Next Article in Journal
Draft Genome and Biological Characteristics of Fusarium solani and Fusarium oxysporum Causing Black Rot in Gastrodia elata
Next Article in Special Issue
The Molecular Mechanism of Polyphenols in the Regulation of Ageing Hallmarks
Previous Article in Journal
Intestinal Transplant Immunology and Intestinal Graft Rejection: From Basic Mechanisms to Potential Biomarkers
Previous Article in Special Issue
Influence of the Mediterranean Diet on Healthy Aging
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Vitamin D as a Shield against Aging

by
Cristina Fantini
1,†,
Clarissa Corinaldesi
2,†,
Andrea Lenzi
3,
Silvia Migliaccio
1 and
Clara Crescioli
1,*
1
Department of Movement, Human and Health Sciences, University of Rome “Foro Italico”, Piazza Lauro de Bosis, 006-00135 Rome, Italy
2
Institute for Cancer Genetics, Columbia University, New York, NY 10027, USA
3
Department of Experimental Medicine, Sapienza University of Rome, Viale del Policlinico, 155-00161 Rome, Italy
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2023, 24(5), 4546; https://doi.org/10.3390/ijms24054546
Submission received: 31 January 2023 / Revised: 17 February 2023 / Accepted: 22 February 2023 / Published: 25 February 2023
(This article belongs to the Special Issue The Molecular Links between Nutrients and Aging)

Abstract

:
Aging can be seen as a physiological progression of biomolecular damage and the accumulation of defective cellular components, which trigger and amplify the process, toward whole-body function weakening. Senescence initiates at the cellular level and consists in an inability to maintain homeostasis, characterized by the overexpression/aberrant expression of inflammatory/immune/stress responses. Aging is associated with significant modifications in immune system cells, toward a decline in immunosurveillance, which, in turn, leads to chronic elevation of inflammation/oxidative stress, increasing the risk of (co)morbidities. Albeit aging is a natural and unavoidable process, it can be regulated by some factors, like lifestyle and diet. Nutrition, indeed, tackles the mechanisms underlying molecular/cellular aging. Many micronutrients, i.e., vitamins and elements, can impact cell function. This review focuses on the role exerted by vitamin D in geroprotection, based on its ability to shape cellular/intracellular processes and drive the immune response toward immune protection against infections and age-related diseases. To this aim, the main biomolecular paths underlying immunosenescence and inflammaging are identified as biotargets of vitamin D. Topics such as heart and skeletal muscle cell function/dysfunction, depending on vitamin D status, are addressed, with comments on hypovitaminosis D correction by food and supplementation. Albeit research has progressed, still limitations exist in translating knowledge into clinical practice, making it necessary to focus attention on the role of vitamin D in aging, especially considering the growing number of older individuals.

1. Introduction

The undeniable increase observed in human life expectancy and longevity in developed countries is too often accompanied by a decrease in the quality of life (QoL) [1]. This means that living longer is not living better, thus, the current challenge is how to maintain and improve a good QoL during a longer lifespan. The aging process is a gradual functional decline, that starts early in adulthood, continues throughout the human lifespan, and might end in chronic diseases, e.g., sarcopenia, metabolic disorders, cardiovascular and neurodegenerative diseases, or cancer [2]. These macroscopic age-related alterations reflect the changes occurring at the cellular and molecular levels: mitochondrial dysfunction, damaged protein accumulation, epigenetic alterations, telomere shortening, aberrant intracellular signaling, and altered nutrient sensing are recognized to be the biomolecular pillars of aging [3,4]. Altogether, these molecular alterations converge in frailty, a multidimensional syndrome, clinically marked by decreased physiological reserve and resistance to stressors, leading to a higher rate of disability and mortality [5]. Frailty is undeniably a multifaceted condition, due to different determinants, however, the reduced capacity in immune defense, known as “immunosenescence”, and the consequent increase in chronic low-grade inflammation, named “inflammaging”, emerge as the main triggers. Whereas aging is a natural and unstoppable process, several adaptable lifestyle factors can favorably impact QoL, by strengthening the immune system and promoting anti-inflammatory processes.
It is well-known that diet can drive human health status toward wellbeing or disease. Indeed, the nutritional pattern per se plays a pivotal role in promoting healthy aging, by modulating several intracellular processes.
Among many factors, vitamin D plays an important role as a nutrient capable of affecting the aging process at a cellular/molecular level, with a wide range of actions.
Indeed, beside bone defects, insufficient vitamin D is associated with the increased risk of developing a wide range of pathologies, from neurological diseases to cancer, obesity, diabetes and metabolic disturbance, cardiovascular diseases, autoimmune diseases, infections, and menopause-related diseases [6,7,8,9,10,11,12,13].
A clear cause–effect relationship still lacks in many cases, nevertheless hypovitaminosis D is a marker for poor health, especially in the elderly, often characterized by co-morbid conditions [14].
Most of the actions of vitamin D are undoubtedly ascribable to its fine-tuned immunomodulatory properties, impacting immune system function and inflammation. This review aims to provide an overview of some of the cellular biomechanisms and biomolecules engaged in immunosenescence and inflammaging, as biotargets for vitamin D interventions during adult life, or even earlier, in order to maintain a level of health as high as possible while aging. Particular attention will be given to the ability of vitamin D to shape the human immune status, acting on the immune system or immune-related activity of some tissues, such as skeletal muscle or heart, whose function highly affects QoL during life, including in old age.

2. Aging Is a Cellular and Molecular Matter

Human aging is a complex process influenced by several factors, including sex, genetics, socioeconomic status, and lifestyle. Age-related changes are associated with higher susceptibility to diseases, depending on macroscopic modifications in the whole-body physiology and organ function, which, in turn, mirror the changes occurring at the cellular and intracellular levels. Accumulating data suggest that dietary pattern, among other factors, can prevent some molecular/cellular changes underlying human aging-related diseases, and maintain the functional abilities associated to wellbeing [World Health Organization Ageing: Healthy Ageing and Functional Ability 2020; https://www.who.int/news-room/questions-and-answers/item/healthy-ageing-and-functional-ability, (accessed on 26 October 2020)]. Some of the main processes triggering and mediating senescence and inflammation are discussed below.

2.1. The Cellular Route to Aging

Understanding the molecular drivers of age-dependent multimorbidity can help to develop new interventions and strategies to delay this condition [15].
The multimodal process of aging includes different biological system remodeling leading to the loss of homeostasis and, consequently, deterioration of several organs and tissues. In this scenario, age-related biomolecular changes in the immune response and immune cells are acknowledged to play a causal role in driving whole-body aging [16]. In humans, the natural decline in immune system function usually starts in the sixth decade, and tends to continuously progress toward immunosenescence and decreased protection against pathogens; at the same time, the inflammatory response increases, in terms of duration/intensity [17]. Age-dependent “re-shaping” of both the innate and adaptive arms of the immune landscape can increase vulnerability to illness, and promote and allow multimorbidity, frailty, and adverse health outcomes [18].
Indeed, whereas the adaptive immunity decreases and leads to weakened antigen-specific response and impaired memory formation, the nonspecific innate immunity overreacts, leaving older individuals unprotected from chronic inflammation. The concurrence of these conditions is defined as “inflammaging” [19,20]. In turn, inflammaging is an additional risk factor in the elderly, since it amplifies other age-associated modifications related to morbidities.
The macroscopic changes reflect microscopic remodeling at the cellular and subcellular levels, involving significant modifications in intracellular paths, especially in the T cell repertoire [21,22,23].
Cellular senescence combines morphological and molecular characteristics, makes cells modify cyclin-dependent kinases (CDKs) regulation, and undergo permanent cell cycle arrest, in a steady status different from terminal differentiation or quiescence [24,25]. The diluted cytoplasmic domain, the enlarged cell size, and increased β-galactosidase lysosomal activity—mostly used as a senescence cell marker—are the major features of senescent cells, along with an altered DNA and chromatin landscape, a damaged protein and lipid profile, and dysfunctional mitochondria and lysosomes [24,26,27]. Remarkably, senescent cells display a “senescence-associated secretory phenotype” (SASP) due to a specific secretory profile or “secretome”, which is characterized by an excessive secretion of proinflammatory molecules, such as cytokines and chemokines, matrix metalloproteinases, and growth and angiogenic modulators [28,29,30]. Figure 1 shows a schematic of the cellular changes associated with senescence.
In this scenario, the production and release of proinflammatory molecules increase, including interleukin (IL)-6, IL-1, tumor necrosis factor(TNF)α, interferon (IFN)γ, acute-phase proteins, reactive oxygen species (ROS), and autoantibodies constantly increase and accumulate. These factors function as mediators of signal communication between cells resident within organs/tissues, and immune cells, and thus amplify inflammaging [31,32].
To date, several in vitro and in vivo studies show that the senescence-related signature is highly variable, but a robust approach to establishing a senescent secretome profile is still lacking [33]. Nevertheless, senescence of the immunocytes is undeniably acknowledged as a critical step in triggering sustained chronic-persistent inflammation, therefore, representing a potential target for interventions.

2.2. Immunosenescence: The First Step toward Aging

The central defect in immunosenescence is the decline in T cell function, as shown by animal and human studies, leading to overactivity, inflammation, and autoantibody production, with quite predictable consequences on health [34,35,36,37].
The main age-dependent changes include a significant decrease in naïve T cells, along with a simultaneous increase in memory cells, impairments in the T cell receptor (TCR) repertoire, defects in NK cells and neutrophil chemotaxis, myeloid skewing, and monocyte dysregulation. Albeit immunosenescence involves most of the immunity-related components, T cells emerge as playing a triggering role in aging, as shown by investigations on rheumatoid arthritis (RA), an autoimmune disease in which the faster progression of this process allowed the characterization of some underlying molecular paths [38,39,40].
Age-related thymus involution is the first event that triggers impairments and defects in T cell differentiation and maturation. The number of new naïve T cells decreases, and circulating naïve T cells live longer, accumulate defects, and cause a decreased TCR diversity [34]. T cell-TCR signaling activation is required for IL-2 release and T cell expansion, which are known to be critical for mounting an effective immune response [41]. In aging, the downtuning of T regulatory (Treg) cells leads to an unbalanced Th17/Treg ratio, with Th17 upregulation and Treg decrease, ending in polarization of the immune response towards inflammation [42].
Furthermore, age-related thymic involution, naïve T cell dysfunction, and aberrant expression of age-associated gene profiles seem to be linked to IL-33/ST2 signaling activation, suggesting that targeting IL-33 or ST2 could be a promising strategy to rejuvenate T cell immunity [43]. IL-33 has been shown to orchestrate the signals within the skeletal muscle/immune system/nervous system in response to injury. Remarkably, the IL-33/ST2 axis is suggested as a promising potential target in managing age-related sarcopenia and muscle repair due to injury or atrophy, both major problems impacting mobility and QoL in the elderly [44].
The dysregulation of T cell signaling includes defects in calcium mobilization, phosphorylation of tyrosine and serine/threonine, mitogen-activated protein kinases (MAPK) activity, and activation of transcription factors such as nuclear factor kB (NF-kB), nuclear factor of activated T cells (NFAT), or activator protein 1 (AP-1) [45,46,47,48,49,50,51,52].
Other age-dependent damage, i.e., the loss of co-stimulatory molecule CD28—which plays a pivotal role in cell activation/proliferation/survival—impacts on the function of B cells (proliferation and Ig production) and the antigenspecific cytotoxic CD8 T cell subset [34].
In addition to T cell deregulation, aberrant increases in NF-kB activation and cyclooxygenase (COX)-2 expression, with a higher production of prostaglandin E2 (PGE2) from macrophages, have been reported in the elderly [53,54,55].
Remarkably, NF-kB dysregulation/hyperactivation represents a potential molecular target for intervention, with particular significance in the elderly, as addressed later in this review.
The number of natural killer (NK) cells, which are known to have cytotoxic/lytic activity against cancers and viruses, seems to not vary with aging, whereas cytokine and chemokine production declines [56,57,58]. Like NK, neutrophil number does not change with age, but neutrophil activity, including oxidative burst, phagocytosis, and chemotaxis (all defense mechanisms) is significantly compromised [57].
In the complex network of aging, oxidative stress is highly determinant, as suggested by the coined term “oxi-inflamm-aging” [59]. Indeed, the high percentage of polyunsaturated fatty acids present in immune cell plasma makes the cells inclined to lipid peroxidation; any damage in the oxidative process can alter signal transmission within/between several types of immunocytes, ending in a defective immune response [60,61].
Thus far, age-related changes in the immune system expose the elderly to a higher risk of disease, first of all, cancer, and infections.
The significant impact of genetic and environmental factors on immune system function is undeniable, and nutrition emerges as an acknowledged tool to regulate the immune status. Indeed, age-related immune alterations seem to be associated with a suboptimal nutrient status, so that nutritional interventions with macronutrients (i.e., polyunsatured fatty acids or PUFA) or micronutrients (i.e., vitamins, elements) are highly recommended for ameliorating QoL in the elderly. Whereas macronutrients provide a substrate for the biosynthesis of molecules engaged in immune response (acute-phase proteins, cytokines, new receptors, amino acids for immunoglobulins), and are fuel for immune cell energy, micronutrients, such as zinc, iron, and vitamins are exquisite regulators of the immune response at the cellular and molecular level [62].
Overall, this review focuses on vitamin D as a fine-tuned regulator of both the innate and adaptive immune responses, playing a pivotal role in the health of individuals of any age, particularly, in the elderly.

3. Vitamin D and Geroprotection

Vitamin D has historically been known as the necessary nutrient to guarantee a correct bone metabolism and health. This molecule is classically defined as a steroid hormone, since it shares with steroids the common progenitor molecule (cyclopentanoperhydrophenanthrene). It is synthesized as the precursor molecule in skin exposed to sufficient UV rays, and is transformed into the biologically active compound by two enzymatic hydroxylations in the liver and kidneys (25- and 1-α-hydroxylase, respectively). Therefore, while “vitamin” is considered an imprecise term [63], this molecule is present in food such as fatty fish, i.e., salmon and trout, beef liver, cheese, egg yolk, albeit the latter ones provide small amounts, mushrooms and some vegetables, and is fully identified as an essential micronutrient with critical regulatory functions [64].
Vitamin D deficiency is a worldwide problem, affecting people of all ages, due to several impacting factors, from ethnicity and skin color, to latitude, habits and lifestyle, sex, and age. It has been reported that the elderly in Europe, the USA, and Australia suffer from insufficient levels of vitamin D, mainly due to less solar exposure, associated with low levels of outdoor activity or with clothing, and also due to reduced synthesis, due to atrophic skin modifications and a lower amount of precursor or reduced renal function [65,66]. Remarkably, a less varied diet, with a lower vitamin D content, is a typical age-associated habit, contributing to vitamin D deficiency in aging populations [67,68].
Thus far, vitamin D intake to a sufficient level is recommended not only for bone homeostasis but to maintain a general status of good heath as well, especially during aging, considering its pleiotropic effects. Vitamin D intake from dietary sources, in particular foods enriched or fortified with this nutrient, is considered an excellent strategy to counteract vitamin D deficiency, a condition currently defined as being a “world problem” [69].
It is acknowledged that this nutrient exerts such an important effect on the immune response that it can be considered a tool to tackle immunosenescence and oxi-inflamm-aging. In general, vitamin D deficiency is associated with a higher risk of infections and autoimmune diseases, involving dysfunctional biological activity of the specific vitamin D receptor (VDR), which is expressed in the majority of immunocytes [70]. Hypovitaminosis D is acknowledged as a pandemic condition, affecting all stages of life, with detrimental consequences on health; but in the elderly, vitamin D deficiency retains even more clinical significance, since this condition often converges with other age-related deficiencies (i.e., hormonal) or diseases, worsening the outcome.
Most of the immune cells, including T and B cells, dendritic cells (DC), macrophages and monocytes, express VDR and respond to vitamin D with fine-tuned modulations in cell signaling, path activation, and molecule production, with significant consequences on immune response [71,72,73,74,75,76,77,78].
Furthermore, many immunocytes express 1-α-hydroxylase and can themselves produce the active metabolite and, thus, control the local cell microenvironment [79,80].
In the elderly, adequate vitamin D levels help to counteract the natural decline in immune surveillance by a fine-tuned orchestration of several effects.

3.1. Vitamin D against Infections

Vitamin D strengthens the first line of host defense, that is particularly relevant during aging, when the risk of infection is higher. Indeed, this molecule can maintain the barrier integrity and induce a set of genes encoding antimicrobial proteins (AMPs), such as cathelicidin, defensins, hepcidin, and neutrophil peptides, which behave as antibiotics against various types of pathogens [81,82].
The vitamin D–cathelicidin axis is the most studied and best characterized among the VDR-dependent signaling engaged in infections and autophagosome formation—the latter one playing a critical role in microorganism clearance and infection resolution. Clinical studies on human tuberculosis, sepsis, viral infection, peritonitis, and pneumonia, whose incidence rise in the elderly [83], document that, after vitamin D supplementation, serum cathelicidin—human cathelicidin LL-37 or human cationic AMP 18 (hCAP-18)—increases, and correlates with improved clinical outcomes [84,85,86].
Evidence in human monocytes/macrophages shows that the activation of vitamin D signaling, triggered by toll-like receptor (TLR)2/1 or TLR8, leads to an increase in antimicrobial response, autophagy, antimicrobial peptide expression, and phagosome–lysosome fusion, in association with an increase in IFNγ, which is likely required to strengthen antibacterial activity, together with IL-12 and IL-18 [87,88,89,90].
Furthermore, IFNγ expression, in combination with CD40–CD40 ligand signaling, increases the activity of the hydroxylase converting 25-hydroxyvitamin D (25D) to the active metabolite, in human monocytes [91].
In Mycobacterium tuberculosis, the most studied infection, vitamin D-induced cathelicidin acts as a second messenger to activate autophagy genes, such as autophagy related 5 (ATG5) and beclin-1 (BECN1), and triggers a downstream wide signaling cascade, including intracellular calcium (Ca2+) release, Ca2+-dependent kinases, extracellular ATP-gated ion channel, purinergic receptor (P2X) 7, mammalian target of rapamycin (mTOR)/AMP-activated protein kinase (AMPK)/phosphoinositide 3-kinase(PI3K) pathway, and reactive ROS signaling [92,93,94].
During Mycobacterium infection, vitamin D boosts cytokines/chemokines production through the induction of IL-1β, that regulates defensin beta 4 gene (DEFB4), encoding for human beta-defensin-2 (HBD2) in macrophages [95].
Thus far, vitamin D/VDR signaling acts as a fine regulator of AMP-dependent autophagy, cytokines/chemokines production, IFN-dependent signaling, and ROS generation [71,96,97].
Since ROS mediates TLR2-induced cathelicidin expression in human monocytes/macrophages [98], and ROS-autophagy events are mutually regulated [99], further understanding the function of the vitamin D–cathelicidin axis in redox homeostasis and autophagy activity seems necessary. The antimicrobial action(s) of host defense proteins induced by vitamin D might be particularly relevant, since infections from drug-resistant pathogens are emerging worldwide [100].
Remarkably, vitamin D-mediated LL-37 induction is specific in human defense system, as this vitamin D function has been reported to fail in a murine system [101]. This observation is in line with the concept that, results from animals are often not translatable to humans, especially concerning studies on immunity [96].
However, the specific role of the vitamin D–cathelicidin axis in the different infections caused by bacteria, viruses, and parasites goes beyond the aim of this paper, and was exhaustively reported in another recent paper [88].
To date, whereas inflammatory cytokines such as TNF-α and IL-1β are central within vitamin D-mediated antibacterial activity during host defense, alterations in the vitamin D–cathelicidin axis, leading to lower vitamin D and excessive cathelicidin, result in overinflammation, and take part in the pathological setting of chronic inflammatory diseases, such as rosacea, a disease whose prevalence increases with age [102,103].
The paradoxical effect of vitamin D emerges; while this molecule works on the one hand to keep active the macrophage phenotype (M1), associated with proinflammatory/anticancer activity, to preserve physiological integrity, on the other hand, it helps to maintain the pro-tolerogenic/anti-inflammatory phenotype (M2), to counteract inflammaging, that is the route opening to a variety of diseases. The main anti-inflammaging effects related to vitamin D are addressed in the following paragraph. Figure 2 depicts the effects of vitamin D against inflammation, as well as some other molecular effects, reported in the following paragraphs.

3.2. Vitamin D against Inflammaging, Cellular Senescence, and Mitochondrial Dysfunction

Following the original concept, inflammaging is a consequence of immunosenescence, even if this model has been re-interpreted in favor of a mutual interplay [104].
The immunomodulatory role of vitamin D in inflammation is widely recognized. Substantially, an adequate level of vitamin D counteracts inflammation with multilevel targeting effects, i.e., inhibiting the expression and signaling of TLR2, 4, and 9, reducing the production of cytokines such as TNF-α, IL-6, IL-23, and IL-1, and repressing the activity of T cells recruiting chemokines [105].
The main inhibitory effects are on CD4+ and CD8+ T cell proliferation, in particular T helper 1 (Th1) cells (a subset of CD4+ effector T cells), in which vitamin D counteracts the release of cytokines, i.e., IL-2 or IFNγ, capable of activating macrophages, and IL-6, IL-8, IL-12, and TNFα, all molecules characterizing inflammaging [106,107].
The vitamin D-induced downregulation of the proinflammatory Th17 cell subset, suppression of dendritic cell (DC) maturation from monocytes, and impaired capacity to present antigens, occur together with Treg enhancement and an increase in anti-inflammatory cytokines such as IL-4, IL-5, IL-10, and CCL2. Altogether, these processes are supposed to be the main mechanisms underlying Th2 protolerogenic subset expansion, which, in turn, is able to mitigate inflammaging and autoimmune disorders [108,109,110,111,112,113].
In lymphocytes, vitamin D has been documented to inhibit IL-6, a critical factor in stimulating the Th17 cell subset, which plays a pivotal role in autoimmune reactions [114].
From in vitro studies on autoimmune diseases, the vitamin D-induced protolerogenic effect on T cells seems undeniable, but less effectiveness in vivo is hypothesized, maybe related to a different ability of T cells to modify their phenotype in response to vitamin D (more phenotypically committed, less responding to vitamin D) [115]. The precise biomolecular mechanism remains to be elucidated. Albeit the etiology of autoimmune diseases is complex and undeniably multifactorial, the role of hypovitaminosis D is acknowledged as highly impacting disease development [116].
In some autoimmune diseases, such as osteoarthritis, psoriasis-associated osteoporosis, and Guillain–Barré syndrome, a pathogenetic immunological crosstalk between vitamin D and IL-33 may be hypothesized, with the combination of hypovitaminosis D–IL-33/ST2 axis activation converging in deleterious effects. Hence, the neutralization of IL-33/ST2 signaling by vitamin D in immunocytes (T cells, DC) is suggested as a therapeutic strategy [117,118,119].
In this scenario, the Th17/Treg ratio tends to decrease simultaneously with a significant increase in the transcription factor fork head box (Fox)P3, the most reliable marker to date for Treg cells, and IL-10 [120].
Remarkably, IL-10 is the anti-inflammatory cytokine known to keep inflammaging and antigenic stress under control, thus representing a good molecular defense mechanism in the elderly [121]. In line with this concept, the poorer vitamin D status—i.e., due to reduced sunlight exposure, declined ability of the skin to produce vitamin D, malnourishment, or decreased vitamin D intake—observed in elderly people, likely contributes to the higher prevalence of a variety of age-related diseases associated with a compromised immune system [122,123,124,125]. As shown in COVID-19 patients, a wide-epigenetic T cell remodeling, promoting VDR expression and enzyme cytochrome P450, family 27, subfamily B, member 1 (CYP27B1) activation in autocrine/paracrine mode, likely underlies the transition from proinflammatory IFN-γ+ Th1 cells (via STAT3, c-JUN, and BACH2) to suppressive IL-10+ cells (via IL-6–STAT3 signaling) [126,127]. In addition to the effect on immunocytes, vitamin D deficiency increases endothelial senescence, allowing vascular dysfunction and atherosclerosis, both processes that increase in prevalence with aging.
This effect is undeniably related to vitamin D’s anti-inflammatory action, as inflammation is a recognized trigger of atherosclerosis initiation, progression, and plaque and thrombus formation, also in the young [128]. However, the vitamin D-induced reduction in cholesterol uptake by macrophages and, in turn, suppression of foam-cell formation, emerges as a key event [129].
So far, hypovitaminosis D supports increased cellular senescence and arterial aging, characterized by the typical dysfunctions as a gradual loss of vascular smooth muscle cells’ contractility, and increased arterial permeability and intima thickness [130].
The delayed cellular senescence by vitamin D/VDR interaction includes different processes, including longer telomere lengths, an increased antioxidant effect via nuclear factor erythroid 2-related factor 2 (Nrf2) transcriptional regulation, decreased oxidative stress, DNA damage, and SASP, downregulation of p16, p53, and p21 (major regulators of the G1/S cell-cycle checkpoint), while upregulating Bmi1(polycomb ring finger oncogene, transcriptional suppressor), whose novel action is reported in cardiac regulation [131,132].
Another important aspect is the regulation of mitochondrial function. Vitamin D deficiency is associated with disorders of mitochondrial function, such as respiratory chain deregulation, with downregulation of mRNA and proteins involved in mitochondrial respiration, inhibition of sirtuin (SIRT) 1, which plays a pivotal role in mitochondrial biogenesis through PGC-1α, and, e.g., in brain aging delay [133,134].
The combination of vitamin D and curcumin, given as a supplement in a continuous way, has been recently hypothesized to counteract neurodegeneration [135]. SIRT signaling and lifespan seem deeply affected by dietary nutrient composition and resveratrol, which suppresses oxidant and inflammatory genes, altering promoter epigenetic status. The immune system is key to a host’s defense against pathogenic organisms. Aging is associated with changes in the immune system, with a decline in protective components (immunosenescence), increased susceptibility to infectious disease, and a chronic elevation in low-grade inflammation (inflammaging), increasing the risk of multiple noncommunicable diseases. Nutrition is a determinant of immune cell function and of the gut microbiota. In turn, the gut microbiota shapes and controls the immune and inflammatory responses. Many older people show changes in their gut microbiota. Age-related changes in immune competence, low-grade inflammation, and gut dysbiosis may be interlinked, and may relate, at least in part, to age-related changes in nutrition. A number of micronutrients (vitamins C, D, and E, and zinc and selenium) play roles in supporting the functions of many immune cell types. Some trials have reported that providing these micronutrients as individual supplements can reverse immune deficits in older people, and/or in those with insufficient intakes. There is inconsistent evidence that this will reduce the risk or severity of infections, including respiratory infections. Probiotic, prebiotic, or synbiotic strategies, that modulate the gut microbiota, especially by promoting the colonization of lactobacilli and bifidobacteria, have been demonstrated to modulate some immune and inflammatory biomarkers in older people, and in some cases, to reduce the risk and severity of gastrointestinal and respiratory infections, although, again, the evidence is inconsistent. Further research, with well-designed and well-powered trials, in at-risk older populations is required to be more certain about the role of micronutrients and of strategies that modify the gut microbiota–host relationship in protecting against infection, especially respiratory infection [136,137]. COX-2 expression has been shown to be dose-dependently inhibited by vitamin D in murine macrophages, through suppression of the protein kinase B (PKB) or Akt/NF-kB/COX-2 pathway [138].
The function of NF-kB is critical in cellular senescence and in inflammaging, since this factor is suggested to serve as a linkage between aging hallmarks in cell–cell communication, and in age-related pathophysiological mechanisms [139,140]. Indeed, while NF-κB unaltered signaling supports the correct interplay between immunocytes and non-immune cells, to maintain a functional host response, NF-kB hyperphosphorylation and enhanced activity are reported in different tissue aging (skin, hypothalamus, and cortical tissues) [141,142]. Similarly, NF-kB constitutive activation is reported in aged skeletal muscle [143].
To date, in vivo data in overweight/obese, but otherwise healthy, individuals failed to show any significant difference in NF-kB activity; however, the inadequacy of sample size to detect such a difference as a primary outcome is indicated by the authors as a possible reason for their null findings [144]. Nevertheless, the function of NF-kB as a regulator of inflammation and immunity continues to emerge as being critical for aging and age-related diseases [145]. NF-kB signaling hyperactivation is associated with pro-aging stimuli, and correlates with the development/progression of most aging-related diseases, whereas NF-kB inhibition can delay or even reverse aging processes [146,147,148,149].
Thus far, inflammaging likely primes inflammatory NF-kB signaling, which is involved in a variety of age-related tissue/organ dysfunctions. Interestingly, vitamin D can target hyperphosphorylation of this transcription factor in several cell types. The following paragraph focuses on this topic.

4. Vitamin D-Dependent NF-kB Regulation: A Putative Anti-Aging Strategy

NF-kB can be considered as the central crossroad where many paths and signals, either promoting or delaying aging, converge, by activating or inhibiting this transcriptional factor, respectively [147]. Biomolecular paths promoting aging such as insulin (I)/insulin growth factor (IGF)-1 signaling, activate NF-kB via the PI3K/AKT cascade and mTOR, a known pro-aging factor [146,150,151,152]. In addition, age-dependent DNA damage and telomere shortening are associated with NF-kB aberrant activation, and with increased levels of COX-2 and ROS [153]. NF-kB activation, besides telomere shortening via telomerase reverse transcriptase (TERT) catalytic subunit activation, intensifies inflammation by macrophage polarization to M1 phenotype, and IL-6 and TNFα upregulation [154,155], and increases SASP and cellular senescence [147]. At variance to this, pro-longevity factors, i.e., sirtuin and FOXO, repress NF-kB transcriptional activity, by directly interacting with the p65 subunit [156,157,158]. To date, chronic activation of NF-kB is found in several age-related diseases, e.g., atherosclerosis, osteoporosis, muscular atrophy, and neurodegeneration [159,160,161,162,163].
Quite remarkably, vitamin D can downregulate and suppress NF-kB expression and activity, significantly limiting inflammation. The mechanism whereby the vitamin D/VDR system targets NF-kB includes a physical interaction between VDR and IκB kinase β (IKKβ), which is enhanced by vitamin D, resulting in IκBα stabilization and p65/p50 nuclear translocation impairment [164]. The blockade of p65 translocation leads to the decrease or suppression of NF-kB activation and transcriptional activity, as shown in different cell types stimulated with TNFα—the prototypic cytokine inducing this nuclear factor—not only in immune cells, i.e., DC or B cells, but also in different types of organ resident cells, like human cardiac cells, skeletal muscle cells, thyrocytes, and nucleus pulposus cells [165,166,167,168,169,170].
Thus far, a well-functioning vitamin D/VDR system represents a helpful tool for tackling inflammation and aging, through the downregulation of the TNF-α/NF-kB/p65 signaling cascade in immune system cells and in different tissue cells. This effect seems particularly relevant at the level of striated cells, such as cardiomyocytes or skeletal muscle cells.

Vitamin D Regulates Heart and Skeletal Muscle Functioning

It is acknowledged that aging per se is a risk factor for heart dysfunction and disease. Age-related cardiac reduced function reflects molecular and cellular modifications, both in non-cardiomyocyte-based components, i.e., vascular cells, fibroblasts, and extracellular matrix, and in cardiac cells, which undergo aberrant processes due to oxidative stress, inflammation, defects in metabolism, cellular repair, telomeres, alteration in gene expression, and post-translational modifications. The dogma that cardiomyocytes are postmitotic cells, terminally differentiated, and unable to undergo cell division, is overcome by the observation that a fraction of cells proliferate and divide in the hearts of the young, adults and the elderly, albeit the cardiomyocyte numbers undeniably decrease with aging, by necrosis [171].
The increased number of necrotic cells likely triggers repair-related aberrant inflammation, and a higher chance of autoantigen generation, due to oxidatively modified proteins [172]. Following oxi-inflamm-aging-induced cell death, garbage molecules such as lipoxidation products and advanced glycation end (AGE) products, accumulate from alterations in the protein degradation machinery and autophagy.
Concerning autophagy and apoptosis, it should be mentioned that a functional vitamin D/VDR system—rather than a sufficient vitamin D level alone—can regulate senescence-induced signaling, either by genomic and non-genomic mechanisms, in immunocytes and non-immune cells, retaining the potentiality to be used as a helpful tool against diseases related to inflammation, oxidative stress, or cancer [173,174,175,176].
These processes would impact gene expression toward aging memorization and progress [177]. During cardiac physiological aging, ROS generation is associated with the activation of the inflammasome NOD-like receptor protein 3 (NLRP3), production of biologically active IL-1β and IL-18, and, in turn, pyroptosis, a process causing cell membrane pore formation/rupture and cell death [178,179,180,181].
The activation of NLRP3 is a multistep process, occurring either with or without a priming signal first, denoted as “canonical” or “noncanonical” activation, respectively [182,183,184]. The difference between the canonical/noncanonical mechanisms are exhaustively reported elsewhere, and are beyond the aim of this review.
Independent of the mechanism of activation, an aberrant inflammasome is widely involved in aging, and in an extraordinary number of human age-related diseases. Of note, vitamin D/VDR signaling directly acts as a negative regulator of NLRP3 oligomerization/assembling/activation and IL-1β release, inhibiting pyroptosis [185,186,187].
There is evidence for IL-1β and IL-18 increase from VDR knockout macrophages, activated either through the canonical or noncanonical path [187]. Indeed, ligand-activated VDR attenuates NRLP3 deubiquitination, increasing the level of mitochondrial membrane uncoupling protein-2 (UCP2), which can directly prevent ROS production [188].
Albeit NLRP3 inflammasome activation is known to play a pivotal role in host defense from viral and bacterial infections, dysregulated or excessive activation of the inflammasome is associated with poor outcomes [189].
Similar to cardiac cells, skeletal muscle cells show age-dependent biomolecular dysregulation, associated with mitochondrial dysfunction, oxi-inflamm-aging, metabolic disturbance protein breakdown, accumulation of senescent cells, and tissue atrophy [190,191,192].
Age-related decline of skeletal muscle can lead to sarcopenia—not limited to muscle area/mass reduction, but extended to tissue functionality—and frailty [5,193]. In muscle as well, aging-related macroscopic changes reflect biomolecular and cellular modifications. Indeed, tissue changes and deficit are associated with an increase in several inflammatory cytokines, including TNFα, IL-1α/β, IL-6, IL-8, IFNγ, and their soluble receptors (sR) IL-1Ra, TNFαsR, and IL-6sR, referred to as “gerokines”, which altogether constitute the “aging secretome”, likely playing a causal role (rather than being simply markers) in inflammaging [19,194,195]. Even though there is still much to understand about how oxi-inflamm-aging translates into skeletal muscle decline, it is undeniable that the reduction in type II muscle fiber number and size, promoted by TNFα-induced apoptosis, and the unbalance in protein synthesis/degradation, induced by ROS and oxidative stress, merge into functional and mechanical muscle impairment [196,197]. The rise in TNFα or IL-1β increases fat mass toward an unfavorable muscle/fat ratio, which typically characterizes the elderly, especially those with sedentary habits [197]. To date, some intracellular cascades, particularly the TNFα/NF-kB/ubiquitin–proteasome system (UPS) cascade/Akt signaling, play a large part in skeletal muscle inflammaging, representing important targets to counteract skeletal muscle decline. Some strategies, such as protein supplementation, can reduce creatine kinase, but cannot affect circulating gerokines [198,199,200,201]. Another feature of aging muscle is the “differential” resistance to I regarding glucose, protein, and lipid metabolism. Indeed, in the elderly it is common that sensitivity to I concerning glucose is maintained, likely related to reduced utilization of peripheral glucose, while almost simultaneously the so-called “anabolic resistance” to I develops, due to protein synthesis/degradation imbalance (proteostasis loss), a condition preceding clinical manifestations [202,203,204,205,206].
The protective function exerted by vitamin D on skeletal muscle is acknowledged: i.e., vitamin D can directly interfere with and counteract the intracellular paths mediating tissue atrophy, such as proto-oncogene tyrosine-protein kinase Src/extracellular signal-regulated kinases 1 and 2 (ERK1/2)/Akt/forkhead box O3 (FOXO3), a signaling cascade, which upon activation, upregulates atrophic markers such as Atrogin-1 and MuRF1 [207].
An adequate intake of vitamin D can delay the aberrant processes associated with aging and age-related diseases, essentially restoring mitochondrial function and counteracting oxi-inflamm-aging. In myocytes, vitamin D can limit or even neutralize oxidative stress and ROS generation, and the accumulation and expression of garbage molecules like AGE/AGE receptor (RAGE), and high-fat diet-induced I resistance and myosteatosis [208,209,210].
Some experimental and human studies have shown that vitamin D is an effective antioxidant, via activation of the Nrf2-Keap1 antioxidant pathway (the main inducible defense against oxidative stress), showing, e.g., higher capacity than vitamin E to reduce zinc-induced oxidative stress in the central nervous system [211,212,213].
Of interest, in skeletal myocytes, vitamin D anti-aging effects are associated with the inhibition of NF-kB, the crossroad where different oxi-inflamm-aging/senescence paths merge, as previously reported [164,167,214].
Thus far, given the evidence on vitamin D’s role in regulating myocyte metabolism/mitochondrial function/ROS generation, vitamin D deficiency (below 25 nmol/l) is of great interest, particularly in the elderly, since this condition is common among community-dwelling elderly, and very common among institutionalized elderly [67].
To date, it should be mentioned that, while undeniably hypovitaminosis D enhances oxi-inflamm-aging and senescence, an overcorrection of vitamin D status may negatively impact on skeletal muscle cell metabolism, similarly to an overdose of antioxidants. Numerous investigations, indeed, have reported on the potential harmful effects of antioxidant (over)supplementation, especially on muscle fiber formation/muscle regeneration, metabolic homeostasis, and mitochondrial biogenesis [207,215,216,217,218].

5. Conclusions

The effect of vitamin D deficiency is classically related to reduced musculo-skeletal functions, and increased risk of disability in locomotion. Besides this undeniable effect, low levels of vitamin D per se strongly affect the aging process, as this molecule regulates cell homeostasis, counteracting oxi-inflamm-aging and cellular senescence with multitargeting actions. In this scenario, the challenge is to restore an adequate vitamin D level.
Some years ago, fortified foods and drinks were introduced, to supplement diet, e.g., some cereals, plant-based beverages like soymilk, orange juice, and some yogurt and cheese. Fortified flour, cornflakes, and juices are more common in the USA, while in Europe, vitamin D enriched margarine, vegetable oil, and milk are more common [219,220,221].
However, from previous simulation studies, fortified foods seem not to be able to provide sufficient vitamin D to the elderly, therefore, individual supplementation is proposed, often in combination with calcium. A supply of about 20 µg vitamin D (800 IU) per day to people over 70, with 50 µg/day recommended as the safe level, and 600 IU, are recommended in subjects from 1 to 70 years of age. Supplementation with 20 µg vitamin D and 1000/1200 mg calcium increases the vitamin D level, suppresses secondary hyperparathyroidism (lowering the parathyroid hormone/PTH), and improves bone and muscle strength [67]. As found in several studies aimed at reducing osteoporotic fractures in the elderly, a lower amount of vitamin D is required (10 µg/day) in the presence of calcium supplementation (1000 mg) [222].
To date, research studies have explored the efficacy of vitamin D supplementation in the prevention and treatment of chronic conditions of aging, nevertheless the translation to standardized application is lacking, for a number of reasons. There is still an urgent need of clear indications as to what is the best dosage for a supplement, considering, e.g., the comorbidities. Moreover, the lack of a universally accepted standard for vitamin D assessment, i.e., due to variability in methods, strictly connects with the absence of clarity in vitamin D status definition (sufficiency/insufficiency/deficiency), based on a standardized reference range. Plasma 25(OH)D (the more stable analyte) is currently used to assess vitamin D status, but it is recognized that the circulating level of this metabolite varies depending upon season, latitude, clothing, dietary habits, race, pigmentation, skin thickness, sex, and age [67]. All the concerns about vitamin D determination and supplementation are exhaustively reported in the latest consensus statement from the 2nd International Conference on Controversies in Vitamin D [223].
So far, the role of vitamin D status in immunosenescence, inflammaging, and whole-body aging is based on scientifically documented data, and is fully recognized by the scientific literature, but still there are important limitations to translate knowledge into clinical practice, with important medical and socioeconomic consequences, considering the high and growing number of individuals aged 65 and older.
The DO-HEALTH multicenter clinical trial, currently running in 2157 community-dwelling European men and women aged 70 and older, combines vitamin D (2000 IU/day) treatment with omega-3 fatty acids intake (1000 mg/day), and a 30-min physical activity (3 times/week home exercise). This trial addresses several health domains (cardiovascular, muscle, bone, brain, and immunity) and hopefully will help to implement clinical practice [224].
This review wants to focus as much attention as possible on these aspects, even though some limitations, such as lack of discussion on sex-dependent variability, or differences between active and sedentary elderly, are present.
Given the importance of this topic, more accurate basic and clinical human research is necessary to approach vitamin D status determination, opening the way to future scenarios of personalized treatment.

Author Contributions

Article research and collection, first draft writing, C.F. and C.C. (Clarissa Corinaldesi); manuscript critical revision, A.L. and S.M.; conceptualization, final manuscript revision and writing, C.C. (Clara Crescioli). All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

Thanks to Riccardo Berti, University of Rome Foro Italico, for his kind assistance in reference preparation.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kirkwood, T.B.L. Why and how are we living longer?: Why and how are we living longer? Exp. Physiol. 2017, 102, 1067–1074. [Google Scholar] [CrossRef] [Green Version]
  2. Marsman, D.; Belsky, D.W.; Gregori, D.; Johnson, M.A.; Low Dog, T.; Meydani, S.; Pigat, S.; Sadana, R.; Shao, A.; Griffiths, J.C. Healthy ageing: The natural consequences of good nutrition—A conference report. Eur. J. Nutr. 2018, 57, 15–34. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Johnson, A.A.; Stolzing, A. The role of lipid metabolism in aging, lifespan regulation, and age-related disease. Aging Cell 2019, 18, e13048. [Google Scholar] [CrossRef] [Green Version]
  4. Johnson, A.A.; Shokhirev, M.N.; Wyss-Coray, T.; Lehallier, B. Systematic review and analysis of human proteomics aging studies unveils a novel proteomic aging clock and identifies key processes that change with age. Ageing Res. Rev. 2020, 60, 101070. [Google Scholar] [CrossRef] [PubMed]
  5. Cesari, M.; Calvani, R.; Marzetti, E. Frailty in Older Persons. Clin. Geriatr. Med. 2017, 33, 293–303. [Google Scholar] [CrossRef] [PubMed]
  6. Uthaiah, C.A.; Beeraka, N.M.; Rajalakshmi, R.; Ramya, C.M.; Madhunapantula, S.V. Role of Neural Stem Cells and Vitamin D Receptor (VDR)–Mediated Cellular Signaling in the Mitigation of Neurological Diseases. Mol. Neurobiol. 2022, 59, 4065–4105. [Google Scholar] [CrossRef] [PubMed]
  7. Mahendra, A.; Choudhury, B.K.; Sharma, T.; Bansal, N.; Bansal, R.; Gupta, S.; Karishma. Vitamin D and gastrointestinal cancer. J. Lab. Physicians 2018, 10, 001–005. [Google Scholar] [CrossRef]
  8. Zhang, H.; Wang, P.; Jie, Y.; Sun, Y.; Wang, X.; Fan, Y. Predictive value of 25-hydroxyvitamin D level in patients with coronary artery disease: A meta-analysis. Front. Nutr. 2022, 9, 984487. [Google Scholar] [CrossRef]
  9. Zhou, A.; Selvanayagam, J.B.; Hyppönen, E. Non-linear Mendelian randomization analyses support a role for vitamin D deficiency in cardiovascular disease risk. Eur. Heart J. 2022, 43, 1731–1739. [Google Scholar] [CrossRef]
  10. Ganmaa, D.; Enkhmaa, D.; Nasantogtokh, E.; Sukhbaatar, S.; Tumur-Ochir, K.; Manson, J. Vitamin D, respiratory infections, and chronic disease: Review of meta-analyses and randomized clinical trials. J. Intern. Med. 2022, 291, 141–164. [Google Scholar] [CrossRef]
  11. Hahn, J.; Cook, N.R.; Alexander, E.K.; Friedman, S.; Walter, J.; Bubes, V.; Kotler, G.; Lee, I.-M.; Manson, J.E.; Costenbader, K.H. Vitamin D and marine omega 3 fatty acid supplementation and incident autoimmune disease: VITAL randomized controlled trial. BMJ 2022, 376, e066452. [Google Scholar] [CrossRef] [PubMed]
  12. Bishop, E.L.; Ismailova, A.; Dimeloe, S.; Hewison, M.; White, J.H. Vitamin D and Immune Regulation: Antibacterial, Antiviral, Anti-Inflammatory. JBMR Plus 2020, 5, e10405. [Google Scholar] [CrossRef] [PubMed]
  13. Anagnostis, P.; Livadas, S.; Goulis, D.G.; Bretz, S.; Ceausu, I.; Durmusoglu, F.; Erkkola, R.; Fistonic, I.; Gambacciani, M.; Geukes, M.; et al. EMAS position statement: Vitamin D and menopausal health. Maturitas 2023, 169, 2–9. [Google Scholar] [CrossRef]
  14. Bergman, P. Low vitamin D is a marker for poor health and increased risk for disease: But causality is still unclear in most cases. J. Intern. Med. 2023, 293, 272–274. [Google Scholar] [CrossRef] [PubMed]
  15. Olshansky, S.J. Articulating the Case for the Longevity Dividend. Cold Spring Harb. Perspect. Med. 2016, 6, a025940. [Google Scholar] [CrossRef] [Green Version]
  16. Yousefzadeh, M.J.; Flores, R.R.; Zhu, Y.; Schmiechen, Z.C.; Brooks, R.W.; Trussoni, C.E.; Cui, Y.; Angelini, L.; Lee, K.-A.; McGowan, S.J.; et al. An aged immune system drives senescence and ageing of solid organs. Nature 2021, 594, 100–105. [Google Scholar] [CrossRef]
  17. Whiting, C.C.; Siebert, J.; Newman, A.M.; Du, H.; Alizadeh, A.A.; Goronzy, J.; Weyand, C.M.; Krishnan, E.; Fathman, C.G.; Maecker, H.T. Large-Scale and Comprehensive Immune Profiling and Functional Analysis of Normal Human Aging. PLoS ONE 2015, 10, e0133627. [Google Scholar] [CrossRef]
  18. Duggal, N.A.; Niemiro, G.; Harridge, S.D.R.; Simpson, R.J.; Lord, J.M. Can physical activity ameliorate immunosenescence and thereby reduce age-related multi-morbidity? Nat. Rev. Immunol. 2019, 19, 563–572. [Google Scholar] [CrossRef]
  19. Franceschi, C.; Bonafè, M.; Valensin, S.; Olivieri, F.; De Luca, M.; Ottaviani, E.; De Benedictis, G. Inflamm-aging: An Evolutionary Perspective on Immunosenescence. Ann. N. Y. Acad. Sci. 2006, 908, 244–254. [Google Scholar] [CrossRef]
  20. Franceschi, C.; Garagnani, P.; Parini, P.; Giuliani, C.; Santoro, A. Inflammaging: A new immune–metabolic viewpoint for age-related diseases. Nat. Rev. Endocrinol. 2018, 14, 576–590. [Google Scholar] [CrossRef]
  21. Goronzy, J.J.; Qi, Q.; Olshen, R.A.; Weyand, C.M. High-throughput sequencing insights into T-cell receptor repertoire diversity in aging. Genome Med. 2015, 7, 117. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Qi, Q.; Liu, Y.; Cheng, Y.; Glanville, J.; Zhang, D.; Lee, J.-Y.; Olshen, R.A.; Weyand, C.M.; Boyd, S.D.; Goronzy, J.J. Diversity and clonal selection in the human T-cell repertoire. Proc. Natl. Acad. Sci. USA 2014, 111, 13139–13144. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Goronzy, J.J.; Fang, F.; Cavanagh, M.M.; Qi, Q.; Weyand, C.M. Naive T Cell Maintenance and Function in Human Aging. J. Immunol. 2015, 194, 4073–4080. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Gorgoulis, V.; Adams, P.D.; Alimonti, A.; Bennett, D.C.; Bischof, O.; Bishop, C.; Campisi, J.; Collado, M.; Evangelou, K.; Ferbeyre, G.; et al. Cellular Senescence: Defining a Path Forward. Cell 2019, 179, 813–827. [Google Scholar] [CrossRef] [PubMed]
  25. He, S.; Sharpless, N.E. Senescence in Health and Disease. Cell 2017, 169, 1000–1011. [Google Scholar] [CrossRef] [Green Version]
  26. Neurohr, G.E.; Terry, R.L.; Lengefeld, J.; Bonney, M.; Brittingham, G.P.; Moretto, F.; Miettinen, T.P.; Vaites, L.P.; Soares, L.M.; Paulo, J.A.; et al. Excessive Cell Growth Causes Cytoplasm Dilution And Contributes to Senescence. Cell 2019, 176, 1083–1097.E18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Lee, B.Y.; Han, J.A.; Im, J.S.; Morrone, A.; Johung, K.; Goodwin, E.C.; Kleijer, W.J.; DiMaio, D.; Hwang, E.S. Senescence-associated β-galactosidase is lysosomal β-galactosidase. Aging Cell 2006, 5, 187–195. [Google Scholar] [CrossRef]
  28. Childs, B.G.; Gluscevic, M.; Baker, D.J.; Laberge, R.-M.; Marquess, D.; Dananberg, J.; van Deursen, J.M. Senescent cells: An emerging target for diseases of ageing. Nat. Rev. Drug Discov. 2017, 16, 718–735. [Google Scholar] [CrossRef] [Green Version]
  29. Coppé, J.-P.; Desprez, P.-Y.; Krtolica, A.; Campisi, J. The Senescence-Associated Secretory Phenotype: The Dark Side of Tumor Suppression. Annu. Rev. Pathol. Mech. Dis. 2010, 5, 99–118. [Google Scholar] [CrossRef] [Green Version]
  30. Faget, D.V.; Ren, Q.; Stewart, S.A. Unmasking senescence: Context-dependent effects of SASP in cancer. Nat. Rev. Cancer 2019, 19, 439–453. [Google Scholar] [CrossRef]
  31. Furman, D.; Campisi, J.; Verdin, E.; Carrera-Bastos, P.; Targ, S.; Franceschi, C.; Ferrucci, L.; Gilroy, D.W.; Fasano, A.; Miller, G.W.; et al. Chronic inflammation in the etiology of disease across the life span. Nat. Med. 2019, 25, 1822–1832. [Google Scholar] [CrossRef]
  32. Di Micco, R.; Krizhanovsky, V.; Baker, D.; d’Adda di Fagagna, F. Cellular senescence in ageing: From mechanisms to therapeutic opportunities. Nat. Rev. Mol. Cell Biol. 2021, 22, 75–95. [Google Scholar] [CrossRef]
  33. Wiley, C.D.; Flynn, J.M.; Morrissey, C.; Lebofsky, R.; Shuga, J.; Dong, X.; Unger, M.A.; Vijg, J.; Melov, S.; Campisi, J. Analysis of individual cells identifies cell-to-cell variability following induction of cellular senescence. Aging Cell 2017, 16, 1043–1050. [Google Scholar] [CrossRef] [Green Version]
  34. Pae, M.; Meydani, S.N.; Wu, D. The role of nutrition in enhancing immunity in aging. Aging Dis. 2012, 3, 91–129. [Google Scholar]
  35. Krabbe, K.S.; Pedersen, M.; Bruunsgaard, H. Inflammatory mediators in the elderly. Exp. Gerontol. 2004, 39, 687–699. [Google Scholar] [CrossRef] [PubMed]
  36. Sarkar, D.; Fisher, P.B. Molecular mechanisms of aging-associated inflammation. Cancer Lett. 2006, 236, 13–23. [Google Scholar] [CrossRef]
  37. Vasto, S.; Candore, G.; Balistreri, C.R.; Caruso, M.; Colonna-Romano, G.; Grimaldi, M.P.; Listi, F.; Nuzzo, D.; Lio, D.; Caruso, C. Inflammatory networks in ageing, age-related diseases and longevity. Mech. Ageing Dev. 2007, 128, 83–91. [Google Scholar] [CrossRef]
  38. Goronzy, J.J.; Shao, L.; Weyand, C.M. Immune Aging and Rheumatoid Arthritis. Rheum. Dis. Clin. North Am. 2010, 36, 297–310. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Goronzy, J.J.; Li, G.; Yang, Z.; Weyand, C.M. The Janus Head of T Cell Aging—Autoimmunity and Immunodeficiency. Front. Immunol. 2013, 4, 131. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Weyand, C.M.; Yang, Z.; Goronzy, J.J. T-cell aging in rheumatoid arthritis. Curr. Opin. Rheumatol. 2014, 26, 93–100. [Google Scholar] [CrossRef] [Green Version]
  41. Smith, K.A. Interleukin-2: Inception, Impact, and Implications. Science 1988, 240, 1169–1176. [Google Scholar] [CrossRef] [PubMed]
  42. Schmitt, V.; Rink, L.; Uciechowski, P. The Th17/Treg balance is disturbed during aging. Exp. Gerontol. 2013, 48, 1379–1386. [Google Scholar] [CrossRef] [PubMed]
  43. Xu, L.; Wei, C.; Chen, Y.; Wu, Y.; Shou, X.; Chen, W.; Lu, D.; Sun, H.; Li, W.; Yu, B.; et al. IL-33 induces thymic involution-associated naive T cell aging and impairs host control of severe infection. Nat. Commun. 2022, 13, 6881. [Google Scholar] [CrossRef] [PubMed]
  44. Kuswanto, W.; Burzyn, D.; Panduro, M.; Wang, K.K.; Jang, Y.C.; Wagers, A.J.; Benoist, C.; Mathis, D. Poor Repair of Skeletal Muscle in Aging Mice Reflects a Defect in Local, Interleukin-33-Dependent Accumulation of Regulatory T Cells. Immunity 2016, 44, 355–367. [Google Scholar] [CrossRef] [Green Version]
  45. Pahlavani, M.A. T cell signaling effect of age. Front. Biosci. 1998, 3, 1120–1133. [Google Scholar] [CrossRef]
  46. Kirk, C.J.; Miller, R.A. Analysis of Raf-1 Activation in Response to TCR Activation and Costimulation in Murine T-Lymphocytes: Effect of Age. Cell. Immunol. 1998, 190, 33–42. [Google Scholar] [CrossRef]
  47. Pahlavani, M.A.; Harris, M.D.; Richardson, A. Activation of p21ras/MAPK Signal Transduction Molecules Decreases with Age in Mitogen-Stimulated T Cells from Rats. Cell. Immunol. 1998, 185, 39–48. [Google Scholar] [CrossRef]
  48. Pahlavani, M.A.; Harris, M.D.; Richardson, A. The Age-Related Decline in the Induction of IL-2 Transcription Is Correlated to Changes in the Transcription Factor NFAT. Cell. Immunol. 1995, 165, 84–91. [Google Scholar] [CrossRef]
  49. Whisler, R.L.; Beiqing, L.; Chen, M. Age-Related Decreases in IL-2 Production by Human T Cells Are Associated with Impaired Activation of Nuclear Transcriptional Factors AP-1 and NF-AT. Cell. Immunol. 1996, 169, 185–195. [Google Scholar] [CrossRef]
  50. Whisler, R.L.; Chen, M.; Beiqing, L.; Carle, K.W. Impaired Induction of c-fos/c-jun Genes and of Transcriptional Regulatory Proteins Binding Distinct c-fos/c-jun Promoter Elements in Activated Human T Cells during Aging. Cell. Immunol. 1997, 175, 41–50. [Google Scholar] [CrossRef]
  51. Trebilcock, G.U.; Ponnappan, U. Evidence for Lowered Induction of Nuclear Factor Kappa B in Activated Human T Lymphocytes during Aging. Gerontology 1996, 42, 137–146. [Google Scholar] [CrossRef] [PubMed]
  52. Trebilcock, G.U.; Ponnappan, U. Nuclear factor-κB induction in CD45RO+ and CD45RA+ T cell subsets during aging. Mech. Ageing Dev. 1998, 102, 149–163. [Google Scholar] [CrossRef] [PubMed]
  53. Claycombe, K.J.; Wu, D.; Nikolova-Karakashian, M.; Palmer, H.; Beharka, A.; Paulson, K.E.; Meydani, S.N. Ceramide Mediates Age-associated Increase in Macrophage Cyclooxygenase-2 Expression. J. Biol. Chem. 2002, 277, 30784–30791. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Hayek, M.G.; Mura, C.; Wu, D.; Beharka, A.A.; Han, S.N.; Paulson, K.E.; Hwang, D.; Meydani, S.N. Enhanced expression of inducible cyclooxygenase with age in murine macrophages. J. Immunol. 1997, 159, 2445–2451. [Google Scholar] [CrossRef] [PubMed]
  55. Wu, D.; Marko, M.; Claycombe, K.; Paulson, K.E.; Meydani, S.N. Ceramide-induced and Age-associated Increase in Macrophage COX-2 Expression Is Mediated through Up-regulation of NF-κB Activity. J. Biol. Chem. 2003, 278, 10983–10992. [Google Scholar] [CrossRef] [Green Version]
  56. Shaw, A.C.; Joshi, S.; Greenwood, H.; Panda, A.; Lord, J.M. Aging of the innate immune system. Curr. Opin. Immunol. 2010, 22, 507–513. [Google Scholar] [CrossRef] [Green Version]
  57. Panda, A.; Arjona, A.; Sapey, E.; Bai, F.; Fikrig, E.; Montgomery, R.R.; Lord, J.M.; Shaw, A.C. Human innate immunosenescence: Causes and consequences for immunity in old age. Trends Immunol. 2009, 30, 325–333. [Google Scholar] [CrossRef] [Green Version]
  58. Mocchegiani, E.; Malavolta, M. NK and NKT cell functions in immunosenescence. Aging Cell 2004, 3, 177–184. [Google Scholar] [CrossRef] [PubMed]
  59. Bauer, M.E.; De la Fuente, M. The role of oxidative and inflammatory stress and persistent viral infections in immunosenescence. Mech. Ageing Dev. 2016, 158, 27–37. [Google Scholar] [CrossRef]
  60. Brodin, P.; Davis, M.M. Human immune system variation. Nat. Rev. Immunol. 2017, 17, 21–29. [Google Scholar] [CrossRef]
  61. Pawelec, G. Does the human immune system ever really become “senescent”? F1000Research 2017, 6, 1323. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Calder, P.C.; Ortega, E.F.; Meydani, S.N.; Adkins, Y.; Stephensen, C.B.; Thompson, B.; Zwickey, H. Nutrition, Immunosenescence, and Infectious Disease: An Overview of the Scientific Evidence on Micronutrients and on Modulation of the Gut Microbiota. Adv. Nutr. 2022, 13, S1–S26. [Google Scholar] [CrossRef]
  63. Cutolo, M.; Pizzorni, C.; Sulli, A. Vitamin D endocrine system involvement in autoimmune rheumatic diseases. Autoimmun. Rev. 2011, 11, 84–87. [Google Scholar] [CrossRef]
  64. Bendik, I.; Friedel, A.; Roos, F.F.; Weber, P.; Eggersdorfer, M. Vitamin D: A critical and essential micronutrient for human health. Front. Physiol. 2014, 5, 248. [Google Scholar] [CrossRef]
  65. Holick, M.F.; Matsuoka, L.Y.; Wortsman, J. Age, vitamin D, and solar ultraviolet. Lancet 1989, 334, 1104–1105. [Google Scholar] [CrossRef] [PubMed]
  66. Lau, K.-H.W.; Baylink, D.J. Vitamin D Therapy of Osteoporosis: Plain Vitamin D Therapy Versus Active Vitamin D Analog (D-Hormone) Therapy. Calcif. Tissue Int. 1999, 65, 295–306. [Google Scholar] [CrossRef] [PubMed]
  67. Mosekilde, L. Vitamin D and the elderly. Clin. Endocrinol. 2005, 62, 265–281. [Google Scholar] [CrossRef]
  68. Lips, P. Vitamin D Deficiency and Secondary Hyperparathyroidism in the Elderly: Consequences for Bone Loss and Fractures and Therapeutic Implications. Endocr. Rev. 2001, 22, 477–501. [Google Scholar] [CrossRef] [PubMed]
  69. Guo, J.; Lovegrove, J.A.; Givens, D.I. 25(OH)D3-enriched or fortified foods are more efficient at tackling inadequate vitamin D status than vitamin D3. Proc. Nutr. Soc. 2018, 77, 282–291. [Google Scholar] [CrossRef] [Green Version]
  70. Aranow, C. Vitamin D and the Immune System. J. Investig. Med. 2011, 59, 881–886. [Google Scholar] [CrossRef] [Green Version]
  71. Liu, P.T.; Stenger, S.; Li, H.; Wenzel, L.; Tan, B.H.; Krutzik, S.R.; Ochoa, M.T.; Schauber, J.; Wu, K.; Meinken, C.; et al. Toll-Like Receptor Triggering of a Vitamin D-Mediated Human Antimicrobial Response. Science 2006, 311, 1770–1773. [Google Scholar] [CrossRef]
  72. Hewison, M. An update on vitamin D and human immunity: An update on vitamin D. Clin. Endocrinol. 2012, 76, 315–325. [Google Scholar] [CrossRef] [PubMed]
  73. Baeke, F.; Takiishi, T.; Korf, H.; Gysemans, C.; Mathieu, C. Vitamin D: Modulator of the immune system. Curr. Opin. Pharmacol. 2010, 10, 482–496. [Google Scholar] [CrossRef] [PubMed]
  74. Di Rosa, M.; Malaguarnera, M.; Nicoletti, F.; Malaguarnera, L. Vitamin D3: A helpful immuno-modulator: Vitamin D3 as immuno-modulator. Immunology 2011, 134, 123–139. [Google Scholar] [CrossRef]
  75. Cantorna, M.; Snyder, L.; Lin, Y.-D.; Yang, L. Vitamin D and 1,25(OH)2D Regulation of T cells. Nutrients 2015, 7, 3011–3021. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Medrano, M.; Carrillo-Cruz, E.; Montero, I.; Perez-Simon, J. Vitamin D: Effect on Haematopoiesis and Immune System and Clinical Applications. IJMS 2018, 19, 2663. [Google Scholar] [CrossRef] [Green Version]
  77. Prietl, B.; Treiber, G.; Pieber, T.; Amrein, K. Vitamin D and Immune Function. Nutrients 2013, 5, 2502–2521. [Google Scholar] [CrossRef]
  78. Sassi, F.; Tamone, C.; D’Amelio, P. Vitamin D: Nutrient, Hormone, and Immunomodulator. Nutrients 2018, 10, 1656. [Google Scholar] [CrossRef] [Green Version]
  79. Fritsche, J.; Mondal, K.; Ehrnsperger, A.; Andreesen, R.; Kreutz, M. Regulation of 25-hydroxyvitamin D3-1α-hydroxylase and production of 1α,25-dihydroxyvitamin D3 by human dendritic cells. Blood 2003, 102, 3314–3316. [Google Scholar] [CrossRef] [Green Version]
  80. Gottfried, E.; Rehli, M.; Hahn, J.; Holler, E.; Andreesen, R.; Kreutz, M. Monocyte-derived cells express CYP27A1 and convert vitamin D3 into its active metabolite. Biochem. Biophys. Res. Commun. 2006, 349, 209–213. [Google Scholar] [CrossRef]
  81. Korf, H.; Decallonne, B.; Mathieu, C. Vitamin D for infections. Curr. Opin. Endocrinol. Diabetes Obes. 2014, 21, 431–436. [Google Scholar] [CrossRef]
  82. Shin, D.-M.; Jo, E.-K. Antimicrobial Peptides in Innate Immunity against Mycobacteria. Immune Netw. 2011, 11, 245–252. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Bellmann-Weiler, R.; Weiss, G. Pitfalls in the Diagnosis and Therapy of Infections in Elderly Patients—A Mini-Review. Gerontology 2009, 55, 241–249. [Google Scholar] [CrossRef] [PubMed]
  84. Rekha, R.S.; Mily, A.; Sultana, T.; Haq, A.; Ahmed, S.; Mostafa Kamal, S.M.; van Schadewijk, A.; Hiemstra, P.S.; Gudmundsson, G.H.; Agerberth, B.; et al. Immune responses in the treatment of drug-sensitive pulmonary tuberculosis with phenylbutyrate and vitamin D3 as host directed therapy. BMC Infect. Dis. 2018, 18, 303. [Google Scholar] [CrossRef] [PubMed]
  85. Bartley, J.; Garrett, J.; Camargo, C.A.; Scragg, R.; Vandal, A.; Sisk, R.; Milne, D.; Tai, R.; Jeon, G.; Cursons, R.; et al. Vitamin D3 supplementation in adults with bronchiectasis: A pilot study. Chronic Respir. Dis. 2018, 15, 384–392. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Mily, A.; Rekha, R.S.; Kamal, S.M.M.; Akhtar, E.; Sarker, P.; Rahim, Z.; Gudmundsson, G.H.; Agerberth, B.; Raqib, R. Oral intake of phenylbutyrate with or without vitamin D3upregulates the cathelicidin LL-37 in human macrophages: A dose finding study for treatment of tuberculosis. BMC Pulm. Med. 2013, 13, 23. [Google Scholar] [CrossRef] [Green Version]
  87. Fabri, M.; Stenger, S.; Shin, D.-M.; Yuk, J.-M.; Liu, P.T.; Realegeno, S.; Lee, H.-M.; Krutzik, S.R.; Schenk, M.; Sieling, P.A.; et al. Vitamin D Is Required for IFN-γ–Mediated Antimicrobial Activity of Human Macrophages. Sci. Transl. Med. 2011, 3, 104ra102. [Google Scholar] [CrossRef] [Green Version]
  88. Chung, C.; Silwal, P.; Kim, I.; Modlin, R.L.; Jo, E.-K. Vitamin D-Cathelicidin Axis: At the Crossroads between Protective Immunity and Pathological Inflammation during Infection. Immune Netw. 2020, 20, e12. [Google Scholar] [CrossRef]
  89. Campbell, G.R.; Spector, S.A. Toll-Like Receptor 8 Ligands Activate a Vitamin D Mediated Autophagic Response that Inhibits Human Immunodeficiency Virus Type 1. PLoS Pathog. 2012, 8, e1003017. [Google Scholar] [CrossRef] [Green Version]
  90. Yang, R.; Yang, E.; Shen, L.; Modlin, R.L.; Shen, H.; Chen, Z.W. IL-12+IL-18 Cosignaling in Human Macrophages and Lung Epithelial Cells Activates Cathelicidin and Autophagy, Inhibiting Intracellular Mycobacterial Growth. J. Immunol. 2018, 200, 2405–2417. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Klug-Micu, G.M.; Stenger, S.; Sommer, A.; Liu, P.T.; Krutzik, S.R.; Modlin, R.L.; Fabri, M. CD40 ligand and interferon-γ induce an antimicrobial response against Mycobacterium tuberculosis in human monocytes. Immunology 2013, 139, 121–128. [Google Scholar] [CrossRef] [PubMed]
  92. Rekha, R.S.; Rao Muvva, S.J.; Wan, M.; Raqib, R.; Bergman, P.; Brighenti, S.; Gudmundsson, G.H.; Agerberth, B. Phenylbutyrate induces LL-37-dependent autophagy and intracellular killing of Mycobacterium tuberculosis in human macrophages. Autophagy 2015, 11, 1688–1699. [Google Scholar] [CrossRef] [Green Version]
  93. Yuk, J.-M.; Shin, D.-M.; Lee, H.-M.; Yang, C.-S.; Jin, H.S.; Kim, K.-K.; Lee, Z.-W.; Lee, S.-H.; Kim, J.-M.; Jo, E.-K. Vitamin D3 Induces Autophagy in Human Monocytes/Macrophages via Cathelicidin. Cell Host Microbe 2009, 6, 231–243. [Google Scholar] [CrossRef] [Green Version]
  94. Jiang, P.; Mizushima, N. Autophagy and human diseases. Cell Res. 2014, 24, 69–79. [Google Scholar] [CrossRef] [Green Version]
  95. Verway, M.; Bouttier, M.; Wang, T.-T.; Carrier, M.; Calderon, M.; An, B.-S.; Devemy, E.; McIntosh, F.; Divangahi, M.; Behr, M.A.; et al. Vitamin D Induces Interleukin-1β Expression: Paracrine Macrophage Epithelial Signaling Controls M. tuberculosis Infection. PLoS Pathog. 2013, 9, e1003407. [Google Scholar] [CrossRef] [Green Version]
  96. Dimitrov, V.; White, J.H. Species-specific regulation of innate immunity by vitamin D signaling. J. Steroid Biochem. Mol. Biol. 2016, 164, 246–253. [Google Scholar] [CrossRef]
  97. Clark, A.; Mach, N. Role of Vitamin D in the Hygiene Hypothesis: The Interplay between Vitamin D, Vitamin D Receptors, Gut Microbiota, and Immune Response. Front. Immunol. 2016, 7, 627. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Yang, C.-S.; Shin, D.-M.; Kim, K.-H.; Lee, Z.-W.; Lee, C.-H.; Park, S.G.; Bae, Y.S.; Jo, E.-K. NADPH Oxidase 2 Interaction with TLR2 Is Required for Efficient Innate Immune Responses to Mycobacteria via Cathelicidin Expression. J. Immunol. 2009, 182, 3696–3705. [Google Scholar] [CrossRef] [Green Version]
  99. Yan, Y.; Finkel, T. Autophagy as a regulator of cardiovascular redox homeostasis. Free. Radic. Biol. Med. 2017, 109, 108–113. [Google Scholar] [CrossRef]
  100. Afacan, N.J.; Yeung, A.T.Y.; Pena, O.M.; Hancock, R.E.W. Therapeutic Potential of Host Defense Peptides in Antibiotic-resistant Infections. Curr. Pharm. Des. 2012, 18, 807–819. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  101. Gombart, A.F.; Borregaard, N.; Koeffler, H.P. Human cathelicidin antimicrobial peptide (CAMP) gene is a direct target of the vitamin D receptor and is strongly up-regulated in myeloid cells by 1,25-dihydroxyvitamin D3. FASEB J. 2005, 19, 1067–1077. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Hilbring, C.; Augustin, M.; Kirsten, N.; Mohr, N. Epidemiology of rosacea in a population-based study of 161,269 German employees. Int. J. Dermatol. 2022, 61, 570–576. [Google Scholar] [CrossRef] [PubMed]
  103. Park, B.W.; Ha, J.M.; Cho, E.B.; Jin, J.K.; Park, E.J.; Park, H.R.; Kang, H.J.; Ko, S.H.; Kim, K.H.; Kim, K.J. A Study on Vitamin D and Cathelicidin Status in Patients with Rosacea: Serum Level and Tissue Expression. Ann. Dermatol. 2018, 30, 136–142. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Fulop, T.; Larbi, A.; Dupuis, G.; Le Page, A.; Frost, E.H.; Cohen, A.A.; Witkowski, J.M.; Franceschi, C. Immunosenescence and Inflamm-Aging As Two Sides of the Same Coin: Friends or Foes? Front. Immunol. 2018, 8, 1960. [Google Scholar] [CrossRef] [Green Version]
  105. Korf, H.; Wenes, M.; Stijlemans, B.; Takiishi, T.; Robert, S.; Miani, M.; Eizirik, D.L.; Gysemans, C.; Mathieu, C. 1,25-Dihydroxyvitamin D3 curtails the inflammatory and T cell stimulatory capacity of macrophages through an IL-10-dependent mechanism. Immunobiology 2012, 217, 1292–1300. [Google Scholar] [CrossRef] [PubMed]
  106. Rigby, W.F.; Stacy, T.; Fanger, M.W. Inhibition of T lymphocyte mitogenesis by 1,25-dihydroxyvitamin D3 (calcitriol). J. Clin. Investig. 1984, 74, 1451–1455. [Google Scholar] [CrossRef] [PubMed]
  107. Lemire, J.M.; Archer, D.C.; Beck, L.; Spiegelberg, H.L. Immunosuppressive actions of 1,25-dihydroxyvitamin D3: Preferential inhibition of Th1 functions. J. Nutr. 1995, 125, 1704S–1708S. [Google Scholar] [CrossRef]
  108. Daniel, C.; Sartory, N.A.; Zahn, N.; Radeke, H.H.; Stein, J.M. Immune Modulatory Treatment of Trinitrobenzene Sulfonic Acid Colitis with Calcitriol Is Associated with a Change of a T Helper (Th) 1/Th17 to a Th2 and Regulatory T Cell Profile. J. Pharmacol. Exp. Ther. 2008, 324, 23–33. [Google Scholar] [CrossRef] [Green Version]
  109. Penna, G.; Adorini, L. 1α,25-Dihydroxyvitamin D3 Inhibits Differentiation, Maturation, Activation, and Survival of Dendritic Cells Leading to Impaired Alloreactive T Cell Activation. J. Immunol. 2000, 164, 2405–2411. [Google Scholar] [CrossRef] [Green Version]
  110. Boonstra, A.; Barrat, F.J.; Crain, C.; Heath, V.L.; Savelkoul, H.F.J.; O’Garra, A. 1α,25-Dihydroxyvitamin D3 Has a Direct Effect on Naive CD4+ T Cells to Enhance the Development of Th2 Cells. J. Immunol. 2001, 167, 4974–4980. [Google Scholar] [CrossRef] [Green Version]
  111. Gorman, S.; Kuritzky, L.A.; Judge, M.A.; Dixon, K.M.; McGlade, J.P.; Mason, R.S.; Finlay-Jones, J.J.; Hart, P.H. Topically Applied 1,25-Dihydroxyvitamin D3 Enhances the Suppressive Activity of CD4+CD25+ Cells in the Draining Lymph Nodes. J. Immunol. 2007, 179, 6273–6283. [Google Scholar] [CrossRef] [Green Version]
  112. Hamzaoui, A.; Berraïes, A.; Hamdi, B.; Kaabachi, W.; Ammar, J.; Hamzaoui, K. Vitamin D reduces the differentiation and expansion of Th17 cells in young asthmatic children. Immunobiology 2014, 219, 873–879. [Google Scholar] [CrossRef] [PubMed]
  113. Penna, G.; Roncari, A.; Amuchastegui, S.; Daniel, K.C.; Berti, E.; Colonna, M.; Adorini, L. Expression of the inhibitory receptor ILT3 on dendritic cells is dispensable for induction of CD4+Foxp3+ regulatory T cells by 1,25-dihydroxyvitamin D3. Blood 2005, 106, 3490–3497. [Google Scholar] [CrossRef]
  114. Arnson, Y.; Amital, H.; Shoenfeld, Y. Vitamin D and autoimmunity: New aetiological and therapeutic considerations. Ann. Rheum. Dis. 2007, 66, 1137–1142. [Google Scholar] [CrossRef] [Green Version]
  115. Fletcher, J.; Bishop, E.L.; Harrison, S.R.; Swift, A.; Cooper, S.C.; Dimeloe, S.K.; Raza, K.; Hewison, M. Autoimmune disease and interconnections with vitamin D. Endocr. Connect. 2022, 11, e210554. [Google Scholar] [CrossRef] [PubMed]
  116. Murdaca, G.; Tonacci, A.; Negrini, S.; Greco, M.; Borro, M.; Puppo, F.; Gangemi, S. Emerging role of vitamin D in autoimmune diseases: An update on evidence and therapeutic implications. Autoimmun. Rev. 2019, 18, 102350. [Google Scholar] [CrossRef]
  117. Sirufo, M.M.; Ginaldi, L.; De Martinis, M. The IL-33/ST2 axis and vitamin D as a possible emerging therapeutic target in osteoarthritis. Rheumatology 2021, 60, e300. [Google Scholar] [CrossRef] [PubMed]
  118. De Martinis, M.; Ginaldi, L.; Sirufo, M.M.; Bassino, E.M.; De Pietro, F.; Pioggia, G.; Gangemi, S. IL-33/Vitamin D Crosstalk in Psoriasis-Associated Osteoporosis. Front. Immunol. 2020, 11, 604055. [Google Scholar] [CrossRef]
  119. Sirufo, M.M.; Magnanimi, L.M.; Ginaldi, L.; De Martinis, M. Guillain-Barré syndrome, the IL-33/ST2 axis, and vitamin D. Eur. J. Neurol. 2022, 29, e20–e21. [Google Scholar] [CrossRef]
  120. Facciotti, F.; Gagliani, N.; Häringer, B.; Alfen, J.S.; Penatti, A.; Maglie, S.; Paroni, M.; Iseppon, A.; Moro, M.; Crosti, M.C.; et al. IL-10–producing forkhead box protein 3–negative regulatory T cells inhibit B-cell responses and are involved in systemic lupus erythematosus. J. Allergy Clin. Immunol. 2016, 137, 318–321.E5. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Castelo-Branco, C.; Soveral, I. The immune system and aging: A review. Gynecol. Endocrinol. 2014, 30, 16–22. [Google Scholar] [CrossRef]
  122. van der Wielen, R.P.J.; de Groot, L.C.P.G.M.; van Staveren, W.A.; Löwik, M.R.H.; van den Berg, H.; Haller, J.; Moreiras, O. Serum vitamin D concentrations among elderly people in Europe. Lancet 1995, 346, 207–210. [Google Scholar] [CrossRef]
  123. Oliveri, B.; Plantalech, L.; Bagur, A.; Wittich, A.C.; Rovai, G.; Pusiol, E.; Giovanelli, J.L.; Ponce, G.; Nieva, A.; Chaperón, A.; et al. High prevalence of vitamin D insufficiency in healthy elderly people living at home in Argentina. Eur. J. Clin. Nutr. 2004, 58, 337–342. [Google Scholar] [CrossRef] [Green Version]
  124. Portela, M.L.P.M.; Mónico, A.; Barahona, A.; Dupraz, H.; Sol Gonzales-Chaves, M.M.; Zeni, S.N. Comparative 25-OH-vitamin D level in institutionalized women older than 65 years from two cities in Spain and Argentina having a similar solar radiation index. Nutrition 2010, 26, 283–289. [Google Scholar] [CrossRef] [PubMed]
  125. Holick, M.F. Vitamin D Deficiency. N. Engl. J. Med. 2007, 357, 266–281. [Google Scholar] [CrossRef]
  126. Chauss, D.; Freiwald, T.; McGregor, R.; Yan, B.; Wang, L.; Nova-Lamperti, E.; Kumar, D.; Zhang, Z.; Teague, H.; West, E.E.; et al. Autocrine vitamin D signaling switches off pro-inflammatory programs of TH1 cells. Nat. Immunol. 2022, 23, 62–74. [Google Scholar] [CrossRef] [PubMed]
  127. Bima, A.I.; Mahdi, A.S.; Al Fayez, F.F.; Khawaja, T.M.; Abo El-Khair, S.M.; Elsamanoudy, A.Z. Cellular Senescence and Vitamin D Deficiency Play a Role in the Pathogenesis of Obesity-Associated Subclinical Atherosclerosis: Study of the Potential Protective Role of Vitamin D Supplementation. Cells 2021, 10, 920. [Google Scholar] [CrossRef] [PubMed]
  128. Satilmis, S.; Celik, O.; Biyik, I.; Ozturk, D.; Celik, K.; Akın, F.; Ayca, B.; Yalcin, B.; Dagdelen, S. Association between serum vitamin D levels and subclinical coronary atherosclerosis and plaque burden/composition in young adult population. Bosn. J. Basic Med Sci. 2015, 15, 67–72. [Google Scholar] [CrossRef] [Green Version]
  129. Oh, J.; Weng, S.; Felton, S.K.; Bhandare, S.; Riek, A.; Butler, B.; Proctor, B.M.; Petty, M.; Chen, Z.; Schechtman, K.B.; et al. 1,25(OH)2 Vitamin D Inhibits Foam Cell Formation and Suppresses Macrophage Cholesterol Uptake in Patients With Type 2 Diabetes Mellitus. Circulation 2009, 120, 687–698. [Google Scholar] [CrossRef] [Green Version]
  130. Katsuumi, G.; Shimizu, I.; Yoshida, Y.; Minamino, T. Vascular Senescence in Cardiovascular and Metabolic Diseases. Front. Cardiovasc. Med. 2018, 5, 18. [Google Scholar] [CrossRef] [Green Version]
  131. Chen, L.; Yang, R.; Qiao, W.; Zhang, W.; Chen, J.; Mao, L.; Goltzman, D.; Miao, D. 1,25-Dihydroxyvitamin D exerts an antiaging role by activation of Nrf2-antioxidant signaling and inactivation of p16/p53-senescence signaling. Aging Cell 2019, 18, e12951. [Google Scholar] [CrossRef]
  132. Yang, D.; Liu, H.-Q.; Yang, Z.; Fan, D.; Tang, Q.-Z. BMI1 in the heart: Novel functions beyond tumorigenesis. eBioMedicine 2021, 63, 103193. [Google Scholar] [CrossRef] [PubMed]
  133. Berridge, M.J. Vitamin D deficiency accelerates ageing and age-related diseases: A novel hypothesis: Vitamin D deficiency and ageing. J. Physiol. 2017, 595, 6825–6836. [Google Scholar] [CrossRef] [PubMed]
  134. Scaini, G.; Rezin, G.T.; Carvalho, A.F.; Streck, E.L.; Berk, M.; Quevedo, J. Mitochondrial dysfunction in bipolar disorder: Evidence, pathophysiology and translational implications. Neurosci. Biobehav. Rev. 2016, 68, 694–713. [Google Scholar] [CrossRef] [PubMed]
  135. Farghali, M.; Ruga, S.; Morsanuto, V.; Uberti, F. Can Brain Health Be Supported by Vitamin D-Based Supplements? A Critical Review. Brain Sci. 2020, 10, 660. [Google Scholar] [CrossRef]
  136. Costantino, S.; Paneni, F.; Cosentino, F. Ageing, metabolism and cardiovascular disease: Mechanisms of cardiovascular ageing. J. Physiol. 2016, 594, 2061–2073. [Google Scholar] [CrossRef]
  137. Fernandez-Medarde, A.; Santos, E. Ras in Cancer and Developmental Diseases. Genes Cancer 2011, 2, 344–358. [Google Scholar] [CrossRef] [Green Version]
  138. Wang, Q.; He, Y.; Shen, Y.; Zhang, Q.; Chen, D.; Zuo, C.; Qin, J.; Wang, H.; Wang, J.; Yu, Y. Vitamin D Inhibits COX-2 Expression and Inflammatory Response by Targeting Thioesterase Superfamily Member 4. J. Biol. Chem. 2014, 289, 11681–11694. [Google Scholar] [CrossRef] [Green Version]
  139. Songkiatisak, P.; Rahman, S.M.T.; Aqdas, M.; Sung, M.-H. NF-κB, a culprit of both inflamm-ageing and declining immunity? Immun. Ageing 2022, 19, 20. [Google Scholar] [CrossRef]
  140. Fafián-Labora, J.A.; O’Loghlen, A. Classical and Nonclassical Intercellular Communication in Senescence and Ageing. Trends Cell Biol. 2020, 30, 628–639. [Google Scholar] [CrossRef]
  141. Jiang, X.; Takahashi, N.; Matsui, N.; Tetsuka, T.; Okamoto, T. The NF-κB Activation in Lymphotoxin β Receptor Signaling Depends on the Phosphorylation of p65 at Serine 536. J. Biol. Chem. 2003, 278, 919–926. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Zhang, G.; Li, J.; Purkayastha, S.; Tang, Y.; Zhang, H.; Yin, Y.; Li, B.; Liu, G.; Cai, D. Hypothalamic programming of systemic ageing involving IKK-β, NF-κB and GnRH. Nature 2013, 497, 211–216. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Thoma, A.; Lightfoot, A.P. NF-kB and Inflammatory Cytokine Signalling: Role in Skeletal Muscle Atrophy. In Muscle Atrophy; Xiao, J., Ed.; Advances in Experimental Medicine and Biology; Springer: Singapore, 2018; Volume 1088, pp. 267–279. ISBN 9789811314346. [Google Scholar]
  144. Mousa, A.; Naderpoor, N.; Johnson, J.; Sourris, K.; de Courten, M.P.J.; Wilson, K.; Scragg, R.; Plebanski, M.; de Courten, B. Effect of vitamin D supplementation on inflammation and nuclear factor kappa-B activity in overweight/obese adults: A randomized placebo-controlled trial. Sci. Rep. 2017, 7, 15154. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Hayden, M.S.; West, A.P.; Ghosh, S. NF-κB and the immune response. Oncogene 2006, 25, 6758–6780. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. García-García, V.A.; Alameda, J.P.; Page, A.; Casanova, M.L. Role of NF-κB in Ageing and Age-Related Diseases: Lessons from Genetically Modified Mouse Models. Cells 2021, 10, 1906. [Google Scholar] [CrossRef]
  147. Tilstra, J.S.; Clauson, C.L.; Niedernhofer, L.J.; Robbins, P.D. NF-κB in Aging and Disease. Aging Dis. 2011, 2, 449–465. [Google Scholar]
  148. López-Otín, C.; Blasco, M.A.; Partridge, L.; Serrano, M.; Kroemer, G. The hallmarks of aging. Cell 2013, 153, 1194–1217. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Salminen, A.; Kaarniranta, K. NF-κB Signaling in the Aging Process. J. Clin. Immunol. 2009, 29, 397–405. [Google Scholar] [CrossRef]
  150. Kenyon, C.J. The genetics of ageing. Nature 2010, 464, 504–512. [Google Scholar] [CrossRef]
  151. Petroulakis, E.; Mamane, Y.; Le Bacquer, O.; Shahbazian, D.; Sonenberg, N. mTOR signaling: Implications for cancer and anticancer therapy. Br. J. Cancer 2006, 94, 195–199. [Google Scholar] [CrossRef] [Green Version]
  152. Antinozzi, C.; Marampon, F.; Corinaldesi, C.; Vicini, E.; Sgrò, P.; Vannelli, G.B.; Lenzi, A.; Crescioli, C.; Di Luigi, L. Testosterone insulin-like effects: An in vitro study on the short-term metabolic effects of testosterone in human skeletal muscle cells. J. Endocrinol. Investig. 2017, 40, 1133–1143. [Google Scholar] [CrossRef] [Green Version]
  153. Choudhury, A.R.; Ju, Z.; Djojosubroto, M.W.; Schienke, A.; Lechel, A.; Schaetzlein, S.; Jiang, H.; Stepczynska, A.; Wang, C.; Buer, J.; et al. Cdkn1a deletion improves stem cell function and lifespan of mice with dysfunctional telomeres without accelerating cancer formation. Nat. Genet. 2007, 39, 99–105. [Google Scholar] [CrossRef]
  154. Lee, W.-P.; Hou, M.-C.; Lan, K.-H.; Li, C.-P.; Chao, Y.; Lin, H.-C.; Lee, S.-D. Helicobacter pylori-induced chronic inflammation causes telomere shortening of gastric mucosa by promoting PARP-1-mediated non-homologous end joining of DNA. Arch. Biochem. Biophys. 2016, 606, 90–98. [Google Scholar] [CrossRef]
  155. Wu, X.; Yang, Y.; Li, W.; Cheng, Y.; Li, X.; Huang, C.; Meng, X.; Wu, B.; Liu, X.; Zhang, L.; et al. Telomerase reverse transcriptase acts in a feedback loop with NF-κB pathway to regulate macrophage polarization in alcoholic liver disease. Sci. Rep. 2016, 6, 18685. [Google Scholar] [CrossRef] [Green Version]
  156. Grabowska, W.; Sikora, E.; Bielak-Zmijewska, A. Sirtuins, a promising target in slowing down the ageing process. Biogerontology 2017, 18, 447–476. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Yeung, F.; Hoberg, J.E.; Ramsey, C.S.; Keller, M.D.; Jones, D.R.; Frye, R.A.; Mayo, M.W. Modulation of NF-κB-dependent transcription and cell survival by the SIRT1 deacetylase. EMBO J. 2004, 23, 2369–2380. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Kawahara, T.L.A.; Michishita, E.; Adler, A.S.; Damian, M.; Berber, E.; Lin, M.; McCord, R.A.; Ongaigui, K.C.L.; Boxer, L.D.; Chang, H.Y.; et al. SIRT6 Links Histone H3 Lysine 9 Deacetylation to NF-κB-Dependent Gene Expression and Organismal Life Span. Cell 2009, 136, 62–74. [Google Scholar] [CrossRef] [Green Version]
  159. Cai, D.; Frantz, J.D.; Tawa, N.E.; Melendez, P.A.; Oh, B.-C.; Lidov, H.G.W.; Hasselgren, P.-O.; Frontera, W.R.; Lee, J.; Glass, D.J.; et al. IKKβ/NF-κB Activation Causes Severe Muscle Wasting in Mice. Cell 2004, 119, 285–298. [Google Scholar] [CrossRef] [Green Version]
  160. Liu, Y.; Drozdov, I.; Shroff, R.; Beltran, L.E.; Shanahan, C.M. Prelamin A Accelerates Vascular Calcification Via Activation of the DNA Damage Response and Senescence-Associated Secretory Phenotype in Vascular Smooth Muscle Cells. Circ. Res. 2013, 112, e99–e109. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  161. Chen, Q.; Liu, K.; Robinson, A.R.; Clauson, C.L.; Blair, H.C.; Robbins, P.D.; Niedernhofer, L.J.; Ouyang, H. DNA damage drives accelerated bone aging via an NF-κB-dependent mechanism. J. Bone Miner. Res. 2013, 28, 1214–1228. [Google Scholar] [CrossRef]
  162. Ghosh, A.; Roy, A.; Liu, X.; Kordower, J.H.; Mufson, E.J.; Hartley, D.M.; Ghosh, S.; Mosley, R.L.; Gendelman, H.E.; Pahan, K. Selective inhibition of NF-κB activation prevents dopaminergic neuronal loss in a mouse model of Parkinson’s disease. Proc. Natl. Acad. Sci. USA 2007, 104, 18754–18759. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Hwang, C.J.; Choi, D.-Y.; Park, M.H.; Hong, J.T. NF-κB as a Key Mediator of Brain Inflammation in Alzheimer’s Disease. CNS Neurol. Disord. Drug Targets 2019, 18, 3–10. [Google Scholar] [CrossRef] [PubMed]
  164. Chen, Y.; Zhang, J.; Ge, X.; Du, J.; Deb, D.K.; Li, Y.C. Vitamin D Receptor Inhibits Nuclear Factor κB Activation by Interacting with IκB Kinase β Protein. J. Biol. Chem. 2013, 288, 19450–19458. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Adorini, L.; Penna, G.; Giarratana, N.; Roncari, A.; Amuchastegui, S.; Daniel, K.C.; Uskokovic, M. Dendritic cells as key targets for immunomodulation by Vitamin D receptor ligands. J. Steroid Biochem. Mol. Biol. 2004, 89–90, 437–441. [Google Scholar] [CrossRef] [PubMed]
  166. Geldmeyer-Hilt, K.; Heine, G.; Hartmann, B.; Baumgrass, R.; Radbruch, A.; Worm, M. 1,25-dihydroxyvitamin D3 impairs NF-κB activation in human naïve B cells. Biochem. Biophys. Res. Commun. 2011, 407, 699–702. [Google Scholar] [CrossRef]
  167. Di Luigi, L.; Sottili, M.; Antinozzi, C.; Vannelli, G.B.; Romanelli, F.; Riccieri, V.; Valesini, G.; Lenzi, A.; Crescioli, C. The Vitamin D Receptor Agonist BXL-01-0029 as a Potential New Pharmacological Tool for the Treatment of Inflammatory Myopathies. PLoS ONE 2013, 8, e77745. [Google Scholar] [CrossRef] [Green Version]
  168. Sottili, M.; Cosmi, L.; Borgogni, E.; Sarchielli, E.; Maggi, L.; Francalanci, M.; Vannelli, G.B.; Ronconi, E.; Adorini, L.; Annunziato, F.; et al. Immunomodulatory effects of BXL-01-0029, a less hypercalcemic vitamin D analogue, in human cardiomyocytes and T cells. Exp. Cell Res. 2009, 315, 264–273. [Google Scholar] [CrossRef]
  169. Borgogni, E.; Sarchielli, E.; Sottili, M.; Santarlasci, V.; Cosmi, L.; Gelmini, S.; Lombardi, A.; Cantini, G.; Perigli, G.; Luconi, M.; et al. Elocalcitol Inhibits Inflammatory Responses in Human Thyroid Cells and T Cells. Endocrinology 2008, 149, 3626–3634. [Google Scholar] [CrossRef] [Green Version]
  170. Zhang, C.; Tong, T.; Miao, D.; Wang, L. Vitamin D inhibits TNF-α induced apoptosis of human nucleus pulposus cells through regulation of NF-kB signaling pathway. J. Orthop. Surg. Res. 2021, 16, 411. [Google Scholar] [CrossRef]
  171. Anversa, P.; Leri, A.; Kajstura, J. Cardiac Regeneration. J. Am. Coll. Cardiol. 2006, 47, 1769–1776. [Google Scholar] [CrossRef] [Green Version]
  172. Aslan, E.; Komurcu, N.; Beji, N.K.; Yalcin, O. Bladder Training and Kegel Exercises for Women with Urinary Complaints Living in a Rest Home. Gerontology 2008, 54, 224–231. [Google Scholar] [CrossRef]
  173. Bhutia, S.K. Vitamin D in autophagy signaling for health and diseases: Insights on potential mechanisms and future perspectives. J. Nutr. Biochem. 2022, 99, 108841. [Google Scholar] [CrossRef]
  174. Li, W.; Che, X.; Chen, X.; Zhou, M.; Luo, X.; Liu, T. Study of calcitriol anti-aging effects on human natural killer cells in vitro. Bioengineered 2021, 12, 6844–6854. [Google Scholar] [CrossRef] [PubMed]
  175. Rashidi, N.; Arefi, S.; Sadri, M.; Delbandi, A.-A. Effect of active vitamin D on proliferation, cell cycle and apoptosis in endometriotic stromal cells. Reprod. Biomed. Online 2022, in press. [Google Scholar] [CrossRef] [PubMed]
  176. Chen, Z.; Huang, D.; Yongyut, P.; Li, G.; Esteban, M.Á.; Jintasataporn, O.; Deng, J.; Zhang, W.; Ai, Q.; Mai, K.; et al. Vitamin D3 deficiency induced intestinal inflammatory response of turbot through nuclear factor-κB/inflammasome pathway, accompanied by the mutually exclusive apoptosis and autophagy. Front. Immunol. 2022, 13, 986593. [Google Scholar] [CrossRef]
  177. Bokov, A.; Chaudhuri, A.; Richardson, A. The role of oxidative damage and stress in aging. Mech. Ageing Dev. 2004, 125, 811–826. [Google Scholar] [CrossRef] [PubMed]
  178. Liao, L.-Z.; Chen, Z.-C.; Wang, S.-S.; Liu, W.-B.; Zhao, C.-L.; Zhuang, X.-D. NLRP3 inflammasome activation contributes to the pathogenesis of cardiocytes aging. Aging 2021, 13, 20534–20551. [Google Scholar] [CrossRef]
  179. Filardi, T.; Ghinassi, B.; Di Baldassarre, A.; Tanzilli, G.; Morano, S.; Lenzi, A.; Basili, S.; Crescioli, C. Cardiomyopathy Associated with Diabetes: The Central Role of the Cardiomyocyte. IJMS 2019, 20, 3299. [Google Scholar] [CrossRef] [Green Version]
  180. de Zoete, M.R.; Palm, N.W.; Zhu, S.; Flavell, R.A. Inflammasomes. Cold Spring Harb. Perspect. Biol. 2014, 6, a016287. [Google Scholar] [CrossRef]
  181. Shalini, S.; Dorstyn, L.; Dawar, S.; Kumar, S. Old, new and emerging functions of caspases. Cell Death Differ. 2015, 22, 526–539. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  182. Guo, H.; Callaway, J.B.; Ting, J.P.-Y. Inflammasomes: Mechanism of action, role in disease, and therapeutics. Nat. Med. 2015, 21, 677–687. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. He, Y.; Hara, H.; Núñez, G. Mechanism and Regulation of NLRP3 Inflammasome Activation. Trends Biochem. Sci. 2016, 41, 1012–1021. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Pellegrini, C.; Antonioli, L.; Lopez-Castejon, G.; Blandizzi, C.; Fornai, M. Canonical and Non-Canonical Activation of NLRP3 Inflammasome at the Crossroad between Immune Tolerance and Intestinal Inflammation. Front. Immunol. 2017, 8, 36. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Sebastian-Valverde, M.; Pasinetti, G.M. The NLRP3 Inflammasome as a Critical Actor in the Inflammaging Process. Cells 2020, 9, 1552. [Google Scholar] [CrossRef]
  186. Youm, Y.-H.; Grant, R.W.; McCabe, L.R.; Albarado, D.C.; Nguyen, K.Y.; Ravussin, A.; Pistell, P.; Newman, S.; Carter, R.; Laque, A.; et al. Canonical Nlrp3 Inflammasome Links Systemic Low-Grade Inflammation to Functional Decline in Aging. Cell Metab. 2013, 18, 519–532. [Google Scholar] [CrossRef] [Green Version]
  187. Rao, Z.; Chen, X.; Wu, J.; Xiao, M.; Zhang, J.; Wang, B.; Fang, L.; Zhang, H.; Wang, X.; Yang, S.; et al. Vitamin D Receptor Inhibits NLRP3 Activation by Impeding Its BRCC3-Mediated Deubiquitination. Front. Immunol. 2019, 10, 2783. [Google Scholar] [CrossRef]
  188. Cao, R.; Ma, Y.; Li, S.; Shen, D.; Yang, S.; Wang, X.; Cao, Y.; Wang, Z.; Wei, Y.; Li, S.; et al. 1,25(OH)2D3 alleviates DSS-induced ulcerative colitis via inhibiting NLRP3 inflammasome activation. J. Leukoc. Biol. 2020, 108, 283–295. [Google Scholar] [CrossRef]
  189. Tate, M.D.; Mansell, A. An update on the NLRP3 inflammasome and influenza: The road to redemption or perdition? Curr. Opin. Immunol. 2018, 54, 80–85. [Google Scholar] [CrossRef]
  190. Crescioli, C. Vitamin D, exercise, and immune health in athletes: A narrative review. Front. Immunol. 2022, 13, 954994. [Google Scholar] [CrossRef]
  191. Marzetti, E.; Calvani, R.; Cesari, M.; Buford, T.W.; Lorenzi, M.; Behnke, B.J.; Leeuwenburgh, C. Mitochondrial dysfunction and sarcopenia of aging: From signaling pathways to clinical trials. Int. J. Biochem. Cell Biol. 2013, 45, 2288–2301. [Google Scholar] [CrossRef] [Green Version]
  192. Picca, A.; Calvani, R.; Leeuwenburgh, C.; Coelho-Junior, H.J.; Bernabei, R.; Landi, F.; Marzetti, E. Targeting mitochondrial quality control for treating sarcopenia: Lessons from physical exercise. Expert Opin. Ther. Targets 2019, 23, 153–160. [Google Scholar] [CrossRef] [PubMed]
  193. Landi, F.; Calvani, R.; Cesari, M.; Tosato, M.; Martone, A.M.; Ortolani, E.; Savera, G.; Salini, S.; Sisto, A.; Picca, A.; et al. Sarcopenia: An Overview on Current Definitions, Diagnosis and Treatment. Curr. Protein Pept. Sci. 2018, 19, 633–638. [Google Scholar] [CrossRef] [PubMed]
  194. Minciullo, P.L.; Catalano, A.; Mandraffino, G.; Casciaro, M.; Crucitti, A.; Maltese, G.; Morabito, N.; Lasco, A.; Gangemi, S.; Basile, G. Inflammaging and Anti-Inflammaging: The Role of Cytokines in Extreme Longevity. Arch. Immunol. Ther. Exp. 2016, 64, 111–126. [Google Scholar] [CrossRef] [PubMed]
  195. Ferrucci, L.; Fabbri, E. Inflammageing: Chronic inflammation in ageing, cardiovascular disease, and frailty. Nat. Rev. Cardiol. 2018, 15, 505–522. [Google Scholar] [CrossRef] [PubMed]
  196. Draganidis, D.; Karagounis, L.G.; Athanailidis, I.; Chatzinikolaou, A.; Jamurtas, A.Z.; Fatouros, I.G. Inflammaging and Skeletal Muscle: Can Protein Intake Make a Difference? J. Nutr. 2016, 146, 1940–1952. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Zembron-Lacny, A.; Dziubek, W.; Wolny-Rokicka, E.; Dabrowska, G.; Wozniewski, M. The Relation of Inflammaging With Skeletal Muscle Properties in Elderly Men. Am. J. Men's Health 2019, 13, 155798831984193. [Google Scholar] [CrossRef] [Green Version]
  198. Brown, E.C.; DiSilvestro, R.A.; Babaknia, A.; Devor, S.T. Soy versus whey protein bars: Effects on exercise training impact on lean body mass and antioxidant status. Nutr. J. 2004, 3, 22. [Google Scholar] [CrossRef]
  199. Pal, S.; Ellis, V. The Chronic Effects of Whey Proteins on Blood Pressure, Vascular Function, and Inflammatory Markers in Overweight Individuals. Obesity 2010, 18, 1354–1359. [Google Scholar] [CrossRef]
  200. Pal, S.; Ellis, V. Acute effects of whey protein isolate on blood pressure, vascular function and inflammatory markers in overweight postmenopausal women. Br. J. Nutr. 2011, 105, 1512–1519. [Google Scholar] [CrossRef]
  201. Rowlands, D.S.; Rössler, K.; Thorp, R.M.; Graham, D.F.; Timmons, B.W.; Stannard, S.R.; Tarnopolsky, M.A. Effect of dietary protein content during recovery from high-intensity cycling on subsequent performance and markers of stress, inflammation, and muscle damage in well-trained men. Appl. Physiol. Nutr. Metab. 2008, 33, 39–51. [Google Scholar] [CrossRef]
  202. Fernando, R.; Drescher, C.; Nowotny, K.; Grune, T.; Castro, J.P. Impaired proteostasis during skeletal muscle aging. Free. Radic. Biol. Med. 2019, 132, 58–66. [Google Scholar] [CrossRef]
  203. Guillet, C.; Zangarelli, A.; Gachon, P.; Morio, B.; Giraudet, C.; Rousset, P.; Boirie, Y. Whole Body Protein Breakdown Is Less Inhibited by Insulin, But Still Responsive to Amino Acid, in Nondiabetic Elderly Subjects. J. Clin. Endocrinol. Metab. 2004, 89, 6017–6024. [Google Scholar] [CrossRef] [Green Version]
  204. Rasmussen, B.B.; Fujita, S.; Wolfe, R.R.; Mittendorfer, B.; Roy, M.; Rowe, V.L.; Volpi, E. Insulin resistance of muscle protein metabolism in aging. FASEB J. 2006, 20, 768–769. [Google Scholar] [CrossRef] [Green Version]
  205. Wilkes, E.A.; Selby, A.L.; Atherton, P.J.; Patel, R.; Rankin, D.; Smith, K.; Rennie, M.J. Blunting of insulin inhibition of proteolysis in legs of older subjects may contribute to age-related sarcopenia. Am. J. Clin. Nutr. 2009, 90, 1343–1350. [Google Scholar] [CrossRef] [Green Version]
  206. Höhn, A.; Weber, D.; Jung, T.; Ott, C.; Hugo, M.; Kochlik, B.; Kehm, R.; König, J.; Grune, T.; Castro, J.P. Happily (n)ever after: Aging in the context of oxidative stress, proteostasis loss and cellular senescence. Redox Biol. 2017, 11, 482–501. [Google Scholar] [CrossRef] [PubMed]
  207. Dzik, K.P.; Kaczor, J.J. Mechanisms of vitamin D on skeletal muscle function: Oxidative stress, energy metabolism and anabolic state. Eur. J. Appl. Physiol. 2019, 119, 825–839. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  208. Bouillon, R.; Verstuyf, A. Vitamin D, Mitochondria, and Muscle. J. Clin. Endocrinol. Metab. 2013, 98, 961–963. [Google Scholar] [CrossRef] [PubMed]
  209. Ceglia, L.; Harris, S.S. Vitamin D and Its Role in Skeletal Muscle. Calcif. Tissue Int. 2013, 92, 151–162. [Google Scholar] [CrossRef]
  210. Benetti, E.; Mastrocola, R.; Chiazza, F.; Nigro, D.; D’Antona, G.; Bordano, V.; Fantozzi, R.; Aragno, M.; Collino, M.; Minetto, M.A. Effects of vitamin D on insulin resistance and myosteatosis in diet-induced obese mice. PLoS ONE 2018, 13, e0189707. [Google Scholar] [CrossRef]
  211. Lin, A.M.Y. Antioxidative Effect of Vitamin D3 on Zinc-Induced Oxidative Stress in CNS. Ann. N. Y. Acad. Sci. 2005, 1053, 319–329. [Google Scholar] [CrossRef]
  212. Nakai, K.; Fujii, H.; Kono, K.; Goto, S.; Kitazawa, R.; Kitazawa, S.; Hirata, M.; Shinohara, M.; Fukagawa, M.; Nishi, S. Vitamin D Activates the Nrf2-Keap1 Antioxidant Pathway and Ameliorates Nephropathy in Diabetic Rats. Am. J. Hypertens. 2014, 27, 586–595. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Tanaka, M.; Tokunaga, K.; Komaba, H.; Itoh, K.; Matsushita, K.; Watanabe, H.; Kadowaki, D.; Maruyama, T.; Otagiri, M.; Fukagawa, M. Vitamin D Receptor Activator Reduces Oxidative Stress in Hemodialysis Patients With Secondary Hyperparathyroidism: VDRA and Oxidative Stress. Ther. Apher. Dial. 2011, 15, 161–168. [Google Scholar] [CrossRef] [PubMed]
  214. Giresi, P.G.; Stevenson, E.J.; Theilhaber, J.; Koncarevic, A.; Parkington, J.; Fielding, R.A.; Kandarian, S.C. Identification of a molecular signature of sarcopenia. Physiol. Genom. 2005, 21, 253–263. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Srikuea, R.; Hirunsai, M. Effects of intramuscular administration of 1α,25(OH)2D3 during skeletal muscle regeneration on regenerative capacity, muscular fibrosis, and angiogenesis. J. Appl. Physiol. 2016, 120, 1381–1393. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  216. Trewin, A.J.; Lundell, L.S.; Perry, B.D.; Patil, K.V.; Chibalin, A.V.; Levinger, I.; McQuade, L.R.; Stepto, N.K. Effect of N-acetylcysteine infusion on exercise-induced modulation of insulin sensitivity and signaling pathways in human skeletal muscle. Am. J. Physiol. Endocrinol. Metab. 2015, 309, E388–E397. [Google Scholar] [CrossRef] [Green Version]
  217. Gomez-Cabrera, M.-C.; Domenech, E.; Romagnoli, M.; Arduini, A.; Borras, C.; Pallardo, F.V.; Sastre, J.; Viña, J. Oral administration of vitamin C decreases muscle mitochondrial biogenesis and hampers training-induced adaptations in endurance performance. Am. J. Clin. Nutr. 2008, 87, 142–149. [Google Scholar] [CrossRef] [Green Version]
  218. Ristow, M.; Zarse, K.; Oberbach, A.; Klöting, N.; Birringer, M.; Kiehntopf, M.; Stumvoll, M.; Kahn, C.R.; Blüher, M. Antioxidants prevent health-promoting effects of physical exercise in humans. Proc. Natl. Acad. Sci. USA 2009, 106, 8665–8670. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  219. Blank, S.; Scanlon, K.S.; Sinks, T.H.; Lett, S.; Falk, H. An outbreak of hypervitaminosis D associated with the overfortification of milk from a home-delivery dairy. Am. J. Public Health 1995, 85, 656–659. [Google Scholar] [CrossRef] [Green Version]
  220. Scanlon, K.S.; Blank, S.; Sinks, T.; Lett, S.; Mueller, P.; Freedman, D.S.; Serdula, M.; Falk, H. Subclinical health effects in a population exposed to excess vitamin D in milk. Am. J. Public Health 1995, 85, 1418–1422. [Google Scholar] [CrossRef]
  221. Holick, M.F.; Shao, Q.; Liu, W.W.; Chen, T.C. The vitamin D content of fortified milk and infant formula. N. Engl. J. Med. 1992, 326, 1178–1181. [Google Scholar] [CrossRef]
  222. Larsen, E.R.; Mosekilde, L.; Foldspang, A. Vitamin D and calcium supplementation prevents osteoporotic fractures in elderly community dwelling residents: A pragmatic population-based 3-year intervention study. J. Bone Miner. Res. 2004, 19, 370–378. [Google Scholar] [CrossRef] [PubMed]
  223. Giustina, A.; Adler, R.A.; Binkley, N.; Bollerslev, J.; Bouillon, R.; Dawson-Hughes, B.; Ebeling, P.R.; Feldman, D.; Formenti, A.M.; Lazaretti-Castro, M.; et al. Consensus statement from 2nd International Conference on Controversies in Vitamin D. Rev. Endocr. Metab. Disord. 2020, 21, 89–116. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Bischoff-Ferrari, H.A.; de Godoi Rezende Costa Molino, C.; Rival, S.; Vellas, B.; Rizzoli, R.; Kressig, R.W.; Kanis, J.A.; Manson, J.E.; Dawson-Hughes, B.; Orav, E.J.; et al. DO-HEALTH: Vitamin D3-Omega-3—Home exercise—Healthy aging and longevity trial—Design of a multinational clinical trial on healthy aging among European seniors. Contemp. Clin. Trials 2021, 100, 106124. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The main cellular modification induced by senescence. Aging induced modifications at the cellular level include morphological, enzymatic, and functional variations.
Figure 1. The main cellular modification induced by senescence. Aging induced modifications at the cellular level include morphological, enzymatic, and functional variations.
Ijms 24 04546 g001
Figure 2. The molecular mechanisms involved in the anti-aging effects of vitamin D. An adequate level of vitamin D counteracts age-dependent changes at multi-target biomolecular levels, ending in effects against infections, inflammation, oxidative stress, garbage molecule accumulation, and aberrant mitochondria functioning.
Figure 2. The molecular mechanisms involved in the anti-aging effects of vitamin D. An adequate level of vitamin D counteracts age-dependent changes at multi-target biomolecular levels, ending in effects against infections, inflammation, oxidative stress, garbage molecule accumulation, and aberrant mitochondria functioning.
Ijms 24 04546 g002
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

Fantini, C.; Corinaldesi, C.; Lenzi, A.; Migliaccio, S.; Crescioli, C. Vitamin D as a Shield against Aging. Int. J. Mol. Sci. 2023, 24, 4546. https://doi.org/10.3390/ijms24054546

AMA Style

Fantini C, Corinaldesi C, Lenzi A, Migliaccio S, Crescioli C. Vitamin D as a Shield against Aging. International Journal of Molecular Sciences. 2023; 24(5):4546. https://doi.org/10.3390/ijms24054546

Chicago/Turabian Style

Fantini, Cristina, Clarissa Corinaldesi, Andrea Lenzi, Silvia Migliaccio, and Clara Crescioli. 2023. "Vitamin D as a Shield against Aging" International Journal of Molecular Sciences 24, no. 5: 4546. https://doi.org/10.3390/ijms24054546

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