Next Article in Journal
Histiocytic Sarcoma: A Review and Update
Previous Article in Journal
Myopia and Metabolomics: A Comparative Study of Aqueous Humor and Serum Metabolites in Myopic Adults Undergoing Cataract Surgery
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Anti-Inflammatory Actions of Soluble Klotho in Brain Aging and Its Main Associated Diseases

by
Héctor E. López-Valdés
1,
Martín Hernández-Lucas
1,
Gustavo D. J. Rodríguez-Fabián
1,
Nadia F. Esteban-Román
1,
Roger Gutiérrez-Juárez
2,
Isabel Arrieta-Cruz
3 and
Hilda Martínez-Coria
1,*
1
Departamento de Fisiología, Facultad de Medicina, Universidad Nacional Autónoma de México, México City 04360, Mexico
2
Departamento de Ciencias Biomédicas, Escuela de Medicina, Facultad de Estudios Superiores Zaragoza, Universidad Nacional Autónoma de México, México City 09230, Mexico
3
Departamento de Investigación Básica, División de Investigación, Instituto Nacional de Geriatría, Secretaría de Salud, México City 10200, Mexico
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2025, 26(17), 8551; https://doi.org/10.3390/ijms26178551
Submission received: 25 June 2025 / Revised: 4 August 2025 / Accepted: 22 August 2025 / Published: 3 September 2025

Abstract

The anti-aging protein α-Klotho has several therapeutic effects on different pathophysiological conditions, mainly its anti-inflammatory and antioxidant effects. Experimental evidence and observational studies suggest that there are several strategies to increase α-Klotho in the brain and enhance its beneficial effects, thus contributing to improving its neuroprotective and neuroplasticity mechanisms in brain aging, Alzheimer’s, Parkinson’s, and ischemic stroke diseases. In this article, we summarize the relevant information on α-Klotho, brain aging, Alzheimer’s, Parkinson’s, and ischemic stroke diseases and analyze the role of α-Klotho in each of these alterations, as well as the effect of physical exercise, exogenous application of α-klotho, and various drugs approved for different human diseases on α-Klotho production.

1. Introduction

Alzheimer’s disease (AD), Parkinson’s disease (PD), and focal ischemic stroke (FIS) are diseases whose incidence increases significantly during aging and are among those that have the greatest impact on brain health in older adults [1]. Most of the diseases associated with aging, such as those mentioned above, have a common characteristic in that they present a chronic and low-intensity neuroinflammation process [1]. In this process, there is an alteration in intracellular communication known as inflammaging, in which there is an increase in the secretion of proinflammatory cytokines and inflammatory markers [2]. Neurodegenerative diseases such as AD and PD have no cure, and in the case of FIS, therapeutic options are limited (thrombolytic drugs and mechanical thrombectomy), have a very short therapeutic window, are accessible to a small percentage of patients, and are not without dangerous side effects, so most patients with FIS only have the option of rehabilitation therapies [3]. On the other hand, the α-Klotho protein is well known for its anti-aging effects, and an extensive body of knowledge has been accumulated over several decades on its various beneficial effects, including reducing oxidative stress and neuroinflammation [4,5]. In the following sections of this paper, we briefly review relevant information about the α-Klotho protein, brain aging, AD, PD, FIS, neuroinflammation, and the modulation of neuroinflammation by α-Klotho in these diseases. Finally, we will discuss the therapeutic potential of this protein in AD, PD, and FIS.

2. Klotho Subtypes

In November 1997, Kuro-o and colleagues reported a new gene found in mice that was related to the suppression of several phenotypic features of aging, and they named this gene klotho. Defects in this gene produced several alterations similar to those occurring in humans, such as osteoporosis, atherosclerosis, infertility, and premature death [6]. Six years later, it was reported that overexpression of this gene produces a protein that, when released into the circulation, acts as a hormone and extends lifespan in mice [7]. Since the publication of these two very important articles, much progress has been made in understanding Klotho. Currently, three Klotho proteins are known, coded by different genes called α, β, and γ [6,8,9], all of which are transmembrane proteins that interact with different types of receptors to promote high-affinity binding of different members of the fibroblast growth factor (FGF) family [8,9]. α-Klotho interacts with the FGF23 receptor known as FGFR1c, which is highly expressed in the kidneys and brain and has been described as an anti-aging protein, while β-Klotho interacts with the same type of receptor expressed in adipocytes and with FGFR4 expressed in the liver, which are receptors for FGF21 and FGF19, respectively [10]. Meanwhile, γ-Klotho interacts with FGFR1c, 2c, and 4, which are expressed in the connective tissue, kidneys, and eyes, interacting with FGF19 [11]. The FGF family members interacting with the Klotho subtypes have different physiological effects. For example, FGF19 is a satiety hormone, as it activates metabolic responses in response to feeding; FGF21 is a hunger hormone that induces stress responses by activating the hypothalamic–pituitary–adrenal axis and the peripheral nervous system; and lastly, FGF23 is a hormone that increases the excretion of phosphate [10]. α-Klotho acts as a co-receptor to increase the affinity of FGF23, which modulates phosphorus and vitamin metabolism in the body, and the main consequence of its down-regulation is chronic renal failure [12,13]. In the following sections, we will only mention α-Klotho because it is the most studied in relation to aging and CNS diseases.

2.1. α-Klotho as a Hormone

The α-klotho gene is located on human chromosome XIII and has two major isoforms, a 135 KDa transmembrane protein that serves as an obligatory co-receptor for FGF23, and a 70 kDa secreted form [14]. α-Klotho is a single-step transmembrane protein, having a short intracellular domain and a long extracellular domain with two segments (KL1 and KL2) which are cleaved by enzymes (ADAM 10 and ADAM17) to produce three soluble fragments (Figure 1) that can circulate in the blood, cerebrospinal fluid, and urine [10,15,16]. Additionally, another isoform of α-Klotho can be produced by alternative mRNA splicing and includes the KL1 domain [17], which is sometimes called secreted Klotho or se-Klotho, whose functionality was previously doubted [18], but is now known to be fully functional [19,20,21]. This isoform is much more abundant (10 times more) in the brain than the cleaved isoform and has only been identified in the mouse brain [22]. There is evidence that both cleaved and secreted α-Klotho are found in human cerebrospinal fluid (CSF) and that the concentrations found in CSF and plasma have no correlation [23]. Together with the fact that α-Klotho does not permeate the blood–brain barrier (BBB), evidence suggests that α-Klotho present in plasma and CSF are independently regulated in the kidney and choroid plexus, respectively [24,25]. Hereafter, we will refer to the soluble and secreted isoforms simply as α-Klotho, unless it is an effect or function characterized in a particular protein isoform. The main source of α-Klotho in the blood comes from the kidneys [6,26,27]. α-Klotho performs many functions through a mechanism that is independent of FGF receptors and involves the modulation of sialic acid-containing cell membrane microdomains, such as monosialogangliosides GM1 and GM3 present in lipid rafts, which function as signaling hubs for trophic events [28,29,30]. Other studies have shown that α-Klotho regulates the activity of several ion channels, ion transporters, and growth factor receptors on the cell membrane surface. For example, it increases transient receptor potential cation channels, subfamily V, member 5 channels (TRPV5), and the ROMK1-mediated potassium current [31,32] in the cell membrane. In these channels, α-Klotho removes the sialic acid terminus of N-glycan, which prevents its endocytosis and thus increases its abundance in renal tubule channels and results in an improvement in Ca2+ and K+ regulation [33]. In the arcuate nucleus (ARC) of the hypothalamus, α-Klotho produces complex effects; in 44% of proopiomelanocortin (POMC) neurons, it produces excitatory effects (depolarizes and increases firing rate), and in 55%, it produces inhibitory effects (decreases membrane potential and firing rate) [34]. Additionally, there are many observational studies that establish a correlation between plasma α-Klotho concentration and different disorders and aging. For example, CSF α-Klotho concentrations are lower in older versus younger adults [35], and higher plasma α-Klotho concentrations were associated with a lower risk of meaningful decline and a smaller average decline according to the Mini-Mental State Examination [36]. Other studies have also shown that low plasma concentrations of α-Klotho in older adults are associated with cognitive function decline [37,38]. Low α-Klotho concentrations have also been associated with various diseases such as cardiovascular diseases [39], metabolic syndrome [40], and chronic kidney disease [41], and due to the inability of α-Klotho to cross the BBB, it has been suggested that the primary source is the kidneys since some studies show that this protein mainly comes from these organs [24]. Since α-Klotho concentration in blood plasma has a negative correlation with health status in observational studies of patients with various diseases including nerve diseases, it is thought to be a useful biomarker of health status. Some research groups actually propose to include it as a biomarker of aging and frailty [42]. Recently, a negative correlation has been found between plasma α-Klotho concentration and the SII or systemic immuno-inflammatory index [43], which provides a quantitative measure of the inflammatory immune response of the human body [44]. However, it is not known whether the association between blood α-Klotho levels and patient status is causal.

2.2. α-klotho Polymorphism

Within the human population, it has been found that the α-klotho gene has slight variations (mutations) in its DNA sequence (polymorphic gene), and several variants (alleles) have been found to vary by a single nucleotide (single-nucleotide polymorphism or SNP). It is known that these variations can influence the function of the protein. α-Klotho can be measured and studied by obtaining the protein present in blood serum and it has been used to study if there is any correlation between any polymorphism and longevity, cognition, or any disease. For example, the best studied polymorphism is the functional variant KL-VS, which contains six SNP sequence variants. A meta-analysis concluded that it is positively associated with exceptional human longevity [45]. However, in another study, two particular polymorphisms (G-395A and F352V) were examined and it was found that the G-395A polymorphism is associated with a high risk of urolithiasis and cardiovascular disease but not with cognitive decline, while F352V is associated with protection against breast cancer and longevity [46]. In another study, the KL-VS polymorphism (rs95363147/g) was associated with an increased incidence of dementia in adults [47]. Other research papers have found differences in their correlation with an elevated or reduced risk of disease, depending on whether the mutations are in both copies of the gene (homozygous genotype) or only in one (heterozygous genotype), e.g., homozygous KL-VS individuals are at a higher risk of stroke than heterozygous individuals [48]. Further studies are needed to more clearly establish the relationship between polymorphic variants of klotho and resistance or susceptibility to a specific disease.

2.3. α-Klotho in the Brain

α-Klotho is expressed in neurons of several brain regions such as the hippocampus, cortex, striatum, and Purkinje cells in the cerebellum, but the highest expression levels are in the choroid plexus cells [6,49,50]. Some of the functions of α-Klotho in the brain were elucidated in the hippomorphic kl/kl mice developed by Kuro-o, showing an aging phenotype [6], and in mice that overexpress α-klotho [7]. From the work with these animals, it can be concluded that the absence of the klotho gene causes disorders in axonal transport, a reduction in the number of synapses in the hippocampus, and neurodegeneration in the same tissue. Moreover, reduced Purkinje cells in the cerebellum were also reported. Meanwhile, overexpression of the gene produces an improvement in cognitive functions and, in particular, memory-related tasks, possibly by increasing the number of glutamatergic synapses in the hippocampus and cortex [51,52,53]. In addition, α-Klotho seems to exert a regulatory effect on oligodendrocytes and myelin, since it has been reported that in mice deficient in this protein, the number of oligodendrocytes is lower and myelin is also reduced, while overexpression of the gene produces an increase in myelin. In addition, there is a correlation between the decrease in α-Klotho and cerebral white matter lesions [54,55,56]. The brain functions regulated by α-Klotho are performed via the same mechanism mentioned above, i.e., by modulating sialic acid-containing microdomains of cell membranes such as monosialogangliosides GM1 and GM3 present in lipid rafts and regulating ion channels that can influence intracellular signaling and cell-to-cell communication [28,29,30].

3. Brain Aging and α-Klotho

3.1. Aging Brain

Aging is a complex and inevitable process that results from the interaction of environmental, epigenetic, and genetic factors that produce a gradual reduction in homeostasis with age [1]. This process is characterized by progressive tissue degeneration and a loss of function in various organs, which contributes to an increased risk of disease and death [57,58]. Aging affects individuals differently, but some changes occur more frequently. At present, twelve hallmarks are recognized in aging and these are interconnected with each other. These hallmarks are telomere attrition, genomic instability, epigenetic alterations, loss of proteostasis, cellular senescence, disabled macroautophagy, deregulated nutrient-sensing, altered intercellular communication, mitochondrial dysfunction, stem cell exhaustion, dysbiosis, and inflammation [59]. The aging human brain experiences both macroscopic and microscopic changes, including a decrease in white and gray matter density, thinning of the cortex, atrophic regions (e.g., hippocampus), enlargement of the ventricles, a decrease in blood flow, and a reduction in synapse and neurotransmitter release [60,61]. Diseases whose incidence increases with age are known as aging-related diseases and include neurodegenerative (e.g., AD and PD), metabolic (e.g., type 2 diabetes), cancer, and cerebrovascular diseases (e.g., FIS). Chronic low-intensity neuroinflammation is a feature of aging and is thought to be an important factor in amplifying damage in different diseases by accelerating neuronal dysfunction and death [62]. There are two aspects of brain aging that warrant a more detailed description. The first is cognitive dysfunction, as it represents a manifestation of an underlying disease and is common in many neurodegenerative diseases. The other is the physical barriers that separate blood from brain tissue and cerebrospinal fluid, due to their implications in the neuroinflammatory response, which will be described separately below.

3.1.1. Aging-Associated Cognitive Impairment

Cognitive dysfunction describes the decline in a person’s mental abilities and is known to increase with advancing age [63]. Cognitive aging is one of the obvious signs of neurodegenerative diseases such as AD [64]. Cognitive functions include learning and memory, information manipulation, attention, perception, decision making, speech and language, and social cognition [65]. Cognitive aging is the result of anatomical and functional changes in different brain regions associated with a particular cognitive function, and the magnitude of the alteration is related to genetic and environmental factors (lifestyle, education, experience, emotional factors, health status, and socio-economic factors), which vary among individuals [65]. Chronic peripheral and central inflammation are characteristics of aging and are linked to cognitive aging. Different studies have shown that markers of inflammation present in blood and CSF are associated with increased cognitive decline and progression of neurodegenerative diseases, especially in older people [66,67,68,69]. In animal studies, high levels of proinflammatory cytokines in blood serum and brain parenchyma are associated with cognitive deficits, particularly learning and memory [70], and the use of neutralizing antibodies against IL-17A, a proinflammatory cytokine, improves memory and reduces neuroinflammation [71]. Similarly, the use of the inhibitor PLX3397 and minocycline, compounds that inactivate reactive microglia, improved learning and memory in an animal model of PD [72].

3.1.2. Age-Related Dysfunctions of the Blood–Brain Barrier and Blood–Brain–Cerebrospinal Fluid Barrier

The BBB and blood–cerebrospinal fluid barrier (BCSFB) deserve special mention as they act as a physical barrier for the peripheral immune system, preventing immune cells and antibodies from entering the brain, which would eventually cause an excessive inflammatory response [73,74]. The BBB is a structure present in the network of cerebral capillaries, which is formed by endothelial cells associated by tight junctions, pericytes, astrocytes, and a basement membrane that functions as a highly selective filter that regulates the passage of substances between the blood and brain tissue [74]. The main functions of the BBB are as follows: (1) protection, which prevents harmful substances from entering brain tissue. (2) maintaining homeostasis, controlling the entry of essential nutrients such as glucose and amino acids and the exit of metabolic waste, as well as limiting the penetration of drugs used to treat CNS diseases such as neurodegenerative diseases and cancer [74]. The BBB also undergoes alterations during healthy aging, and these alterations are considerably aggravated when neuroinflammation occurs, as in neurodegenerative diseases such as AD, PD, and stroke, where there is degradation of endothelial cells, a decrease in proteins that form tight junctions (e.g., claudin, occludin), and increased permeability, which, combined with endothelial cell senescence, greatly contribute to damage to the BBB and neurovascular coupling dysfunction. All of these events result in an increase in proinflammatory cytokines (e.g., IL-1β, IL-2, IL-6, TNF), activation of microglia and astrocytes, leukocyte infiltration, and increased oxidative stress [75]. On the other hand, the blood–cerebrospinal fluid barrier (BCSFB) is a structure specialized in regulating the movement of substances between the blood and cerebrospinal fluid. It is in cerebral ventricles and is formed by the choroid plexuses, which form a barrier between the blood transported in the fenestrated capillaries and the cerebrospinal fluid [76]. In addition, the choroid plexuses produce cerebrospinal fluid and α-Klotho [73]. The choroid plexuses are made up of epithelial cells, fibroblasts, endothelial cells, and macrophages, the latter located in the stroma and epiplexus [77]. In aging and neurodegenerative diseases such as AD, this barrier is dysfunctional and its integrity is compromised, facilitating the migration of circulating immune cells to the CSF, which is a marker of neuroinflammation in age-related neurodegenerative diseases [76]. Few studies have investigated α-Klotho on cerebral microvasculature, and it has been shown that endothelial cells in the human cerebral microvasculature differ from those in other parts of the body; for example, they do not express the protein and do not show some responses to FGF-23 stimulation that are dependent on α-Klotho, such as cell proliferation and nitric oxide production [78]. On the contrary, there is evidence in mice that a decrease in α-Klotho causes BCSFB dysfunction, as it has been shown that a decrease in α-Klotho secretion in the choroid plexuses causes macrophage infiltration, an increase in proinflammatory mediators, and microglia activation [79]. In the following sections, we will briefly describe neuroinflammation in aging, the most common diseases associated with aging, and the relevance of neuroinflammation in these diseases, and lastly, we will analyze the modulatory effect of α-Klotho on this pathological process in each of these diseases.

3.2. Neuroinflammation

Neuroinflammation refers to the response of the innate immune system of the central nervous system (CNS) when tissue damage or infection is detected. Acutely, it is a response that promotes tissue recovery, but chronically, it can promote damage [80]. Microglia and astrocytes are the main cells of the immune response in the brain and maintain bidirectional communication. They perform different functions during neuroinflammation and can promote protective or damaging actions, depending on the context of the inflammation. The main immune cell in the CNS is the microglia and these cells continuously monitor the parenchymal environment; if tissue damage occurs, they are the first to respond [81]. The immune response begins when the damaged or dead cells present in the CNS of non-communicable diseases release molecules (damage-associated molecular patterns: DAMPs) that are recognized by a family of receptors present in the cells, known as pattern recognition receptors (PRRs), which include different subfamilies, one of which is the Toll-like (TLR) family [62,82]. DAMPs include several molecules such as cytoplasmic and nuclear proteins released in necrosis, cytokines of the interleukin 1 (IL-1) family, ATP, and proteins present in diseases such as AD (Amyloid β) and PD (α-synuclein) [83]. The different DAMPs are usually recognized by different subtypes of TLRs and some of these are widely expressed by microglia and astrocytes and to a lesser extent by neurons and oligodendrocytes [84]. The interaction of DAMPs with their respective receptors in the cells triggers the activation of an intracellular signaling cascade such as mitogen-activated protein kinase (MAPK), phosphoinositide 3-kinase/protein kinase B (PI3K/AKT), and the mammalian target of rapamycin (mTOR), which in turn leads to the activation of different transcription factors such as nuclear factor kappa-B (NF-κB), activator protein 1 (AP-1), nitric oxide synthase (iNOS), and cyclooxygenase-2 (COX-2). These transcription factors regulate the production of proinflammatory cytokines and other mediators of the immune response, forming a cellular protection mechanism, but if stimulus clearance fails, this response can become chronic and harmful [1,85]. In chronic inflammation, once PRRs are activated, microglia undergo morphofunctional changes (reactive microglia) that induce the release of proinflammatory cytokines such as interleukin 1β (IL-1β), interleukin 6 (IL-6), tumor necrosis factor α (TNF-α), and interferon γ (IFN-γ), stimulating astrocytes in the area. Similar to what occurs in the microglia, the activation of PRRs triggers a series of morphofunctional changes in astrocytes (reactive astrocytes) that, in turn, induce a second inflammatory response because these cells also release proinflammatory cytokines such as IL-1β, TNF-α, and IL-6 [86]. Importantly, the changes experienced by microglia and astrocytes following PRR activation are multiple and are intimately related to the particular microenvironment in which the cells are found; so, the changes experienced vary between different physiological stages and across diseases, and even during the progression of a particular disease [87,88].

3.3. Neuroinflammation in Aging (Inflammaging)

In aging, there is an alteration in intercellular communication known as inflammaging, which consists of a state of low-intensity inflammation with an increase in the secretion of proinflammatory cytokines and inflammatory markers, but this alteration generally does not initiate the process of neurodegeneration, though it does contribute to magnifying the disease and is believed to participate in neuronal dysfunction and death [81]. This state of chronic low-intensity inflammation is in addition to another important feature of aging, cellular senescence, which was initially referred to as permanent proliferative arrest and therefore included only proliferating cells, but more recently, it has been shown that except for proliferative arrest, most of the other features of senescence (DNA damage, mitochondrial dysfunction, increased level of senescence-associated beta-galactosidase (SA-β-gal), senescence-associated secretory phenotypes (SASPs), etc.) are present in post-mitotic cells such as neurons [89]. In some neurodegenerative diseases and FIS, senescent cells are found, and it has been proposed that this contributes to the pathology of these disorders [90]. This contribution can be illustrated by the mitochondrial dysfunction that occurs in senescent cells which induces the secretion of several DAMPs that, in turn, activate multiple inflammatory cascades in neighboring cells, causing a chronic inflammatory response [62]. Several in vivo and in vitro studies show that microglia show important changes that suggest a reduction in their ability to perform their homeostatic functions and also that they may be senescent, e.g., they show an exaggerated inflammatory response, secrete more ROS, and their phagocytic and chemotactic capacity is reduced [91,92,93,94]. Furthermore, in aging and also in neurodegenerative diseases and FIS, there are alterations in the BBB that provoke the infiltration of cells of the peripheral immune system, favoring tissue regeneration or increasing neuroinflammation. In light of this, it is important to point out the role of the adaptive immune system in aging and these diseases. Recently, it has become evident that the gut–brain microbiota axis plays a very important role in aging and neuroinflammation. Intestinal dysbiosis may cause damage to the intestinal epithelium and provoke an immune response in the gut with the consequent release of proinflammatory cytokines that may translocate into the circulation and promote a systemic inflammation that may contribute to inflammaging and damage to the BBB, which would favor the infiltration of immune cells and contribute to increasing neuroinflammation and neurodegeneration [95]. Research suggests that the microbiota plays a role in neuroinflammation in AD, PD, and FIS. For example, short-chain fatty acids (SCFAs) from the microbiota increased microglial reactivity and promoted amyloid-beta deposition in an amyloidosis animal model [96]. Intestinal dysbiosis was found in a PD model, and transplantation of microbiota from these animals to healthy animals resulted in motor dysfunctions [97]. Similarly, in a PD animal model, transplantation of the microbiota from PD patients increased disorders in the animals compared to those who received microbiota transplants from healthy people [98].

3.4. α-Klotho Effects in Brain Aging

Although α-Klotho is a protein with anti-aging effects, it is important to note that its expression decreases with age [99], and this decline is associated with poor cognitive performance. For example, in one study, it was found that in women and men with decreased Klotho, low blood serum α-Klotho levels were associated with poor performance on cognitive tests [38]. Decades ago, it was observed that overexpression of this protein increased longevity and improved health, while its inhibition caused an acceleration in the presentation of a phenotype associated with aging [6,100]. α-Klotho has a protective effect on cognitive decline [25,53], through modulation of different mechanisms, including increased glutamatergic synapses [53], decreased endoplasmic reticulum (ER) stress, apoptosis, and vascular senescence [101,102], increased antioxidant mechanisms [103], the promotion of autophagy [104], the elimination of misfolded proteins, increased oligodendrocyte maturation and myelination of the CNS [105], and reduced gliosis and neuroinflammation [106,107,108]. In the last process, α-Klotho decreases the activation of NF-κB, reducing the production of proinflammatory cytokines [106,108,109]. α-Klotho deficiency causes activation of the Wnt/β catenin signaling pathway, which contributes to the aging process and germ cell exhaustion [110,111]. In contrast, α-Klotho interacts with soluble agonists as an antagonist of this pathway and blocks their interaction with receptors in such a way that it acts as an antagonist of Wnt/β catenin signaling [112]. In senescence-accelerated transgenic mice (SAMP 8), α-Klotho upregulation improved memory deficits and decreased the PI3/Akt pathway [113]. In a recent article, treatment with α-Klotho in old non-human primates significantly improved memory [114]. Using cultured human astrocytes and endothelial cells, it was observed that the secretion of proinflammatory substances such as Activin A and IL-1α secreted by senescent cells decreased mRNA in these cells, and that senescent cells transplanted into young mice produced a significant decrease in α-Klotho in the brain, kidneys, and urine [115]. Using an in vitro model of neuroinflammation (stimulation with lipopolysaccharide or LPS), it was found that treatment with α-Klotho significantly decreased TNF-α and IL-6 production, reversed NF-κB activation in glial cells, and decreased neuronal death caused by the secretion of these substances [116]. Despite the relevance of BBB dysfunction in neuroinflammation, there are no studies investigating the direct effect of α-Klotho on the cerebral vasculature, However, there is evidence that exposure of peripheral blood mononuclear cells from AD patients exposed to α-Klotho reduces plasma levels of IL-1β, IL-6 and TNF-α [107], suggesting that increased peripheral α-Klotho may decrease the effects of peripheral inflammation and thereby reduce BBB damage and neuroinflammation. Furthermore, there is evidence in mice that a decrease in Klotho causes BCSFB dysfunction, as it has been shown that a decrease in α-Klotho secretion in the choroid plexuses causes macrophage infiltration, an increase in proinflammatory mediators, and microglia activation [79].

4. Main Neurological Disorders Associated with Aging

The most common aging-related diseases include FIS and incurable neurodegenerative diseases such as AD and PD. Neurodegenerative diseases are different neurological disorders where modifications occur in specific areas, primarily in the synapses and neuronal networks, and in some cases, abnormal proteins accumulate and lead to cell death. The most common neurodegenerative disorders include, in addition to AD and PD, Huntington’s disease, and amyotrophic lateral sclerosis, and they all share a characteristic with other diseases associated with aging, such as FIS, this being that they all present chronic inflammation [117].

4.1. Alzheimer’s Disease (AD)

AD is a disease resulting from neuronal loss and atrophy of large areas of the hippocampus and cerebral cortex manifested by memory loss and difficulties in communication, problem solving, and behavior [118]. AD is generally grouped into two types, familial (FAD) and sporadic (SAD), also called idiopathic. FAD accounts for about 5% of reported cases and is caused by dominant mutations in the gene coding for Presynilin 1 (PSEN1), Presynilin 2 (PRESN 2), and the beta-amyloid protein precursor (APP). SAD, on the other hand, represents the remaining 95% of reported cases and has no genetic cause. Another frequently used classification, based on the age at disease onset, is division into late onset (≥65 years of age), which includes mostly people with SAD, and early onset (<65 years of age), which is mostly FAD [118,119]. The underlying cause of AD is still unknown but it is associated with synaptic loss and the presence of senile plaques, which are the accumulation of insoluble forms of the amyloid-beta (Aβ) protein that form extracellular plaques, and also intracellular accumulation of neurofibrillary tangles (NTF), which are the abnormal accumulation of hyperphosphorylated tau protein filaments, resulting in a loss of cytoskeletal microtubules and tubulin-associated proteins [120]. Among the hypotheses proposed to explain the causes of the disease, the three most widely accepted are as follows: (1) the cholinergic; (2) the amyloid, and (3) the Tau hypotheses. The first hypothesis essentially states that cognitive deficit is due to severe neurodegeneration of the cholinergic innervation, while the amyloid hypothesis suggests that increased Aβ42/Aβ40 induces neurodegeneration and cell death, and finally, the Tau hypothesis proposes that fibrillary aggregates of misfolded NTF proteins accumulate intracellularly in neurons and spread to other cells in a prion-like manner and disseminate throughout the brain [118,121].

4.1.1. Neuroinflammation in AD

Microglia and astrocytes play a very important role in the pathology of AD. For example, there is evidence that in the early stage of the disease, neuroinflammation is already present. It has been proposed that this process is present several decades before the onset of severe cognitive impairment and that it is strongly linked to the accumulation of Aβ and NTF [122]. Activated microglia are involved in the protection of neurons and the degradation of Aβ, but if they fail to clear plaques, they become chronically activated, resulting in the secretion of proinflammatory cytokines and may also contribute to promoting plaque formation, as shown in one publication, where pharmacological removal of microglia resulted in a significant reduction in plaque deposition and intraneuronal amyloid [123,124]. In postmortem studies of brains from AD patients, alterations in the morphology of microglia and increased immunoreactivity to ionized calcium-binding adaptor molecule 1 (Iba-1) were found, suggesting the presence of activated microglia [125]. Some PET studies in humans using markers for activated microglia such as 11PK11195 and 11 PBR28 [126,127,128] have found these cells in several regions (e.g., frontal, temporal, parietal, occipital, and cingulate cortices). Other studies have used these markers for activated microglia in combination with markers for Aβ (e.g., 18 F-flutemetamol) or Tau (e.g., 18 F-AV145) and found a convergence of signals of the association cortex [129].

4.1.2. α-Klotho Effects in AD

In an animal model (APP/PS1) of AD, α-Klotho overexpression significantly decreased the percentage of Aβ plaques in the cortex and hippocampus and soluble Aβ1-40 and Aβ1-42, and also decreased NLRP3/caspase-1 inflammasome signaling [108]. In another study using the same animal model, α-klotho overexpression was found to decrease cognitive deficits, inhibit AKT and mTOR pathways, and decrease Aβ1-40 in the brain [130]. In another animal model (SAMP8), Ligustilide (a molecule extracted in Chinese medicine plants such as Angelica sinensis) was found to decrease severe memory deficits, AB1-40 accumulation, and neuronal loss, among other beneficial effects, and also increased α-klotho expression in choroid plexuses and serum [131]. Subsequently, this natural compound was also found to increase soluble APP processing and autophagy [132]. In a study conducted in people of both sexes and in a wide age range, it was found that α-Klotho levels in serum and CSF levels are positively correlated and that both levels are negatively correlated with clinical dementia rating [133]. In a study using MRI in elderly people with cognitive decline, reduced α-Klotho levels were found to correlate with brain lesions in deep white matter [54].

4.2. Parkinson’s Disease (PD)

PD is a multifactorial condition manifested by bradykinesia, resting tremors, and rigidity, and it shares these symptoms with other disorders that together are known as parkinsonism [134]. In PD, other disorders such as sleep disturbance, depression, and dementia are also observed [135]. The main microscopic features of PD are degeneration of the dopaminergic neurons of the substantia nigra pars compacta and intraneuronal protein aggregates, known as Lewy bodies [136]. Most PD patients correspond to the idiopathic type and very few cases are of genetic origin. In the latter type, several mutated genes are involved, such as α-synuclein (the main component of Lewy bodies), in which point mutations have been found in addition to gene duplication or triplication. On the other hand, in the idiopathic form, several risk factors have been identified, such as aging, exposure to pesticides, and traumatic brain injury [136]. In non-pathological conditions, α-synuclein has two conformations, a soluble unfolded monomer and membrane-bound multimeric α-helical form. α-synuclein is involved in various cellular functions in neurons, such as dopamine synthesis and vesicle trafficking. Under pathological conditions, the soluble unfolded monomeric form generates oligomers called protofibrils that transform into insoluble fibrils, which can spread from neuron to neuron via a transcellular mechanism [136]. In addition to the neurons of the substantia nigra, other structures such as the basal nucleus of Meynert, the raphe nuclei, the amygdala, and the neocortex also show Lewy body inclusions [137].

4.2.1. Neuroinflammation in PD

The presence of neuroinflammation in the pathology of Parkinson’s disease is observed in various studies, such as postmortem or in vivo PET studies in patients. In postmortem studies, activated astrocytes and microglia have been found in the substantia nigra of brains of patients with the disease [138,139,140]. In a PET study, activated microglia were found to be present in several regions, including the substantia nigra, caudate nucleus, and frontal lobe [141]. Some studies using animal models and postmortem studies of patients have provided evidence for the possible involvement of the adaptive immune system; for example, T cells (CD4+ and CD8+) were found in the substantia nigra [142]. Some in vitro studies show that α-synuclein induces the activation of microglia and increases the production of TNF α and IL-1B [143]. α-synuclein induces the activation of the NF-κB signaling pathway in microglia and this signaling pathway leads to the activation of the NLRP3 inflammasome, which in turn promotes the release of IL-1β [144].

4.2.2. α-Klotho Effects in PD

In klotho-deficient mice, a lower number of dopaminergic neurons are observed in the substantia nigra and ventral tegmental area of the brain, and in the same type of animals, but older, a striatal reduction in dopamine is observed [145]. In a transgenic mouse model overexpressing α-synuclein, involving both young and old mice, it was observed that I.P. application of an α-Klotho fragment produced a significant improvement in learning and memory tests, but this improvement was blocked by treatment with GluN2B subunit blockers of the NMDA-type glutamate receptor, although the mechanism by which this process occurs is not known [25]. In a different animal model of PD (Intrastriatal 6-OHDA-lesioned rats), pretreatment with α-Klotho produced a decrease in reactive oxygen species, GFAP, DNA fragmentation, and α-synuclein, among other indicators, thus collectively providing neuroprotection [106]. In a study of PD patients, high levels of α-Klotho were found to be associated with a slower onset of cognitive decline [146]. Furthermore, in a small study involving patients with PD, it was found that α-Klotho is decreased in blood serum but increased in cerebrospinal fluid, and in the latter compartment, α-synuclein is inversely related to α-Klotho, suggesting that α-Klotho in cerebrospinal fluid is associated with the neurodegenerative process and α-Klotho in serum is related to peripheral and systemic changes [147].

4.3. Focal Ischemic Stroke (FIS)

Stroke is the most common cerebrovascular disease. According to the World Stroke Organization, each year, approximately 12.2 million people suffer a stroke, half of whom may die, while survivors may experience severe impairment in functions such as mobility, language, and emotions. In addition, it also states that most cases occur between the ages of 50 and 70 [148]. There are two main types of stroke, hemorrhagic, which represents ~15%, and ischemic, which represents the remaining ~85%; within the latter type, the most common is focal ischemic infarction, which is caused by the occlusion of a cerebral artery from either the local formation of a thrombus or an embolus that originated distally [3]. Stroke can be caused by a number of factors but the most important are small vessel disease (including arteriosclerosis and cerebral amyloid angiopathy), cardioembolism, and large artery disease, caused by stenosis or the occlusion of a large artery in the brain [149].
Currently, there are two types of treatments for this condition, one is treatment with a fibrinolytic agent and the other is via mechanical removal. Both are effective, but their therapeutic window is very short, and they are not without side effects. Treatment with a fibrinolytic agent such as recombinant tissue plasminogen activator (rt-PA, alteplase) is effective but its therapeutic window is four and a half hours from the onset of stroke and there is a potential risk of hemorrhage. Mechanical removal, or endovascular thrombectomy, is also effective and its therapeutic window can be extended up to six hours from the onset of stroke. In both cases, it is necessary to evaluate patients and determine the suitability of the treatment options in order to minimize side effects. Generally, these treatments are only applied in places where trained personnel and imaging units are available [150,151].
The reduction or cessation of blood perfusion in a specific area of the brain decreases the oxygen and glucose supply necessary to meet metabolic demand and can cause different levels of damage depending on different factors, for example, the magnitude and time of the reduction in blood flow and the susceptibility of the damaged tissue. In general, two different areas are observed at the injury site, one necrotic and the other with viable tissue surrounding the necrotic zone (peri-infarct), in which different events take place such as excitotoxicity, ionic alterations, cytotoxic edema, peri-infarct depolarization, neuroinflammation, etc., which together constitute the ischemic cascade [3].

4.3.1. Neuroinflammation in FIS

Excitotoxicity and other damage mechanisms harm and kill cells in the stroke zone, releasing DAMPs that eventually interact with their respective receptors on the different cell types present in the peri-infarct zone and within these cells. The first to respond is the microglia, which quickly move towards the damaged zone and release different proinflammatory cytokines (e.g., TNFα, IL-1α and C1q) to stimulate the generation of reactive astrocytes with a proinflammatory secretory profile; they also secrete other compounds (e.g., MMP-9) that contribute to the weakening of the BBB and facilitate the infiltration of leukocytes, exacerbating cell death. The weakening of the BBB continues, its permeability is further increased, and vasogenic edema occurs; the activated astrocytes release proinflammatory cytokines (IL-6, TNFα, IL-1α, IL-1β, IFNϒ) and free radicals. Also, a subset of reactive astrocytes proliferate and together with other cell types (e.g., fibroblasts, microglia) begin to form the glial scar that delimits the area of injury and restricts the diffusion of neuroinflammation [3,152]. After stroke, neuroinflammation may persist for days or months, as long as tissue damage is not resolved, and if this is the case, the activated microglia that initially secreted anti-inflammatory cytokines undergo morphophysiological changes and secrete proinflammatory cytokines [3,152]. In recent years, evidence has been accumulating in humans and animals linking the development of cognitive disorders and neuroinflammation. Post-stroke secondary inflammation has been found at sites distant from the primary lesion but related to the altered function [153,154,155,156,157]. Cognitive disturbances or dementia are common among stroke survivors, and these disorders together occur in approximately 70% of patients [158,159]. The pathophysiology of these alterations is not well known, but it is known that several factors favor their presentation, including the severity of the stroke, the volume of the lesion, the site, and pre-existing diseases [160]. Post-stroke cognitive impairment (PSCI) commonly occurs during the first year and may be completely reversible in some cases (20%), but for the majority, complete recovery does not occur and even about 30% develop dementia within the first 5 years [161]. PSCI manifests heterogeneously and some of the common disturbances include selective attention, language learning, episodic memory, and memory impairment [162].

4.3.2. α-Klotho Effects in FIS

As mentioned above, Klotho decreases with age and Klotho deficiency is known to be associated with vascular calcification, impaired angiogenesis and vasculogenesis, and loss of arterial wall elasticity, which contributes to increased blood pressure and cardiovascular diseases [6,163]. Endothelial dysfunction and increased arterial stiffness contribute to increased blood pressure, cardiovascular problems, and cerebral infarction [164]. A low concentration of α-Klotho is also associated with cerebral small vessel disease [165], vascular cognitive impairment [166], and poor outcomes after three months of acute stroke [167]. In an observational study, the klotho KL:VS variant was found to be associated with stroke in young adults in India [168]. In animal models, overexpression of klotho significantly improves neurobehavioral deficits [169,170] and inhibits the inflammatory response by activating the Nrf2 pathway, which decreases the activity of the NLRP3 inflammasome, causing decreased neuroinflammation [171], while it also improves endothelial function by increasing nitric oxide [172] and reduces infarct size in MCAO rats via a mechanism involving a reduction in aquaporin 4 (AQP4). In a mice model, α-Klotho decreased GFAP and Iba-1 immunoreactivity in the ischemic tissue, inhibiting RIG-I/NF-kB signaling activation and proinflammatory cytokine production [169]. Another study showed similar results, finding decreased reactive astrocytes and proinflammatory cytokines [170]. Some in vitro studies on endothelial cells showed that α-Klotho increased nitric oxide and the anti-oxidative effect by enhancing the expression of manganese superoxide dismutase (MnSOD) [173] and suppressing the proinflammatory cytokine TNF-α [174].

5. Discussion

Chronic neuroinflammation is present in AD, PD, and FIS and induces alterations ranging from the molecular level to the dysfunction of neuronal networks that manifest themselves in a decline in complex functions such as learning, memory, and neuroplasticity processes. AD and PD cannot be cured, and in the case of FIS, existing treatments have a very short therapeutic window and only a very small percentage of patients receive them; therefore, in most cases, they depend on rehabilitation therapy. In addition, post-stroke cognitive impairments are common and contribute significantly to increased mortality, permanent disability, and a decreased quality of life. Therefore, it is essential to continue the search for new compounds that can help reduce the harmful effects of these diseases.
α-Klotho has several beneficial effects on the brain, including decreasing oxidative stress, increasing synaptic plasticity, memory, and learning, and modulating the inflammatory response. The evidence that α-Klotho can decrease the inflammatory response comes from multiple sources and different approaches, such as overexpression of α-klotho in different animal models (AD, PD, etc.), direct administration in preclinic studies, and positive correlation in observational studies in humans. However, the mechanism of these effects is not fully understood. There is evidence that upregulation of α-klotho blocks important molecules in the development of the inflammatory response such as the inflammasome NLRP3 and the transcription factor NF-κB, which are involved in the production of proinflammatory cytokines. In different cell types, α-Klotho inhibits the translocation of NF-κB, and since this transcription factor is responsible for the expression of several proinflammatory proteins such as IL-1, IL-2, IL-6, IL-8, IL-12, and TNF-α, it inhibits their synthesis [174,175,176,177]. Moreover, it was also found to inhibit TLR4, and it is known that the activation of this receptor signals through NF-κB, and in this way, it could also inhibit inflammation [178]. α-Klotho inhibits NF-κB inflammasome activation and thus prevents the conversion of pro-IL-1β, pro-IL-18, and gasdermin D to their active forms, thereby inhibiting the proinflammatory effects of interleukins and pyroptosis, a type of gasdermin D-mediated cell death [108,179]. On the other hand, the synthesis and release of α-Klotho declines in aging, and because the aforementioned diseases occur mainly during aging, it is to be expected that α-Klotho levels in blood serum and CSF are decreased, as has been shown in several studies involving animal models and in observational studies with humans, where there is a negative association between plasma and CSF concentrations and cognitive impairment. Moreover, there is evidence that α-Klotho does not cross the BBB [24], so it is possible that this correlation reflects a decline in both peripheral and central α-Klotho responsiveness. However, experimental evidence suggests that there are several viable ways to increase α-Klotho synthesis and secretion, such as physical exercise, exogenous administration of α-Klotho, and some flavonoids and various approved drugs for different human diseases. We will discuss these options below, except for overexpression using molecular biological methods (e.g., gene editing) because it is not yet feasible to apply them to humans.

5.1. Age, Gender, and Serum α-Klotho

It is widely recognized that serum α-Klotho concentrations decrease with age in both sexes, and some studies show that adult women of all ages have higher concentrations than men [180,181]. However, there is a large variability in serum α-Klotho that is related to multiple factors that influence this variability such as physiological changes (e.g., hormonal fluctuations, Klotho metabolism), diseases, and lifestyle [181,182].

5.2. Physical Exercise and Exogenous Administration of α-Klotho

It is widely recognized that physical exercise, both aerobic and anaerobic, produces beneficial effects in animal models of neurodegenerative and traumatic diseases as well as in patients with these diseases. However, these benefits vary in relation to the many factors involved in each situation (age, gender, duration, frequency, type and intensity of training, health condition, exercise modality, etc.). Studies evaluating plasma α-Klotho show varied results, as do the protocols used, but for few exceptions [183], the results show a significant increase [184,185]. Aerobic exercise training (12 weeks) increases α-Klotho levels in sedentary middle-aged adults [185]. A single high-intensity physical exercise event produces a transient increase in α-Klotho levels in healthy volunteers [186]. A single maximal 20-min aerobic exercise session induces an increase in α-Klotho protein levels [187]. Acute physical exercise (45 min of treadmill running) significantly increases circulating α-Klotho in both young and old mice, although the response in aged animals is lower, and similar results have been observed in humans [188]. Twelve weeks of moderate aerobic exercise training increase plasma α-Klotho levels in postmenopausal women [189]. A meta-analysis concluded that, in general, physical training increases blood α-Klotho levels, but there is no consensus about which protocol produces the highest blood α-Klotho concentration [184]. In addition, within this type of study, some have found that there is a positive correlation between increased blood α-Klotho and cognitive improvement. Twelve weeks of training produced a significant increase in α-Klotho, brain blood flow, and hippocampus volume, and correlated with an improvement in episodic memory in older adults [190], but another study with a longer duration (26 weeks) in middle-aged adults at risk of AD found no significant increase in α-Klotho and no correlation with cognitive changes [183]. A simple explanation for this discrepancy could be due to the difference in age and duration of training between the studies, but both studies have a very small sample size, so more studies are needed to determine if physical exercise modifies cognition and elucidate the mechanisms involved. However, physical exercise increases plasma α-Klotho, and exogenous application of this protein has shown cognitive enhancement in preclinical models, suggesting that physical exercise may have modulatory effects on cognition. Administration of α-Klotho to young and old healthy rodents in a PD model significantly improved cognition [21,25]. In addition, subcutaneous injection of a single low dose of α-Klotho in old non-human primates significantly improved memory and this effect persisted for at least two weeks [114].

5.3. Repurposing Approved Drugs

Repurposing drugs refers to using drugs approved for one disease or target for a different therapeutic target. This strategy offers many advantages over the development of a new drug; for example, these drugs have already passed the development and clinical testing phases, so their approval for a new use would be faster and cheaper. Preclinical studies have shown that different statin (Atorvastatin, Pitavastatin and Simvastatin), drugs that regulate cholesterol (HDL and LDL) and triglycerides increase the expression of klotho in kidneys [191,192,193]. Combining the antileukemic drug Dasatinib and the flavonoid quercetin increases Klotho in the brain, kidneys, and in mice and humans urine [115]. It has been shown in clinical studies that some drugs used to treat diseases such as chronic renal failure and hypertension and whose mechanism of action is to block the renin angiotensin and aldosterone system (RAAS), such as losartan and valsartan, increase α-Klotho in blood serum [194,195,196]. Also, there are reports that PPARϒ (peroxisome proliferator-activated receptor gamma) antagonists, pioglitazone and rosiglitazone, which are used to treat type 2 diabetes and metabolic syndrome, equally show an increase in circulating α-Klotho in animal models [197,198].

5.4. Other Compounds

Vitamin D supplementation and systemic use of the neurotransmitter GABA in a model of diabetes also produced an increase in α-Klotho [199,200]. In the kidney, vitamin D induces the synthesis of α-Klotho and fibroblast growth factor 23 (FGF23), and both proteins are involved in phosphate homeostasis [201]. Systemic treatment with GABA produces beneficial effects such as pancreatic β-cell regeneration in animal models of type 1 diabetes. One of the mechanisms found is through increased synthesis and secretion of α-Klotho, which decreases apoptosis, inflammation, and oxidative stress [202].

6. Research Limitations

The evidence on the neuroprotective and neuroplastic role of α-Klotho in neurodegenerative diseases and stroke is consistent; however, most of the evidence comes from preclinical studies and, to a lesser extent, from observational studies in humans. The anti-inflammatory effect of α-Klotho is one of the main beneficial effects but many basic aspects of α-Klotho modulation and immune response in the brain are unknown, and it is also unknown how increased peripheral α-Klotho is involved in the reduction in neuroinflammation. More observational studies in healthy humans and in humans with neurodegenerative diseases are needed to obtain a more solid picture of the relationship between peripheral α-Klotho, its genetic variants, and different neurodegenerative diseases and stroke.

7. Conclusions

Low levels of α-Klotho are associated with aging, cognitive impairment, and other disorders common to neurodegenerative diseases. α-Klotho augmentation has significant potential to improve brain function in older adults and in AD, PD, and FIS patients, probably through the activation of anti-inflammatory and antioxidant mechanisms, neurogenesis, synaptic plasticity, and cognitive functions related to the hippocampus (Figure 2). However, there are some basic unanswered questions, such as what signaling pathways modulate its synthesis, secretion, or inhibition in the brain, and how exogenous α-Klotho could participate in influencing brain physiopathology, among many other basic questions that need to be answered to allow us to better understand the physiology of this protein, which would help us to develop a therapeutic option.

8. Future Directions

Several biopharmaceutical companies are developing strategies to increase α-Klotho through gene therapy, such as the application of CRISP/Casp, but it is uncertain when this technology could be applied in humans. On the other hand, recombinant α-Klotho therapy is also very promising, but it is not yet available for use in humans and variants that can cross the BBB and exert their effects in the brain are yet to be found. Other strategies to increase endogenous α-Klotho production that have shown positive results should be further investigated, such as physical exercise, caloric restriction under medical supervision, the use of flavonoids, proopiomelanocortin, and the repurposing of some drugs. Finally, we can conclude that there are many possible therapeutic alternatives to treat neurodegenerative diseases and strokes, but much work remains to be done.

Author Contributions

All authors contributed data collection, data interpretation, and revising the paper. H.E.L.-V. and H.M.-C. designed, wrote, and edited the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

ADAlzheimer’s disease
ADAMa disintegrin and metalloproteinase
Apoapolipoprotein
APPamyloid precursor protein
BACE1β-site amyloid precursor protein cleaving enzyme 1
BBBblood–brain barrier
BCSFBBlood–cerebrospinal fluid barrier
CNScentral nervous system
CPchoroid plexus
CSFcerebrospinal fluid
DAMPsdamage-associated molecular patterns
Erkextracellular signal-regulated kinase
FADfamilial Alzheimer’s disease
FGFfibroblast growth factor
FGFRFGF receptor
FoxO1forkhead box protein O1
HEKhuman embryonic kidney
IGFinsulin-like growth factor
ILinterleukin
KLKlotho
KL-VSKlotho with valine and serine substitutions
MHC IImajor histocompatibility complex class II
MAPKmitogen-activated protein kinase
MMSEMini-Mental Status Examination
NFκBnuclear factor κ-light-chain-enhancer of activated B cells
NLRP3NACHT-, LRR- and pyrin domain-containing protein 3
NMDAN-methyl-d-aspartate
PERKprotein kinase R (PKR)-like endoplasmic reticulum kinase
PI3Kphosphoinositide 3-kinase
rDLPFCright dorsolateral prefrontal cortex
PRRspattern recognition receptors
SAMPsenescence-accelerated prone mouse
SNPsingle-nucleotide polymorphism

References

  1. Martínez-Coria, H.; Arrieta-Cruz, I.; Gutiérrez-Juárez, R.; López-Valdés, H.E. Anti-Inflammatory Effects of Flavonoids in Common Neurological Disorders Associated with Aging. Int. J. Mol. Sci. 2023, 24, 4297. [Google Scholar] [CrossRef]
  2. Teissier, T.; Boulanger, E.; Cox, L.S. Interconnections between Inflammageing and Immunosenescence during Ageing. Cells 2022, 11, 359. [Google Scholar] [CrossRef]
  3. Martínez-Coria, H.; Arrieta-Cruz, I.; Cruz, M.-E.; López-Valdés, H.E. Physiopathology of Ischemic Stroke and Its Modulation Using Memantine: Evidence from Preclinical Stroke. Neural Regen. Res. 2021, 16, 433–439. [Google Scholar] [CrossRef]
  4. Liu, F.; Wu, S.; Ren, H.; Gu, J. Klotho Suppresses RIG-I-Mediated Senescence-Associated Inflammation. Nat. Cell Biol. 2011, 13, 254–262. [Google Scholar] [CrossRef]
  5. Zeldich, E.; Chen, C.-D.; Colvin, T.A.; Bove-Fenderson, E.A.; Liang, J.; Zhou, T.B.T.; Harris, D.A.; Abraham, C.R. The Neuroprotective Effect of Klotho Is Mediated via Regulation of Members of the Redox System*. J. Biol. Chem. 2014, 289, 24700–24715. [Google Scholar] [CrossRef]
  6. Kuro-o, M.; Matsumura, Y.; Aizawa, H.; Kawaguchi, H.; Suga, T.; Utsugi, T.; Ohyama, Y.; Kurabayashi, M.; Kaname, T.; Kume, E.; et al. Mutation of the Mouse Klotho Gene Leads to a Syndrome Resembling Ageing. Nature 1997, 390, 45–51. [Google Scholar] [CrossRef] [PubMed]
  7. Kurosu, H.; Yamamoto, M.; Clark, J.D.; Pastor, J.V.; Nandi, A.; Gurnani, P.; McGuinness, O.P.; Chikuda, H.; Yamaguchi, M.; Kawaguchi, H.; et al. Suppression of Aging in Mice by the Hormone Klotho. Science 2005, 309, 1829–1833. [Google Scholar] [CrossRef]
  8. Ito, S.; Kinoshita, S.; Shiraishi, N.; Nakagawa, S.; Sekine, S.; Fujimori, T.; Nabeshima, Y. Molecular Cloning and Expression Analyses of Mouse Βklotho, Which Encodes a Novel Klotho Family Protein. Mech. Dev. 2000, 98, 115–119. [Google Scholar] [CrossRef] [PubMed]
  9. Ito, S.; Fujimori, T.; Hayashizaki, Y.; Nabeshima, Y. Identification of a Novel Mouse Membrane-Bound Family 1 Glycosidase-like Protein, Which Carries an Atypical Active Site Structure. Biochim. Biophys. Acta (BBA)—Gene Struct. Expr. 2002, 1576, 341–345. [Google Scholar] [CrossRef]
  10. Kuro-O, M. The Klotho Proteins in Health and Disease. Nat. Rev. Nephrol. 2019, 15, 27–44. [Google Scholar] [CrossRef]
  11. Fon Tacer, K.; Bookout, A.L.; Ding, X.; Kurosu, H.; John, G.B.; Wang, L.; Goetz, R.; Mohammadi, M.; Kuro-o, M.; Mangelsdorf, D.J.; et al. Research Resource: Comprehensive Expression Atlas of the Fibroblast Growth Factor System in Adult Mouse. Mol. Endocrinol. 2010, 24, 2050–2064. [Google Scholar] [CrossRef] [PubMed]
  12. Erben, R.G. α-Klotho’s Effects on Mineral Homeostasis Are Fibroblast Growth Factor-23 Dependent. Curr. Opin. Nephrol. Hypertens. 2018, 27, 229–235. [Google Scholar] [CrossRef] [PubMed]
  13. Prié, D.; Friedlander, G. Reciprocal Control of 1,25-Dihydroxyvitamin D and FGF23 Formation Involving the FGF23/Klotho System. Clin. J. Am. Soc. Nephrol. 2010, 5, 1717–1722. [Google Scholar] [CrossRef]
  14. Chen, C.-D.; Tung, T.Y.; Liang, J.; Zeldich, E.; Tucker Zhou, T.B.; Turk, B.E.; Abraham, C.R. Identification of Cleavage Sites Leading to the Shed Form of the Anti-Aging Protein Klotho. Biochemistry 2014, 53, 5579–5587. [Google Scholar] [CrossRef] [PubMed]
  15. Wu, S.-E.; Chen, W.-L. Soluble Klotho as an Effective Biomarker to Characterize Inflammatory States. Ann. Med. 2022, 54, 1520–1529. [Google Scholar] [CrossRef]
  16. Chen, C.-D.; Podvin, S.; Gillespie, E.; Leeman, S.E.; Abraham, C.R. Insulin Stimulates the Cleavage and Release of the Extracellular Domain of Klotho by ADAM10 and ADAM17. Proc. Natl. Acad. Sci. USA 2007, 104, 19796–19801. [Google Scholar] [CrossRef]
  17. Shiraki-Iida, T.; Aizawa, H.; Matsumura, Y.; Sekine, S.; Iida, A.; Anazawa, H.; Nagai, R.; Kuro-o, M.; Nabeshima, Y. Structure of the Mouse Klotho Gene and Its Two Transcripts Encoding Membrane and Secreted Protein. FEBS Lett. 1998, 424, 6–10. [Google Scholar] [CrossRef]
  18. Mencke, R.; Harms, G.; Moser, J.; van Meurs, M.; Diepstra, A.; Leuvenink, H.G.; Hillebrands, J.-L. Human Alternative Klotho mRNA Is a Nonsense-Mediated mRNA Decay Target Inefficiently Spliced in Renal Disease. JCI Insight 2017, 2, e94375. [Google Scholar] [CrossRef]
  19. Massó, A.; Sánchez, A.; Bosch, A.; Giménez-Llort, L.; Chillón, M. Secreted αKlotho Isoform Protects against Age-Dependent Memory Deficits. Mol. Psychiatry 2018, 23, 1937–1947. [Google Scholar] [CrossRef]
  20. Roig-Soriano, J.; Griñán-Ferré, C.; Espinosa-Parrilla, J.F.; Abraham, C.R.; Bosch, A.; Pallàs, M.; Chillón, M. AAV-Mediated Expression of Secreted and Transmembrane αKlotho Isoforms Rescues Relevant Aging Hallmarks in Senescent SAMP8 Mice. Aging Cell 2022, 21, e13581. [Google Scholar] [CrossRef]
  21. Gupta, S.; Moreno, A.J.; Wang, D.; Leon, J.; Chen, C.; Hahn, O.; Poon, Y.; Greenberg, K.; David, N.; Wyss-Coray, T.; et al. KL1 Domain of Longevity Factor Klotho Mimics the Metabolome of Cognitive Stimulation and Enhances Cognition in Young and Aging Mice. J. Neurosci. 2022, 42, 4016–4025. [Google Scholar] [CrossRef] [PubMed]
  22. Massó, A.; Sánchez, A.; Gimen ez-Llort, L.; Lizcano, J.M.; Cañete, M.; García, B.; Torres-Lista, V.; Puig, M.; Bosch, A.; Chillon, M. Secreted and Transmembrane αKlotho Isoforms Have Different Spatio-Temporal Profiles in the Brain during Aging and Alzheimer’s Disease Progression. PLoS ONE 2015, 10, e0143623. [Google Scholar] [CrossRef]
  23. Kunert, S.K.; Hartmann, H.; Haffner, D.; Leifheit-Nestler, M. Klotho and Fibroblast Growth Factor 23 in Cerebrospinal Fluid in Children. J. Bone Min. Metab. 2017, 35, 215–226. [Google Scholar] [CrossRef] [PubMed]
  24. Hu, M.C.; Shi, M.; Zhang, J.; Addo, T.; Cho, H.J.; Barker, S.L.; Ravikumar, P.; Gillings, N.; Bian, A.; Sidhu, S.S.; et al. Renal Production, Uptake, and Handling of Circulating αKlotho. J. Am. Soc. Nephrol. 2016, 27, 79–90. [Google Scholar] [CrossRef]
  25. Leon, J.; Moreno, A.J.; Garay, B.I.; Chalkley, R.J.; Burlingame, A.L.; Wang, D.; Dubal, D.B. Peripheral Elevation of a Klotho Fragment Enhances Brain Function and Resilience in Young, Aging, and α-Synuclein Transgenic Mice. Cell Rep. 2017, 20, 1360–1371. [Google Scholar] [CrossRef] [PubMed]
  26. Lindberg, K.; Amin, R.; Moe, O.W.; Hu, M.-C.; Erben, R.G.; Wernerson, A.Ö.; Lanske, B.; Olauson, H.; Larsson, T.E. The Kidney Is the Principal Organ Mediating Klotho Effects. J. Am. Soc. Nephrol. 2014, 25, 2169–2175. [Google Scholar] [CrossRef]
  27. Li, S.-A.; Watanabe, M.; Yamada, H.; Nagai, A.; Kinuta, M.; Takei, K. Immunohistochemical Localization of Klotho Protein in Brain, Kidney, and Reproductive Organs of Mice. Cell Struct. Funct. 2004, 29, 91–99. [Google Scholar] [CrossRef]
  28. Dalton, G.; An, S.-W.; Al-Juboori, S.I.; Nischan, N.; Yoon, J.; Dobrinskikh, E.; Hilgemann, D.W.; Xie, J.; Luby-Phelps, K.; Kohler, J.J.; et al. Soluble Klotho Binds Monosialoganglioside to Regulate Membrane Microdomains and Growth Factor Signaling. Proc. Natl. Acad. Sci. USA 2017, 114, 752–757. [Google Scholar] [CrossRef]
  29. Cha, S.-K.; Ortega, B.; Kurosu, H.; Rosenblatt, K.P.; Kuro-O, M.; Huang, C.-L. Removal of Sialic Acid Involving Klotho Causes Cell-Surface Retention of TRPV5 Channel via Binding to Galectin-1. Proc. Natl. Acad. Sci. USA 2008, 105, 9805–9810. [Google Scholar] [CrossRef]
  30. Wright, J.D.; An, S.-W.; Xie, J.; Lim, C.; Huang, C.-L. Soluble Klotho Regulates TRPC6 Calcium Signaling via Lipid Rafts, Independent of the FGFR-FGF23 Pathway. FASEB J. 2019, 33, 9182–9193. [Google Scholar] [CrossRef]
  31. Chang, Q.; Hoefs, S.; van der Kemp, A.W.; Topala, C.N.; Bindels, R.J.; Hoenderop, J.G. The Beta-Glucuronidase Klotho Hydrolyzes and Activates the TRPV5 Channel. Science 2005, 310, 490–493. [Google Scholar] [CrossRef]
  32. Cha, S.-K.; Hu, M.-C.; Kurosu, H.; Kuro-O, M.; Moe, O.; Huang, C.-L. Regulation of Renal Outer Medullary Potassium Channel and Renal K(+) Excretion by Klotho. Mol. Pharmacol. 2009, 76, 38–46. [Google Scholar] [CrossRef]
  33. Huang, C.-L. Regulation of Ion Channels by Secreted Klotho: Mechanisms and Implications. Kidney Int. 2010, 77, 855–860. [Google Scholar] [CrossRef]
  34. Landry, T.; Li, P.; Shookster, D.; Jiang, Z.; Li, H.; Laing, B.T.; Bunner, W.; Langton, T.; Tong, Q.; Huang, H. Centrally Circulating α-Klotho Inversely Correlates with Human Obesity and Modulates Arcuate Cell Populations in Mice. Mol. Metab. 2021, 44, 101136. [Google Scholar] [CrossRef]
  35. Semba, R.D.; Moghekar, A.R.; Hu, J.; Sun, K.; Turner, R.; Ferrucci, L.; O’Brien, R. Klotho in the Cerebrospinal Fluid of Adults with and without Alzheimer’s Disease. Neurosci. Lett. 2014, 558, 37–40. [Google Scholar] [CrossRef]
  36. Shardell, M.; Semba, R.D.; Rosano, C.; Kalyani, R.R.; Bandinelli, S.; Chia, C.W.; Ferrucci, L. Plasma Klotho and Cognitive Decline in Older Adults: Findings from the InCHIANTI Study. J. Gerontol. A Biol. Sci. Med. Sci. 2016, 71, 677–682. [Google Scholar] [CrossRef] [PubMed]
  37. Sanz, B.; Arrieta, H.; Rezola-Pardo, C.; Fernández-Atutxa, A.; Garin-Balerdi, J.; Arizaga, N.; Rodriguez-Larrad, A.; Irazusta, J. Low Serum Klotho Concentration Is Associated with Worse Cognition, Psychological Components of Frailty, Dependence, and Falls in Nursing Home Residents. Sci. Rep. 2021, 11, 9098. [Google Scholar] [CrossRef] [PubMed]
  38. Ge, S.; Dong, F.; Tian, C.; Yang, C.-H.; Liu, M.; Wei, J. Serum Soluble Alpha-Klotho Klotho and Cognitive Functioning in Older Adults Aged 60 and 79: An Analysis of Cross-Sectional Data of the National Health and Nutrition Examination Survey 2011 to 2014. BMC Geriatr. 2024, 24, 245. [Google Scholar] [CrossRef]
  39. Sun, X.; Chen, L.; He, Y.; Zheng, L. Circulating α-Klotho Levels in Relation to Cardiovascular Diseases: A Mendelian Randomization Study. Front. Endocrinol. 2022, 13, 842846. [Google Scholar] [CrossRef]
  40. Cheng, Y.-W.; Hung, C.-C.; Fang, W.-H.; Chen, W.-L. Association between Soluble α-Klotho Protein and Metabolic Syndrome in the Adult Population. Biomolecules 2022, 12, 70. [Google Scholar] [CrossRef] [PubMed]
  41. Wang, Q.; Su, W.; Shen, Z.; Wang, R. Correlation between Soluble α-Klotho and Renal Function in Patients with Chronic Kidney Disease: A Review and Meta-Analysis. Biomed. Res. Int. 2018, 2018, 9481475. [Google Scholar] [CrossRef]
  42. Cardoso, A.L.; Fernandes, A.; Aguilar-Pimentel, J.A.; de Angelis, M.H.; Guedes, J.R.; Brito, M.A.; Ortolano, S.; Pani, G.; Athanasopoulou, S.; Gonos, E.S.; et al. Towards Frailty Biomarkers: Candidates from Genes and Pathways Regulated in Aging and Age-Related Diseases. Ageing Res. Rev. 2018, 47, 214–277. [Google Scholar] [CrossRef] [PubMed]
  43. Wen, Z.; Liu, X.; Zhang, T. L-Shaped Association of Systemic Immune-Inflammation Index (SII) with Serum Soluble α-Klotho in the Prospective Cohort Study from the NHANES Database. Sci. Rep. 2024, 14, 13189. [Google Scholar] [CrossRef]
  44. Ye, C.; Yuan, L.; Wu, K.; Shen, B.; Zhu, C. Association between Systemic Immune-Inflammation Index and Chronic Obstructive Pulmonary Disease: A Population-Based Study. BMC Pulm. Med. 2023, 23, 295. [Google Scholar] [CrossRef]
  45. Revelas, M.; Thalamuthu, A.; Oldmeadow, C.; Evans, T.-J.; Armstrong, N.J.; Kwok, J.B.; Brodaty, H.; Schofield, P.R.; Scott, R.J.; Sachdev, P.S.; et al. Review and Meta-Analysis of Genetic Polymorphisms Associated with Exceptional Human Longevity. Mech. Ageing Dev. 2018, 175, 24–34. [Google Scholar] [CrossRef]
  46. Zhu, Z.; Xia, W.; Cui, Y.; Zeng, F.; Li, Y.; Yang, Z.; Hequn, C. Klotho Gene Polymorphisms Are Associated with Healthy Aging and Longevity: Evidence from a Meta-Analysis. Mech. Ageing Dev. 2019, 178, 33–40. [Google Scholar] [CrossRef]
  47. Almeida, O.P.; Morar, B.; Hankey, G.J.; Yeap, B.B.; Golledge, J.; Jablensky, A.; Flicker, L. Longevity Klotho Gene Polymorphism and the Risk of Dementia in Older Men. Maturitas 2017, 101, 1–5. [Google Scholar] [CrossRef]
  48. Arking, D.E.; Atzmon, G.; Arking, A.; Barzilai, N.; Dietz, H.C. Association between a Functional Variant of the KLOTHO Gene and High-Density Lipoprotein Cholesterol, Blood Pressure, Stroke, and Longevity. Circ. Res. 2005, 96, 412–418. [Google Scholar] [CrossRef] [PubMed]
  49. Clinton, S.M.; Glover, M.E.; Maltare, A.; Laszczyk, A.M.; Mehi, S.J.; Simmons, R.K.; King, G.D. Expression of Klotho mRNA and Protein in Rat Brain Parenchyma from Early Postnatal Development into Adulthood. Brain Res. 2013, 1527, 1–14. [Google Scholar] [CrossRef]
  50. German, D.C.; Khobahy, I.; Pastor, J.; Kuro-O, M.; Liu, X. Nuclear Localization of Klotho in Brain: An Anti-Aging Protein. Neurobiol. Aging 2012, 33, 1483.e25–1483.e30. [Google Scholar] [CrossRef] [PubMed]
  51. Li, D.; Jing, D.; Liu, Z.; Chen, Y.; Huang, F.; Behnisch, T. Enhanced Expression of Secreted α-Klotho in the Hippocampus Alters Nesting Behavior and Memory Formation in Mice. Front. Cell Neurosci. 2019, 13, 133. [Google Scholar] [CrossRef] [PubMed]
  52. Shiozaki, M.; Yoshimura, K.; Shibata, M.; Koike, M.; Matsuura, N.; Uchiyama, Y.; Gotow, T. Morphological and Biochemical Signs of Age-Related Neurodegenerative Changes in Klotho Mutant Mice. Neuroscience 2008, 152, 924–941. [Google Scholar] [CrossRef] [PubMed]
  53. Dubal, D.B.; Yokoyama, J.S.; Zhu, L.; Broestl, L.; Worden, K.; Wang, D.; Sturm, V.E.; Kim, D.; Klein, E.; Yu, G.-Q.; et al. Life Extension Factor Klotho Enhances Cognition. Cell Rep. 2014, 7, 1065–1076. [Google Scholar] [CrossRef]
  54. Kuriyama, N.; Ozaki, E.; Mizuno, T.; Ihara, M.; Mizuno, S.; Koyama, T.; Matsui, D.; Watanabe, I.; Akazawa, K.; Takeda, K.; et al. Association between α-Klotho and Deep White Matter Lesions in the Brain: A Pilot Case Control Study Using Brain MRI. J. Alzheimers Dis. 2018, 61, 145–155. [Google Scholar] [CrossRef]
  55. Chen, C.-D.; Li, H.; Liang, J.; Hixson, K.; Zeldich, E.; Abraham, C.R. The Anti-Aging and Tumor Suppressor Protein Klotho Enhances Differentiation of a Human Oligodendrocytic Hybrid Cell Line. J. Mol. Neurosci. 2015, 55, 76–90. [Google Scholar] [CrossRef]
  56. Zeldich, E.; Chen, C.-D.; Avila, R.; Medicetty, S.; Abraham, C.R. The Anti-Aging Protein Klotho Enhances Remyelination Following Cuprizone-Induced Demyelination. J. Mol. Neurosci. 2015, 57, 185–196. [Google Scholar] [CrossRef]
  57. MacNee, W.; Rabinovich, R.A.; Choudhury, G. Ageing and the Border between Health and Disease. Eur. Respir. J. 2014, 44, 1332–1352. [Google Scholar] [CrossRef]
  58. López-Otín, C.; Blasco, M.A.; Partridge, L.; Serrano, M.; Kroemer, G. The Hallmarks of Aging. Cell 2013, 153, 1194–1217. [Google Scholar] [CrossRef]
  59. López-Otín, C.; Blasco, M.A.; Partridge, L.; Serrano, M.; Kroemer, G. Hallmarks of Aging: An Expanding Universe. Cell 2023, 186, 243–278. [Google Scholar] [CrossRef]
  60. Lee, J.; Kim, H.-J. Normal Aging Induces Changes in the Brain and Neurodegeneration Progress: Review of the Structural, Biochemical, Metabolic, Cellular, and Molecular Changes. Front. Aging Neurosci. 2022, 14, 931536. [Google Scholar] [CrossRef] [PubMed]
  61. Blinkouskaya, Y.; Caçoilo, A.; Gollamudi, T.; Jalalian, S.; Weickenmeier, J. Brain Aging Mechanisms with Mechanical Manifestations. Mech. Ageing Dev. 2021, 200, 111575. [Google Scholar] [CrossRef]
  62. Mishra, A.; Bandopadhyay, R.; Singh, P.K.; Mishra, P.S.; Sharma, N.; Khurana, N. Neuroinflammation in Neurological Disorders: Pharmacotherapeutic Targets from Bench to Bedside. Metab. Brain Dis. 2021, 36, 1591–1626. [Google Scholar] [CrossRef] [PubMed]
  63. Ownby, R.L. Neuroinflammation and Cognitive Aging. Curr. Psychiatry Rep. 2010, 12, 39–45. [Google Scholar] [CrossRef] [PubMed]
  64. National Institute on Aging What Are the Signs of Alzheimer’s Disease? Available online: https://www.nia.nih.gov/health/alzheimers-symptoms-and-diagnosis/what-are-signs-alzheimers-disease (accessed on 30 July 2025).
  65. Baghel, M.S.; Singh, P.; Srivas, S.; Thakur, M.K. Cognitive Changes with Aging. Proc. Natl. Acad. Sci. India Sect. B Biol. Sci. 2019, 89, 765–773. [Google Scholar] [CrossRef]
  66. Mekhora, C.; Lamport, D.J.; Spencer, J.P.E. An Overview of the Relationship between Inflammation and Cognitive Function in Humans, Molecular Pathways and the Impact of Nutraceuticals. Neurochem. Int. 2024, 181, 105900. [Google Scholar] [CrossRef]
  67. Janelidze, S.; Mattsson, N.; Stomrud, E.; Lindberg, O.; Palmqvist, S.; Zetterberg, H.; Blennow, K.; Hansson, O. CSF Biomarkers of Neuroinflammation and Cerebrovascular Dysfunction in Early Alzheimer Disease. Neurology 2018, 91, e867–e877. [Google Scholar] [CrossRef]
  68. Alley, D.E.; Crimmins, E.M.; Karlamangla, A.; Hu, P.; Seeman, T.E. Inflammation and Rate of Cognitive Change in High-Functioning Older Adults. J. Gerontol. A Biol. Sci. Med. Sci. 2008, 63, 50–55. [Google Scholar] [CrossRef]
  69. Yaffe, K.; Lindquist, K.; Penninx, B.W.; Simonsick, E.M.; Pahor, M.; Kritchevsky, S.; Launer, L.; Kuller, L.; Rubin, S.; Harris, T. Inflammatory Markers and Cognition in Well-Functioning African-American and White Elders. Neurology 2003, 61, 76–80. [Google Scholar] [CrossRef] [PubMed]
  70. d’Avila, J.C.; Siqueira, L.D.; Mazeraud, A.; Azevedo, E.P.; Foguel, D.; Castro-Faria-Neto, H.C.; Sharshar, T.; Chrétien, F.; Bozza, F.A. Age-Related Cognitive Impairment Is Associated with Long-Term Neuroinflammation and Oxidative Stress in a Mouse Model of Episodic Systemic Inflammation. J. Neuroinflamm. 2018, 15, 28. [Google Scholar] [CrossRef]
  71. IL-17A Is Implicated in Lipopolysaccharide-Induced Neuroinflammation and Cognitive Impairment in Aged Rats Via Microglial Activation-PubMed. Available online: https://pubmed.ncbi.nlm.nih.gov/26373740/ (accessed on 30 July 2025).
  72. Zhang, D.; Li, S.; Hou, L.; Jing, L.; Ruan, Z.; Peng, B.; Zhang, X.; Hong, J.-S.; Zhao, J.; Wang, Q. Microglial Activation Contributes to Cognitive Impairments in Rotenone-Induced Mouse Parkinson’s Disease Model. J. Neuroinflamm. 2021, 18, 4. [Google Scholar] [CrossRef]
  73. Gelb, S.; Lehtinen, M.K. Snapshot: Choroid Plexus Brain Barrier. Cell 2023, 186, 3522–3522.e1. [Google Scholar] [CrossRef] [PubMed]
  74. Wu, D.; Chen, Q.; Chen, X.; Han, F.; Chen, Z.; Wang, Y. The Blood–Brain Barrier: Structure, Regulation and Drug Delivery. Sig. Transduct. Target. Ther. 2023, 8, 1–27. [Google Scholar] [CrossRef]
  75. Knox, E.G.; Aburto, M.R.; Clarke, G.; Cryan, J.F.; O’Driscoll, C.M. The Blood-Brain Barrier in Aging and Neurodegeneration. Mol. Psychiatry 2022, 27, 2659–2673. [Google Scholar] [CrossRef]
  76. Vandenbroucke, R.E. A Hidden Epithelial Barrier in the Brain with a Central Role in Regulating Brain Homeostasis. Implications for Aging. Ann. ATS 2016, 13, S407–S410. [Google Scholar] [CrossRef]
  77. Xu, H.; Hehnly, C.; Lehtinen, M.K. The Choroid Plexus: A Command Center for Brain–Body Communication during Inflammation. Curr. Opin. Immunol. 2025, 93, 102540. [Google Scholar] [CrossRef]
  78. Chung, C.-P.; Chang, Y.-C.; Ding, Y.; Lim, K.; Liu, Q.; Zhu, L.; Zhang, W.; Lu, T.-S.; Molostvov, G.; Zehnder, D.; et al. α-Klotho Expression Determines Nitric Oxide Synthesis in Response to FGF-23 in Human Aortic Endothelial Cells. PLoS ONE 2017, 12, e0176817. [Google Scholar] [CrossRef]
  79. Zhu, L.; Stein, L.R.; Kim, D.; Ho, K.; Yu, G.-Q.; Zhan, L.; Larsson, T.E.; Mucke, L. Klotho Controls the Brain–Immune System Interface in the Choroid Plexus. Proc. Natl. Acad. Sci. USA 2018, 115, E11388–E11396. [Google Scholar] [CrossRef]
  80. López-Valdés, H.E.; Martínez-Coria, H. The Role of Neuroinflammation in Age-Related Dementias. Rev. Investig. Clin. 2016, 68, 40–48. [Google Scholar] [CrossRef]
  81. Fulop, T.; Larbi, A.; Pawelec, G.; Khalil, A.; Cohen, A.A.; Hirokawa, K.; Witkowski, J.M.; Franceschi, C. Immunology of Aging: The Birth of Inflammaging. Clin. Rev. Allergy Immunol. 2023, 64, 109–122. [Google Scholar] [CrossRef] [PubMed]
  82. Li, D.; Wu, M. Pattern Recognition Receptors in Health and Diseases. Signal Transduct. Target. Ther. 2021, 6, 291. [Google Scholar] [CrossRef] [PubMed]
  83. Rea, I.M.; Gibson, D.S.; McGilligan, V.; McNerlan, S.E.; Alexander, H.D.; Ross, O.A. Age and Age-Related Diseases: Role of Inflammation Triggers and Cytokines. Front. Immunol. 2018, 9, 586. [Google Scholar] [CrossRef]
  84. Kumar, V. Toll-like Receptors in the Pathogenesis of Neuroinflammation. J. Neuroimmunol. 2019, 332, 16–30. [Google Scholar] [CrossRef]
  85. Pascual, M.; Calvo-Rodriguez, M.; Núñez, L.; Villalobos, C.; Ureña, J.; Guerri, C. Toll-like Receptors in Neuroinflammation, Neurodegeneration, and Alcohol-Induced Brain Damage. IUBMB Life 2021, 73, 900–915. [Google Scholar] [CrossRef]
  86. Vidal-Itriago, A.; Radford, R.A.W.; Aramideh, J.A.; Maurel, C.; Scherer, N.M.; Don, E.K.; Lee, A.; Chung, R.S.; Graeber, M.B.; Morsch, M. Microglia Morphophysiological Diversity and Its Implications for the CNS. Front. Immunol. 2022, 13, 997786. [Google Scholar] [CrossRef]
  87. Paolicelli, R.C.; Sierra, A.; Stevens, B.; Tremblay, M.-E.; Aguzzi, A.; Ajami, B.; Amit, I.; Audinat, E.; Bechmann, I.; Bennett, M.; et al. Microglia States and Nomenclature: A Field at Its Crossroads. Neuron 2022, 110, 3458–3483. [Google Scholar] [CrossRef] [PubMed]
  88. Escartin, C.; Galea, E.; Lakatos, A.; O’Callaghan, J.P.; Petzold, G.C.; Serrano-Pozo, A.; Steinhäuser, C.; Volterra, A.; Carmignoto, G.; Agarwal, A.; et al. Reactive Astrocyte Nomenclature, Definitions, and Future Directions. Nat. Neurosci. 2021, 24, 312–325. [Google Scholar] [CrossRef] [PubMed]
  89. Chou, S.-M.; Yen, Y.-H.; Yuan, F.; Zhang, S.-C.; Chong, C.-M. Neuronal Senescence in the Aged Brain. Aging Dis. 2023, 14, 1618–1632. [Google Scholar] [CrossRef]
  90. Martínez-Cué, C.; Rueda, N. Cellular Senescence in Neurodegenerative Diseases. Front. Cell Neurosci. 2020, 14, 16. [Google Scholar] [CrossRef] [PubMed]
  91. Godbout, J.P.; Chen, J.; Abraham, J.; Richwine, A.F.; Berg, B.M.; Kelley, K.W.; Johnson, R.W. Exaggerated Neuroinflammation and Sickness Behavior in Aged Mice Following Activation of the Peripheral Innate Immune System. FASEB J. 2005, 19, 1329–1331. [Google Scholar] [CrossRef]
  92. Caldeira, C.; Oliveira, A.F.; Cunha, C.; Vaz, A.R.; Falcão, A.S.; Fernandes, A.; Brites, D. Microglia Change from a Reactive to an Age-like Phenotype with the Time in Culture. Front. Cell Neurosci. 2014, 8, 152. [Google Scholar] [CrossRef]
  93. Njie, E.G.; Boelen, E.; Stassen, F.R.; Steinbusch, H.W.M.; Borchelt, D.R.; Streit, W.J. Ex Vivo Cultures of Microglia from Young and Aged Rodent Brain Reveal Age-Related Changes in Microglial Function. Neurobiol. Aging 2012, 33, 195.e1–195.e12. [Google Scholar] [CrossRef] [PubMed]
  94. Sierra, A.; Encinas, J.M.; Deudero, J.J.P.; Chancey, J.H.; Enikolopov, G.; Overstreet-Wadiche, L.S.; Tsirka, S.E.; Maletic-Savatic, M. Microglia Shape Adult Hippocampal Neurogenesis through Apoptosis-Coupled Phagocytosis. Cell Stem Cell 2010, 7, 483–495. [Google Scholar] [CrossRef]
  95. Mou, Y.; Du, Y.; Zhou, L.; Yue, J.; Hu, X.; Liu, Y.; Chen, S.; Lin, X.; Zhang, G.; Xiao, H.; et al. Gut Microbiota Interact with the Brain Through Systemic Chronic Inflammation: Implications on Neuroinflammation, Neurodegeneration, and Aging. Front. Immunol. 2022, 13, 796288. [Google Scholar] [CrossRef]
  96. Colombo, A.V.; Sadler, R.K.; Llovera, G.; Singh, V.; Roth, S.; Heindl, S.; Sebastian Monasor, L.; Verhoeven, A.; Peters, F.; Parhizkar, S.; et al. Microbiota-Derived Short Chain Fatty Acids Modulate Microglia and Promote Aβ Plaque Deposition. eLife 2021, 10, e59826. [Google Scholar] [CrossRef]
  97. Sun, M.-F.; Zhu, Y.-L.; Zhou, Z.-L.; Jia, X.-B.; Xu, Y.-D.; Yang, Q.; Cui, C.; Shen, Y.-Q. Neuroprotective Effects of Fecal Microbiota Transplantation on MPTP-Induced Parkinson’s Disease Mice: Gut Microbiota, Glial Reaction and TLR4/TNF-α Signaling Pathway. Brain Behav. Immun. 2018, 70, 48–60. [Google Scholar] [CrossRef]
  98. Sampson, T.R.; Debelius, J.W.; Thron, T.; Janssen, S.; Shastri, G.G.; Ilhan, Z.E.; Challis, C.; Schretter, C.E.; Rocha, S.; Gradinaru, V.; et al. Gut Microbiota Regulate Motor Deficits and Neuroinflammation in a Model of Parkinson’s Disease. Cell 2016, 167, 1469–1480.e12. [Google Scholar] [CrossRef]
  99. Abraham, C.R.; Li, A. Aging-Suppressor Klotho: Prospects in Diagnostics and Therapeutics. Ageing Res. Rev. 2022, 82, 101766. [Google Scholar] [CrossRef] [PubMed]
  100. Kurosu, H.; Ogawa, Y.; Miyoshi, M.; Yamamoto, M.; Nandi, A.; Rosenblatt, K.P.; Baum, M.G.; Schiavi, S.; Hu, M.-C.; Moe, O.W.; et al. Regulation of Fibroblast Growth Factor-23 Signaling by Klotho. J. Biol. Chem. 2006, 281, 6120–6123. [Google Scholar] [CrossRef]
  101. Ikushima, M.; Rakugi, H.; Ishikawa, K.; Maekawa, Y.; Yamamoto, K.; Ohta, J.; Chihara, Y.; Kida, I.; Ogihara, T. Anti-Apoptotic and Anti-Senescence Effects of Klotho on Vascular Endothelial Cells. Biochem. Biophys. Res. Commun. 2006, 339, 827–832. [Google Scholar] [CrossRef] [PubMed]
  102. Banerjee, S.; Zhao, Y.; Sarkar, P.S.; Rosenblatt, K.P.; Tilton, R.G.; Choudhary, S. Klotho Ameliorates Chemically Induced Endoplasmic Reticulum (ER) Stress Signaling. Cell. Physiol. Biochem. 2013, 31, 659–672. [Google Scholar] [CrossRef]
  103. Yamamoto, M.; Clark, J.D.; Pastor, J.V.; Gurnani, P.; Nandi, A.; Kurosu, H.; Miyoshi, M.; Ogawa, Y.; Castrillon, D.H.; Rosenblatt, K.P.; et al. Regulation of Oxidative Stress by the Anti-Aging Hormone Klotho*,♦. J. Biol. Chem. 2005, 280, 38029–38034. [Google Scholar] [CrossRef] [PubMed]
  104. Zhou, H.; Pu, S.; Zhou, H.; Guo, Y. Klotho as Potential Autophagy Regulator and Therapeutic Target. Front. Pharmacol. 2021, 12, 755366. [Google Scholar] [CrossRef]
  105. Chen, C.-D.; Sloane, J.A.; Li, H.; Aytan, N.; Giannaris, E.L.; Zeldich, E.; Hinman, J.D.; Dedeoglu, A.; Rosene, D.L.; Bansal, R.; et al. The Antiaging Protein Klotho Enhances Oligodendrocyte Maturation and Myelination of the CNS. J. Neurosci. 2013, 33, 1927–1939. [Google Scholar] [CrossRef]
  106. Baluchnejadmojarad, T.; Eftekhari, S.-M.; Jamali-Raeufy, N.; Haghani, S.; Zeinali, H.; Roghani, M. The Anti-Aging Protein Klotho Alleviates Injury of Nigrostriatal Dopaminergic Pathway in 6-Hydroxydopamine Rat Model of Parkinson’s Disease: Involvement of PKA/CaMKII/CREB Signaling. Exp. Gerontol. 2017, 100, 70–76. [Google Scholar] [CrossRef]
  107. Sedighi, M.; Baluchnejadmojarad, T.; Fallah, S.; Moradi, N.; Afshin-Majdd, S.; Roghani, M. Klotho Ameliorates Cellular Inflammation via Suppression of Cytokine Release and Upregulation of miR-29a in the PBMCs of Diagnosed Alzheimer’s Disease Patients. J. Mol. Neurosci. 2019, 69, 157–165. [Google Scholar] [CrossRef]
  108. Zhao, Y.; Zeng, C.-Y.; Li, X.-H.; Yang, T.-T.; Kuang, X.; Du, J.-R. Klotho Overexpression Improves Amyloid-β Clearance and Cognition in the APP/PS1 Mouse Model of Alzheimer’s Disease. Aging Cell 2020, 19, e13239. [Google Scholar] [CrossRef]
  109. Xiang, T.; Luo, X.; Ye, L.; Huang, H.; Wu, Y. Klotho Alleviates NLRP3 Inflammasome-Mediated Neuroinflammation in a Temporal Lobe Epilepsy Rat Model by Activating the Nrf2 Signaling Pathway. Epilepsy Behav. 2022, 128, 108509. [Google Scholar] [CrossRef]
  110. Lehmann, M.; Hu, Q.; Hu, Y.; Costa, R.; Ansari, M.; Pineda, R.; Janssen, W.; Theis, F.; Schiller, H.; Königshoff, M. Prolonged WNT/ß-Catenin Signaling Induces Cellular Senescence in Aging and Pulmonary Fibrosis. ERJ Open Res. 2020, 6, 81. [Google Scholar] [CrossRef]
  111. Zhou, L.; Li, Y.; Zhou, D.; Tan, R.J.; Liu, Y. Loss of Klotho Contributes to Kidney Injury by Derepression of Wnt/β-Catenin Signaling. J. Am. Soc. Nephrol. 2013, 24, 771–785. [Google Scholar] [CrossRef] [PubMed]
  112. Liu, H.; Fergusson, M.M.; Castilho, R.M.; Liu, J.; Cao, L.; Chen, J.; Malide, D.; Rovira, I.I.; Schimel, D.; Kuo, C.J.; et al. Augmented Wnt Signaling in a Mammalian Model of Accelerated Aging. Science 2007, 317, 803–806. [Google Scholar] [CrossRef]
  113. Zhou, H.-J.; Zeng, C.-Y.; Yang, T.-T.; Long, F.-Y.; Kuang, X.; Du, J.-R. Lentivirus-Mediated Klotho up-Regulation Improves Aging-Related Memory Deficits and Oxidative Stress in Senescence-Accelerated Mouse Prone-8 Mice. Life Sci. 2018, 200, 56–62. [Google Scholar] [CrossRef]
  114. Castner, S.A.; Gupta, S.; Wang, D.; Moreno, A.J.; Park, C.; Chen, C.; Poon, Y.; Groen, A.; Greenberg, K.; David, N.; et al. Longevity Factor Klotho Enhances Cognition in Aged Nonhuman Primates. Nat. Aging 2023, 3, 931–937. [Google Scholar] [CrossRef]
  115. Zhu, Y.; Prata, L.G.P.L.; Gerdes, E.O.W.; Netto, J.M.E.; Pirtskhalava, T.; Giorgadze, N.; Tripathi, U.; Inman, C.L.; Johnson, K.O.; Xue, A.; et al. Orally-Active, Clinically-Translatable Senolytics Restore α-Klotho in Mice and Humans. EBioMedicine 2022, 77, 103912. [Google Scholar] [CrossRef] [PubMed]
  116. Nakao, V.W.; Mazucanti, C.H.Y.; de Sá Lima, L.; de Mello, P.S.; de Souza Port’s, N.M.; Kinoshita, P.F.; Leite, J.A.; Kawamoto, E.M.; Scavone, C. Neuroprotective Action of α-Klotho against LPS-Activated Glia Conditioned Medium in Primary Neuronal Culture. Sci. Rep. 2022, 12, 18884. [Google Scholar] [CrossRef]
  117. Mayne, K.; White, J.A.; McMurran, C.E.; Rivera, F.J.; de la Fuente, A.G. Aging and Neurodegenerative Disease: Is the Adaptive Immune System a Friend or Foe? Front. Aging Neurosci. 2020, 12, 572090. [Google Scholar] [CrossRef] [PubMed]
  118. Pons, V.; Rivest, S. Targeting Systemic Innate Immune Cells as a Therapeutic Avenue for Alzheimer Disease. Pharmacol. Rev. 2022, 74, 1–17. [Google Scholar] [CrossRef]
  119. Alzheimer’s Association. 2019 Alzheimer’s Disease Facts and Figures. Alzheimer’s Dement. 2019, 15, 321–387. [Google Scholar] [CrossRef]
  120. Breijyeh, Z.; Karaman, R. Comprehensive Review on Alzheimer’s Disease: Causes and Treatment. Molecules 2020, 25, 5789. [Google Scholar] [CrossRef] [PubMed]
  121. Andrade-Guerrero, J.; Santiago-Balmaseda, A.; Jeronimo-Aguilar, P.; Vargas-Rodríguez, I.; Cadena-Suárez, A.R.; Sánchez-Garibay, C.; Pozo-Molina, G.; Méndez-Catalá, C.F.; Cardenas-Aguayo, M.-C.; Diaz-Cintra, S.; et al. Alzheimer’s Disease: An Updated Overview of Its Genetics. Int. J. Mol. Sci. 2023, 24, 3754. [Google Scholar] [CrossRef]
  122. Hampel, H.; Caraci, F.; Cuello, A.C.; Caruso, G.; Nisticò, R.; Corbo, M.; Baldacci, F.; Toschi, N.; Garaci, F.; Chiesa, P.A.; et al. A Path Toward Precision Medicine for Neuroinflammatory Mechanisms in Alzheimer’s Disease. Front. Immunol. 2020, 11, 456. [Google Scholar] [CrossRef]
  123. Sosna, J.; Philipp, S.; Albay, R.; Reyes-Ruiz, J.M.; Baglietto-Vargas, D.; LaFerla, F.M.; Glabe, C.G. Early Long-Term Administration of the CSF1R Inhibitor PLX3397 Ablates Microglia and Reduces Accumulation of Intraneuronal Amyloid, Neuritic Plaque Deposition and Pre-Fibrillar Oligomers in 5XFAD Mouse Model of Alzheimer’s Disease. Mol. Neurodegener. 2018, 13, 11. [Google Scholar] [CrossRef]
  124. McQuade, A.; Blurton-Jones, M. Microglia in Alzheimer’s Disease: Exploring How Genetics and Phenotype Influence Risk. J. Mol. Biol. 2019, 431, 1805–1817. [Google Scholar] [CrossRef]
  125. Davies, D.S.; Ma, J.; Jegathees, T.; Goldsbury, C. Microglia Show Altered Morphology and Reduced Arborization in Human Brain during Aging and Alzheimer’s Disease. Brain Pathol. 2017, 27, 795–808. [Google Scholar] [CrossRef]
  126. Cagnin, A.; Brooks, D.J.; Kennedy, A.M.; Gunn, R.N.; Myers, R.; Turkheimer, F.E.; Jones, T.; Banati, R.B. In-Vivo Measurement of Activated Microglia in Dementia. Lancet 2001, 358, 461–467. [Google Scholar] [CrossRef]
  127. Edison, P.; Archer, H.A.; Gerhard, A.; Hinz, R.; Pavese, N.; Turkheimer, F.E.; Hammers, A.; Tai, Y.F.; Fox, N.; Kennedy, A.; et al. Microglia, Amyloid, and Cognition in Alzheimer’s Disease: An [11C](R)PK11195-PET and [11C]PIB-PET Study. Neurobiol. Dis. 2008, 32, 412–419. [Google Scholar] [CrossRef]
  128. Schuitemaker, A.; Kropholler, M.A.; Boellaard, R.; van der Flier, W.M.; Kloet, R.W.; van der Doef, T.F.; Knol, D.L.; Windhorst, A.D.; Luurtsema, G.; Barkhof, F.; et al. Microglial Activation in Alzheimer’s Disease: An (R)-[11C]PK11195 Positron Emission Tomography Study. Neurobiol. Aging 2013, 34, 128–136. [Google Scholar] [CrossRef] [PubMed]
  129. Dani, M.; Wood, M.; Mizoguchi, R.; Fan, Z.; Walker, Z.; Morgan, R.; Hinz, R.; Biju, M.; Kuruvilla, T.; Brooks, D.J.; et al. Microglial Activation Correlates in Vivo with Both Tau and Amyloid in Alzheimer’s Disease. Brain 2018, 141, 2740–2754. [Google Scholar] [CrossRef] [PubMed]
  130. Zeng, C.-Y.; Yang, T.-T.; Zhou, H.-J.; Zhao, Y.; Kuang, X.; Duan, W.; Du, J.-R. Lentiviral Vector–Mediated Overexpression of Klotho in the Brain Improves Alzheimer’s Disease–like Pathology and Cognitive Deficits in Mice. Neurobiol. Aging 2019, 78, 18–28. [Google Scholar] [CrossRef] [PubMed]
  131. Kuang, X.; Chen, Y.-S.; Wang, L.-F.; Li, Y.-J.; Liu, K.; Zhang, M.-X.; Li, L.-J.; Chen, C.; He, Q.; Wang, Y.; et al. Klotho Upregulation Contributes to the Neuroprotection of Ligustilide in an Alzheimer’s Disease Mouse Model. Neurobiol. Aging 2014, 35, 169–178. [Google Scholar] [CrossRef]
  132. Kuang, X.; Zhou, H.-J.; Thorne, A.H.; Chen, X.-N.; Li, L.-J.; Du, J.-R. Neuroprotective Effect of Ligustilide through Induction of α-Secretase Processing of Both APP and Klotho in a Mouse Model of Alzheimer’s Disease. Front. Aging Neurosci. 2017, 9, 353. [Google Scholar] [CrossRef]
  133. Kundu, P.; Zimmerman, B.; Quinn, J.F.; Kaye, J.; Mattek, N.; Westaway, S.K.; Raber, J. Serum Levels of α-Klotho Are Correlated with Cerebrospinal Fluid Levels and Predict Measures of Cognitive Function. J. Alzheimers Dis. 2022, 86, 1471–1481. [Google Scholar] [CrossRef] [PubMed]
  134. Bologna, M.; Truong, D.; Jankovic, J. The Etiopathogenetic and Pathophysiological Spectrum of Parkinsonism. J. Neurol. Sci. 2022, 433, 120012. [Google Scholar] [CrossRef]
  135. Bloem, B.R.; Okun, M.S.; Klein, C. Parkinson’s Disease. Lancet 2021, 397, 2284–2303. [Google Scholar] [CrossRef] [PubMed]
  136. Simon, D.K.; Tanner, C.M.; Brundin, P. Parkinson Disease Epidemiology, Pathology, Genetics, and Pathophysiology. Clin. Geriatr. Med. 2020, 36, 1–12. [Google Scholar] [CrossRef] [PubMed]
  137. Dickson, D.W. Neuropathology of Parkinson Disease. Park. Relat. Disord. 2018, 46 (Suppl. S1), S30–S33. [Google Scholar] [CrossRef]
  138. Banati, R.B.; Daniel, S.E.; Blunt, S.B. Glial Pathology but Absence of Apoptotic Nigral Neurons in Long-Standing Parkinson’s Disease. Mov. Disord. 1998, 13, 221–227. [Google Scholar] [CrossRef]
  139. Vawter, M.P.; Dillon-Carter, O.; Tourtellotte, W.W.; Carvey, P.; Freed, W.J. TGFbeta1 and TGFbeta2 Concentrations Are Elevated in Parkinson’s Disease in Ventricular Cerebrospinal Fluid. Exp. Neurol. 1996, 142, 313–322. [Google Scholar] [CrossRef]
  140. Boka, G.; Anglade, P.; Wallach, D.; Javoy-Agid, F.; Agid, Y.; Hirsch, E.C. Immunocytochemical Analysis of Tumor Necrosis Factor and Its Receptors in Parkinson’s Disease. Neurosci. Lett. 1994, 172, 151–154. [Google Scholar] [CrossRef]
  141. Gerhard, A.; Watts, J.; Trender-Gerhard, I.; Turkheimer, F.; Banati, R.B.; Bhatia, K.; Brooks, D.J. In Vivo Imaging of Microglial Activation with [11C](R)-PK11195 PET in Corticobasal Degeneration. Mov. Disord. 2004, 19, 1221–1226. [Google Scholar] [CrossRef]
  142. Brochard, V.; Combadière, B.; Prigent, A.; Laouar, Y.; Perrin, A.; Beray-Berthat, V.; Bonduelle, O.; Alvarez-Fischer, D.; Callebert, J.; Launay, J.-M.; et al. Infiltration of CD4+ Lymphocytes into the Brain Contributes to Neurodegeneration in a Mouse Model of Parkinson Disease. J. Clin. Investig. 2009, 119, 182–192. [Google Scholar] [CrossRef]
  143. Klegeris, A.; Pelech, S.; Giasson, B.I.; Maguire, J.; Zhang, H.; McGeer, E.G.; McGeer, P.L. Alpha-Synuclein Activates Stress Signaling Protein Kinases in THP-1 Cells and Microglia. Neurobiol. Aging 2008, 29, 739–752. [Google Scholar] [CrossRef]
  144. Li, Y.; Xia, Y.; Yin, S.; Wan, F.; Hu, J.; Kou, L.; Sun, Y.; Wu, J.; Zhou, Q.; Huang, J.; et al. Targeting Microglial α-Synuclein/TLRs/NF-kappaB/NLRP3 Inflammasome Axis in Parkinson’s Disease. Front. Immunol. 2021, 12, 719807. [Google Scholar] [CrossRef]
  145. Kosakai, A.; Ito, D.; Nihei, Y.; Yamashita, S.; Okada, Y.; Takahashi, K.; Suzuki, N. Degeneration of Mesencephalic Dopaminergic Neurons in Klotho Mouse Related to Vitamin D Exposure. Brain Res. 2011, 1382, 109–117. [Google Scholar] [CrossRef] [PubMed]
  146. Zimmermann, M.; Fandrich, M.; Jakobi, M.; Röben, B.; Wurster, I.; Lerche, S.; Schulte, C.; Zimmermann, S.; Deuschle, C.; Schneiderhan-Marra, N.; et al. Association of Elevated Cerebrospinal Fluid Levels of the Longevity Protein α-Klotho with a Delayed Onset of Cognitive Impairment in Parkinson’s Disease Patients. Eur. J. Neurol. 2024, 31, e16388. [Google Scholar] [CrossRef] [PubMed]
  147. Sancesario, G.M.; Di Lazzaro, G.; Grillo, P.; Biticchi, B.; Giannella, E.; Alwardat, M.; Pieri, M.; Bernardini, S.; Mercuri, N.B.; Pisani, A.; et al. Biofluids Profile of α-Klotho in Patients with Parkinson’s Disease. Park. Relat. Disord. 2021, 90, 62–64. [Google Scholar] [CrossRef]
  148. World Stroke Organization Impact of Stroke. Available online: https://www.world-stroke.org/world-stroke-day-campaign/about-stroke/impact-of-stroke (accessed on 29 January 2025).
  149. Murphy, S.J.X.; Werring, D.J. Stroke: Causes and Clinical Features. Medicine 2020, 48, 561–566. [Google Scholar] [CrossRef]
  150. Möhlenbruch, M.A.; Bendszus, M. Technical standards for the interventional treatment of acute ischemic stroke. Nervenarzt 2015, 86, 1209–1216. [Google Scholar] [CrossRef]
  151. Asadi, H.; Dowling, R.; Yan, B.; Wong, S.; Mitchell, P. Advances in Endovascular Treatment of Acute Ischaemic Stroke. Intern. Med. J. 2015, 45, 798–805. [Google Scholar] [CrossRef] [PubMed]
  152. Alsbrook, D.L.; Di Napoli, M.; Bhatia, K.; Biller, J.; Andalib, S.; Hinduja, A.; Rodrigues, R.; Rodriguez, M.; Sabbagh, S.Y.; Selim, M.; et al. Neuroinflammation in Acute Ischemic and Hemorrhagic Stroke. Curr. Neurol. Neurosci. Rep. 2023, 23, 407–431. [Google Scholar] [CrossRef]
  153. Sanchez-Bezanilla, S.; Hood, R.J.; Collins-Praino, L.E.; Turner, R.J.; Walker, F.R.; Nilsson, M.; Ong, L.K. More than Motor Impairment: A Spatiotemporal Analysis of Cognitive Impairment and Associated Neuropathological Changes Following Cortical Photothrombotic Stroke. J. Cereb. Blood Flow. Metab. 2021, 41, 2439–2455. [Google Scholar] [CrossRef]
  154. Walberer, M.; Jantzen, S.U.; Backes, H.; Rueger, M.A.; Keuters, M.H.; Neumaier, B.; Hoehn, M.; Fink, G.R.; Graf, R.; Schroeter, M. In-Vivo Detection of Inflammation and Neurodegeneration in the Chronic Phase after Permanent Embolic Stroke in Rats. Brain Res. 2014, 1581, 80–88. [Google Scholar] [CrossRef] [PubMed]
  155. Rodriguez-Grande, B.; Blackabey, V.; Gittens, B.; Pinteaux, E.; Denes, A. Loss of Substance P and Inflammation Precede Delayed Neurodegeneration in the Substantia Nigra after Cerebral Ischemia. Brain Behav. Immun. 2013, 29, 51–61. [Google Scholar] [CrossRef]
  156. Pappata, S.; Levasseur, M.; Gunn, R.N.; Myers, R.; Crouzel, C.; Syrota, A.; Jones, T.; Kreutzberg, G.W.; Banati, R.B. Thalamic Microglial Activation in Ischemic Stroke Detected In Vivo by PET and [11C]PK1195. Neurology 2000, 55, 1052–1054. [Google Scholar] [CrossRef]
  157. Kulesh, A.; Drobakha, V.; Kuklina, E.; Nekrasova, I.; Shestakov, V. Cytokine Response, Tract-Specific Fractional Anisotropy, and Brain Morphometry in Post-Stroke Cognitive Impairment. J. Stroke Cerebrovasc. Dis. 2018, 27, 1752–1759. [Google Scholar] [CrossRef]
  158. Pendlebury, S.T.; Rothwell, P.M. Oxford Vascular Study Incidence and Prevalence of Dementia Associated with Transient Ischaemic Attack and Stroke: Analysis of the Population-Based Oxford Vascular Study. Lancet Neurol. 2019, 18, 248–258. [Google Scholar] [CrossRef]
  159. Sexton, E.; McLoughlin, A.; Williams, D.J.; Merriman, N.A.; Donnelly, N.; Rohde, D.; Hickey, A.; Wren, M.-A.; Bennett, K. Systematic Review and Meta-Analysis of the Prevalence of Cognitive Impairment No Dementia in the First Year Post-Stroke. Eur. Stroke J. 2019, 4, 160–171. [Google Scholar] [CrossRef]
  160. Rost, N.S.; Brodtmann, A.; Pase, M.P.; van Veluw, S.J.; Biffi, A.; Duering, M.; Hinman, J.D.; Dichgans, M. Post-Stroke Cognitive Impairment and Dementia. Circ. Res. 2022, 130, 1252–1271. [Google Scholar] [CrossRef]
  161. El Husseini, N.; Katzan, I.L.; Rost, N.S.; Blake, M.L.; Byun, E.; Pendlebury, S.T.; Aparicio, H.J.; Marquine, M.J.; Gottesman, R.F.; Smith, E.E.; et al. Cognitive Impairment After Ischemic and Hemorrhagic Stroke: A Scientific Statement from the American Heart Association/American Stroke Association. Stroke 2023, 54, e272–e291. [Google Scholar] [CrossRef] [PubMed]
  162. Gallucci, L.; Sperber, C.; Guggisberg, A.G.; Kaller, C.P.; Heldner, M.R.; Monsch, A.U.; Hakim, A.; Silimon, N.; Fischer, U.; Arnold, M.; et al. Post-Stroke Cognitive Impairment Remains Highly Prevalent and Disabling despite State-of-the-Art Stroke Treatment. Int. J. Stroke 2024, 19, 888–897. [Google Scholar] [CrossRef] [PubMed]
  163. Donate-Correa, J.; Martín-Núñez, E.; Mora-Fernández, C.; Muros-de-Fuentes, M.; Pérez-Delgado, N.; Navarro-González, J.F. Klotho in Cardiovascular Disease: Current and Future Perspectives. World J. Biol. Chem. 2015, 6, 351–357. [Google Scholar] [CrossRef]
  164. Mencke, R.; Hillebrands, J.-L. NIGRAM consortium The Role of the Anti-Ageing Protein Klotho in Vascular Physiology and Pathophysiology. Ageing Res. Rev. 2017, 35, 124–146. [Google Scholar] [CrossRef] [PubMed]
  165. Woo, H.G.; Chang, Y.; Ryu, D.-R.; Song, T.-J. Plasma Klotho Concentration Is Associated with the Presence, Burden and Progression of Cerebral Small Vessel Disease in Patients with Acute Ischaemic Stroke. PLoS ONE 2019, 14, e0220796. [Google Scholar] [CrossRef] [PubMed]
  166. Brombo, G.; Bonetti, F.; Ortolani, B.; Morieri, M.L.; Bosi, C.; Passaro, A.; Vigna, G.B.; Borgna, C.; Arcidicono, M.V.; Tisato, V.; et al. Lower Plasma Klotho Concentrations Are Associated with Vascular Dementia but Not Late-Onset Alzheimer’s Disease. Gerontology 2018, 64, 414–421. [Google Scholar] [CrossRef]
  167. Lee, J.-B.; Woo, H.G.; Chang, Y.; Jin, Y.M.; Jo, I.; Kim, J.; Song, T.-J. Plasma Klotho Concentrations Predict Functional Outcome at Three Months after Acute Ischemic Stroke Patients. Ann. Med. 2019, 51, 262–269. [Google Scholar] [CrossRef]
  168. Majumdar, V.; Nagaraja, D.; Christopher, R. Association of the Functional KL-VS Variant of Klotho Gene with Early-Onset Ischemic Stroke. Biochem. Biophys. Res. Commun. 2010, 403, 412–416. [Google Scholar] [CrossRef]
  169. Zhu, G.; Xiang, T.; Liang, S.; Liu, K.; Xiao, Z.; Ye, Q. Klotho Gene Might Antagonize Ischemic Injury in Stroke Rats by Reducing the Expression of AQP4 via P38MAPK Pathway. J. Stroke Cerebrovasc. Dis. 2023, 32, 107205. [Google Scholar] [CrossRef]
  170. Zhou, H.-J.; Li, H.; Shi, M.-Q.; Mao, X.-N.; Liu, D.-L.; Chang, Y.-R.; Gan, Y.-M.; Kuang, X.; Du, J.-R. Protective Effect of Klotho against Ischemic Brain Injury Is Associated with Inhibition of RIG-I/NF-κB Signaling. Front. Pharmacol. 2017, 8, 950. [Google Scholar] [CrossRef]
  171. Liu, X.-Y.; Zhang, L.-Y.; Wang, X.-Y.; Li, S.-C.; Hu, Y.-Y.; Zhang, J.-G.; Xian, X.-H.; Li, W.-B.; Zhang, M. STAT4-Mediated Klotho Up-Regulation Contributes to the Brain Ischemic Tolerance by Cerebral Ischemic Preconditioning via Inhibiting Neuronal Pyroptosis. Mol. Neurobiol. 2024, 61, 2336–2356. [Google Scholar] [CrossRef]
  172. Saito, Y.; Yamagishi, T.; Nakamura, T.; Ohyama, Y.; Aizawa, H.; Suga, T.; Matsumura, Y.; Masuda, H.; Kurabayashi, M.; Kuro-o, M.; et al. Klotho Protein Protects against Endothelial Dysfunction. Biochem. Biophys. Res. Commun. 1998, 248, 324–329. [Google Scholar] [CrossRef]
  173. Rakugi, H.; Matsukawa, N.; Ishikawa, K.; Yang, J.; Imai, M.; Ikushima, M.; Maekawa, Y.; Kida, I.; Miyazaki, J.; Ogihara, T. Anti-Oxidative Effect of Klotho on Endothelial Cells through cAMP Activation. Endocrine 2007, 31, 82–87. [Google Scholar] [CrossRef] [PubMed]
  174. Maekawa, Y.; Ishikawa, K.; Yasuda, O.; Oguro, R.; Hanasaki, H.; Kida, I.; Takemura, Y.; Ohishi, M.; Katsuya, T.; Rakugi, H. Klotho Suppresses TNF-Alpha-Induced Expression of Adhesion Molecules in the Endothelium and Attenuates NF-kappaB Activation. Endocrine 2009, 35, 341–346. [Google Scholar] [CrossRef]
  175. Wang, Y.; Wang, K.; Bao, Y.; Zhang, T.; Ainiwaer, D.; Xiong, X.; Wang, G.; Sun, Z. The Serum Soluble Klotho Alleviates Cardiac Aging and Regulates M2a/M2c Macrophage Polarization via Inhibiting TLR4/Myd88/NF-κB Pathway. Tissue Cell 2022, 76, 101812. [Google Scholar] [CrossRef]
  176. He, J.; Cui, J.; Shi, Y.; Wang, T.; Xin, J.; Li, Y.; Shan, X.; Zhu, Z.; Gao, Y. Astragaloside IV Attenuates High-Glucose-Induced Impairment in Diabetic Nephropathy by Increasing Klotho Expression via the NF-κB/NLRP3 Axis. J. Diabetes Res. 2023, 2023, 7423661. [Google Scholar] [CrossRef]
  177. Liu, T.; Zhang, L.; Joo, D.; Sun, S.-C. NF-κB Signaling in Inflammation. Signal Transduct. Target. Ther. 2017, 2, 17023. [Google Scholar] [CrossRef]
  178. Romero, A.; Dongil, P.; Valencia, I.; Vallejo, S.; Hipólito-Luengo, Á.S.; Díaz-Araya, G.; Bartha, J.L.; González-Arlanzón, M.M.; Rivilla, F.; de la Cuesta, F.; et al. Pharmacological Blockade of NLRP3 Inflammasome/IL-1β-Positive Loop Mitigates Endothelial Cell Senescence and Dysfunction. Aging Dis. 2022, 13, 284–297. [Google Scholar] [CrossRef] [PubMed]
  179. Swanson, K.V.; Deng, M.; Ting, J.P.-Y. The NLRP3 Inflammasome: Molecular Activation and Regulation to Therapeutics. Nat. Rev. Immunol. 2019, 19, 477–489. [Google Scholar] [CrossRef] [PubMed]
  180. Wang, Z.; Zhang, H.; Zheng, G.; Wang, Z.; Shi, L. Gender-Specific Association between Circulating Serum Klotho and Metabolic Components in Adults. BMC Endocr. Disord. 2024, 24, 198. [Google Scholar] [CrossRef] [PubMed]
  181. Schweizer, J.R.d.O.L.; Schilbach, K.; Haenelt, M.; Gagliardo, A.; Peters, A.; Thorand, B.; Störmann, S.; Schopohl, J.; Reincke, M.; Lauseker, M.; et al. Soluble Alpha Klotho—Impact of Biological Variables and Reference Intervals for Adults. Eur. J. Endocrinol. 2025, 192, 631–640. [Google Scholar] [CrossRef]
  182. Yu, L.-X.; Li, S.-S.; Sha, M.-Y.; Kong, J.-W.; Ye, J.-M.; Liu, Q.-F. The Controversy of Klotho as a Potential Biomarker in Chronic Kidney Disease. Front. Pharmacol. 2022, 13, 931746. [Google Scholar] [CrossRef]
  183. Gaitán, J.M.; Moon, H.Y.; Stremlau, M.; Dubal, D.B.; Cook, D.B.; Okonkwo, O.C.; van Praag, H. Effects of Aerobic Exercise Training on Systemic Biomarkers and Cognition in Late Middle-Aged Adults at Risk for Alzheimer’s Disease. Front. Endocrinol 2021, 12, 660181. [Google Scholar] [CrossRef]
  184. Corrêa, H.d.L.; Raab, A.T.O.; Araújo, T.M.; Deus, L.A.; Reis, A.L.; Honorato, F.S.; Rodrigues-Silva, P.L.; Neves, R.V.P.; Brunetta, H.S.; Mori, M.A.d.S.; et al. A Systematic Review and Meta-Analysis Demonstrating Klotho as an Emerging Exerkine. Sci. Rep. 2022, 12, 17587. [Google Scholar] [CrossRef] [PubMed]
  185. Amaro-Gahete, F.J.; De-la-O, A.; Jurado-Fasoli, L.; Espuch-Oliver, A.; de Haro, T.; Gutierrez, A.; Ruiz, J.R.; Castillo, M.J. Exercise Training Increases the S-Klotho Plasma Levels in Sedentary Middle-Aged Adults: A Randomised Controlled Trial. The FIT-AGEING Study. J. Sports Sci. 2019, 37, 2175–2183. [Google Scholar] [CrossRef]
  186. Tan, S.-J.; Chu, M.M.; Toussaint, N.D.; Cai, M.M.; Hewitson, T.D.; Holt, S.G. High-Intensity Physical Exercise Increases Serum α-Klotho Levels in Healthy Volunteers. J. Circ. Biomark. 2018, 7, 1849454418794582. [Google Scholar] [CrossRef] [PubMed]
  187. Santos-Dias, A.; MacKenzie, B.; Oliveira-Junior, M.C.; Moyses, R.M.; Consolim-Colombo, F.M.; Vieira, R.P. Longevity Protein Klotho Is Induced by a Single Bout of Exercise. Br. J. Sports Med. 2017, 51, 549–550. [Google Scholar] [CrossRef]
  188. Avin, K.G.; Coen, P.M.; Huang, W.; Stolz, D.B.; Sowa, G.A.; Dubé, J.J.; Goodpaster, B.H.; O’Doherty, R.M.; Ambrosio, F. Skeletal Muscle as a Regulator of the Longevity Protein, Klotho. Front. Physiol. 2014, 5, 189. [Google Scholar] [CrossRef] [PubMed]
  189. Matsubara, T.; Miyaki, A.; Akazawa, N.; Choi, Y.; Ra, S.-G.; Tanahashi, K.; Kumagai, H.; Oikawa, S.; Maeda, S. Aerobic Exercise Training Increases Plasma Klotho Levels and Reduces Arterial Stiffness in Postmenopausal Women. Am. J. Physiol. Heart Circ. Physiol. 2014, 306, H348–H355. [Google Scholar] [CrossRef]
  190. Becke, A.; Maass, A.; Kreutz, M.R.; Duezel, E. Serum α-Klotho Levels Correlate with Episodic Memory Changes Related to Cardiovascular Exercise in Older Adults. bioRxiv 2020. [Google Scholar] [CrossRef]
  191. Yoon, H.E.; Lim, S.W.; Piao, S.G.; Song, J.-H.; Kim, J.; Yang, C.W. Statin Upregulates the Expression of Klotho, an Anti-Aging Gene, in Experimental Cyclosporine Nephropathy. Nephron Exp. Nephrol. 2012, 120, e123–e133. [Google Scholar] [CrossRef]
  192. Kuwahara, N.; Sasaki, S.; Kobara, M.; Nakata, T.; Tatsumi, T.; Irie, H.; Narumiya, H.; Hatta, T.; Takeda, K.; Matsubara, H.; et al. HMG-CoA Reductase Inhibition Improves Anti-Aging Klotho Protein Expression and Arteriosclerosis in Rats with Chronic Inhibition of Nitric Oxide Synthesis. Int. J. Cardiol. 2008, 123, 84–90. [Google Scholar] [CrossRef]
  193. Narumiya, H.; Sasaki, S.; Kuwahara, N.; Irie, H.; Kusaba, T.; Kameyama, H.; Tamagaki, K.; Hatta, T.; Takeda, K.; Matsubara, H. HMG-CoA Reductase Inhibitors up-Regulate Anti-Aging Klotho mRNA via RhoA Inactivation in IMCD3 Cells. Cardiovasc. Res. 2004, 64, 331–336. [Google Scholar] [CrossRef]
  194. Janić, M.; Lunder, M.; Novaković, S.; Škerl, P.; Šabovič, M. Expression of Longevity Genes Induced by a Low-Dose Fluvastatin and Valsartan Combination with the Potential to Prevent/Treat “Aging-Related Disorders”. Int. J. Mol. Sci. 2019, 20, 1844. [Google Scholar] [CrossRef]
  195. Lim, S.C.; Liu, J.-J.; Subramaniam, T.; Sum, C.F. Elevated Circulating Alpha-Klotho by Angiotensin II Receptor Blocker Losartan Is Associated with Reduction of Albuminuria in Type 2 Diabetic Patients. J. Renin Angiotensin Aldosterone Syst. 2014, 15, 487–490. [Google Scholar] [CrossRef] [PubMed]
  196. Karalliedde, J.; Maltese, G.; Hill, B.; Viberti, G.; Gnudi, L. Effect of Renin-Angiotensin System Blockade on Soluble Klotho in Patients with Type 2 Diabetes, Systolic Hypertension, and Albuminuria. Clin. J. Am. Soc. Nephrol. 2013, 8, 1899–1905. [Google Scholar] [CrossRef] [PubMed]
  197. Cheng, L.; Zhang, L.; Yang, J.; Hao, L. Activation of Peroxisome Proliferator-Activated Receptor γ Inhibits Vascular Calcification by Upregulating Klotho. Exp. Ther. Med. 2017, 13, 467–474. [Google Scholar] [CrossRef]
  198. Chen, L.-J.; Cheng, M.-F.; Ku, P.-M.; Lin, J.-W. Rosiglitazone Increases Cerebral Klotho Expression to Reverse Baroreflex in Type 1-like Diabetic Rats. Biomed. Res. Int. 2014, 2014, 309151. [Google Scholar] [CrossRef]
  199. Haussler, M.R.; Livingston, S.; Sabir, Z.L.; Haussler, C.A.; Jurutka, P.W. Vitamin D Receptor Mediates a Myriad of Biological Actions Dependent on Its 1,25-Dihydroxyvitamin D Ligand: Distinct Regulatory Themes Revealed by Induction of Klotho and Fibroblast Growth Factor-23. JBMR Plus 2021, 5, e10432. [Google Scholar] [CrossRef]
  200. Prud’homme, G.J.; Glinka, Y.; Kurt, M.; Liu, W.; Wang, Q. The Anti-Aging Protein Klotho Is Induced by GABA Therapy and Exerts Protective and Stimulatory Effects on Pancreatic Beta Cells. Biochem. Biophys. Res. Commun. 2017, 493, 1542–1547. [Google Scholar] [CrossRef]
  201. Razzaque, M.S. FGF23, Klotho and Vitamin D Interactions: What Have We Learned from in Vivo Mouse Genetics Studies? Adv. Exp. Med. Biol. 2012, 728, 84–91. [Google Scholar] [CrossRef]
  202. Wang, Q.; Ren, L.; Wan, Y.; Prud’homme, G.J. GABAergic Regulation of Pancreatic Islet Cells: Physiology and Antidiabetic Effects. J. Cell Physiol. 2019, 234, 14432–14444. [Google Scholar] [CrossRef]
Figure 1. Klotho structure. The chart shows how the secreted and trimmed forms of α-Klotho are produced. The Klotho protein is membrane-bound and consists of three domains: a cytoplasmic domain (CYT), a transmembrane domain (TM) and an ectodomain. The latter is in turn formed by two internal repeats, KL1 and KL2. The ectodomain can be trimmed at two different levels by the enzyme.
Figure 1. Klotho structure. The chart shows how the secreted and trimmed forms of α-Klotho are produced. The Klotho protein is membrane-bound and consists of three domains: a cytoplasmic domain (CYT), a transmembrane domain (TM) and an ectodomain. The latter is in turn formed by two internal repeats, KL1 and KL2. The ectodomain can be trimmed at two different levels by the enzyme.
Ijms 26 08551 g001
Figure 2. Different effects of α-Klotho increase in pathology of AD, PD and FIS.
Figure 2. Different effects of α-Klotho increase in pathology of AD, PD and FIS.
Ijms 26 08551 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

López-Valdés, H.E.; Hernández-Lucas, M.; Rodríguez-Fabián, G.D.J.; Esteban-Román, N.F.; Gutiérrez-Juárez, R.; Arrieta-Cruz, I.; Martínez-Coria, H. The Anti-Inflammatory Actions of Soluble Klotho in Brain Aging and Its Main Associated Diseases. Int. J. Mol. Sci. 2025, 26, 8551. https://doi.org/10.3390/ijms26178551

AMA Style

López-Valdés HE, Hernández-Lucas M, Rodríguez-Fabián GDJ, Esteban-Román NF, Gutiérrez-Juárez R, Arrieta-Cruz I, Martínez-Coria H. The Anti-Inflammatory Actions of Soluble Klotho in Brain Aging and Its Main Associated Diseases. International Journal of Molecular Sciences. 2025; 26(17):8551. https://doi.org/10.3390/ijms26178551

Chicago/Turabian Style

López-Valdés, Héctor E., Martín Hernández-Lucas, Gustavo D. J. Rodríguez-Fabián, Nadia F. Esteban-Román, Roger Gutiérrez-Juárez, Isabel Arrieta-Cruz, and Hilda Martínez-Coria. 2025. "The Anti-Inflammatory Actions of Soluble Klotho in Brain Aging and Its Main Associated Diseases" International Journal of Molecular Sciences 26, no. 17: 8551. https://doi.org/10.3390/ijms26178551

APA Style

López-Valdés, H. E., Hernández-Lucas, M., Rodríguez-Fabián, G. D. J., Esteban-Román, N. F., Gutiérrez-Juárez, R., Arrieta-Cruz, I., & Martínez-Coria, H. (2025). The Anti-Inflammatory Actions of Soluble Klotho in Brain Aging and Its Main Associated Diseases. International Journal of Molecular Sciences, 26(17), 8551. https://doi.org/10.3390/ijms26178551

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