeletal muscle microcirculation increases resistance to perfusion in obese Zucker rats. Am. J. Physiol. Heart Circ. Physiol. 2003, 285, H104–H111. [Google Scholar] [CrossRef] [PubMed]
  324. Frisbee, J.C. Reduced nitric oxide bioavailability contributes to skeletal muscle microvessel rarefaction in the metabolic syndrome. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2005, 289, R307–R316. [Google Scholar] [CrossRef] [PubMed]
  325. Frisbee, J.C. Hypertension-independent microvascular rarefaction in the obese Zucker rat model of the metabolic syndrome. Microcirculation 2005, 12, 383–392. [Google Scholar] [CrossRef]
  326. Hoffmann, J.; Haendeler, J.; Aicher, A.; Rössig, L.; Vasa, M.; Zeiher, A.M.; Dimmeler, S. Aging enhances the sensitivity of endothelial cells toward apoptotic stimuli: Important role of nitric oxide. Circ. Res. 2001, 89, 709–715. [Google Scholar] [CrossRef]
  327. Kobayashi, N.; DeLano, F.A.; Schmid-Schönbein, G.W. Oxidative stress promotes endothelial cell apoptosis and loss of microvessels in the spontaneously hypertensive rats. Arterioscler. Thromb. Vasc. Biol. 2005, 25, 2114–2121. [Google Scholar] [CrossRef]
  328. Villar-Cheda, B.; Sousa-Ribeiro, D.; Rodriguez-Pallares, J.; Rodriguez-Perez, A.I.; Guerra, M.J.; Labandeira-Garcia, J.L. Aging and sedentarism decrease vascularization and VEGF levels in the rat substantia nigra. Implications for Parkinson’s disease. J. Cereb. Blood Flow Metab. 2009, 29, 230–234. [Google Scholar] [CrossRef]
  329. Galli, G.; Pinchera, A.; Piaggi, P.; Fierabracci, P.; Giannetti, M.; Querci, G.; Scartabelli, G.; Manetti, L.; Ceccarini, G.; Martinelli, S.; et al. Serum insulin-like growth factor-1 concentrations are reduced in severely obese women and raise after weight loss induced by laparoscopic adjustable gastric banding. Obes. Surg. 2012, 22, 1276–1280. [Google Scholar] [CrossRef]
  330. Bancu, I.; Navarro Díaz, M.; Serra, A.; Granada, M.; Lopez, D.; Romero, R.; Bonet, J. Low Insulin-Like Growth Factor-1 Level in Obesity Nephropathy: A New Risk Factor? PLoS ONE 2016, 11, e0154451. [Google Scholar] [CrossRef]
  331. Breese, C.R.; Ingram, R.L.; Sonntag, W.E. Influence of age and long-term dietary restriction on plasma insulin-like growth factor-1 (IGF-1), IGF-1 gene expression, and IGF-1 binding proteins. J. Gerontol. 1991, 46, B180–B187. [Google Scholar] [CrossRef]
  332. Csiszar, A.; Sosnowska, D.; Tucsek, Z.; Gautam, T.; Toth, P.; Losonczy, G.; Colman, R.J.; Weindruch, R.; Anderson, R.M.; Sonntag, W.E.; et al. Circulating factors induced by caloric restriction in the nonhuman primate Macaca mulatta activate angiogenic processes in endothelial cells. J. Gerontol. A Biol. Sci. Med. Sci. 2013, 68, 235–249. [Google Scholar] [CrossRef] [PubMed]
  333. Ungvari, Z.; Tarantini, S.; Kiss, T.; Wren, J.D.; Giles, C.B.; Griffin, C.T.; Murfee, W.L.; Pacher, P.; Csiszar, A. Endothelial dysfunction and angiogenesis impairment in the ageing vasculature. Nat. Rev. Cardiol. 2018, 15, 555–565. [Google Scholar] [CrossRef] [PubMed]
  334. Ungvari, Z.; Tucsek, Z.; Sosnowska, D.; Toth, P.; Gautam, T.; Podlutsky, A.; Csiszar, A.; Losonczy, G.; Valcarcel-Ares, M.N.; Sonntag, W.E. Aging-Induced Dysregulation of Dicer1-Dependent MicroRNA Expression Impairs Angiogenic Capacity of Rat Cerebromicrovascular Endothelial Cells. J. Gerontol. A Biol. Sci. Med. Sci. 2013, 68, 877–891. [Google Scholar] [CrossRef] [PubMed]
  335. Valcarcel-Ares, M.N.; Gautam, T.; Warrington, J.P.; Bailey-Downs, L.; Sosnowska, D.; de Cabo, R.; Losonczy, G.; Sonntag, W.E.; Ungvari, Z.; Csiszar, A. Disruption of Nrf2 signaling impairs angiogenic capacity of endothelial cells: Implications for microvascular aging. J. Gerontol. A Biol. Sci. Med. Sci. 2012, 67, 821–829. [Google Scholar] [CrossRef]
  336. Bohannon, D.G.; Okhravi, H.R.; Kim, J.; Kuroda, M.J.; Didier, E.S.; Kim, W.K. A subtype of cerebrovascular pericytes is associated with blood-brain barrier disruption that develops during normal aging and simian immunodeficiency virus infection. Neurobiol. Aging 2020, 96, 128–136. [Google Scholar] [CrossRef]
  337. Costea, L.; Meszaros, A.; Bauer, H.; Bauer, H.C.; Traweger, A.; Wilhelm, I.; Farkas, A.E.; Krizbai, I.A. The Blood-Brain Barrier and Its Intercellular Junctions in Age-Related Brain Disorders. Int. J. Mol. Sci. 2019, 20, 5472. [Google Scholar] [CrossRef]
  338. Erdo, F.; Denes, L.; de Lange, E. Age-associated physiological and pathological changes at the blood-brain barrier: A review. J. Cereb. Blood Flow Metab. 2017, 37, 4–24. [Google Scholar] [CrossRef]
  339. Farrall, A.J.; Wardlaw, J.M. Blood-brain barrier: Ageing and microvascular disease--systematic review and meta-analysis. Neurobiol. Aging 2009, 30, 337–352. [Google Scholar] [CrossRef]
  340. Chu, M.; Teng, J.; Guo, L.; Wang, Y.; Zhang, L.; Gao, J.; Liu, L. Mild hyperhomocysteinemia induces blood-brain barrier dysfunction but not neuroinflammation in the cerebral cortex and hippocampus of wild-type mice. Can. J. Physiol. Pharmacol. 2021, 99, 847–856. [Google Scholar] [CrossRef]
  341. Tchantchou, F.; Hsia, R.C.; Puche, A.; Fiskum, G. Hippocampal vulnerability to hyperhomocysteinemia worsens pathological outcomes of mild traumatic brain injury in rats. J. Cent. Nerv. Syst. Dis. 2023, 15, 11795735231160025. [Google Scholar] [CrossRef]
  342. Tawfik, A.; Elsherbiny, N.M.; Zaidi, Y.; Rajpurohit, P. Homocysteine and Age-Related Central Nervous System Diseases: Role of Inflammation. Int. J. Mol. Sci. 2021, 22, 6259. [Google Scholar] [CrossRef] [PubMed]
  343. Tawfik, A.; Samra, Y.A.; Elsherbiny, N.M.; Al-Shabrawey, M. Implication of Hyperhomocysteinemia in Blood Retinal Barrier (BRB) Dysfunction. Biomolecules 2020, 10, 1119. [Google Scholar] [CrossRef] [PubMed]
  344. Kim, J.M.; Lee, H.; Chang, N. Hyperhomocysteinemia due to short-term folate deprivation is related to electron microscopic changes in the rat brain. J. Nutr. 2002, 132, 3418–3421. [Google Scholar] [CrossRef] [PubMed]
  345. Tawfik, A.; Markand, S.; Al-Shabrawey, M.; Mayo, J.N.; Reynolds, J.; Bearden, S.E.; Ganapathy, V.; Smith, S.B. Alterations of retinal vasculature in cystathionine-beta-synthase heterozygous mice: A model of mild to moderate hyperhomocysteinemia. Am. J. Pathol. 2014, 184, 2573–2585. [Google Scholar] [CrossRef] [PubMed]
  346. Gulej, R.; Nyul-Toth, A.; Ahire, C.; DelFavero, J.; Balasubramanian, P.; Kiss, T.; Tarantini, S.; Benyo, Z.; Pacher, P.; Csik, B.; et al. Elimination of senescent cells by treatment with Navitoclax/ABT263 reverses whole brain irradiation-induced blood-brain barrier disruption in the mouse brain. Geroscience 2023. [Google Scholar] [CrossRef]
  347. Preininger, M.K.; Kaufer, D. Blood-Brain Barrier Dysfunction and Astrocyte Senescence as Reciprocal Drivers of Neuropathology in Aging. Int. J. Mol. Sci. 2022, 23, 6217. [Google Scholar] [CrossRef]
  348. Tucsek, Z.; Toth, P.; Sosnowsk, D.; Gautam, T.; Mitschelen, M.; Koller, A.; Szalai, G.; Sonntag, W.E.; Ungvari, Z.; Csiszar, A. Obesity in aging exacerbates blood brain barrier disruption, neuroinflammation and oxidative stress in the mouse hippocampus: Effects on expression of genes involved in beta-amyloid generation and Alzheimer’s disease. J. Gerontol. A Biol. Sci. Med. Sci. 2014, 69, 1212–1226. [Google Scholar] [CrossRef]
  349. 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. 2018, 74, 290–298. [Google Scholar] [CrossRef]
  350. Baydas, G.; Ozer, M.; Yasar, A.; Tuzcu, M.; Koz, S.T. Melatonin improves learning and memory performances impaired by hyperhomocysteinemia in rats. Brain Res. 2005, 1046, 187–194. [Google Scholar] [CrossRef]
  351. Postnikova, T.Y.; Amakhin, D.V.; Trofimova, A.M.; Tumanova, N.L.; Dubrovskaya, N.M.; Kalinina, D.S.; Kovalenko, A.A.; Shcherbitskaia, A.D.; Vasilev, D.S.; Zaitsev, A.V. Maternal Hyperhomocysteinemia Produces Memory Deficits Associated with Impairment of Long-Term Synaptic Plasticity in Young Rats. Cells 2022, 12, 58. [Google Scholar] [CrossRef]
  352. Suszynska-Zajczyk, J.; Luczak, M.; Marczak, L.; Jakubowski, H. Hyperhomocysteinemia and bleomycin hydrolase modulate the expression of mouse brain proteins involved in neurodegeneration. J. Alzheimers Dis. 2014, 40, 713–726. [Google Scholar] [CrossRef] [PubMed]
  353. Wang, M.; Liang, X.; Zhang, Q.; Luo, S.; Liu, H.; Wang, X.; Sai, N.; Zhang, X. Homocysteine can aggravate depressive like behaviors in a middle cerebral artery occlusion/reperfusion rat model: A possible role for NMDARs-mediated synaptic alterations. Nutr. Neurosci. 2023, 26, 483–495. [Google Scholar] [CrossRef]
  354. Viggiano, A.; Viggiano, E.; Monda, M.; Ingrosso, D.; Perna, A.F.; De Luca, B. Methionine-enriched diet decreases hippocampal antioxidant defences and impairs spontaneous behaviour and long-term potentiation in rats. Brain Res. 2012, 1471, 66–74. [Google Scholar] [CrossRef] [PubMed]
  355. Chai, G.S.; Jiang, X.; Ni, Z.F.; Ma, Z.W.; Xie, A.J.; Cheng, X.S.; Wang, Q.; Wang, J.Z.; Liu, G.P. Betaine attenuates Alzheimer-like pathological changes and memory deficits induced by homocysteine. J. Neurochem. 2013, 124, 388–396. [Google Scholar] [CrossRef] [PubMed]
  356. Gao, L.; Zeng, X.N.; Guo, H.M.; Wu, X.M.; Chen, H.J.; Di, R.K.; Wu, Y. Cognitive and neurochemical alterations in hyperhomocysteinemic rat. Neurol. Sci. 2012, 33, 39–43. [Google Scholar] [CrossRef] [PubMed]
  357. Almeida, M.R.; Mabasa, L.; Crane, C.; Park, C.S.; Venancio, V.P.; Bianchi, M.L.; Antunes, L.M. Maternal vitamin B6 deficient or supplemented diets on expression of genes related to GABAergic, serotonergic, or glutamatergic pathways in hippocampus of rat dams and their offspring. Mol. Nutr. Food Res. 2016, 60, 1615–1624. [Google Scholar] [CrossRef] [PubMed]
  358. Kulminski, A.M.; Loiko, E.; Loika, Y.; Culminskaya, I. Pleiotropic predisposition to Alzheimer’s disease and educational attainment: Insights from the summary statistics analysis. Geroscience 2022, 44, 265–280. [Google Scholar] [CrossRef]
  359. Lu, W.H.; de Souto Barreto, P.; Rolland, Y.; Bouyahia, A.; Fischer, C.; Mangin, J.F.; Giudici, K.V.; Vellas, B.; Group, M.D. Biological and Neuroimaging Markers as Predictors of 5-Year Incident Frailty in Older Adults: A Secondary Analysis of the MAPT Study. J. Gerontol. A Biol. Sci. Med. Sci. 2021, 76, e361–e369. [Google Scholar] [CrossRef]
  360. Guillotin, S.; Vallet, A.; Lorthois, S.; Cestac, P.; Schmidt, E.; Delcourt, N. Association Between Homocysteine, Frailty and Biomechanical Response of the CNS in NPH-Suspected Patients. J. Gerontol. A Biol. Sci. Med. Sci. 2022, 77, 1335–1343. [Google Scholar] [CrossRef]
  361. 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]
  362. Gallucci, M.; Zanardo, A.; Bendini, M.; Di Paola, F.; Boldrini, P.; Grossi, E. Serum folate, homocysteine, brain atrophy, and auto-CM system: The Treviso Dementia (TREDEM) study. J. Alzheimers Dis. 2014, 38, 581–587. [Google Scholar] [CrossRef] [PubMed]
  363. Braun, D.J.; Dimayuga, E.; Morganti, J.M.; Van Eldik, L.J. Microglial-associated responses to comorbid amyloid pathology and hyperhomocysteinemia in an aged knock-in mouse model of Alzheimer’s disease. J. Neuroinflamm. 2020, 17, 274. [Google Scholar] [CrossRef] [PubMed]
  364. Bae, J.B.; Han, J.W.; Song, J.; Lee, K.; Kim, T.H.; Kwak, K.P.; Kim, B.J.; Kim, S.G.; Kim, J.L.; Moon, S.W.; et al. Hypohomocysteinemia may increases the risk of dementia and Alzheimer’s disease: A nationwide population-based prospective cohort study. Clin. Nutr. 2021, 40, 4579–4584. [Google Scholar] [CrossRef] [PubMed]
  365. Elsherbiny, N.M.; Sharma, I.; Kira, D.; Alhusban, S.; Samra, Y.A.; Jadeja, R.; Martin, P.; Al-Shabrawey, M.; Tawfik, A. Homocysteine Induces Inflammation in Retina and Brain. Biomolecules 2020, 10, 393. [Google Scholar] [CrossRef]
  366. Farina, N.; Jerneren, F.; Turner, C.; Hart, K.; Tabet, N. Homocysteine concentrations in the cognitive progression of Alzheimer’s disease. Exp. Gerontol. 2017, 99, 146–150. [Google Scholar] [CrossRef]
  367. Kovalska, M.; Hnilicova, P.; Kalenska, D.; Adamkov, M.; Kovalska, L.; Lehotsky, J. Alzheimer’s Disease-like Pathological Features in the Dorsal Hippocampus of Wild-Type Rats Subjected to Methionine-Diet-Evoked Mild Hyperhomocysteinaemia. Cells 2023, 12, 2087. [Google Scholar] [CrossRef]
  368. Lee, C.C.; Hsu, S.W.; Huang, C.W.; Chang, W.N.; Chen, S.F.; Wu, M.K.; Chang, C.C.; Hwang, L.C.; Chen, P.C. Effects of Homocysteine on white matter diffusion parameters in Alzheimer’s disease. BMC Neurol. 2017, 17, 192. [Google Scholar] [CrossRef]
  369. Luzzi, S.; Cherubini, V.; Falsetti, L.; Viticchi, G.; Silvestrini, M.; Toraldo, A. Homocysteine, Cognitive Functions, and Degenerative Dementias: State of the Art. Biomedicines 2022, 10, 2741. [Google Scholar] [CrossRef]
  370. Shirafuji, N.; Hamano, T.; Yen, S.H.; Kanaan, N.M.; Yoshida, H.; Hayashi, K.; Ikawa, M.; Yamamura, O.; Kuriyama, M.; Nakamoto, Y. Homocysteine Increases Tau Phosphorylation, Truncation and Oligomerization. Int. J. Mol. Sci. 2018, 19, 891. [Google Scholar] [CrossRef]
  371. Song, Y.; Quan, M.; Li, T.; Jia, J. Serum Homocysteine, Vitamin B12, Folate, and Their Association with Mild Cognitive Impairment and Subtypes of Dementia. J. Alzheimers Dis. 2022, 90, 681–691. [Google Scholar] [CrossRef]
  372. Weekman, E.M.; Johnson, S.N.; Rogers, C.B.; Sudduth, T.L.; Xie, K.; Qiao, Q.; Fardo, D.W.; Bottiglieri, T.; Wilcock, D.M. Atorvastatin rescues hyperhomocysteinemia-induced cognitive deficits and neuroinflammatory gene changes. J. Neuroinflamm. 2023, 20, 199. [Google Scholar] [CrossRef]
  373. Weekman, E.M.; Sudduth, T.L.; Price, B.R.; Woolums, A.E.; Hawthorne, D.; Seaks, C.E.; Wilcock, D.M. Time course of neuropathological events in hyperhomocysteinemic amyloid depositing mice reveals early neuroinflammatory changes that precede amyloid changes and cerebrovascular events. J. Neuroinflamm. 2019, 16, 284. [Google Scholar] [CrossRef] [PubMed]
  374. Weekman, E.M.; Woolums, A.E.; Sudduth, T.L.; Wilcock, D.M. Hyperhomocysteinemia-Induced Gene Expression Changes in the Cell Types of the Brain. ASN Neuro 2017, 9, 1759091417742296. [Google Scholar] [CrossRef] [PubMed]
  375. Zuin, M.; Cervellati, C.; Brombo, G.; Trentini, A.; Roncon, L.; Zuliani, G. Elevated Blood Homocysteine and Risk of Alzheimer’s Dementia: An Updated Systematic Review and Meta-Analysis Based on Prospective Studies. J. Prev. Alzheimers Dis. 2021, 8, 329–334. [Google Scholar] [CrossRef]
  376. Vecchio, F.; Miraglia, F.; Judica, E.; Cotelli, M.; Alu, F.; Rossini, P.M. Human brain networks: A graph theoretical analysis of cortical connectivity normative database from EEG data in healthy elderly subjects. Geroscience 2020, 42, 575–584. [Google Scholar] [CrossRef] [PubMed]
  377. Bray, N.W.; Pieruccini-Faria, F.; Witt, S.T.; Bartha, R.; Doherty, T.J.; Nagamatsu, L.S.; Almeida, Q.J.; Liu-Ambrose, T.; Middleton, L.E.; Bherer, L.; et al. Combining exercise with cognitive training and vitamin D(3) to improve functional brain connectivity (FBC) in older adults with mild cognitive impairment (MCI). Results from the SYNERGIC trial. Geroscience 2023, 45, 1967–1985. [Google Scholar] [CrossRef]
  378. Bray, N.W.; Pieruccini-Faria, F.; Witt, S.T.; Rockwood, K.; Bartha, R.; Doherty, T.J.; Nagamatsu, L.S.; Almeida, Q.J.; Liu-Ambrose, T.; Middleton, L.E.; et al. Frailty and functional brain connectivity (FBC) in older adults with mild cognitive impairment (MCI): Baseline results from the SYNERGIC Trial. Geroscience 2023, 45, 1033–1048. [Google Scholar] [CrossRef]
  379. Czoch, A.; Kaposzta, Z.; Mukli, P.; Stylianou, O.; Eke, A.; Racz, F.S. Resting-state fractal brain connectivity is associated with impaired cognitive performance in healthy aging. Geroscience 2023. [Google Scholar] [CrossRef]
  380. Hardcastle, C.; Hausman, H.K.; Kraft, J.N.; Albizu, A.; Evangelista, N.D.; Boutzoukas, E.M.; O’Shea, A.; Langer, K.; Van Van Etten, E.; Bharadwaj, P.K.; et al. Higher-order resting state network association with the useful field of view task in older adults. Geroscience 2022, 44, 131–145. [Google Scholar] [CrossRef]
  381. Hausman, H.K.; Hardcastle, C.; Albizu, A.; Kraft, J.N.; Evangelista, N.D.; Boutzoukas, E.M.; Langer, K.; O’Shea, A.; Van Etten, E.J.; Bharadwaj, P.K.; et al. Cingulo-opercular and frontoparietal control network connectivity and executive functioning in older adults. Geroscience 2022, 44, 847–866. [Google Scholar] [CrossRef]
  382. Huang, C.Y.; Chen, L.C.; Wu, R.M.; Hwang, I.S. Effects of task prioritization on a postural-motor task in early-stage Parkinson’s disease: EEG connectivity and clinical implication. Geroscience 2022, 44, 2061–2075. [Google Scholar] [CrossRef] [PubMed]
  383. Kramer, C.; Stumme, J.; da Costa Campos, L.; Dellani, P.; Rubbert, C.; Caspers, J.; Caspers, S.; Jockwitz, C. Prediction of cognitive performance differences in older age from multimodal neuroimaging data. Geroscience 2023. [Google Scholar] [CrossRef] [PubMed]
  384. Miraglia, F.; Pappalettera, C.; Guglielmi, V.; Cacciotti, A.; Manenti, R.; Judica, E.; Vecchio, F.; Rossini, P.M. The combination of hyperventilation test and graph theory parameters to characterize EEG changes in mild cognitive impairment (MCI) condition. Geroscience 2023, 45, 1857–1867. [Google Scholar] [CrossRef] [PubMed]
  385. Pappalettera, C.; Miraglia, F.; Cotelli, M.; Rossini, P.M.; Vecchio, F. Analysis of complexity in the EEG activity of Parkinson’s disease patients by means of approximate entropy. Geroscience 2022, 44, 1599–1607. [Google Scholar] [CrossRef]
  386. Ren, P.; Ma, M.; Zhuang, Y.; Huang, J.; Tan, M.; Wu, D.; Luo, G. Dorsal and ventral fronto-amygdala networks underlie risky decision-making in age-related cognitive decline. Geroscience 2023. [Google Scholar] [CrossRef]
  387. Wang, Q.; Qi, L.; He, C.; Feng, H.; Xie, C.; Depression Imaging, R.C. Age- and gender-related dispersion of brain networks across the lifespan. Geroscience 2023. [Google Scholar] [CrossRef]
  388. Williamson, J.; James, S.A.; Mukli, P.; Yabluchanskiy, A.; Wu, D.H.; Sonntag, W.; Ciro, C.; Yang, Y. Sex difference in brain functional connectivity of hippocampus in Alzheimer’s disease. Front. Aging Neurosci. 2022, 14, 959394. [Google Scholar] [CrossRef]
  389. Garcia-Casares, N.; Bernal-Lopez, M.R.; Roe-Vellve, N.; Gutierrez-Bedmar, M.; Fernandez-Garcia, J.C.; Garcia-Arnes, J.A.; Ramos-Rodriguez, J.R.; Alfaro, F.; Santamaria-Fernandez, S.; Steward, T.; et al. Brain Functional Connectivity Is Modified by a Hypocaloric Mediterranean Diet and Physical Activity in Obese Women. Nutrients 2017, 9, 685. [Google Scholar] [CrossRef]
  390. Gaynor, A.M.; Varangis, E.; Song, S.; Gazes, Y.; Noofoory, D.; Babukutty, R.S.; Habeck, C.; Stern, Y.; Gu, Y. Diet moderates the effect of resting state functional connectivity on cognitive function. Sci. Rep. 2022, 12, 16080. [Google Scholar] [CrossRef]
  391. Li, T.; Willette, A.A.; Wang, Q.; Pollpeter, A.; Larsen, B.A.; Mohammadiarvejeh, P.; Fili, M. Alzheimer’s Disease Genetic Influences Impact the Associations between Diet and Resting-State Functional Connectivity: A Study from the UK Biobank. Nutrients 2023, 15, 3390. [Google Scholar] [CrossRef]
  392. Jiang, L.; Geng, W.; Chen, H.; Zhang, H.; Bo, F.; Mao, C.N.; Chen, Y.C.; Yin, X. Decreased functional connectivity within the default-mode network in acute brainstem ischemic stroke. Eur. J. Radiol. 2018, 105, 221–226. [Google Scholar] [CrossRef] [PubMed]
  393. Kong, Y.; Li, X.; Chang, L.; Liu, Y.; Jia, L.; Gao, L.; Ren, L. Hypertension with High Homocysteine Is Associated with Default Network Gray Matter Loss. Front. Neurol. 2021, 12, 740819. [Google Scholar] [CrossRef] [PubMed]
  394. 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 2022, 15, 47. [Google Scholar] [CrossRef] [PubMed]
  395. Fekete, M.; Szarvas, Z.; Fazekas-Pongor, V.; Lehoczki, A.; Tarantini, S.; Varga, J.T. Effects of omega-3 supplementation on quality of life, nutritional status, inflammatory parameters, lipid profile, exercise tolerance and inhaled medications in chronic obstructive pulmonary disease. Ann. Palliat. Med. 2022, 11, 2819–2829. [Google Scholar] [CrossRef] [PubMed]
  396. Li, B.; Yabluchanskiy, A.; Tarantini, S.; Allu, S.R.; Sencan-Egilmez, I.; Leng, J.; Alfadhel, M.A.H.; Porter, J.E.; Fu, B.; Ran, C.; et al. Measurements of cerebral microvascular blood flow, oxygenation, and morphology in a mouse model of whole-brain irradiation-induced cognitive impairment by two-photon microscopy and optical coherence tomography: Evidence for microvascular injury in the cerebral white matter. Geroscience 2023, 45, 1491–1510. [Google Scholar] [CrossRef]
  397. Negri, S.; Sanford, M.; Shi, H.; Tarantini, S. The role of endothelial TRP channels in age-related vascular cognitive impairment and dementia. Front. Aging Neurosci. 2023, 15, 1149820. [Google Scholar] [CrossRef]
  398. Sanford, M.; Negri, S.; Tarantini, S. Editorial: New developments in understanding brain and cerebromicrovascular aging: Toward prevention of vascular cognitive impairment and Alzheimer’s disease. Front. Aging Neurosci. 2022, 14, 1020271. [Google Scholar] [CrossRef]
  399. Tarantini, S.; Subramanian, M.; Butcher, J.T.; Yabluchanskiy, A.; Li, X.; Miller, R.A.; Balasubramanian, P. Revisiting adipose thermogenesis for delaying aging and age-related diseases: Opportunities and challenges. Ageing Res. Rev. 2023, 87, 101912. [Google Scholar] [CrossRef]
  400. Maroto-Rodriguez, J.; Delgado-Velandia, M.; Ortola, R.; Carballo-Casla, A.; Garcia-Esquinas, E.; Rodriguez-Artalejo, F.; Sotos-Prieto, M. Plant-based diets and risk of frailty in community-dwelling older adults: The Seniors-ENRICA-1 cohort. Geroscience 2023, 45, 221–232. [Google Scholar] [CrossRef]
  401. Dobreva, I.; Marston, L.; Mukadam, N. Which components of the Mediterranean diet are associated with dementia? A UK Biobank cohort study. Geroscience 2022, 44, 2541–2554. [Google Scholar] [CrossRef]
  402. Raffin, J.; Rolland, Y.; Aggarwal, G.; Nguyen, A.D.; Morley, J.E.; Li, Y.; Bateman, R.J.; Vellas, B.; Barreto, P.S.; Group, M.D. Associations Between Physical Activity, Blood-Based Biomarkers of Neurodegeneration, and Cognition in Healthy Older Adults: The MAPT Study. J. Gerontol. A Biol. Sci. Med. Sci. 2021, 76, 1382–1390. [Google Scholar] [CrossRef] [PubMed]
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

Ungvari, A.; Gulej, R.; Csik, B.; Mukli, P.; Negri, S.; Tarantini, S.; Yabluchanskiy, A.; Benyo, Z.; Csiszar, A.; Ungvari, Z. The Role of Methionine-Rich Diet in Unhealthy Cerebrovascular and Brain Aging: Mechanisms and Implications for Cognitive Impairment. Nutrients 2023, 15, 4662. https://doi.org/10.3390/nu15214662

AMA Style

Ungvari A, Gulej R, Csik B, Mukli P, Negri S, Tarantini S, Yabluchanskiy A, Benyo Z, Csiszar A, Ungvari Z. The Role of Methionine-Rich Diet in Unhealthy Cerebrovascular and Brain Aging: Mechanisms and Implications for Cognitive Impairment. Nutrients. 2023; 15(21):4662. https://doi.org/10.3390/nu15214662

Chicago/Turabian Style

Ungvari, Anna, Rafal Gulej, Boglarka Csik, Peter Mukli, Sharon Negri, Stefano Tarantini, Andriy Yabluchanskiy, Zoltan Benyo, Anna Csiszar, and Zoltan Ungvari. 2023. "The Role of Methionine-Rich Diet in Unhealthy Cerebrovascular and Brain Aging: Mechanisms and Implications for Cognitive Impairment" Nutrients 15, no. 21: 4662. https://doi.org/10.3390/nu15214662

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