Skip to main content

Can exercise benefits be harnessed with drugs? A new way to combat neurodegenerative diseases by boosting neurogenesis

Abstract

Adult hippocampal neurogenesis (AHN) is affected by multiple factors, such as enriched environment, exercise, ageing, and neurodegenerative disorders. Neurodegenerative disorders can impair AHN, leading to progressive neuronal loss and cognitive decline. Compelling evidence suggests that individuals engaged in regular exercise exhibit higher production of proteins that are essential for AHN and memory. Interestingly, specific molecules that mediate the effects of exercise have shown effectiveness in promoting AHN and cognition in different transgenic animal models. Despite these advancements, the precise mechanisms by which exercise mimetics induce AHN remain partially understood. Recently, some novel exercise molecules have been tested and the underlying mechanisms have been proposed, involving intercommunications between multiple organs such as muscle-brain crosstalk, liver-brain crosstalk, and gut-brain crosstalk. In this review, we will discuss the current evidence regarding the effects and potential mechanisms of exercise mimetics on AHN and cognition in various neurological disorders. Opportunities, challenges, and future directions in this research field are also discussed.

Introduction

With the ageing of society, the prevalence of neurodegenerative disorders, such as Alzheimer's disease (AD), Parkinson's disease (PD), and Huntington's disease (HD), is rapidly increasing [1,2,3,4,5], but effective treatments for these diseases are currently lacking. Physical exercise has long been known to have multiple benefits for brain health, including enhancing cognitive function, memory, learning, and attention [6, 7]. These benefits are believed to be linked in part to increased adult hippocampal neurogenesis (AHN) [8,9,10]. The subgranular zone of the hippocampal dentate gyrus (DG) contains neural stem cells (NSCs) that continuously produce dentate granule cells (DGCs) across the mammalian lifespan [11]. The process of AHN is primarily driven by experience, allowing the brain to adapt to environmental demands and affecting pre-existing circuitry and hippocampus-dependent memory function. Recent evidence from postmortem brain specimens revealed that AHN is persistent throughout life in humans, and its magnitude is affected by the presence of a variety of neurodegenerative diseases, such as PD, AD, and HD [12,13,14,15]. Interestingly, a recent study applied the single nucleus RNA sequencing (snRNA-seq) technique to profile neurogenic trajectory in the adult human hippocampus and revealed persistence of neurogenesis in humans across divergent ages [16, 17]. In addition, the percentage of immature granule cells is lower in AD patients than that in matched controls [16]. Increasing evidence has indicated that AHN is impaired in various neurodegenerative conditions such as AD, PD, and HD [18, 19]. While those diseases are not exclusively hippocampal disorders, they frequently affect the hippocampus and lead to cognitive symptoms [20, 21]. For example, the hippocampus is one of the earliest regions affected by dispersed aggregates of amyloid beta (Aβ) and tau proteins, leading to impaired hippocampal neurogenesis and memory deficits in AD [20, 21]. Enhancing neurogenesis in the hippocampus, such as through reduction of tau [22] and APP accumulation [23] and increment of neurotrophic factors [24], could potentially mitigate the adverse effects of AD pathology on the hippocampus. This may offer a compensatory mechanism to counteract cognitive dysfunction, the loss of neuronal connectivity, and cell death associated with AD [20, 25]. Similarly, the hippocampus is vulnerable to the accumulation of fibril α-synuclein, resulting in non-motor symptoms such as cognitive decline and mood disorders in PD [26, 27]. HD is characterised by cognitive deficits and affective disorders [28], and hippocampal disorders are involved in the cognitive and mood symptoms associated with HD [29]. Targeting AHN through interventions such as environmental enrichment and exercise might counteract the negative impacts of neurodegenerative conditions on the hippocampus, resulting in improved memory processing, reduced depression, and enhanced brain plasticity in PD [27, 30] and HD [31, 32]. Taken together, as the hippocampus is a common area impaired in neurodegenerative conditions, a growing body of evidence has suggested the potential of AHN-targeting strategies for neurodegenerative conditions [20, 25, 27, 30,31,32].

Physical exercise promotes adult neurogenesis, modifies synaptic connections, and enhances cognitive functions by activating a variety of molecular and cellular processes [33,34,35]. Interestingly, the exercise-induced pro-neurogenic effects have been replicated through administration of certain compounds such as brain-derived neurotrophic factor (BDNF), irisin, and clusterin derived from ‘runner plasma’, which boost neurogenesis and enhance cognitive function in rodents [34, 36, 37]. The concept of ‘exercise mimetics’ indeed has attracted much interest; it represents a new approach to ameliorating neurodegeneration and improving cognitive function in ageing and neurodegenerative disorders by replicating, to some extent, the effects of exercise [38, 39]. It is particularly important for the elderly or the diseased people who have difficulty in conducting regular exercises. Additionally, evidence indicates that the newborn neurons do not always work efficiently with local neural networks; instead, they induce disturbance in the existing circuits, resulting in forgetting. This disturbing impact could be ameliorated by exercise intervention [40], indicating that exercise mimetics may also facilitate the integration of newly generated neurons into the existing circuitry.

Significant progresses have been made in the regulation of adult neurogenesis, such as using pro-neurogenic factors to promote neurogenesis, re-activation of endogenous neurogenesis or directed reprogramming in situ. Among various strategies, targeting AHN by administering specific compounds that replicate the effects of exercise is a promising new approach [34, 36, 37]. This review aims to address critical questions and emerging evidence related to the impact of exercise mimetics on adult neurogenesis and its relevance in neurodegenerative disorders. Particularly, the following aspects are focused on in this review: (i) the generation and modulation of adult neurogenesis and its relevance in neurodegenerative diseases; (ii) the effects of exercise mimetics on AHN and their potential benefits in animal models and patients with brain disorders; (iii) the mechanisms underlying the action of exercise mimetics; and (iv) the challenges and future directions of putative molecules of exercise mimetics.

Neurogenic deficit is an essential feature of neurodegeneration

Impaired hippocampal neurogenesis is an essential feature of neurodegeneration in both ageing and diseases. The ageing-associated neurodegeneration is characterised by a decline in AHN and impairment of cognition and memory [2]. Aged mice show a significant reduction of neurogenesis in both the subventricular zone and DG niches, leading to a marked decrease of newborn neurons [41]. Similar to rodents, decreased hippocampal neurogenesis has also been observed in aged humans. One study examined doublecortin (DCX)-positive immature neurons in 13 healthy individuals aged between 43 and 87 years, and showed decreased number of DCX-positive cells in the DG of older adults [12]. Older individuals also have a smaller quiescent NSC pool in the DG and reduced neuroplasticity and angiogenesis [14]. Neuroinflammation in the aged central nervous system (CNS) may contribute to reduced neurogenesis [42]. Accordingly, exposure of older animals to the young blood can restore neurogenesis and cognition [43]. Additionally, the maintenance of neurogenesis depends on a closely integrated network of signals, and changes to these signalling pathways during ageing have been associated with the reduction in neurogenesis [44].

The neurogenic niche is susceptible to neurodegenerative conditions [45, 46]. Neuroinflammation is one of the key features of neurodegenerative disorders, such as AD, PD, and HD. It can be stimulated by Aβ plaques, fibrillar tau, or α-synuclein aggregates [45, 46]. Postmortem brain tissue analysis of PD patients revealed hippocampal atrophy, reduced NSCs in the DG region, as well as an increased number of reactive microglia and elevated levels of pro-inflammatory factors, such as tumor necrosis factor-alpha (TNF-α), interleukin (IL)-1β, and IL-6 in the midbrain and cerebrospinal fluid [47]. Moreover, a recent study examining postmortem brain samples of 45 AD patients aged 52 to 97 years suggested a decreased number of immature neurons at all stages of the disease compared to healthy individuals of similar age [12]. Similarly, transgenic AD models, such as the 3 × Tg, 5 × FAD, and APP overexpression models, show a clear neurogenic impairment [21]. HD also shows obvious AHN deficits. In HD patients, newly generated neurons display early maturation impairments, morphological abnormalities, and reduced expression of NeuN [13], consistent with the observed impairments in AHN in transgenic HD mouse models [48]. DG astrogliosis and elevated microglial activations are recognized as detrimental to the homeostasis of the DG neurogenic niche [13]. Together, neurodegenerative conditions have a profound impact on AHN and memory function, which also offer a potential therapeutic strategy for neurodegenerative disorders, i.e., boosting neurogenesis by exercise and agents mimicking the effects of exercise. The correlation between AHN and neurodegenerative diseases and the potential therapeutic approach will be discussed in the sections below.

AHN shapes the plasticity of mammal brains

Over the past two decades, significant progress has been made in understanding the origin, formation, regulation, and integration of adult-born DGCs in the mammalian brain [49, 50]. NSCs serve as the main origin of new hippocampal neurons in the adult brain. They are located in the neurogenic niche, an area enriched with various types of cells, extracellular matrix, and cortical and subcortical neuronal projections, which collectively provide diverse signals essential for maintaining the function of NSCs [51] (Fig. 1). NSCs reside mainly in a dormant, quiescent state, and the quiescent NSCs proliferate as an adaption to environmental cues such as feeding and exercise [52,53,54]. Dormant NSCs can be activated and differentiate into DGCs in response to changes of neurogenic niche signals. During integration and maturation, approximately 75% of newborn granule cells will die within three weeks after division [55, 56]. The integration of adult-born neurons into the hippocampal circuit is accomplished by their competition with mature granule cells for the perforant pathway input. Newborn neurons with more spines may undergo an accelerated integration process [57,58,59]. Those circuits connecting anatomically and physiologically between the entorhinal cortex, DG, and CA3 facilitate the resolution of memory interference.

Fig. 1
figure 1

Proliferation, division, maturation, and integration of adult hippocampal neurogenesis in mammalians. Radial glia-like neural stem cells (RGLs) reside in the subgranular zone (SGZ) of the hippocampal dentate gyrus (DG) and give rise to new granule cells. RGLs can be activated and differentiate into dentate granule cells (DGCs) in response to neurogenic niche signals, making subsequent fate choices: (i) self-replicate to expand RGL pools; (ii) generate neurons; (iii) generate one RGL and one non-RGL; (iv) divide into two non-RGLs; (v) give rise to astrocytes [382]. Activated RGLs experience remarkable genetic and metabolic changes, such as upregulated expression of achaete-scute homologue (ASCL1), and cell metabolism-associated cell activation [383, 384]. Adult-born neurons in the rodent hippocampal DG ultimately develop into excitatory, glutamatergic granule neurons within the following several weeks after birth. However, not all newly generated DGCs in rodents can survive; finally, a substantial fraction of newborn DGCs die within three weeks after division, leaving less than one-quarter of newborn neurons to survive and integrate into the existing circuits [55, 56]. EC, Entorhinal cortex; GFAP, Glial fibrillary acidic protein; DCX, Doublecortin; NeuN, Neuronal nuclei; Prox1, Prospero homeobox protein 1; PSA-NCAM, Polysialylated-neural cell adhesion molecule

The newly generated neurons are critical for the resolution of memory interference by improving hippocampal memory discrimination, consolidation, and clearance. Depleting hippocampal neurogenesis increases memory interference during reversal learning tests [60,61,62], whereas genetically expanding the number of newborn DGCs had the opposite effect [59]. Newly generated DGCs are also crucial for expediting the consolidation of prior experiences and improving the precision of long-term memory [61, 63, 64]. As a result, they facilitate discrimination between subsequent experiences. Furthermore, the ongoing formation and integration of newly generated neurons into the preexisting circuitry facilitate a unique memory phenomenon called ‘active forgetting’. This process is essential for memory encoding and consolidation by degrading or eliminating existing information stored in hippocampal circuits, thereby creating space for new learning [40, 65]. However, AHN is often impaired in various neurological disorders. It can also be enhanced by favourable conditions such as exercise. Therefore, AHN may hold the potential as a therapeutic target for many neurodegenerative diseases.

New evidence for the presence of AHN in humans

The existence of AHN has been validated in non-human mammalians, but remains debated in humans since the first report of hippocampal neurogenesis in rodents [11]. Eriksson et al. [66] led pioneering work to reveal the existence of adult neurogenesis in the human hippocampus by incorporating a thymidine analogue, bromodeoxyuridine (BrdU), into the DNA of cells undergoing division. Later, another prominent work conducted by Spalding et al. [67], measured the immediate increment of incorporated 14C in dividing cells, and validated the presence of adult-born granule cells in the human hippocampus. Following those studies, a growing range of studies extended the evidence to support the existence of adult neurogenesis in humans [68,69,70,71,72,73,74,75].

However, there are still discrepancies in the findings related to this phenomenon [12,13,14, 76,77,78]. Recent emerging techniques, such as snRNA-seq analysis, have brought a novel perspective on this issue. Habib et al. [79] first generated high-quality single-nuclei transcriptomic profiles from preserved adult human postmortem tissues and detected a cluster of 201 hippocampal cells identified as NSCs based upon known marker gene expression [79]. Subsequently, two research groups applying snRNA-seq analysis [16, 17], revealed the cell populations, transcriptomic markers, and neurogenic trajectory within the adult human hippocampus. Interestingly, one study [16] utilized an enhanced snRNA-seq protocol, incorporating a machine learning-based analysis program, to reanalyse several previously published snRNA-seq datasets [79,80,81]. This study confirmed the presence of immature granule cells in all three snRNA-seq datasets, including the dataset [81] that had originally reported an extremely low rate of neurogenesis. Moreover, Zhou et al. [16] assessed the impact of neurodegenerative diseases on hippocampal neurogenesis and found that the number of newborn neurons was reduced by 50% in patients with AD compared to control subjects. Together, by applying advanced analysis approaches, researchers have now been able to better delineate NSCs, progenitors and immature neurons in postmortem human brain samples and reason out their neurogenic lineage and maturation. However, the snRNA-seq data are extremely scarce in human studies. To fully understand the neurogenic profiles associated with various neurodegenerative disorders, more studies are urgently needed.

Neurodegenerative diseases and AHN: evidence and therapeutic implications

AD and hippocampal neurogenesis

Individuals with AD exhibit a marked reduction in the number of viable precursor cells within the hippocampus [82]. Musashi-1- and Ki-67-positive precursor cells from AD patients are capable of self-renewal but reach senescence early [82]. Moreover, a recent study analyzed the number of adult-born neurons among 45 AD autopsy brain samples from subjects aged 52–97 years. The results showed that in contrast to the detection of thousands of immature neurons in the hippocampal DG of healthy subjects across the age, the number and the maturation of those cells markedly decreased in AD patients [12]. Additionally, in the study by Zhou et al. [16], the impact of neurodegenerative diseases on AHN was examined by snRNA-seq analysis on postmortem brain samples from AD patients. The findings indicated a 50% reduction in the proportion of immature granule cells in AD patients compared to the control group. Our current understanding of the mechanism of how AD pathology affects AHN is mainly based on the evidence from various animal models. 3 × Tg-AD mice, a most popular AD model carrying APP, PSEN1 M146V, and MAPT P301L transgenes [83], display obvious protein clumps, tangles, neuroinflammation, cognitive dysfunction, and impaired hippocampal neurogenesis at an early stage (4 months) [84,85,86]. In an alternative transgenic rodent model of tauopathy, which specifically overexpresses human tau (htau) in gamma-aminobutyric acid (GABAergic) interneurons in DG, the hippocampal neurogenesis and morphology of immature neurons are impaired via suppression of GABAergic transmission and disruption of local neural circuitry [87]. Importantly, increasing GABAergic transmission by THIP restores hippocampal neurogenesis and improves cognitive function, providing a potential treatment strategy for AD [87]. Taken together, these studies indicate the vulnerability of the neurogenic niche and the maturation processes of newborn neurons in response to AD pathology.

PD and hippocampal neurogenesis

Generally, postmortem analysis data on adult neurogenesis with respect to PD pathology are limited. Terreros-Roncal and colleagues [13] demonstrated that PD patients display an increased density of radial glia-like cells with NSC properties but abnormal morphology of DCX-positive granule cells with impaired maturation. Hoglinger et al. [88] determined the cells expressing nestin and β-III-tubulin in postmortem hippocampal DG from PD patients, and they showed that those immature neuronal markers were significantly down-regulated in the hippocampal DG compared to controls. A pathological hallmark of PD is the deposits of misfolded α-synuclein into Lewy bodies, which are cytotoxic to neural cells [89]. Winner et al. [90] generated α/β-synuclein double knockout (KO) murine models and detected a significant increment in newborn neurons in the hippocampus. Subsequently, in transgenic mice overexpressing human wild-type α-synuclein (Hwt-α-syn) under the control of platelet-derived growth factor-β promoter, they found reduced AHN and elevated cell death as well as impaired dendrite outgrowth, decreased dendritic length and lesioned dendritic branching in immature newborn neurons [90, 91]. Meanwhile, Lrrk2 transgenic rodents expressing mutant G2019S display enhanced anxiety-associated behaviours and decreased hippocampal neurogenesis, leading to a 77% reduction in the number of immature adult-born neurons [92]. Generally, regional and cell-specific overexpression of α-synuclein or mutant Lrrk2 results in the degeneration of selective circuitries, as well as motor deficits and inclusion formation featured in PD, ultimately leading to impairment of proliferation, differentiation and survival of newborn neurons [93,94,95]. Collectively, evidence from both clinical data and animal models of PD indicates that impaired AHN is a hallmark of PD pathology.

HD and hippocampal neurogenesis

Early maturation defects, morphological changes, and reduced expression of NeuN have been detected in newborn neurons in the postmortem brain samples of HD patients [13]. DG astrogliosis and increased microglial activation may disrupt the homeostasis of the DG neurogenic niche. R6 lines are the most widely used HD transgenic animal models carrying 115 (R6/1) or 150 (R6/2) CAG repeats via introducing exon 1 of the human HD gene into the rodent germline [32, 96]. The R6/1 mice show a remarkable decrease (64%) in the number of newborn granule cells at age 20 weeks [96]. Moreover, both differentiation and survival of newborn neurons in the hippocampus are severely impaired in these mice, leading to an overall decrease in the population of mature newborn neurons [96]. In contrast, the R6/2 mice exhibit impaired hippocampal neurogenesis at an early stage (about 2 weeks) and a 70% decrease in the number of newly generated neuroblasts at 12 weeks of age [32]. Additionally, both R6 lines showed decreased expression of polysialylated-neural cell adhesion molecule in the hippocampus, indicating impairment in synaptic plasticity, axonal growth, dendritic branching and cell migration [97, 98]. Consistently, in another HD transgenic model, the YAC128 mice, newly generated granule cells in the DG are particularly impaired [99]. The decrease of neurogenesis is persistent from three months (early symptom) to 18 months of age (end-stage) compared to WT controls [99].

Together, the persistent decrease in hippocampal neurogenesis represents a featured neuropathological process in AD, PD, and HD. Given that hippocampal neurogenesis plays a critical role in hippocampal-dependent memory function, the impairment in neurogenesis might account for the cognitive dysfunction in those transgenic rodents.

Exercise: an approach to combating neurodegeneration and boosting neurogenesis

Exercise prevents neurodegenerative disorders in human studies

Neurodegenerative disorders are a group of diseases characterized by progressive loss of neurons and cognitive impairment, leading to disability and death [5, 100, 101]. These diseases pose a significant challenge to public health and social care for there is currently no effective treatment for those diseases. Exercise is a well-established therapeutic strategy to promote brain health and prevent the onset and progression of neurodegenerative disorders [6, 102, 103]. The latest report on AD in 2022 has proposed a range of risk factors for AD, particularly the lack of physical activity [104]. Accordingly, a large-scale study involving 160,000 participants revealed that regular exercisers have a 45% lower chance of developing AD [105]; a similar finding (with a 53% reduction in AD risk) was reported in a longitudinal study of 716 older adults on the association of physical activity with dementia status [106]. Moreover, a recent meta-analysis incorporating eight prospective studies with 544,336 participants and 2192 patients with PD evaluated the effect of physical activity on the risk of PD [107]. The results showed that physical activity, particularly moderate to vigorous physical activity, is associated with a significant reduction in PD risk (relative risk, 0.71; 95% CI, 0.58–0.87) [107], with a stronger association among men than women. Additionally, a Cochrane review provided robust evidence of beneficial effects of most types of physical exercise on the severity of motor signs and quality of life in people with PD, compared to the control group [108]. However, the results also revealed little evidence of differences between intervention protocols, leading to the difficulty of recommending the best type of exercise for individuals [108]. Collectively, the current findings have provided robust evidence for the preventive role of physical exercise in neurodegenerative disorders [102, 109, 110].

Exercise can yield various benefits for brain function, such as increasing brain volume, cerebral perfusion, hippocampal functions, and synaptic connectivity in healthy adults and patients with neurodegenerative disorders [111,112,113] (Fig. 2). These beneficial effects are partly involved in the up-regulation of biomarker production. Exercise can increase the levels of various biomarkers that are involved in adult neurogenesis, in peripheral tissues and cerebrospinal fluid (CSF) [114,115,116], such as BDNF, insulin-like growth factor (IGF)-1, irisin, vascular endothelial growth factor (VEGF), fibroblast growth factor (FGF)-2, glial cell line-derived neurotrophic factor (GDNF), and DCX. Furthermore, physical exercise also influences the levels and activities of various neurotransmitters and neuropeptides in the human brain (Fig. 2), such as glutamate, GABA, serotonin (5-HT), dopamine (DA), norepinephrine, and acetylcholine (ACh). These neurotransmitters play crucial roles in regulating adult neurogenesis by influencing neural stem cell proliferation, differentiation, migration, and integration [117, 118]. Additionally, exercise has been shown to decrease the levels of pro-inflammatory cytokines and improve the levels of anti-inflammatory cytokines in the peripheral blood or CSF of humans [34, 119, 120], which can be advantageous for adult neurogenesis by modulating neuroinflammation and neuroimmune interactions. Together, growing evidence suggests that physical exercise can improve brain blood flow, oxygen delivery, glucose metabolism, and neurotrophic factors within the brain, as well as regulating inflammation, oxidative stress, mitochondrial function, and epigenetic modification. These processes are governed by various molecular and cellular signalling pathways, but the precise mechanisms remain unclear.

Fig. 2
figure 2

Exercise for hippocampal neurogenesis in neurodegenerative conditions. Physical exercise can exert multiple positive effects on AD and PD brains, such as enhancing cerebral blood flow, neurogenesis, synaptic plasticity, neurotrophic factors, antioxidant defense, and cognitive function. Exercise can inhibit Aβ formation and deposition, abnormal phosphorylation of Tau and α-synuclein aggregates. More importantly, physical exercise can increase hippocampal neurogenesis and memory function by modulating mitochondrial dysfunction, neuronal apoptosis, and neuroinflammation. Physical exercise can impact the activation state and phenotype of microglia and astrocytes in AD, resulting in the shift of the polarization of microglia and astrocytes from pro-inflammatory (M1 or A1) to anti-inflammatory (M2 or A2) pattern. Additionally, physical exercise can strengthen the connection of BBB, which can prevent the infiltration of peripheral immune cells and inflammatory molecules into the brain. AD, Alzheimer's disease; PD, Parkinson’s disease; Aβ, Amyloid beta; BBB, the blood–brain barrier; BDNF, Brain-derived neurotrophic factor; IGF, Insulin-like growth factor; Drp1, Dynamin-related protein 1; Mtf1/2, Mitofusion protein 1; 5-HT, 5-hydroxytryptamine; DA, Dopamine; FNDC5, fibronectin type III domain-containing protein 5; mROS, Mitochondrial reactive oxygen species

Exercise restores adult neurogenesis in neurodegenerative disorders

One of the mechanisms by which exercise alleviates neurodegeneration and improves cognition is through enhancing AHN [10, 121] (Fig. 2). Growing evidence from in vivo and in vitro experiments has demonstrated that physical exercise can protect brain plasticity and function in ageing rodents and those with neurological conditions. These changes include promotion of the number, the survival, the differentiation, and the integration of newly generated neurons in the brain, as well as the increase of their synaptic connections [10, 122, 123] (Fig. 2). Recent studies have found that different types of exercise training, such as wheel running, treadmill running, and swimming, can promote the proliferation, survival, differentiation, and integration of new neurons in the DG region of the adult rodent brain [124,125,126,127]. Moreover, exercise can also improve synaptic plasticity and the function of these new neurons, resulting in more complex dendrites, more spines, stronger LTP, and higher electrophysiological activity [102, 6, 112]. Accordingly, rodents that received exercise showed improved cognition and mood in various hippocampal-dependent tasks, such as spatial learning, memory consolidation, pattern separation (PS), and depression-like behaviour [128,129,130].

The regulatory mechanisms by which exercise promotes adult neurogenesis in rodents are complex and multifactorial, involving neurotrophic factors, neurotransmitters, inflammation, oxidative stress, epigenetics, and mitochondrial functions [9, 34, 114, 116, 131,132,133] (Fig. 2). Exercise can increase the expression and secretion of various growth factors, such as BDNF, irisin, IGF-1, VEGF, FGF-2, and GDNF, which can support the survival, growth, and differentiation of new neurons [114,115,116, 134]. Moreover, the levels and activities of neurotransmitters that influence adult neurogenesis are also elevated in rodents after exercise, such as GABA, 5-HT, DA, NE, ACh, endocannabinoids, and neuropeptides [117, 118]. Additionally, recent studies have found that exercise can reduce the production of pro-inflammatory cytokines and increase the production of anti-inflammatory cytokines (IL-4, IL-10, IL-1 receptor antagonist, etc.) in the brain, which can have beneficial effects on adult neurogenesis by modulating microglial activation, astrocyte function, blood–brain barrier (BBB) permeability, and immune cell infiltration [34, 120, 135, 136]. Similarly, exercise has been shown to enhance the antioxidant defence system in the brain by increasing the expression and activity of antioxidant enzymes, such as superoxide dismutase, catalase, glutathione peroxidase, and glutathione reductase (GR) [137,138,139,140,141]. Exercise can also reduce brain oxidative damage to DNA and proteins [142], and alter epigenetic regulation of gene expression by modulating DNA methylation, histone modifications, and non-coding RNAs [143,144,145,146,147]. Together, physical exercise is capable of producing various neurobiological advantages that promote brain health and counteract neurodegeneration in both ageing individuals and those with neurodegenerative diseases. These findings indicate that exercise could be a promising strategy for managing brain disorders.

Exercise mimetics: a pill for neurogenesis and neurological disorders

Exercise mimetic is a new strategy for neurological disorders by boosting neurogenesis

There is an increasing consensus on the positive effects of physical exercise in a broad spectrum of models of human diseases [148, 149], particularly CNS disorders such as AD and PD [150]. Besides the therapeutic effects on the pathological process of diseases, exercise also has a favourable impact on cognitive function in neurodegenerative disorders [6, 8, 33, 151, 152]. It is fairly clear, as many studies have discussed [38, 153, 154], that exercise could be developed as a pharmacological strategy for neurodegeneration and cognitive dysfunction, presenting as exercise mimetics [38, 39, 155] (Table 1). This becomes particularly important for older adults and patients with injuries or severe neurological diseases who have difficulties in conducting exercise regimens. Exercise mimetics are an attempt to let ‘exercise effects’ become available to the whole population through pharmacological proteins or factors (such as BDNF, Clusterin, irisin, and IGF-1) (Table 1). Interestingly, several featured studies have been conducted recently to explore the potential of ‘exercise mimetics’ in improving brain functions in neuropathological or ageing conditions [34, 36, 37]. Understanding the molecular and cellular processes underlying the communications between the periphery and the brain (such as muscle-brain crosstalk, liver-brain axis, and gut-brain axis) in response to exercise can facilitate development of exercise mimetics. We here particularly focus on the molecular effects of exercise mimetics on adult neurogenesis and cognitive function in neurological disorders.

Table 1 Candidate molecules of exercise mimetics for the treatment of neurodegenerative disorders

BDNF

BDNF is a member of the neurotrophin family and is highly expressed in the brain. Peripherally secreted BDNF can cross the BBB and reach the brain [156, 157]. The levels of BDNF can be affected by various neurodegenerative disorders, including PD, AD, and HD. Patients with these conditions often have reduced circulating levels of BDNF compared to matched healthy controls [158,159,160]. BDNF is one of the most intensively investigated exercise mimetics due to its close relation to physical exercise and crucial role in mediating neuronal proliferation, maturation, survival, and integration into existing circuits [161,162,163] (Fig. 3). The exercise-induced increase of adult neurogenesis and improvement of brain function are proposed to be associated with up-regulation of BDNF levels in the hippocampus [9, 164, 165]. BDNF exerts neurobiological impact primarily through its receptor, tropomyosin receptor kinase B (TrkB) [166]. Interestingly, exercise interventions effectively protect against neuropathology and cognitive dysfunction and increase BDNF expression in animal models of AD [24], HD [167], and other neurodegenerative disorders [163]. Moreover, BDNFMet/Met mutant mice display impairments in BDNF expression, hippocampal neurogenesis, and behavioural performance compared to BDNFVal/Val wild-type mice, which could not be reversed by exercise intervention, indicating that the presence of BDNF is essential for the neurobiological effects of exercise [168]. According to those data, pharmacologically manipulating hippocampal BDNF expression may replicate in part the neurobiological effects of exercise. Parrini and colleagues [155] increased hippocampal BDNF levels in Ts65Dn mice (a transgenic model of Down syndrome) via chronic administration (5 mg/kg body weight, 4 weeks) of 7, 8-dihydroxyflavone (DHF). Chronic DHF intervention, in contrast to acute treatment [169], directly induced a 26% increase in phosphorylated TrkB levels and successfully restored hippocampal synaptic plasticity and cognitive function [155].

Fig. 3
figure 3

Impact of exercise on brain plasticity and function through intercommunications between muscle, gut, liver, and brain. During exercise, a wide range of molecules, factors, or cytokines are secreted from different organs or tissues, such as muscle, liver, and intestinal tract, and enter blood flow, directly or indirectly affecting the central nervous system (CNS). It is fairly clear that exercise could be developed as a pharmacological strategy for alleviating neurodegeneration and cognitive dysfunction, presenting as exercise mimetics [38, 39, 155]. MARK, Adenosine monophosphate-activated protein kinase; SIRT1, Sirtuin 1; PGC-1α, Peroxisome proliferator-activated receptor coactivator 1α; FNDC5, Fibronectin type III domain-containing protein 5; Gpld1, Glycosylphosphatidylinositol (GPI)–specific phospholipase D1; BDNF, Brain-derived neurotrophic factor; DHF, 7, 8-dihydroxyflavone; SCFAs, Short-chain fatty acids; RGLs, Radial glia-like stem cells

Recently, one prominent study conducted by Choi and colleagues [36] combined BDNF protein and drug-induced neurogenesis to recapitulate the effect of exercise on cognitive performance in a transgenic rodent model of AD. They reported that only the combined protocols, instead of either alone, produced the exercise-related neurobiological effects (particularly cognitive function), demonstrating that the application of exercise mimetics in certain animal models is more complicated than expected. Additionally, apart from BDNF, exercise also increases expression of several other proteins, including postsynaptic density protein 95, synaptophysin, IL-6, and fibronectin type III domain containing 5 (FNDC5), some of which have been recognized to be associated with the neuropathology of AD [170, 171]. Given the critical role of BDNF in the effects of exercise on cognitive function, anxiety, and depression-like behaviours via mediating neurogenesis and synaptic plasticity [172,173,174], BDNF has the potential to be a promising molecular target of exercise mimetics.

FNDC5/Irisin

Exercise improves brain function partly through production of myokines from muscles to target the CNS, such as irisin. Irisin is a novel myokine initially identified in muscles and adipose, consequently extending to the heart, the liver, the bone, and the brain. It generates the new molecular basis for the crosstalk between muscles and the brain [134, 175, 176]. Irisin secretion is stimulated by exercise and regulated by peroxisome proliferator-activated receptor-γ (PPARγ) coactivator 1α (PGC-1α), a transcriptional co-regulator involved in energy metabolism [177]. Compelling studies involving individuals with AD and PD have implied a close relationship between circulating irisin levels and the risk of developing AD and PD [178,179,180,181,182,183]. Moreover, the preventive effects of irisin treatment against neuropathology have been extensively investigated in various rodent models of neurological disorders [181, 184,185,186,187,188,189]. Bretland et al., [190] showed that a 4-week injection of recombinant irisin (100 µg/kg, weekly) significantly decreased the tau phosphorylation load and inflammatory cytokine levels such as TNF-α in the hippocampus of female htau mice. Similarly, the recombinant irisin also improves novel object recognition and fear conditioning memory in AD mice, and over-expression of FNDC5/irisin rescues synaptic plasticity and memory defects in AD mice [191]. A recent study on rodent PD models showed that irisin intervention improves motor function and prevents dopaminergic neurodegeneration by reducing oxidative stress, restoring mitochondrial function, and normalizing mitochondrial dynamics and morphology [181]. Moreover, the presence of irisin is indispensable for exercise-associated improvement of brain plasticity and function. Islam and colleagues generated Fndc5 KO mice and housed each in a cage with free access to a voluntary running wheel. Fndc5 KO mice did not show exercise-related beneficial effects in spatial learning and memory compared with wild-type mice [192]. Moreover, while Fndc5 KO rodents displayed difficulty in PS tasks, overexpression of irisin by injecting adeno-associated virus (AAV) 8-irisin-FLAG in hippocampal DG restored PS in Fndc5 mutant mice [192]. To further test the therapeutic potential of irisin, several studies have been conducted by peripherally overexpressing or blocking irisin to test its impact on cognitive function [191, 192]. Lourenco et al. reported that swimming exercise (1 h per day, 5 days per week for 5 weeks) improved memory and hippocampal Fndc5 expression in mice with Aβ oligomer infusion [191]. Interestingly, the effects of exercise could be recapitulated, to some extent, by peripherally overexpressing FNDC5/irisin, which resulted in increased hippocampal irisin level and reversed AD-related memory deficits and neural pathology [191]. Additionally, the effects of exercise on synaptic plasticity and memory of AD mice could be inhibited by peripherally or cerebrally blocking FNDC5/irisin [191]. Those findings suggest that irisin could recapitulate partial ‘exercise effects’, indicating that it is a potential molecular exercise mimetic to improving brain plasticity and function (Fig. 3).

Irisin is implicated in exercise-related cerebral plasticity and function probably via modulating the process of neurogenesis through up-regulating hippocampal BDNF expression. Exercise cannot rescue the impairment of the morphology and maturation of newborn neurons in the hippocampus of Fndc5 KO mice [192]. Additionally, Fndc5 KO during differentiation of rodent embryonic stem cells into neurons resulted in remarkable down-regulation of neural progenitor markers, such as Sox1, Sox3, paired box 6, and Nestin [193]. It is hinted that the presence of irisin is critical for exercise-induced beneficial effects on neuronal proliferation and maturation. The capacity of irisin to improve brain plasticity and function is partly through increasing BDNF expression in the brain. Wrann et al. [194] demonstrated that forced expression of Fndc5 remarkably increased Bdnf expression in primary cortical neurons, whereas Fndc5 deletion via lentiviral delivery of shRNA markedly decreased Bdnf expression. Interestingly, peripheral overexpression of Fndc5 in the liver resulted in elevated serum level of irisin and enhanced hippocampal Bdnf expression in mice [194]. Recently, Choi et al. [36] demonstrated that exercise remarkably increased adult neurogenesis in transgenic AD mouse models together with elevated levels of hippocampal FNDC5 and BDNF. Recent evidence proposes an important role of irisin in mediating muscle-brain communications in exercise via a PGC-1α/irisin/BDNF-dependent pathway [134, 195]. Given the close relationship between irisin and BDNF, pharmacological or exercise-induced expression of irisin is likely to increase cerebral BDNF levels and consequently favor hippocampal neurogenesis and cognition in various brain conditions [9, 194, 196].

Gut microbiota

The microbiota-gut-brain axis is an emerging field of interest and a promising therapeutic target for CNS disorders [197]. Changes in gut microbial profiles have been reported in diverse brain disorders, including PD [198,199,200,201], AD [202, 203], and major depressive disorder [204, 205]. More interestingly, transplantation of faecal microbiota from patients with PD [206] or depression [207] led to phenotypes of disease in the recipient mice. On the contrary, transplantation of the intestinal microbiota from wild-type mice to ADLPAPT mice (a transgenic mouse model of AD) alleviated Aβ plaque deposition, neurofibrillary tangle formation, and cognitive deficits [208]. Growing evidence has suggested that changes in intestinal microbiota can potentially affect AHN [209,210,211], neuroplasticity [212], and cognition [213, 214] in neurodegenerative diseases and in ageing. In a recent study, Kim et al. [211] transferred fecal microbiota from 5 × FAD transgenic mice to normal C57BL/6 mice, which resulted in reduced AHN and BDNF expression, increased p21 expression, and memory impairment. The gut microbiota composition of the 5 × FAD mice differed from that of the control or wild-type mice, accompanied by elevated microglial activation in the hippocampus and increased pro-inflammatory cytokines in both the colon and the plasma [211]. Another study transplanted fecal microbiota from AD patients into microbiota-depleted young adult rats, which resulted in impairments in AHN-related behaviours in the recipient rats [209]. Moreover, human neural cell culture showed reduced neurogenesis when exposed to serum from AD patients. Importantly, the extent of neurogenesis reduction correlated with the cognitive scores of the AD patients, highlighting the significant influence of gut microbiota on hippocampal neurogenesis in AD [209]. These data suggest that altering the gut microbiota, possibly through exercise or products targeting its enrichment, could offer a new strategy for enhancing AHN and cognitive function in neurodegenerative conditions [215].

There is increasing evidence that physical exercise benefits both the body and the brain. One mechanism is through modification of gut microbiota, including decrease in gut transit time that leads to reduced colonization of pathogenic bacteria in the gut, and increases in diverse microbiota profiles, production of short-chain fatty acids (SCFAs), and gut microbiome composition [216,217,218,219,220,221]. In an American Gut Project, gut microbiota from individuals with regular exercise (3–5 times per week) showed a greater diversity, with enrichment in certain members of the Firmicutes phylum, such as Faecalibacterium prausnitzii, Lachnospira, Oscillospira, and Coprococcus [222]. The gut microbiota has been proposed to account for the effects of exercise on neurodegenerative disorders via enhancing AHN and cognitive function. Accordingly, exercise intervention dramatically increases the microbial enrichment in high-fat diet mice, leading to cognitive improvement [223]. In a mouse model of AD, exercise significantly ameliorated cognitive dysfunction and neuropathological biomarkers of AD, companied by microbiome alterations favouring SCFA-producing bacteria [224]. As mentioned previously, the enrichment of gut microbiota is important for AHN. The changes in gut microbiota in response to exercise may therefore benefit hippocampal neurogenesis and cognition. Indeed, research has shown that sedentary rats with disrupted microbiota display impaired performance in hippocampal neurogenesis-dependent tasks, such as the modified spontaneous location recognition task and the novelty-suppressed feeding test [225]. However, voluntary exercise is able to alleviate these effects and leads to increased hippocampal neurogenesis, which is linked to changes in caecal metabolomics [225]. Moreover, Mohle et al. [210] found that exercise training led to a moderate improvement of AHN in antibiotic-treated mice. Interestingly, when the antibiotic-treated mice received a fecal transplant from wild-type murine, the exercise-induced neurogenesis was increased by 47%, equivalent to the effect observed in control mice. This highlights the essential role of gut microbiota in mediating the effects of exercise on AHN. Therefore, the potential of physical exercise in increasing AHN by modulating the gut microbiome offers a promising protocol for the treatment of cognitive deficits associated with ageing and neurodegenerative diseases [225, 226].

Although the exact regulators driving such beneficial effects remain largely unclear, SCFAs, as the major microbial metabolites of dietary fibre in the gut, are speculated as an essential mediator. Evidence from animal models suggests that SCFAs might directly cross the BBB and affect brain function via their receptors, free fatty acid receptors [227]. Additionally, SCFAs can also interact with diverse immune cells to modulate systematic inflammation and control microglial integrity and activation that are involved in neuroinflammation [227]. PD patients often have a reduced enrichment in SCFA-producing bacteria and lower faecal SCFA contents compared with healthy controls [198, 200, 228]. Interestingly, butyrate administration ameliorated motor deficits and dopamine deficiency in rodent PD models [229, 230] and restored cognitive performance and expression of transcripts involving memory function in an AD mouse model [231]. To investigate the effects of SCFAs on neurogenesis, Xiong et al. [232] performed a single-cell mRNA analysis on samples from mice with traumatic brain injury. The results showed that SCFAs increased the frequency of immature neurons and downregulated the expression of genes associated with neurodegenerative diseases during the differentiation of immature neurons in the injured mice. Moreover, SCFAs can rescue hippocampal neurogenesis, improve BBB damage, and suppress microglial activation and neuroinflammation in mice fed a high-fructose diet and exposed to chronic stress, ultimately alleviating depressive-like behaviours [233]. The beneficial effects of SCFAs on neurogenesis, BBB integrity, and neuroinflammation suggest that these metabolites may have therapeutic potential for treating cognitive deficits and other neurological disorders.

Another mediator is kynurenine, which is strongly associated with depression. Kynurenine is proposed to be involved in the central effects of gut microbiota, which is linked with the capacity of gut microbiota to control host tryptophan metabolism and circulating kynurenine levels [234]. Physical activity increases PGC-1α expression in muscles and consequently up-regulates the kynurenine signalling cascade in the brain [235], which consequently results in resilience to stress-induced depression and CNS inflammation [235]. Taken together, the management of gut microbiota is a promising approach to recapitulating the beneficial effects of exercise on brain function (Fig. 3). One unique advantage of this strategy is the availability of a range of therapeutic products, such as probiotics and prebiotics (the applications of those products have been reviewed elsewhere [197, 227, 236]). In a recent study, Mohle et al. conducted an interesting experiment that may provide insights into the relationship between gut microbiome management by probiotics and AHN [210]. They found that simply normalizing the overall species distribution of the gut flora in mice with antibiotic-induced impairment of AHN had limited effect in restoring hippocampal neurogenesis. However, when the antibiotic-treated mice received a probiotic intervention, the hippocampal neurogenesis was completely restored to control levels [210]. This suggests that introducing beneficial probiotic bacteria may be an effective approach to restoring hippocampal neurogenesis. However, questions remain to be resolved in this field. For example, which particular microbial species are mostly affected by exercise in relation to cerebral function? Understanding this helps to discern novel candidate targets of exercise mimetics. In addition, identification of relevant mediators and elucidation of how those factors work in microbiota-gut-brain communications in response to exercise remain to be addressed in future studies.

S-Adenosyl methionine (SAM)

SAM is known as a methyl donor that provides a methyl group to another molecule through the process of methylation [237, 238]. Methyl donors are involved in diverse biological processes, including DNA methylation. DNA methylation has the potential to impact gene expression, protein synthesis, neurogenesis, and neurotransmitter metabolism [239], which are frequently disrupted in neurological conditions like AD and PD [240, 241]. 5-Methylcytosine (5mC) is an important epigenetic pattern of DNA methylation. It functions through the recognition by the methyl-binding domain (MBD) 1, which serves as an essential partner in the process of adult neurogenesis [242]. In vitro, cultured MBD1-deficient NSCs exhibit reduced neuronal differentiation and increased genomic instability [243]. Furthermore, adult mice lacking MBD1 show impaired neurogenesis and spatial learning abilities [243]. Additionally, another study found that the stimulation of mature dentate neurons in the hippocampus in adult mice via electroconvulsive therapy resulted in a prolonged elevation of hippocampal neurogenesis [244, 245]. This phenomenon is linked to the upregulation of the Gadd45b (growth arrest and DNA-damage-inducible protein 45) gene in the brains of these mice. Gadd45b plays a pivotal role in the demethylation of particular gene promoters and the activation of genes that are fundamental to the process of neurogenesis in adults, such as BDNF and FGF-1 [244, 245]. Growing evidence has revealed significant alterations in the methylation profiles of specific genes in AD patients, indicating the potential involvement of methylation in AD pathogenesis. A recent meta-analysis of methylation data in various brain regions of AD highlighted the enrichment of methylation changes in genes crucial for neurodevelopment and neurogenesis [241]. DNA methylation is under dynamic regulation by various factors, including DNA repair, oxidative stress, inflammation, environmental stimuli, and exercise [246], which may subsequently influence brain functions such as neurogenesis, synaptic plasticity, and cognition. These hint at potential therapeutic implications of DNA methylation in regulating the development and progression of AD pathology [246].

Exercise is a dynamic mediator for methylation patterns in brain health and disorders [247]. For instance, among a cohort of elderly African Americans with mild cognitive impairment, a 6-month regimen of aerobic exercise (40 min per day, 3 days per week) led to global DNA methylation changes [143], including methylation of VSP52, SACRB1, ARTN, NR1H2, and PPPLR5D, which play roles in diverse processes such as amyloid formation, intracellular protein transport, and lipoprotein mediation [143]. Additionally, exercise alters the activity of DNA methyltransferases, which are involved in the regulation of neuronal survival and methylation processes associated with ageing and disease-related neurodegeneration [248]. Swimming for 1 h a day, 26 days, increased the protein level and the DNA-binding activity of Nrf2 (nuclear factor erythroid 2-related factor 2) in a rodent model of AD, which subsequently led to decreased expression of antioxidant genes [249]. Considering the important role of gene methylation in the beneficial effects of exercise on brain function, DNA methylation and associated signalling pathways, including methyl donors, are potential therapeutic targets for brain disorders like AD and PD.

As an essential methyl donor, SAM can be used as both a dietary supplement and a prescription drug for various conditions, including depression, anxiety, liver disease, osteoarthritis, PD, and AD. Di Rocco et al. [250] treated PD patients with SAM at doses of 800 to 3600 mg/day for 10 weeks and observed a significant improvement on the 17-point Hamilton Depression Scale. Furthermore, SAM improved AD pathology and cognitive function. In 3 × Tg-AD mice, SAM administration (100 mg/kg) reduced intracellular Aβ deposits and phosphorylated tau in the hippocampus [251]. SAM therapy decreased AD neuropathology in a time-dependent manner, showing an 80% decrease in extracellular Aβ deposition in 11-month-old mice following 1-month treatment, but only a 24% reduction in 15.5-month-old mice after 3-month treatment [251]. SAM potentially influences Aβ metabolism by controlling its production, removal, and clumping. As a methyl donor, SAM can affect the functions of enzymes like secretases, which are involved in preventing generation or breakdown of APP [252]. Additionally, SAM may boost Aβ elimination by raising the levels of ApoE [253] and hinder Aβ clumping by interrupting its interaction with metal ions or changing its structure [254].

SAM treatment can also regulate certain disease processes such as oxidative damage, inflammation, mitochondrial dysfunction, and cholinergic deficits. For instance, SAM exerts antioxidant properties by increasing glutathione (GSH) and its transferase, which are crucial for neutralizing oxidative substances and removing harmful foreign substances. SAM supplementation increases glutathione S-transferase activity and the removal of reactive oxygen species (ROS) in ApoE−/− mice [255, 256]. Furthermore, in rats with D-galactose-induced brain ageing, SAM treatment (16 mg/kg for 4 weeks) reverses cognitive decline, rescues neuron loss, increases BDNF levels in the hippocampus, suppresses microglial activation, and decreases the levels of pro-inflammatory cytokines in the hippocampus and serum [257]. Additionally, SAM (10 mg/kg daily for 6 weeks) attenuated Aβ-induced cellular damage by suppressing neuroinflammation and oxidative stress [258]. Mitochondrial impairment is an essential feature of brain disorders [259] and represents a promising treatment target. Recently, Lam et al. [260] found that addition of vitamin B12 in the diet can alleviate mitochondrial fragmentation, bioenergetic defects, and oxidative stress, delaying Aβ-induced paralysis in a methionine/SAM-dependent manner. The attenuation of neuropathological conditions, such as neuroinflammation, oxidative stress, and mitochondrial dysfunction, as well as the increased expression of BDNF and other genes that regulate the process of hippocampal neurogenesis, is proposed to enhance hippocampal neurogenesis and cognition [36, 261,262,263]. In folate-deficient mice, intraperitoneal injection of SAM at a dose of 50 mg/kg per day for 14 days, starting at the age of 7 weeks, prevented an excessive number of immature neurons (DCX+ cells) while simultaneously promoting the maturation of new neurons in the DG [264]. Furthermore, SAM supplementation improves dendritic complexity and increases the number of mature spines in the DG [264]. Proper intake of methyl-donor nutrients is important for brain development and cognition. Decreased methyl-donor nutrients in the developing fetus can adversely affect brain development, significantly increasing offspring risk for metabolic and neurological disease development [265]. For example, mice fed a low folate diet show depression-like behaviour [266]. These mice display neuronal immaturities, such as increased immature neurons, decreased newborn mature neurons, reduced dendrite complexity and mature dendritic spines in the dentate gyrus (DG), accompanied by down-regulation of transcription factors involved in neuronal differentiation and maturation in the DG [266, 267]. In contrast, as demonstrated in a study by Wang [268], supplementation of the methyl donor folic acid (4–12 mg/kg, 3 times per week, for 4 weeks) improved cognition and restored the expression of BDNF in the hippocampus of mice with inhibited DNA methylation. Moreover, this treatment also improved hippocampal neurons and increased expression of pro-neurogenic genes, including DNA methyltransferase 1, histone deacetylase (HDAC) 6, and HDAC8 [268]. Collectively, SAM and other methyl-donors have demonstrated beneficial effects on neuropathological conditions associated with neurodegenerative diseases [269]. However, SAM treatment may face some limitations, including the absence of a clearly defined optimal dosage, duration, and method of administration, as well as the potential side effects. Additionally, there is a lack of well-designed, population-based studies evaluating the impact of SAM administration on hippocampal neurogenesis and cognition in various neurodegenerative diseases. Therefore, future studies are needed to address these issues.

MicroRNAs (miRNAs)

MiRNAs are short non-coding RNAs that bind mRNAs and affect their translation or degradation. MiRNAs affect diverse physiological and pathological aspects of brain health and diseases such as AD, by controlling the expression of genes related to Aβ formation, tau phosphorylation, neuroinflammation, synaptic plasticity, and neuronal viability [270,271,272,273]. Growing evidence has indicated a close relationship between miRNAs and neurological diseases [270, 271, 274,275,276]. Recently, Walgrave et al. [277] have suggested that in AD, the expression of miR-132 is down-regulated by Aβ deposition. The decrease of miR-132 is associated with impaired AHN in the DG of the hippocampus in transgenic AD mice. In addition, restoring miR-132 level in adult AD mouse hippocampus rescues AHN and memory deficits, suggesting that miR-132 is a potential biomarker and therapeutic target for brain disorders. Besides regulating gene expression in brain disorders, miRNAs are also key signalling molecules involved in brain response to physical exercise by forming regulatory complexes that influence various neurobiological processes [278,279,280,281,282]. For example, the levels of miR-106a-5p, miR-103a-3p and miR-29a-3p are significantly elevated after an 8-week exercise intervention that improves cognitive performance in PD patients [283]. In a similar study, voluntary physical exercise ameliorated the degradation of the pro-neurogenic environment in the hippocampus of a mouse model of AD by mediating the expression of miR-132 [278]. The exercise-associated improvement in the pro-neurogenic environment was associated with enhanced cognitive function in AD mice [278]. Additionally, volunteer exercise increases the proliferation of NPCs in mouse DG by down-regulating miR-135a [147]. Genetic suppression of miR-135a promotes NPC proliferation, leading to enhanced hippocampal neurogenesis [147]. The beneficial impact is linked to the modulation of 17 proteins, which are predicted targets of miR-135a in NPCs. Subsequent studies have identified a range of miRNAs, such as miR-34, miR-144-5p, miR-708-5p, miR-129–1-3p and miR-482, that are implicated in the exercise-induced enhancement of brain plasticity and cognitive function by stimulating hippocampal neurogenesis [279, 284, 285]. Interestingly, Improta-Caria et al. conducted a novel pooled analysis to evaluate the relationship between exercise-associated changes in miRNAs and AD symptoms by comparing the lists of miRNAs that are altered in AD and by exercise [286]. The analysis yielded 27 overlapping miRNAs between AD and exercise, of which 10 miRNAs displayed opposite expression patterns in AD versus exercise, which are mainly involved in cell metabolic, proteolytic, and inflammatory pathways [286]. miRNAs are a particular cluster of compounds that play a crucial role in the interaction between exercise and neurological disorders. miRNAs regulate exercise-associated neurobiological effects by binding a variety of target genes that are involved in the regulation of neuronal plasticity and function [280, 281, 287]. Recent studies have shown great interest in the therapeutic potential of miRNAs as exercise mimetics for brain disorders. Accordingly, a number of miRNAs have been identified as exercise mediators in various rodent models of neurological diseases, which may have the potential for treating brain disorders [283, 286, 288,289,290,291].

Considering the essential role of miRNAs in exercise and brain health, several studies have developed novel therapeutic approaches targeting miRNAs and relevant signalling to treat or delay neuropathological processes. MiR-146 is among the most investigated, commonly dysregulated miRNAs with known or potential roles in the neuroimmune interface in AD [292]. MiR-146 is highly expressed in microglia [293,294,295] and miR-146 knock-out mice do not show effective microglial phagocytosis in response to lipopolysaccharide (LPS), indicating that miR-146 is essential for microglial response to neuropathological stimuli [296]. Exercise can alter the expression and function of miR-146 in various tissues [297,298,299]. To test the effects of miR-146a on AD pathology, Mai et al. intranasally administered a miR-146a agomir (M146AG) to transgenic AD mice. They found alleviation of amyloid and tau pathologies and improvement of neuroinflammation and cognitive function [300]. MiR-146 plays an important role in promoting hippocampal neurogenesis, and restoring the miR-146 level improves AHN and cognition [301]. Moreover, neuroinflammation is considered a major contributor to the impairment of AHN in various neurodegenerative diseases [47, 302, 303]. Therefore, reducing neuroinflammation linked to miR-146 expression may alleviate the impairment of AHN in these neurological disorders. MiR-23b-3p is recognised as a robust modulator of α-synuclein, showing therapeutic potential in treating PD pathology [276]. MiR-23b-3p is significantly downregulated in PD patients as well as in NSCs and rats treated with 6-hydroxydopamine. Cai et al. [276] found that the miR-23b-3p mimic decreased α-synuclein mRNA by 65%, while the miR-23b-3p inhibitor increased α-synuclein by 1.4 fold. Other miRNAs such as miR-155, miR-21, miR-7, and miR-129 also show therapeutic potential for brain disorders such as AD, PD, and stroke in animal studies [304,305,306,307,308]. Despite the therapeutic potentials by targeting miRNAs, there are challenges in the use of those compounds to treat neurological disorders. Current evidence may shed some light on the directions for future research to elucidate the roles of these molecules in mediating the effects of exercise on the brain.

Glycosylphosphatidylinositol (GPI)–specific phospholipase D1 (Gpld1)

Apart from the above-mentioned exercise mimetics, several novel molecules have been tested recently and shown potential in mimicking the effects of exercise on brain plasticity and function [34, 37] (Fig. 3). Free running leads to elevated adult neurogenesis, up-regulation of BDNF expression, and enhanced hippocampal-dependent memory performance in aged rodents. These effects could be transferred to sedentary aged mice via intravenous injection (8 times for 3 weeks) of circulating blood factors in plasma isolated from the aged running rodents [37]. Liquid chromatography–tandem mass spectrometry and functional enrichment analysis identified a liver-originated, exercise-stimulated circulating factor, Gpld1, to correlate with the exercise-induced memory improvement [37]. In addition, increased plasma concentrations of Gpld1 were also identified among active older adults compared to sedentary age-matched individuals. Hepatic overexpression of Gpld1 markedly increased circulating GPI levels and was sufficient to enhance hippocampal neurogenesis and cognitive performance in aged rodent hippocampus by triggering signalling pathways downstream of GPI-targeted substrate cleavage [37]. This study identifies a liver–brain communication by which the hepatic-derived blood compound confers effects of exercise to aged mice, raising the potential of manipulating GPI expression to recapitulate the beneficial effects of exercise.

Clusterin

Mice that received ‘runner plasma’ collected from exercised mice show a markedly increased number of neural stem/progenitor cells and DCX+ neuroblasts, enhanced contextual learning and memory function, and decreased expression of inflammatory genes in the hippocampus [34]. Interestingly, the capacity of blood factors from ‘runner plasma’ to recapitulate the exercise effects shows an exercise duration-dependent manner. The plasma collected from mice with a running-wheel exercise for 28 days is more likely to replicate ‘exercise effects’ compared to plasma from mice with wheel-running for 7 or 14 days [34]. The ‘runner plasma’ shows a potent anti-inflammatory effect on neural response to LPS in the hippocampus, and subsequently, an anti-inflammatory factor–clusterin was identified. As a putative ‘blood component’, depletion of clusterin largely abolished the anti-neuroinflammatory effect of runner plasma, whereas treatment with recombinant clusterin inversed the cerebral expressions of immune and inflammatory transcripts in the mouse hippocampal endothelial cells with inoculation of LPS [34]. Moreover, serum levels of clusterin were markedly elevated after a 6-month exercise intervention in the patients with amnestic mild cognitive impairment compared to basic values before interventions, demonstrating the possibility of translating exercise effects to humans via management of clusterin [34]. While this study proposed clusterin as a key factor contributing to the positive effect of exercise on AHN and cognition, the direct impact of clusterin on hippocampal neurogenesis was not determined. This underscores the need for future investigations to address specifically this issue.

Recently, a meta-analysis encompassing a total of 28 studies found that the levels of clusterin are significantly elevated in both the plasma and the brain tissues of individuals with dementia compared to healthy control subjects [309]. Similar increases in clusterin concentrations were also observed in individuals with PD and dementia with Lewy bodies [310]. These data suggest that changes in clusterin levels may play an essential role in mediating the pathology of those diseases. Early research has indicated that overexpression of α-synuclein in SH-SY5Y human neuroblastoma cells increases clusterin expression, whereas decreasing clusterin expression aggravates α-synuclein deposition, highlighting the role of clusterin in modulating α-synuclein accumulation [311]. Further research has found that clusterin interacts with the α-synuclein complex in the lysosomal systems for α-synuclein degradation [312]. Additionally, clusterin can directly interact with Aβ and tau proteins, preventing the formation of Aβ oligomers and fibril tau and promoting their clearance [313, 314]. Clusterin shows multiple impacts on neuronal plasticity, such as promoting the survival and differentiation of neurons derived from NPCs [315] and increasing glutamatergic synaptic transmission and dendritic spine density [312]. The protective role of clusterin in facilitating the removal of misfolded proteins like α-synuclein, Aβ and tau proteins may offer potential therapeutic strategies, such as physical exercise or agents targeting clusterin expression, for alleviating neuropathology and enhancing brain plasticity and function. However, the function of clusterin in regulating Aβ and tau in AD pathology is complicated and may differ across the different stages of the disease [312, 316]. This indicates that the present understanding of the precise mechanisms mediating the interaction between clusterin and the pathology of misfolded proteins is not fully addressed, highlighting a critical area for future research endeavours. Nonetheless, as suggested by Palihati and colleagues [312], clusterin plays a crucial role in regulating neuronal plasticity and cognitive functions associated with neurodegenerative disorders; thus, therapeutic strategies targeting this protein could be a promising approach to managing these disorders.

Serotonin

Serotonin has attracted particular interest as studies have revealed increased adult neurogenesis in the hippocampus after treatment of selective serotonin reuptake inhibitors, which is regarded as indispensable for antidepressant effects [317, 318]. Research has shown a decrease in serotonin expression in the brains of AD patients compared to matched controls, with similar reductions observed in peripheral and CSF levels [319, 320]. Furthermore, dysregulated plasma serotonin levels have been linked to non-motor symptoms in PD patients [321, 322], suggesting a regulatory role for serotonin dysregulation in neurodegenerative diseases. Moreover, serotonin-mediated adult neurogenesis is involved in exercise-related beneficial effects on brain plasticity and function [323, 324]. Tryptophan hydroxylase 2-KO mice, a rodent model absolutely deficient in brain serotonin, display impaired NSC proliferation in response to exercise [324]. Additionally, Htr3a−/− mice with deletion of 5-HT type 3A receptor subunit, show resistance to exercise-induced hippocampal neurogenesis, whereas administration of serotonin receptor agonist increases newly generated neurons in the rodent hippocampus, which partly recapitulates exercise effects [325]. Those data reinforce the crucial role of serotonin-induced hippocampal neurogenesis in mediating the beneficial effects of antidepressant agents and exercise interventions [317, 318, 324].

5-Aminoimidazole-4-carboxamide ribonucleotide (AICAR): mimetics targeting metabolic pathway

The AMP-activated protein kinase (AMPK)–Sirtuin 1–PGC-1α–PPARγ signalling network is an important metabolic pathway that regulates energy consumption and mitochondrial function in response to exercise (Fig. 3). Pharmacological targeting of metabolic networks has generated several molecular agents, such as AICAR, metformin and GW501516. AICAR is an analogue of AMP that can affect multiple organ functions. It mediates a plethora of metabolic processes partly via recapitulating exercise effects, including up-regulating glucose transport protein (GLUT)-4 expression, hexokinase activities, muscle mitochondrial homeostasis and VEGF expression, as well as ameliorating inflammatory process [326,327,328]. Its featured function is the capacity to replicate, in part, exercise effects on brain health. Administration of AICAR (500 mg/kg) in adult rodents for 1 week led to increased hippocampal neurogenesis and PS performance [329]. Even in old mice (2 years), the treatment with AICAR resulted in improvement of memory and motor coordination [329]. Moreover, the administration of AICAR could promote the expression of insulin-degrading enzymes and reduce Aβ deposition in mice with AD, resulting in enhanced spatial learning and recognition performance [330]. Additionally, AICAR treatment also alleviates AD-like pathological changes including biochemistry and cognitive function in rodent models [331]. The beneficial impact of AICAR in the context of AD may be correlated to its anti-neuroinflammatory effects. Ayasolla et al. revealed that AICAR administration has inhibitory effects on LPS/Aβ-induced inflammatory response by reducing the production of pro-inflammatory cytokines (such as TNF-α, IL-1β, and IL-6) and attenuating ROS generation and glutathione depletion in glial cells [332].

Metformin

Metformin, a key compound from the biguanidine class, is among the primary drugs prescribed to patients with type 2 diabetes to control hyperglycemia by up-regulating GLUT-4 (glucose transporter 4) expression and membrane translocation in skeletal muscle and adipose tissues [333, 334]. Administration of metformin (200 mg/kg per day, 38 days) markedly promoted adult neurogenesis and spatial memory function in rodent hippocampus [335]. Moreover, in a rodent stroke model, 30-day treatment of metformin (50 mg/kg per day) led to memory improvement and increased adult-born neurons in the hippocampus via elevating AMPK activity [336]. It suggests that the exercise-induced favourable impact on brain plasticity and function could be recapitulated in part by metformin administration. The relationship between long-term metformin use and the risk of AD has been extensively investigated. Several studies have shown that metformin use may reduce the risk of AD and improve cognitive function in diabetes patients [337, 338]. For example, Hsu et al. [339] matched 800 diabetic patients who developed dementia with 3200 controls. They reported that metformin use reduced the risk of dementia (adjusted odds ratio [AOR] 0.46; 95% CI 0.35–0.61), especially AD (AOR 0.38; 95% CI 0.25–0.58), compared with controls without metformin use. However, the results are not always consistent. Using a UK primary care database, Imfeld et al. [340] reported that metformin use increased the risk of dementia (AOR 1.23, 95% CI 1.13–1.34), especially AD (AOR 1.29, 95% CI 1.15–1.44), compared with non-use of metformin. The underlying reasons for the discrepancy across studies may include variations in study design, population characteristics, confounding factors, and outcome measures. For instance, Ng et al. [341] performed a meta-analysis of observational studies and reported that metformin use reduced the risk of AD in patients with diabetes (risk ratio [RR] 0.76, 95% CI 0.63–0.92), but not in patients without diabetes (1.05, 95% CI 0.76–1.46). Since metformin use has been shown to reduce Aβ deposition and abnormal tau phosphorylation, alleviate inflammation, increase insulin sensitivity, and enhance neurogenesis [342,343,344,345], all of which are key mechanisms of AD, the relationship between metformin use and AD risk warrants further investigation.

GW501516 and cautions on metabolic pathway mimetics

Other compounds targeting AMPK pathways, such as GW501516, also show neuroprotective effects against neuropathology in a murine model of PD [346]. The neuronal protective effects of this compound may involve its anti-inflammatory effects since GW501516 treatment could reduce the interferon-α-induced inflammation in brain cells [347]. Moreover, GW501516 administration (5 mg/kg per day, for 7 days) could markedly promote memory function and hippocampal neurogenesis in young female mice [348]. Collectively, those compounds could replicate, to some degree, exercise effects and likely have clinical potential in counteracting neurodegeneration and cognitive dysfunction. However, there are some essential questions or limitations that need to be addressed for those compounds in treatment. For example, due to the low permeability across BBB, the precise doses for AICAR and GW501516 treatment are still not established [349]. Furthermore, long-term administration of those compounds may induce side effects. For example, mice with long-term AICAR treatment do not show expected enhancement in markers or genes associated with neurogenesis and cognition [329], and there might be unintended or off-target effects due to the varying distribution of AMPK receptor subtypes in different tissues of the body [350]. Moreover, long-term treatment with metformin may induce an adverse impact on cognition [340, 351]. Additionally, some side effects have been reported for the treatment of GW501516, such as gastrointestinal discomfort, muscle cramps, joint pains, and cancer risk [352, 353]. Therefore, while these compounds show promise, more research is needed to address the questions or limitations regarding the dosage, long-term safety, and potential side effects before they can be considered viable treatments for enhancing AHN and cognition.

Conclusions

In the ageing society, there is a growing burden of diseases and medication or healthcare costs. Novel, effective approaches for dealing with those issues are urgently needed. An active lifestyle is a preferred choice for improving body and brain health. Compounds derived from runner plasma, have proven valuable in achieving this goal, especially for people with limited mobility due to diseases, injuries, or ageing-associated frailty.

Blood components frequently show particular effects, which are somewhat different from the general impact of exercise. The influence of certain specific molecules (such as BDNF and irisin) on hippocampal neurogenesis and cognition presents great interest due to the fact that some neurodegenerative disorders (such as AD and PD) are regarded as incurable. Additionally, the tested molecules are often administered systematically and show marked cerebral effects [34, 36, 37], such as neurogenesis, synaptic plasticity, and cognitive function, without crossing the BBB. Moreover, exercise mimetics have one unique advantage: some biomarkers are available in a range of therapeutic products, such as probiotics and prebiotics.

However, there are some questions that need to be addressed in future studies. Current research methods are primarily conducted on animal models; therefore, the data obtained from these studies should be applied with caution to patients with neurodegenerative disorders. For instance, AHN has been demonstrated in animal hippocampi, but efficient methods for determining AHN in humans are currently lacking. Consequently, assessing the impact of exercise on AHN in patients with neurodegenerative diseases remains challenging. Moreover, the optimal exercise protocol for enhancing AHN and cognitive function has not yet been established, including the types and duration of exercise interventions. Considering the diverse pathological origins of neurodegenerative diseases, the specific exercises designed to address these varying pathologies and enhance AHN could potentially vary across different conditions. Therefore, the urgent priority is to construct disease-specific, comprehensive models (at different levels: molecule, cell, and system) that can help elucidate the underlying mechanisms of how exercise impacts the brain. Moreover, the following step is to discriminate the causal from correlative compounds involved in the molecular process of brain response to exercise intervention, and those mediating molecules could be targets for pharmacological and clinical development. The treatment of neurological disorders with exercise mimetics is still in its infancy, and the optimal dose, timing, and duration of compounds are not well established. The safety and efficacy of exercise mimetics in humans need to be evaluated in clinical trials.

Availability of data and materials

Not applicable.

Abbreviations

AHN:

Adult hippocampal neurogenesis

HD:

Huntington's disease

NSC:

Neural stem Cell

DG:

Dentate gyrus

DGCs:

Dentate gyrus granule cells

snRNA-seq:

Single-nucleus RNA sequencing

BrdU:

Bromodeoxyuridine

PSANCAM:

Polysialic acid-neural cell adhesion molecule

APP:

Amyloid precursor protein

PS1:

Presenilin 1

CSF:

Cerebrospinal fluid

IGF-1:

Insulin-like growth factor-1

GABA:

Gamma-aminobutyric acid

5HT:

Serotonin

DA:

Dopamine

ACh:

Acetylcholine

FGF-2:

Fibroblast growth factor-2

TrkB:

Tropomyosin receptor kinase B

FNDC5:

Fibronectin type III domain containing 5

IL-6:

Interleukin 6

BDNF:

Brain-derived neurotrophic factor

AD:

Alzheimer's disease

PPARγ:

Peroxisome proliferator-activated receptor gamma

PGC1α:

Peroxisome proliferator-activated receptor gamma coactivator 1 alpha

PD:

Parkinson's disease

TNF-α:

Tumor necrosis factor alpha

CNS:

Central nervous system

SCFAs:

Short-chain fatty acids

BBB:

Blood-brain barrier

miRNAs:

MicroRNAs

Gpld1:

Glycosylphosphatidylinositol specific phospholipase D1

LPS:

Lipopolysaccharide

AMPK:

AMP-activated protein kinase

AICAR:

5-Aminoimidazole-4-carboxamide ribonucleotide

VEGF:

Vascular endothelial growth factor

References

  1. Hou Y, Dan X, Babbar M, Wei Y, Hasselbalch SG, Croteau DL, et al. Ageing as a risk factor for neurodegenerative disease. Nat Rev Neurol. 2019;15:565–81.

    Article  PubMed  Google Scholar 

  2. Alzheimer’s disease facts and figures. Alzheimers Dement. 2023;19:1598–695.

    Article  Google Scholar 

  3. Medina A, Mahjoub Y, Shaver L, Pringsheim T. Prevalence and incidence of Huntington’s disease: an updated systematic review and meta-analysis. Mov Disord. 2022;37:2327–35.

    Article  PubMed  PubMed Central  Google Scholar 

  4. Bloem BR, Okun MS, Klein C. Parkinson’s disease. Lancet. 2021;397:2284–303.

    Article  CAS  PubMed  Google Scholar 

  5. Scheltens P, De Strooper B, Kivipelto M, Holstege H, Chetelat G, Teunissen CE, et al. Alzheimer’s disease. Lancet. 2021;397:1577–90.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Alkadhi KA. Exercise as a positive modulator of brain function. Mol Neurobiol. 2018;55:3112–30.

    Article  CAS  PubMed  Google Scholar 

  7. McDonnell MN, Smith AE, Mackintosh SF. Aerobic exercise to improve cognitive function in adults with neurological disorders: a systematic review. Arch Phys Med Rehabil. 2011;92:1044–52.

    Article  PubMed  Google Scholar 

  8. Ma C-L, Ma X-T, Wang J-J, Liu H, Chen Y-F, Yang Y. Physical exercise induces hippocampal neurogenesis and prevents cognitive decline. Behav Brain Res. 2017;317:332–9.

    Article  PubMed  Google Scholar 

  9. Liu PZ, Nusslock R. Exercise-mediated neurogenesis in the hippocampus via BDNF. Front Neurosci. 2018;12:52.

    Article  PubMed  PubMed Central  Google Scholar 

  10. van Praag H, Shubert T, Zhao C, Gage FH. Exercise enhances learning and hippocampal neurogenesis in aged mice. J Neurosci. 2005;25:8680–5.

    Article  PubMed  PubMed Central  Google Scholar 

  11. Altman J, Das GD. Post-natal origin of microneurones in the rat brain. Nature. 1965;207:953–6.

    Article  CAS  PubMed  Google Scholar 

  12. Moreno-Jimenez EP, Flor-Garcia M, Terreros-Roncal J, Rabano A, Cafini F, Pallas-Bazarra N, et al. Adult hippocampal neurogenesis is abundant in neurologically healthy subjects and drops sharply in patients with Alzheimer’s disease. Nat Med. 2019;25:554–60.

    Article  CAS  PubMed  Google Scholar 

  13. Terreros-Roncal J, Moreno-Jiménez EP, Flor-García M, Rodríguez-Moreno CB, Trinchero MF, Cafini F, et al. Impact of neurodegenerative diseases on human adult hippocampal neurogenesis. Science. 2021;374:1106–13.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Boldrini M, Fulmore CA, Tartt AN, Simeon LR, Pavlova I, Poposka V, et al. Human hippocampal neurogenesis persists throughout aging. Cell Stem Cell. 2018;22(589–99): e5.

    Google Scholar 

  15. Flor-Garcia M, Terreros-Roncal J, Moreno-Jimenez EP, Avila J, Rabano A, Llorens-Martin M. Unraveling human adult hippocampal neurogenesis. Nat Protoc. 2020;15:668–93.

    Article  CAS  PubMed  Google Scholar 

  16. Zhou Y, Su Y, Li S, Kennedy BC, Zhang DY, Bond AM, et al. Molecular landscapes of human hippocampal immature neurons across lifespan. Nature. 2022;607:527–33.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Wang W, Wang M, Yang M, Zeng B, Qiu W, Ma Q, et al. Transcriptome dynamics of hippocampal neurogenesis in macaques across the lifespan and aged humans. Cell Res. 2022;32:729–43.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Abrous DN, Koehl M, Lemoine M. A Baldwin interpretation of adult hippocampal neurogenesis: from functional relevance to physiopathology. Mol Psychiatry. 2021;27:383–402.

    Article  PubMed  PubMed Central  Google Scholar 

  19. Welberg L. Adult neurogenesis is altered in neurodegenerative disease. Nat Neurosci. 2021;24:1640.

    Article  CAS  PubMed  Google Scholar 

  20. Liu H, Zhang H, Ma Y. Molecular mechanisms of altered adult hippocampal neurogenesis in Alzheimer's disease. Mech Ageing Dev. 2021;195:111452.

  21. Babcock KR, Page JS, Fallon JR, Webb AE. Adult hippocampal neurogenesis in aging and Alzheimer’s disease. Stem Cell Reports. 2021;16:681–93.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Criado-Marrero M, Sabbagh JJ, Jones MR, Chaput D, Dickey CA, Blair LJ. Hippocampal neurogenesis is enhanced in adult tau deficient mice. Cells. 2020;9:210.

  23. Yu H, Zhang C, Xia J, Xu B. Treadmill exercise ameliorates adult hippocampal neurogenesis possibly by adjusting the APP proteolytic pathway in APP/PS1 transgenic mice. Int J Mol Sci. 2021;22:9570.

  24. Wang R, Holsinger RMD. Exercise-induced brain-derived neurotrophic factor expression: Therapeutic implications for Alzheimer’s dementia. Ageing Res Rev. 2018;48:109–21.

    Article  PubMed  Google Scholar 

  25. Mu Y, Gage FH. Adult hippocampal neurogenesis and its role in Alzheimer’s disease. Mol Neurodegener. 2011;6:85.

    Article  PubMed  PubMed Central  Google Scholar 

  26. Marxreiter F, Regensburger M, Winkler J. Adult neurogenesis in Parkinson’s disease. Cell Mol Life Sci. 2012;70:459–73.

    Article  PubMed  PubMed Central  Google Scholar 

  27. Lim J, Bang Y, Choi HJ. Abnormal hippocampal neurogenesis in Parkinson’s disease: relevance to a new therapeutic target for depression with Parkinson’s disease. Arch Pharm Res. 2018;41:943–54.

    Article  CAS  PubMed  Google Scholar 

  28. Jellinger KA. The pathobiology of depression in Huntington's disease: an unresolved puzzle. J Neural Transm (Vienna). 2024. https://doi.org/10.1007/s00702-024-02750-w.

  29. Ransome MI, Renoir T, Hannan AJ. Hippocampal neurogenesis, cognitive deficits and affective disorder in Huntington’s disease. Neural Plast. 2012;2012: 874387.

    Article  PubMed  PubMed Central  Google Scholar 

  30. Ahmad MH, Rizvi MA, Ali M, Mondal AC. Neurobiology of depression in Parkinson’s disease: Insights into epidemiology, molecular mechanisms and treatment strategies. Ageing Res Rev. 2023;85: 101840.

    Article  CAS  PubMed  Google Scholar 

  31. Grote HE, Bull ND, Howard ML, van Dellen A, Blakemore C, Bartlett PF, et al. Cognitive disorders and neurogenesis deficits in Huntington’s disease mice are rescued by fluoxetine. Eur J Neurosci. 2005;22:2081–8.

    Article  PubMed  Google Scholar 

  32. Gil-Mohapel J, Simpson JM, Ghilan M, Christie BR. Neurogenesis in Huntington’s disease: can studying adult neurogenesis lead to the development of new therapeutic strategies? Brain Res. 2011;1406:84–105.

    Article  CAS  PubMed  Google Scholar 

  33. Ben-Zeev T, Shoenfeld Y, Hoffman JR. The effect of exercise on neurogenesis in the brain. Isr Med Assoc J. 2022;24:533–8.

    PubMed  Google Scholar 

  34. De Miguel Z, Khoury N, Betley MJ, Lehallier B, Willoughby D, Olsson N, et al. Exercise plasma boosts memory and dampens brain inflammation via clusterin. Nature. 2021;600:494–9.

    Article  PubMed  PubMed Central  Google Scholar 

  35. Reddy I, Yadav Y, Dey CS. Cellular and molecular regulation of exercise-a neuronal perspective. Cell Mol Neurobiol. 2022;43:1551–71.

  36. Choi SH, Bylykbashi E, Chatila ZK, Lee SW, Pulli B, Clemenson GD, et al. Combined adult neurogenesis and BDNF mimic exercise effects on cognition in an Alzheimer's mouse model. Science. 2018;361:eaan8821.

  37. Horowitz AM, Fan X, Bieri G, Smith LK, Sanchez-Diaz CI, Schroer AB, et al. Blood factors transfer beneficial effects of exercise on neurogenesis and cognition to the aged brain. Science. 2020;369:167–73.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Gubert C, Hannan AJ. Exercise mimetics: harnessing the therapeutic effects of physical activity. Nat Rev Drug Discov. 2021;20:862–79.

    Article  CAS  PubMed  Google Scholar 

  39. Jang YJ, Byun S. Molecular targets of exercise mimetics and their natural activators. BMB Rep. 2021;54:581–91.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Akers KG, Martinez-Canabal A, Restivo L, Yiu AP, De Cristofaro A, Hsiang HL, et al. Hippocampal neurogenesis regulates forgetting during adulthood and infancy. Science. 2014;344:598–602.

    Article  CAS  PubMed  Google Scholar 

  41. Enwere E, Shingo T, Gregg C, Fujikawa H, Ohta S, Weiss S. Aging results in reduced epidermal growth factor receptor signaling, diminished olfactory neurogenesis, and deficits in fine olfactory discrimination. J Neurosci. 2004;24:8354–65.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Barrientos RM, Kitt MM, Watkins LR, Maier SF. Neuroinflammation in the normal aging hippocampus. Neuroscience. 2015;309:84–99.

    Article  CAS  PubMed  Google Scholar 

  43. Villeda SA, Plambeck KE, Middeldorp J, Castellano JM, Mosher KI, Luo J, et al. Young blood reverses age-related impairments in cognitive function and synaptic plasticity in mice. Nat Med. 2014;20:659–63.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Audesse AJ, Webb AE. Mechanisms of enhanced quiescence in neural stem cell aging. Mech Ageing Dev. 2020;191: 111323.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Kwon HS, Koh SH. Neuroinflammation in neurodegenerative disorders: the roles of microglia and astrocytes. Can J Appl Sport Sci. 2020;9:42.

    Google Scholar 

  46. Covacu R, Brundin L. Effects of neuroinflammation on neural stem cells. Neuroscientist. 2017;23:27–39.

    Article  CAS  PubMed  Google Scholar 

  47. Amanollahi M, Jameie M, Heidari A, Rezaei N. The dialogue between neuroinflammation and adult neurogenesis: mechanisms involved and alterations in neurological diseases. Mol Neurobiol. 2022;60:923–59.

    Article  PubMed  Google Scholar 

  48. Gil JM, Mohapel P, Araujo IM, Popovic N, Li JY, Brundin P, et al. Reduced hippocampal neurogenesis in R6/2 transgenic Huntington’s disease mice. Neurobiol Dis. 2005;20:744–51.

    Article  CAS  PubMed  Google Scholar 

  49. van Praag H, Schinder AF, Christie BR, Toni N, Palmer TD, Gage FH. Functional neurogenesis in the adult hippocampus. Nature. 2002;415:1030–4.

    Article  PubMed  PubMed Central  Google Scholar 

  50. Dhaliwal J, Lagace DC. Visualization and genetic manipulation of adult neurogenesis using transgenic mice. Eur J Neurosci. 2011;33:1025–36.

    Article  PubMed  Google Scholar 

  51. Guo NN, Sahay A. Neural circuits serve as periscopes for NSCs. Cell Stem Cell. 2017;21:557–9.

    Article  CAS  PubMed  Google Scholar 

  52. Paul A, Chaker Z, Doetsch F. Hypothalamic regulation of regionally distinct adult neural stem cells and neurogenesis. Science. 2017;356:1383–6.

    Article  CAS  PubMed  Google Scholar 

  53. Furutachi S, Matsumoto A, Nakayama KI, Gotoh Y. p57 controls adult neural stem cell quiescence and modulates the pace of lifelong neurogenesis. EMBO J. 2013;32:970–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Bouchard-Cannon P, Lowden C, Trinh D, Cheng H-YM. Dexras1 is a homeostatic regulator of exercise-dependent proliferation and cell survival in the hippocampal neurogenic niche. Sci Rep. 2018;8:5294.

  55. Sierra A, Encinas JM, Deudero JJP, Chancey JH, Enikolopov G, Overstreet-Wadiche LS, et al. Microglia shape adult hippocampal neurogenesis through apoptosis-coupled phagocytosis. Cell Stem Cell. 2010;7:483–95.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Tashiro A, Sandler VM, Toni N, Zhao C, Gage FH. NMDA-receptor-mediated, cell-specific integration of new neurons in adult dentate gyrus. Nature. 2006;442:929–33.

    Article  CAS  PubMed  Google Scholar 

  57. Adlaf EW, Vaden RJ, Niver AJ, Manuel AF, Onyilo VC, Araujo MT, et al. Adult-born neurons modify excitatory synaptic transmission to existing neurons. Elife. 2017;6:e19886.

  58. Krzisch M, Fülling C, Jabinet L, Armida J, Gebara E, Cassé F, et al. Synaptic adhesion molecules regulate the integration of new granule neurons in the postnatal mouse hippocampus and their impact on spatial memory. Cereb Cortex. 2016;27:4048–59.

  59. McAvoy KM, Scobie KN, Berger S, Russo C, Guo N, Decharatanachart P, et al. Modulating neuronal competition dynamics in the dentate gyrus to rejuvenate aging memory circuits. Neuron. 2016;91:1356–73.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Clelland CD, Choi M, Romberg C, Clemenson GD Jr, Fragniere A, Tyers P, et al. A functional role for adult hippocampal neurogenesis in spatial pattern separation. Science. 2009;325:210–3.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Pan YW, Chan GC, Kuo CT, Storm DR, Xia Z. Inhibition of adult neurogenesis by inducible and targeted deletion of ERK5 mitogen-activated protein kinase specifically in adult neurogenic regions impairs contextual fear extinction and remote fear memory. J Neurosci. 2012;32:6444–55.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Zhang J, Ji F, Liu Y, Lei X, Li H, Ji G, et al. Ezh2 regulates adult hippocampal neurogenesis and memory. J Neurosci. 2014;34:5184–99.

    Article  PubMed  PubMed Central  Google Scholar 

  63. Arruda-Carvalho M, Sakaguchi M, Akers KG, Josselyn SA, Frankland PW. Posttraining ablation of adult-generated neurons degrades previously acquired memories. J Neurosci. 2011;31:15113–27.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Wang W, Pan YW, Zou J, Li T, Abel GM, Palmiter RD, et al. Genetic activation of ERK5 MAP kinase enhances adult neurogenesis and extends hippocampus-dependent long-term memory. J Neurosci. 2014;34:2130–47.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Frankland PW, Kohler S, Josselyn SA. Hippocampal neurogenesis and forgetting. Trends Neurosci. 2013;36:497–503.

    Article  CAS  PubMed  Google Scholar 

  66. Eriksson PS, Perfilieva E, Björk-Eriksson T, Alborn A-M, Nordborg C, Peterson DA, et al. Neurogenesis in the adult human hippocampus. Nat Med. 1998;4:1313–7.

    Article  CAS  PubMed  Google Scholar 

  67. Spalding Kirsty L, Bergmann O, Alkass K, Bernard S, Salehpour M, Huttner Hagen B, et al. Dynamics of hippocampal neurogenesis in adult humans. Cell. 2013;153:1219–27.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Boldrini M, Underwood MD, Hen R, Rosoklija GB, Dwork AJ, John Mann J, et al. Antidepressants increase neural progenitor cells in the human hippocampus. Neuropsychopharmacology. 2009;34:2376–89.

    Article  CAS  PubMed  Google Scholar 

  69. Crews L, Adame A, Patrick C, Delaney A, Pham E, Rockenstein E, et al. Increased BMP6 levels in the brains of Alzheimer’s disease patients and APP transgenic mice are accompanied by impaired neurogenesis. J Neurosci. 2010;30:12252–62.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Knoth R, Singec I, Ditter M, Pantazis G, Capetian P, Meyer RP, et al. Murine features of neurogenesis in the human hippocampus across the lifespan from 0 to 100 years. PLoS ONE. 2010;5: e8809.

    Article  PubMed  PubMed Central  Google Scholar 

  71. Ernst A, Alkass K, Bernard S, Salehpour M, Perl S, Tisdale J, et al. Neurogenesis in the striatum of the adult human brain. Cell. 2014;156:1072–83.

    Article  CAS  PubMed  Google Scholar 

  72. Palmer TD, Schwartz PH, Taupin P, Kaspar B, Stein SA, Gage FH. Cell culture. Progenitor cells from human brain after death. Nature. 2001;411:42–3.

    Article  CAS  PubMed  Google Scholar 

  73. Dennis CV, Suh LS, Rodriguez ML, Kril JJ, Sutherland GT. Human adult neurogenesis across the ages: An immunohistochemical study. Neuropathol Appl Neurobiol. 2016;42:621–38.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Liu YWJ, Curtis MA, Gibbons HM, Mee EW, Bergin PS, Teoh HH, et al. Doublecortin expression in the normal and epileptic adult human brain. Eur J Neurosci. 2008;28:2254–65.

    Article  CAS  PubMed  Google Scholar 

  75. Mathews KJ, Allen KM, Boerrigter D, Ball H, Shannon Weickert C, Double KL. Evidence for reduced neurogenesis in the aging human hippocampus despite stable stem cell markers. Aging Cell. 2017;16:1195–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Cipriani S, Ferrer I, Aronica E, Kovacs GG, Verney C, Nardelli J, et al. Hippocampal radial glial subtypes and their neurogenic potential in human fetuses and healthy and Alzheimer’s disease adults. Cereb Cortex. 2018;28:2458–78.

    Article  PubMed  Google Scholar 

  77. Sorrells SF, Paredes MF, Cebrian-Silla A, Sandoval K, Qi D, Kelley KW, et al. Human hippocampal neurogenesis drops sharply in children to undetectable levels in adults. Nature. 2018;555:377–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Seki T, Hori T, Miyata H, Maehara M, Namba T. Analysis of proliferating neuronal progenitors and immature neurons in the human hippocampus surgically removed from control and epileptic patients. Sci Rep. 2019;9:18194.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Habib N, Avraham-Davidi I, Basu A, Burks T, Shekhar K, Hofree M, et al. Massively parallel single-nucleus RNA-seq with DroNc-seq. Nat Methods. 2017;14:955–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Ayhan F, Kulkarni A, Berto S, Sivaprakasam K, Douglas C, Lega BC, et al. Resolving cellular and molecular diversity along the hippocampal anterior-to-posterior axis in humans. Neuron. 2021;109(2091–105): e6.

    Google Scholar 

  81. Franjic D, Skarica M, Ma S, Arellano JI, Tebbenkamp ATN, Choi J, et al. Transcriptomic taxonomy and neurogenic trajectories of adult human, macaque, and pig hippocampal and entorhinal cells. Neuron. 2022;110:452-69.e14.

    Article  CAS  PubMed  Google Scholar 

  82. Lovell MA, Geiger H, Van Zant GE, Lynn BC, Markesbery WR. Isolation of neural precursor cells from Alzheimer’s disease and aged control postmortem brain. Neurobiol Aging. 2006;27:909–17.

    Article  PubMed  Google Scholar 

  83. Oddo S, Caccamo A, Shepherd JD, Murphy MP, Golde TE, Kayed R, et al. Triple-transgenic model of Alzheimer’s disease with plaques and tangles. Neuron. 2003;39:409–21.

    Article  CAS  PubMed  Google Scholar 

  84. Hamilton LK, Aumont A, Julien C, Vadnais A, Calon F, Fernandes KJ. Widespread deficits in adult neurogenesis precede plaque and tangle formation in the 3xTg mouse model of Alzheimer’s disease. Eur J Neurosci. 2010;32:905–20.

    Article  PubMed  Google Scholar 

  85. Gwinn K, Rodríguez JJ, Jones VC, Tabuchi M, Allan SM, Knight EM, et al. Impaired adult neurogenesis in the dentate gyrus of a triple transgenic mouse model of Alzheimer's disease. PLoS ONE. 2008;3:e2935.

  86. Rodriguez JJ, Jones VC, Verkhratsky A. Impaired cell proliferation in the subventricular zone in an Alzheimer’s disease model. NeuroReport. 2009;20:907–12.

    Article  CAS  PubMed  Google Scholar 

  87. Zheng J, Li HL, Tian N, Liu F, Wang L, Yin Y, et al. Interneuron accumulation of phosphorylated tau impairs adult hippocampal neurogenesis by suppressing GABAergic transmission. Cell Stem Cell. 2020;26:462–6.

    Article  CAS  PubMed  Google Scholar 

  88. Hoglinger GU, Rizk P, Muriel MP, Duyckaerts C, Oertel WH, Caille I, et al. Dopamine depletion impairs precursor cell proliferation in Parkinson disease. Nat Neurosci. 2004;7:726–35.

    Article  PubMed  Google Scholar 

  89. Villar-Piqué A, Lopes da Fonseca T, Outeiro TF. Structure, function and toxicity of alpha-synuclein: the Bermuda triangle in synucleinopathies. J Neurochem. 2016;139:240–55.

    Article  PubMed  Google Scholar 

  90. Winner B, Regensburger M, Schreglmann S, Boyer L, Prots I, Rockenstein E, et al. Role of alpha-synuclein in adult neurogenesis and neuronal maturation in the dentate gyrus. J Neurosci. 2012;32:16906–16.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Winner B, Lie DC, Rockenstein E, Aigner R, Aigner L, Masliah E, et al. Human wild-type alpha-synuclein impairs neurogenesis. J Neuropathol Exp Neurol. 2004;63:1155–66.

    Article  CAS  PubMed  Google Scholar 

  92. Winner B, Kohl Z, Gage FH. Neurodegenerative disease and adult neurogenesis. Eur J Neurosci. 2011;33:1139–51.

    Article  PubMed  Google Scholar 

  93. Hashimoto M, Rockenstein E, Masliah E. Transgenic models of alpha-synuclein pathology: past, present, and future. Ann N Y Acad Sci. 2003;991:171–88.

    Article  CAS  PubMed  Google Scholar 

  94. Crews L, Mizuno H, Desplats P, Rockenstein E, Adame A, Patrick C, et al. Alpha-synuclein alters Notch-1 expression and neurogenesis in mouse embryonic stem cells and in the hippocampus of transgenic mice. J Neurosci. 2008;28:4250–60.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Kohl Z, Winner B, Ubhi K, Rockenstein E, Mante M, Munch M, et al. Fluoxetine rescues impaired hippocampal neurogenesis in a transgenic A53T synuclein mouse model. Eur J Neurosci. 2012;35:10–9.

    Article  PubMed  PubMed Central  Google Scholar 

  96. Lazic SE, Grote H, Armstrong RJ, Blakemore C, Hannan AJ, van Dellen A, et al. Decreased hippocampal cell proliferation in R6/1 Huntington’s mice. NeuroReport. 2004;15:811–3.

    Article  PubMed  Google Scholar 

  97. Cremer H, Chazal G, Lledo PM, Rougon G, Montaron MF, Mayo W, et al. PSA-NCAM: an important regulator of hippocampal plasticity. Int J Dev Neurosci. 2000;18:213–20.

    Article  CAS  PubMed  Google Scholar 

  98. Kiss JZ, Troncoso E, Djebbara Z, Vutskits L, Muller D. The role of neural cell adhesion molecules in plasticity and repair. Brain Res Brain Res Rev. 2001;36:175–84.

    Article  CAS  PubMed  Google Scholar 

  99. Simpson JM, Gil-Mohapel J, Pouladi MA, Ghilan M, Xie Y, Hayden MR, et al. Altered adult hippocampal neurogenesis in the YAC128 transgenic mouse model of Huntington disease. Neurobiol Dis. 2011;41:249–60.

    Article  CAS  PubMed  Google Scholar 

  100. Armstrong MJ, Okun MS. Diagnosis and treatment of Parkinson disease. JAMA. 2020;323:548–60.

  101. Fyfe I. Neurogenesis altered in multiple neurodegenerative diseases. Nat Rev Neurol. 2021;17:726.

    PubMed  Google Scholar 

  102. Wang Z, van Praag H. Exercise and the Brain: Neurogenesis, Synaptic Plasticity, Spine Density, and Angiogenesis. In: Boecker H, Hillman CH, Scheef L, editors. Functional Neuroimaging in Exercise and Sport Sciences. New York: Springer; 2012. p. 3–24.

  103. Ahlskog JE, Geda YE, Graff-Radford NR, Petersen RC. Physical exercise as a preventive or disease-modifying treatment of dementia and brain aging. Mayo Clin Proc. 2011;86:876–84.

    Article  PubMed  PubMed Central  Google Scholar 

  104. Livingston G, Huntley J, Sommerlad A, Ames D, Ballard C, Banerjee S, et al. Dementia prevention, intervention, and care: 2020 report of the Lancet Commission. Lancet. 2020;396:413–46.

    Article  PubMed  PubMed Central  Google Scholar 

  105. Hamer M, Chida Y. Physical activity and risk of neurodegenerative disease: a systematic review of prospective evidence. Psychol Med. 2009;39:3–11.

    Article  CAS  PubMed  Google Scholar 

  106. Buchman AS, Boyle PA, Yu L, Shah RC, Wilson RS, Bennett DA. Total daily physical activity and the risk of AD and cognitive decline in older adults. Neurology. 2012;78:1323–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  107. Fang X, Han D, Cheng Q, Zhang P, Zhao C, Min J, et al. Association of levels of physical activity with risk of Parkinson disease. JAMA Network Open. 2018;1.

  108. Ernst M, Folkerts A-K, Gollan R, Lieker E, Caro-Valenzuela J, Adams A, et al. Physical exercise for people with Parkinson’s disease: a systematic review and network meta-analysis. Cochrane Database Syst Rev. 2023;1:CD013856.

  109. Mak MKY, Wong-Yu ISK. Exercise for Parkinson’s disease. Int Rev Neurobiol. 2019;147:1–44.

    Article  CAS  PubMed  Google Scholar 

  110. Paillard T, Rolland Y, de Souto BP. Protective effects of physical exercise in Alzheimer’s disease and Parkinson’s disease: a narrative review. J Clin Neurol. 2015;11:212–9.

    Article  PubMed  PubMed Central  Google Scholar 

  111. Chen F-T, Hopman RJ, Huang C-J, Chu C-H, Hillman CH, Hung T-M, et al. The effect of exercise training on brain structure and function in older adults: a systematic review based on evidence from randomized control trials. J Clin Med. 2020;9:914.

  112. Augusto-Oliveira M, Arrifano GP, Leal-Nazaré CG, Santos-Sacramento L, Lopes-Araújo A, Royes LFF, et al. Exercise reshapes the brain: molecular, cellular, and structural changes associated with cognitive improvements. Mol Neurobiol. 2023;60:6950–74.

  113. Cabral DF, Rice J, Morris TP, Rundek T, Pascual-Leone A, Gomes-Osman J. Exercise for brain health: an investigation into the underlying mechanisms guided by dose. Neurotherapeutics. 2019;16:580–99.

    Article  PubMed  PubMed Central  Google Scholar 

  114. Nay K, Smiles WJ, Kaiser J, McAloon LM, Loh K, Galic S, et al. Molecular mechanisms underlying the beneficial effects of exercise on brain function and neurological disorders. Int J Mol Sci. 2021;22:4052.

  115. Ma Q. Beneficial effects of moderate voluntary physical exercise and its biological mechanisms on brain health. Neurosci Bull. 2008;24:265–70.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  116. Cotman CW, Berchtold NC, Christie LA. Exercise builds brain health: key roles of growth factor cascades and inflammation. Trends Neurosci. 2007;30:464–72.

    Article  CAS  PubMed  Google Scholar 

  117. Bhattacharya P, Chatterjee S, Roy D. Impact of exercise on brain neurochemicals: a comprehensive review. Sport Sci Health. 2023;19:405–52.

    Article  Google Scholar 

  118. Kondo M. Molecular mechanisms of experience-dependent structural and functional plasticity in the brain. Anat Sci Int. 2016;92:1–17.

    Article  PubMed  Google Scholar 

  119. Wang M, Zhang H, Liang J, Huang J, Chen N. Exercise suppresses neuroinflammation for alleviating Alzheimer’s disease. J Neuroinflammation. 2023;20:76.

    Article  PubMed  PubMed Central  Google Scholar 

  120. Mee-Inta O, Zhao ZW, Kuo YM. Physical exercise inhibits inflammation and microglial activation. Cells. 2019;8:691.

  121. Pereira AC, Huddleston DE, Brickman AM, Sosunov AA, Hen R, McKhann GM, et al. An in vivo correlate of exercise-induced neurogenesis in the adult dentate gyrus. Proc Natl Acad Sci U S A. 2007;104:5638–43.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  122. van Praag H, Christie BR, Sejnowski TJ, Gage FH. Running enhances neurogenesis, learning, and long-term potentiation in mice. Proc Natl Acad Sci U S A. 1999;96:13427–31.

    Article  PubMed  PubMed Central  Google Scholar 

  123. Ridler C. Exercise wards off Alzheimer disease by boosting neurogenesis and neuroprotective factors. Nat Rev Neurol. 2018;14:632.

    CAS  PubMed  Google Scholar 

  124. Huang YQ, Wu C, He XF, Wu D, He X, Liang FY, et al. Effects of voluntary wheel-running types on hippocampal neurogenesis and spatial cognition in middle-aged mice. Front Cell Neurosci. 2018;12:177.

    Article  PubMed  PubMed Central  Google Scholar 

  125. van Praag H, Kempermann G, Gage FH. Running increases cell proliferation and neurogenesis in the adult mouse dentate gyrus. Nat Neurosci. 1999;2:266–70.

    Article  PubMed  Google Scholar 

  126. Uda M, Ishido M, Kami K, Masuhara M. Effects of chronic treadmill running on neurogenesis in the dentate gyrus of the hippocampus of adult rat. Brain Res. 2006;1104:64–72.

    Article  CAS  PubMed  Google Scholar 

  127. Park H-S, Kim T-W, Park S-S, Lee S-J. Swimming exercise ameliorates mood disorder and memory impairment by enhancing neurogenesis, serotonin expression, and inhibiting apoptosis in social isolation rats during adolescence. J Exerc Rehabil. 2020;16:132–40.

    Article  PubMed  PubMed Central  Google Scholar 

  128. Yau S-Y, Lau BW-M, So K-F. Adult hippocampal neurogenesis: a possible way how physical exercise counteracts stress. Cell Transplant. 2011;20:99–111.

    Article  PubMed  Google Scholar 

  129. Lucassen PJ, Oomen CA. Stress, hippocampal neurogenesis and cognition: functional correlations. Front Biol. 2016;11:182–92.

    Article  CAS  Google Scholar 

  130. Lu Y, Bu FQ, Wang F, Liu L, Zhang S, Wang G, et al. Recent advances on the molecular mechanisms of exercise-induced improvements of cognitive dysfunction. Transl Neurodegener. 2023;12:9.

    Article  PubMed  PubMed Central  Google Scholar 

  131. Connolly MG, Bruce SR, Kohman RA. Exercise duration differentially effects age-related neuroinflammation and hippocampal neurogenesis. Neuroscience. 2022;490:275–86.

    Article  CAS  PubMed  Google Scholar 

  132. Scheffer DDL, Latini A. Exercise-induced immune system response: Anti-inflammatory status on peripheral and central organs. Biochim Biophys Acta Mol Basis Dis. 2020;1866: 165823.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  133. Knaepen K, Goekint M, Heyman EM, Meeusen R. Neuroplasticity - exercise-induced response of peripheral brain-derived neurotrophic factor: a systematic review of experimental studies in human subjects. Sports Med. 2010;40:765–801.

    Article  PubMed  Google Scholar 

  134. Pedersen BK. Physical activity and muscle-brain crosstalk. Nat Rev Endocrinol. 2019;15:383–92.

    Article  PubMed  Google Scholar 

  135. Wang J, Liu S, Li G, Xiao J. Exercise regulates the immune system. Adv Exp Med Biol. 2020;1228:395–408.

    Article  CAS  PubMed  Google Scholar 

  136. Małkiewicz MA, Szarmach A, Sabisz A, Cubała WJ, Szurowska E, Winklewski PJ. Blood-brain barrier permeability and physical exercise. J Neuroinflammation. 2019;16:15.

  137. de Sousa CV, Sales MM, Rosa TS, Lewis JE, de Andrade RV, Simões HG. The antioxidant effect of exercise: a systematic review and meta-analysis. Sport Med. 2016;47:277–93.

    Article  Google Scholar 

  138. Li G. The positive and negative aspects of reactive oxygen species in sports performance. In: Current Issues in Sports and Exercise Medicine. 2013.

    Google Scholar 

  139. Kawamura T, Muraoka I. Exercise-induced oxidative stress and the effects of antioxidant intake from a physiological viewpoint. Antioxidants. 2018;7:119.

  140. Somani SM, Husain K, Diaz-Phillips L, Lanzotti DJ, Kareti KR, Trammell GL. Interaction of exercise and ethanol on antioxidant enzymes in brain regions of the rat. Alcohol. 1996;13:603–10.

    Article  CAS  PubMed  Google Scholar 

  141. Mazzola PN, Terra M, Rosa AP, Mescka CP, Moraes TB, Piccoli B, et al. Regular exercise prevents oxidative stress in the brain of hyperphenylalaninemic rats. Metab Brain Dis. 2011;26:291–7.

    Article  CAS  PubMed  Google Scholar 

  142. Thirupathi A, Pinho RA, Ugbolue UC, He Y, Meng Y, Gu Y. Effect of running exercise on oxidative stress biomarkers: a systematic review. Front Physioly. 2021;11:610112.

  143. Ngwa JS, Nwulia E, Ntekim O, Bedada FB, Kwabi-Addo B, Nadarajah S, et al. Aerobic exercise training-induced changes on DNA methylation in mild cognitively impaired elderly African Americans: gene, exercise, and memory study - GEMS-I. Front Mol Neurosci. 2022;14:752403.

  144. Sellami M, Bragazzi N, Prince MS, Denham J, Elrayess M. Regular, intense exercise training as a healthy aging lifestyle strategy: preventing DNA damage, telomere shortening and adverse DNA methylation changes over a lifetime. Front Genet. 2021;12:652497.

  145. Tomiga Y, Sakai K, Ra SG, Kusano M, Ito A, Uehara Y, et al. Short-term running exercise alters DNA methylation patterns in neuronal nitric oxide synthase and brain-derived neurotrophic factor genes in the mouse hippocampus and reduces anxiety-like behaviors. FASEB J. 2021;35: e21767.

    Article  CAS  PubMed  Google Scholar 

  146. Chen Y, Sun Y, Luo Z, Lin J, Qi B, Kang X, et al. Potential mechanism underlying exercise upregulated circulating blood exosome miR-215-5p to prevent necroptosis of neuronal cells and a model for early diagnosis of Alzheimer’s disease. Front Aging Neurosci. 2022;14: 860364.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  147. Pons-Espinal M, Gasperini C, Marzi MJ, Braccia C, Armirotti A, Pötzsch A, et al. MiR-135a-5p is critical for exercise-induced adult neurogenesis. Stem Cell Reports. 2019;12:1298–312.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  148. Fiuza-Luces C, Santos-Lozano A, Joyner M, Carrera-Bastos P, Picazo O, Zugaza JL, et al. Exercise benefits in cardiovascular disease: beyond attenuation of traditional risk factors. Nat Rev Cardiol. 2018;15:731–43.

    Article  CAS  PubMed  Google Scholar 

  149. Gleeson M, Bishop NC, Stensel DJ, Lindley MR, Mastana SS, Nimmo MA. The anti-inflammatory effects of exercise: mechanisms and implications for the prevention and treatment of disease. Nat Rev Immunol. 2011;11:607–15.

    Article  CAS  PubMed  Google Scholar 

  150. Hillman CH, Erickson KI, Kramer AF. Be smart, exercise your heart: exercise effects on brain and cognition. Nat Rev Neurosci. 2008;9:58–65.

    Article  CAS  PubMed  Google Scholar 

  151. De la Rosa A, Olaso-Gonzalez G, Arc-Chagnaud C, Millan F, Salvador-Pascual A, García-Lucerga C, et al. Physical exercise in the prevention and treatment of Alzheimer’s disease. J Sport Health Sci. 2020;9:394–404.

    Article  PubMed  PubMed Central  Google Scholar 

  152. Murray DK, Sacheli MA, Eng JJ, Stoessl AJ. The effects of exercise on cognition in Parkinson’s disease: a systematic review. Transl Neurodegener. 2014;3:5.

    Article  PubMed  PubMed Central  Google Scholar 

  153. Guerrieri D, Moon HY, van Praag H. Exercise in a pill: the latest on exercise-mimetics. Brain Plast. 2017;2:153–69.

    Article  PubMed  PubMed Central  Google Scholar 

  154. Hunter P. Exercise in a bottle: Elucidating how exercise conveys health benefits might lead to new therapeutic options for a range of diseases from cancer to metabolic syndrome. EMBO Rep. 2016;17:136–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  155. Parrini M, Ghezzi D, Deidda G, Medrihan L, Castroflorio E, Alberti M, et al. Aerobic exercise and a BDNF-mimetic therapy rescue learning and memory in a mouse model of Down syndrome. Sci Rep. 2017;7:16825.

    Article  PubMed  PubMed Central  Google Scholar 

  156. Reichardt LF. Neurotrophin-regulated signalling pathways. Philos Trans R Soc Lond B Biol Sci. 2006;361:1545–64.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  157. Pan W, Banks WA, Fasold MB, Bluth J, Kastin AJ. Transport of brain-derived neurotrophic factor across the blood-brain barrier. Neuropharmacology. 1998;37:1553–61.

    Article  CAS  PubMed  Google Scholar 

  158. Rahmani F, Saghazadeh A, Rahmani M, Teixeira AL, Rezaei N, Aghamollaii V, et al. Plasma levels of brain-derived neurotrophic factor in patients with Parkinson disease: A systematic review and meta-analysis. Brain Res. 2019;1704:127–36.

    Article  CAS  PubMed  Google Scholar 

  159. Qin XY, Cao C, Cawley NX, Liu TT, Yuan J, Loh YP, et al. Decreased peripheral brain-derived neurotrophic factor levels in Alzheimer’s disease: a meta-analysis study (N=7277). Mol Psychiatry. 2017;22:312–20.

    Article  CAS  PubMed  Google Scholar 

  160. Ciammola A, Sassone J, Cannella M, Calza S, Poletti B, Frati L, et al. Low brain-derived neurotrophic factor (BDNF) levels in serum of Huntington’s disease patients. Am J Med Genet B Neuropsychiatr Genet. 2007;144B:574–7.

    Article  CAS  PubMed  Google Scholar 

  161. Lu B, Nagappan G, Lu Y. BDNF and synaptic plasticity, cognitive function, and dysfunction. In: Lewin GR, Carter BD, editors. Neurotrophic Factors. Berlin, Heidelberg: Springer; 2014. p. 223–50.

  162. Tapia-Arancibia L, Aliaga E, Silhol M, Arancibia S. New insights into brain BDNF function in normal aging and Alzheimer disease. Brain Res Rev. 2008;59:201–20.

    Article  CAS  PubMed  Google Scholar 

  163. Phillips C, Baktir MA, Srivatsan M, Salehi A. Neuroprotective effects of physical activity on the brain: a closer look at trophic factor signaling. Front Cellular Neurosci. 2014;8:170.

  164. Vaynman S, Ying Z, Gomez-Pinilla F. Hippocampal BDNF mediates the efficacy of exercise on synaptic plasticity and cognition. Eur J Neurosci. 2004;20:2580–90.

    Article  PubMed  Google Scholar 

  165. Ieraci A, Mallei A, Musazzi L, Popoli M. Physical exercise and acute restraint stress differentially modulate hippocampal brain-derived neurotrophic factor transcripts and epigenetic mechanisms in mice. Hippocampus. 2015;25:1380–92.

    Article  CAS  PubMed  Google Scholar 

  166. Squinto SP, Stitt TN, Aldrich TH, Davis S, Bianco SM, Radziejewski C, et al. trkB encodes a functional receptor for brain-derived neurotrophic factor and neurotrophin-3 but not nerve growth factor. Cell. 1991;65:885–93.

    Article  CAS  PubMed  Google Scholar 

  167. Pang TYC, Stam NC, Nithianantharajah J, Howard ML, Hannan AJ. Differential effects of voluntary physical exercise on behavioral and brain-derived neurotrophic factor expression deficits in huntington’s disease transgenic mice. Neuroscience. 2006;141:569–84.

    Article  CAS  PubMed  Google Scholar 

  168. Ieraci A, Madaio AI, Mallei A, Lee FS, Popoli M. Brain-derived neurotrophic factor Val66Met human polymorphism impairs the beneficial exercise-induced neurobiological changes in mice. Neuropsychopharmacology. 2016;41:3070–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  169. Gao L, Tian M, Zhao HY, Xu QQ, Huang YM, Si QC, et al. TrkB activation by 7, 8-dihydroxyflavone increases synapse AMPA subunits and ameliorates spatial memory deficits in a mouse model of Alzheimer’s disease. J Neurochem. 2016;136:620–36.

    Article  CAS  PubMed  Google Scholar 

  170. Proctor DT, Coulson EJ, Dodd PR. Reduction in post-synaptic scaffolding PSD-95 and SAP-102 protein levels in the Alzheimer inferior temporal cortex is correlated with disease pathology. J Alzheimers Dis. 2010;21:795–811.

    Article  CAS  PubMed  Google Scholar 

  171. Hamos JE, DeGennaro LJ, Drachman DA. Synaptic loss in Alzheimer’s disease and other dementias. Neurology. 1989;39:355–61.

    Article  CAS  PubMed  Google Scholar 

  172. Pietrelli A, Matkovic L, Vacotto M, Lopez-Costa JJ, Basso N, Brusco A. Aerobic exercise upregulates the BDNF-Serotonin systems and improves the cognitive function in rats. Neurobiol Learn Mem. 2018;155:528–42.

    Article  CAS  PubMed  Google Scholar 

  173. Alomari MA, Khabour OF, Alzoubi KH, Alzubi MA. Forced and voluntary exercises equally improve spatial learning and memory and hippocampal BDNF levels. Behav Brain Res. 2013;247:34–9.

    Article  CAS  PubMed  Google Scholar 

  174. Seo JH, Park HS, Park SS, Kim CJ, Kim DH, Kim TW. Physical exercise ameliorates psychiatric disorders and cognitive dysfunctions by hippocampal mitochondrial function and neuroplasticity in post-traumatic stress disorder. Exp Neurol. 2019;322: 113043.

    Article  CAS  PubMed  Google Scholar 

  175. Maak S, Norheim F, Drevon CA, Erickson HP. Progress and challenges in the biology of FNDC5 and Irisin. Endocr Rev. 2021;42:436–56.

  176. Zhao R. Irisin at the crossroads of inter-organ communications: Challenge and implications. Front Endocrinol. 2022;13:989135.

  177. Bostrom P, Wu J, Jedrychowski MP, Korde A, Ye L, Lo JC, et al. A PGC1-alpha-dependent myokine that drives brown-fat-like development of white fat and thermogenesis. Nature. 2012;481:463–8.

    Article  PubMed  PubMed Central  Google Scholar 

  178. Lourenco MV, Ribeiro FC, Sudo FK, Drummond C, Assunção N, Vanderborght B, et al. Cerebrospinal fluid irisin correlates with amyloid-β, BDNF, and cognition in Alzheimer’s disease. Alzheimers Dement (Amst). 2020;12: e12034.

    PubMed  Google Scholar 

  179. Zhang F, Hou G, Hou G, Wang C, Shi B, Zheng Y. Serum irisin as a potential biomarker for cognitive decline in vascular dementia. Front Neurol. 2021;12:755046.

  180. Tsai C-L, Pai M-C. Circulating levels of Irisin in obese individuals at genetic risk for Alzheimer’s disease: Correlations with amyloid-β, metabolic, and neurocognitive indices. Behav Brain Res. 2021;400:113013.

  181. Zhang X, Xu S, Hu Y, Liu Q, Liu C, Chai H, et al. Irisin exhibits neuroprotection by preventing mitochondrial damage in Parkinson’s disease. NPJ Parkinsons Dis. 2023;9:13.

    Article  PubMed  PubMed Central  Google Scholar 

  182. Chen K, Wang K, Wang T. Protective effect of irisin against Alzheimer’s disease. Front Psychiatry. 2022;13: 967683.

    Article  PubMed  PubMed Central  Google Scholar 

  183. Kim OY, Song J. The role of irisin in Alzheimer's disease. J Clin Med. 2018;7:407.

  184. Noda Y, Kuzuya A, Tanigawa K, Araki M, Kawai R, Ma B, et al. Fibronectin type III domain-containing protein 5 interacts with APP and decreases amyloid β production in Alzheimer’s disease. Mol Brain. 2018;11:61.

  185. Peng J, Deng X, Huang W, Yu JH, Wang JX, Wang JP, et al. Irisin protects against neuronal injury induced by oxygen-glucose deprivation in part depends on the inhibition of ROS-NLRP3 inflammatory signaling pathway. Mol Immunol. 2017;91:185–94.

    Article  CAS  PubMed  Google Scholar 

  186. Pignataro P, Dicarlo M, Zerlotin R, Zecca C, Dell'Abate MT, Buccoliero C, et al. FNDC5/Irisin system in neuroinflammation and neurodegenerative diseases: update and novel perspective. Int J Mol Sci. 2021;22:1605.

  187. Wang K, Song F, Xu K, Liu Z, Han S, Li F, et al. Irisin attenuates neuroinflammation and prevents the memory and cognitive deterioration in streptozotocin-induced diabetic mice. Mediators Inflamm. 2019;2019:1567179.

    Article  PubMed  PubMed Central  Google Scholar 

  188. Wang Y, Tian M, Tan J, Pei X, Lu C, Xin Y, et al. Irisin ameliorates neuroinflammation and neuronal apoptosis through integrin alphaVbeta5/AMPK signaling pathway after intracerebral hemorrhage in mice. J Neuroinflammation. 2022;19:82.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  189. Kam T-I, Park H, Chou S-C, Van Vranken JG, Mittenbühler MJ, Kim H, et al. Amelioration of pathologic α-synuclein-induced Parkinson’s disease by irisin. Proc Natl Acad Sci USA. 2022;119:e2204835119.

  190. Bretland KA, Lin L, Bretland KM, Smith MA, Fleming SM, Dengler-Crish CM. Irisin treatment lowers levels of phosphorylated tau in the hippocampus of pre-symptomatic female but not male htau mice. Neuropathol Appl Neurobiol. 2021;47:967–78.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  191. Lourenco MV, Frozza RL, de Freitas GB, Zhang H, Kincheski GC, Ribeiro FC, et al. Exercise-linked FNDC5/irisin rescues synaptic plasticity and memory defects in Alzheimer’s models. Nat Med. 2019;25:165–75.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  192. Islam MR, Valaris S, Young MF, Haley EB, Luo R, Bond SF, et al. Exercise hormone irisin is a critical regulator of cognitive function. Nat Metab. 2021;3:1058–70.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  193. Hashemi MS, Ghaedi K, Salamian A, Karbalaie K, Emadi-Baygi M, Tanhaei S, et al. Fndc5 knockdown significantly decreased neural differentiation rate of mouse embryonic stem cells. Neuroscience. 2013;231:296–304.

    Article  CAS  PubMed  Google Scholar 

  194. Wrann CD, White JP, Salogiannnis J, Laznik-Bogoslavski D, Wu J, Ma D, et al. Exercise induces hippocampal BDNF through a PGC-1alpha/FNDC5 pathway. Cell Metab. 2013;18:649–59.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  195. Zhao R. Exercise mimetics: a novel strategy to combat neuroinflammation and Alzheimer’s disease. J Neuroinflammation. 2024;21:40.

  196. Zhao C, Deng W, Gage FH. Mechanisms and functional implications of adult neurogenesis. Cell. 2008;132:645–60.

    Article  CAS  PubMed  Google Scholar 

  197. Cryan JF, O’Riordan KJ, Cowan CSM, Sandhu KV, Bastiaanssen TFS, Boehme M, et al. The microbiota-gut-brain axis. Physiol Rev. 2019;99:1877–2013.

    Article  CAS  PubMed  Google Scholar 

  198. Unger MM, Spiegel J, Dillmann K-U, Grundmann D, Philippeit H, Bürmann J, et al. Short chain fatty acids and gut microbiota differ between patients with Parkinson’s disease and age-matched controls. Parkinsonism Relat Disord. 2016;32:66–72.

    Article  PubMed  Google Scholar 

  199. Bedarf JR, Hildebrand F, Coelho LP, Sunagawa S, Bahram M, Goeser F, et al. Functional implications of microbial and viral gut metagenome changes in early stage L-DOPA-naïve Parkinson’s disease patients. Genome Med. 2017;9:39.

  200. Keshavarzian A, Green SJ, Engen PA, Voigt RM, Naqib A, Forsyth CB, et al. Colonic bacterial composition in Parkinson’s disease. Mov Disord. 2015;30:1351–60.

    Article  CAS  PubMed  Google Scholar 

  201. Sampson TR, Debelius JW, Thron T, Janssen S, Shastri GG, Ilhan ZE, et al. Gut microbiota regulate motor deficits and neuroinflammation in a model of Parkinson’s disease. Cell. 2016;167(1469–80): e12.

    Google Scholar 

  202. D'Argenio V, Veneruso I, Gong C, Cecarini V, Bonfili L, Eleuteri AM. Gut microbiome and mycobiome alterations in an in vivo model of Alzheimer's disease. Genes (Basel). 2022;13:1564.

  203. Vogt NM, Kerby RL, Dill-McFarland KA, Harding SJ, Merluzzi AP, Johnson SC, et al. Gut microbiome alterations in Alzheimer’s disease. Sci Rep. 2017;7:13537.

    Article  PubMed  PubMed Central  Google Scholar 

  204. Naseribafrouei A, Hestad K, Avershina E, Sekelja M, Linlokken A, Wilson R, et al. Correlation between the human fecal microbiota and depression. Neurogastroenterol Motil. 2014;26:1155–62.

    Article  CAS  PubMed  Google Scholar 

  205. Zheng P, Zeng B, Zhou C, Liu M, Fang Z, Xu X, et al. Gut microbiome remodeling induces depressive-like behaviors through a pathway mediated by the host’s metabolism. Mol Psychiatry. 2016;21:786–96.

    Article  CAS  PubMed  Google Scholar 

  206. Scheperjans F, Aho V, Pereira PA, Koskinen K, Paulin L, Pekkonen E, et al. Gut microbiota are related to Parkinson’s disease and clinical phenotype. Mov Disord. 2015;30:350–8.

    Article  PubMed  Google Scholar 

  207. Kelly JR, Borre Y, Patterson E, El Aidy S, Deane J, et al. Transferring the blues: Depression-associated gut microbiota induces neurobehavioural changes in the rat. J Psychiatr Res. 2016;82:109–18.

    Article  PubMed  Google Scholar 

  208. Kim MS, Kim Y, Choi H, Kim W, Park S, Lee D, et al. Transfer of a healthy microbiota reduces amyloid and tau pathology in an Alzheimer’s disease animal model. Gut. 2020;69:283–94.

    Article  CAS  PubMed  Google Scholar 

  209. Grabrucker S, Marizzoni M, Silajdzic E, Lopizzo N, Mombelli E, Nicolas S, et al. Microbiota from Alzheimer’s patients induce deficits in cognition and hippocampal neurogenesis. Brain. 2023;146:4916–34.

    Article  PubMed  PubMed Central  Google Scholar 

  210. Mohle L, Mattei D, Heimesaat MM, Bereswill S, Fischer A, Alutis M, et al. Ly6C(hi) monocytes provide a link between antibiotic-induced changes in gut microbiota and adult hippocampal neurogenesis. Cell Rep. 2016;15:1945–56.

    Article  PubMed  Google Scholar 

  211. Kim N, Jeon SH, Ju IG, Gee MS, Do J, Oh MS, et al. Transplantation of gut microbiota derived from Alzheimer’s disease mouse model impairs memory function and neurogenesis in C57BL/6 mice. Brain Behav Immun. 2021;98:357–65.

    Article  CAS  PubMed  Google Scholar 

  212. Lupori L, Cornuti S, Mazziotti R, Borghi E, Ottaviano E, Cas MD, et al. The gut microbiota of environmentally enriched mice regulates visual cortical plasticity. Cell Rep. 2022;38: 110212.

    Article  CAS  PubMed  Google Scholar 

  213. Rei D, Saha S, Haddad M, Rubio AH, Perlaza BL, Berard M, et al. Age-associated gut microbiota impair hippocampus-dependent memory in a vagus-dependent manner. JCI Insight. 2022;7:e147700.

  214. Hoban AE, Moloney RD, Golubeva AV, McVey Neufeld KA, O’Sullivan O, Patterson E, et al. Behavioural and neurochemical consequences of chronic gut microbiota depletion during adulthood in the rat. Neuroscience. 2016;339:463–77.

    Article  CAS  PubMed  Google Scholar 

  215. Long-Smith C, O’Riordan KJ, Clarke G, Stanton C, Dinan TG, Cryan JF. Microbiota-gut-brain axis: new therapeutic opportunities. Annu Rev Pharmacol Toxicol. 2020;60:477–502.

    Article  CAS  PubMed  Google Scholar 

  216. Dalton A, Mermier C, Zuhl M. Exercise influence on the microbiome-gut-brain axis. Gut Microbes. 2019;10:555–68.

    Article  PubMed  PubMed Central  Google Scholar 

  217. Cronin O, O’Sullivan O, Barton W, Cotter PD, Molloy MG, Shanahan F. Gut microbiota: implications for sports and exercise medicine. Br J Sports Med. 2017;51:700–1.

    Article  PubMed  Google Scholar 

  218. Monda V, Villano I, Messina A, Valenzano A, Esposito T, Moscatelli F, et al. Exercise modifies the gut microbiota with positive health effects. Oxid Med Cell Longev. 2017;2017:3831972.

    Article  PubMed  PubMed Central  Google Scholar 

  219. Allen JM, Mailing LJ, Niemiro GM, Moore R, Cook MD, White BA, et al. Exercise alters gut microbiota composition and function in lean and obese humans. Med Sci Sports Exerc. 2018;50:747–57.

    Article  PubMed  Google Scholar 

  220. Allen JM, Mailing LJ, Cohrs J, Salmonson C, Fryer JD, Nehra V, et al. Exercise training-induced modification of the gut microbiota persists after microbiota colonization and attenuates the response to chemically-induced colitis in gnotobiotic mice. Gut Microbes. 2018;9:115–30.

    Article  CAS  PubMed  Google Scholar 

  221. Batacan RB, Fenning AS, Dalbo VJ, Scanlan AT, Duncan MJ, Moore RJ, et al. A gut reaction: the combined influence of exercise and diet on gastrointestinal microbiota in rats. J Appl Microbiol. 2017;122:1627–38.

    Article  CAS  PubMed  Google Scholar 

  222. McFadzean R. Exercise can help modulate human gut microbiota. Undergraduate Honors Thesis. Boulder: University of Colorado; 2014. p. 1–24.

  223. Kang SS, Jeraldo PR, Kurti A, Miller ME, Cook MD, Whitlock K, et al. Diet and exercise orthogonally alter the gut microbiome and reveal independent associations with anxiety and cognition. Mol Neurodegener. 2014;9:36.

    Article  PubMed  PubMed Central  Google Scholar 

  224. Abraham D, Feher J, Scuderi GL, Szabo D, Dobolyi A, Cservenak M, et al. Exercise and probiotics attenuate the development of Alzheimer’s disease in transgenic mice: Role of microbiome. Exp Gerontol. 2019;115:122–31.

    Article  CAS  PubMed  Google Scholar 

  225. Nicolas S, Dohm-Hansen S, Lavelle A, Bastiaanssen TFS, English JA, Cryan JF, et al. Exercise mitigates a gut microbiota-mediated reduction in adult hippocampal neurogenesis and associated behaviours in rats. Transl Psychiatry. 2024;14:195.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  226. Sarubbo F, Cavallucci V, Pani G. The influence of gut microbiota on neurogenesis: evidence and hopes. Cells. 2022;11:382.

  227. Dalile B, Van Oudenhove L, Vervliet B, Verbeke K. The role of short-chain fatty acids in microbiota–gut–brain communication. Nat Rev Gastroenterol Hepatol. 2019;16:461–78.

    Article  PubMed  Google Scholar 

  228. Hopfner F, Kunstner A, Muller SH, Kunzel S, Zeuner KE, Margraf NG, et al. Gut microbiota in Parkinson disease in a northern German cohort. Brain Res. 2017;1667:41–5.

    Article  CAS  PubMed  Google Scholar 

  229. Paiva I, Pinho R, Pavlou MA, Hennion M, Wales P, Schutz AL, et al. Sodium butyrate rescues dopaminergic cells from alpha-synuclein-induced transcriptional deregulation and DNA damage. Hum Mol Genet. 2017;26:2231–46.

    Article  CAS  PubMed  Google Scholar 

  230. Sharma S, Taliyan R, Singh S. Beneficial effects of sodium butyrate in 6-OHDA induced neurotoxicity and behavioral abnormalities: Modulation of histone deacetylase activity. Behav Brain Res. 2015;291:306–14.

    Article  CAS  PubMed  Google Scholar 

  231. Govindarajan N, Agis-Balboa RC, Walter J, Sananbenesi F, Fischer A. Sodium butyrate improves memory function in an alzheimer’s disease mouse model when administered at an advanced stage of disease progression. J Alzheimers Dis. 2011;26:187–97.

    Article  CAS  PubMed  Google Scholar 

  232. Xiong Z, Nelson B, Sneiderman C, Janesko-Feldman K, Kochanek P, Rajasundaram D, et al. Microbiota-derived SCFAs promote chronic neurogenesis and anti-inflammatory gene expression after traumatic brain injury (P1–6.008). Neurology. 2023;100:3218.

  233. Tang C-F, Wang C-Y, Wang J-H, Wang Q-N, Li S-J, Wang H-O, et al. Short-chain fatty acids ameliorate depressive-like behaviors of high fructose-fed mice by rescuing hippocampal neurogenesis decline and blood–brain barrier damage. Nutrients. 2022;14:1882.

  234. Claes S, Myint A-M, Domschke K, Del-Favero J, Entrich K, Engelborghs S, et al. The kynurenine pathway in major depression: Haplotype analysis of three related functional candidate genes. Psychiatry Res. 2011;188:355–60.

    Article  CAS  PubMed  Google Scholar 

  235. Agudelo Leandro Z, Femenía T, Orhan F, Porsmyr-Palmertz M, Goiny M, Martinez-Redondo V, et al. Skeletal muscle PGC-1α1 modulates kynurenine metabolism and mediates resilience to stress-induced depression. Cell. 2014;159:33–45.

    Article  CAS  PubMed  Google Scholar 

  236. Liu X, Cao S, Zhang X. Modulation of gut microbiota-brain axis by probiotics, prebiotics, and diet. J Agric Food Chem. 2015;63:7885–95.

    Article  CAS  PubMed  Google Scholar 

  237. Mato J, Alvarez L, Ortiz P, Pajares MA. S-adenosylmethionine synthesis: Molecular mechanisms and clinical implications. Pharmacol Therapeutics. 1997;73:265–80.

    Article  CAS  Google Scholar 

  238. Chavez M. SAMe: S-Adenosylmethionine. Am J Health Syst Pharm. 2000;57:119–23.

    Article  CAS  PubMed  Google Scholar 

  239. Robertson KD, Wolffe AP. DNA methylation in health and disease. Nat Rev Genet. 2000;1:11–9.

    Article  CAS  PubMed  Google Scholar 

  240. Altuna M, Urdanoz-Casado A, Sanchez-Ruiz de Gordoa J, Zelaya MV, Labarga A, Lepesant JMJ, et al. DNA methylation signature of human hippocampus in Alzheimer’s disease is linked to neurogenesis. Clin Epigenetics. 2019;11:91.

    Article  PubMed  PubMed Central  Google Scholar 

  241. Pellegrini C, Pirazzini C, Sala C, Sambati L, Yusipov I, Kalyakulina A, et al. A meta-analysis of brain DNA methylation across sex, age, and Alzheimer’s disease points for accelerated epigenetic aging in neurodegeneration. Front Aging Neurosci. 2021;13: 639428.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  242. Wang Z, Tang B, He Y, Jin P. DNA methylation dynamics in neurogenesis Epigenomics. 2016;8:401–14.

    CAS  PubMed  Google Scholar 

  243. Zhao X, Ueba T, Christie BR, Barkho B, McConnell MJ, Nakashima K, et al. Mice lacking methyl-CpG binding protein 1 have deficits in adult neurogenesis and hippocampal function. Proc Natl Acad Sci U S A. 2003;100:6777–82.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  244. Ma DK, Jang MH, Guo JU, Kitabatake Y, Chang ML, Pow-Anpongkul N, et al. Neuronal activity-induced Gadd45b promotes epigenetic DNA demethylation and adult neurogenesis. Science. 2009;323:1074–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  245. Madsen TM, Treschow A, Bengzon J, Bolwig TG, Lindvall O, Tingstrom A. Increased neurogenesis in a model of electroconvulsive therapy. Biol Psychiatry. 2000;47:1043–9.

    Article  CAS  PubMed  Google Scholar 

  246. Xie J, Xie L, Wei H, Li X-J, Lin L. Dynamic regulation of DNA methylation and brain functions. Biology. 2023;12:152.

  247. Swiatowy WJ, Drzewiecka H, Kliber M, Sasiadek M, Karpinski P, Plawski A, et al. Physical activity and DNA methylation in humans. Int J Mol Sci. 2021;22:12989.

  248. Xu M, Zhu J, Liu X-D, Luo M-Y, Xu N-J. Roles of physical exercise in neurodegeneration: reversal of epigenetic clock. Transl Neurodegen. 2021;10:30.

  249. Wu C, Yang L, Tucker D, Dong Y, Zhu L, Duan R, et al. Beneficial effects of exercise pretreatment in a sporadic Alzheimer’s rat model. Med Sci Sports Exerc. 2018;50:945–56.

    Article  PubMed  PubMed Central  Google Scholar 

  250. Di Rocco A, Rogers JD, Brown R, Werner P, Bottiglieri T. S-Adenosyl-Methionine improves depression in patients with Parkinson’s disease in an open-label clinical trial. Mov Disord. 2000;15:1225–9.

    Article  PubMed  Google Scholar 

  251. Lee S, Lemere CA, Frost JL, Shea TB. Dietary supplementation with S-adenosyl methionine delayed amyloid-β and tau pathology in 3xTg-AD mice. J Alzheimers Dis. 2012;28:423–31.

    Article  CAS  PubMed  Google Scholar 

  252. Coppede F. Epigenetic regulation in Alzheimer’s disease: is it a potential therapeutic target? Expert Opin Ther Targets. 2021;25:283–98.

    Article  CAS  PubMed  Google Scholar 

  253. Chan A, Tchantchou F, Graves V, Rozen R, Shea TB. Dietary and genetic compromise in folate availability reduces acetylcholine, cognitive performance and increases aggression: critical role of S-adenosyl methionine. J Nutr Health Aging. 2008;12:252–61.

    Article  CAS  PubMed  Google Scholar 

  254. Fuso A, Nicolia V, Ricceri L, Cavallaro RA, Isopi E, Mangia F, et al. S-adenosylmethionine reduces the progress of the Alzheimer-like features induced by B-vitamin deficiency in mice. Neurobiol Aging. 2012;33:1482.e1-82.e16.

    Article  CAS  PubMed  Google Scholar 

  255. Tchantchou F, Graves M, Ortiz D, Chan A, Rogers E, Shea TB. S-adenosyl methionine: A connection between nutritional and genetic risk factors for neurodegeneration in Alzheimer’s disease. J Nutr Health Aging. 2006;10:541–4.

    CAS  PubMed  Google Scholar 

  256. Tchantchou F, Graves M, Falcone D, Shea TB. S-adenosylmethionine mediates glutathione efficacy by increasing glutathione S-transferase activity: implications for S-adenosyl methionine as a neuroprotective dietary supplement. J Alzheimers Dis. 2008;14:323–8.

    Article  CAS  PubMed  Google Scholar 

  257. Zhang Y, Ma R, Deng Q, Wang W, Cao C, Yu C, et al. S-adenosylmethionine improves cognitive impairment in D-galactose-induced brain aging by inhibiting oxidative stress and neuroinflammation. J Chem Neuroanat. 2023;128:102232.

  258. Li Q, Cui J, Fang C, Liu M, Min G, Li L, et al. S-Adenosylmethionine attenuates oxidative stress and neuroinflammation induced by amyloid-β through modulation of glutathione metabolism. J Alzheimers Dis. 2017;58:549–58.

    Article  CAS  PubMed  Google Scholar 

  259. Wang W, Zhao F, Ma X, Perry G, Zhu X. Mitochondria dysfunction in the pathogenesis of Alzheimer’s disease: recent advances. Mol Neurodegenn. 2020;15:30.

  260. Lam AB, Kervin K, Tanis JE. Vitamin B12 impacts amyloid beta-induced proteotoxicity by regulating the methionine/S-adenosylmethionine cycle. Cell Rep. 2021;36:109753.

  261. Wu A, Zhang J. Neuroinflammation, memory, and depression: new approaches to hippocampal neurogenesis. J Neuroinflammation. 2023;20:283.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  262. Taupin P. A dual activity of ROS and oxidative stress on adult neurogenesis and Alzheimer’s disease. Cent Nerv Syst Agents Med Chem. 2010;10:16–21.

    Article  CAS  PubMed  Google Scholar 

  263. Richetin K, Moulis M, Millet A, Arrazola MS, Andraini T, Hua J, et al. Amplifying mitochondrial function rescues adult neurogenesis in a mouse model of Alzheimer’s disease. Neurobiol Dis. 2017;102:113–24.

    Article  CAS  PubMed  Google Scholar 

  264. Araki R, Nishida S, Oishi Y, Tachioka H, Kita A, Yabe T. Methyl donor supplementation prevents a folate deficiency-induced depression-like state and neuronal immaturity of the dentate gyrus in mice. Neuroscience. 2022;485:12–22.

    Article  CAS  PubMed  Google Scholar 

  265. McKee SE, Reyes TM. Effect of supplementation with methyl-donor nutrients on neurodevelopment and cognition: considerations for future research. Nutr Rev. 2018;76:497–511.

    Article  PubMed  PubMed Central  Google Scholar 

  266. Nishida S, Araki R, Baba A, Asari S, Tachibana S, Nakajima Y, et al. Post-weaning folate deficiency induces a depression-like state via neuronal immaturity of the dentate gyrus in mice. J Pharmacol Sci. 2020;143:97–105.

    Article  CAS  PubMed  Google Scholar 

  267. El Hajj CS, Pourie G, Martin N, Alberto JM, Daval JL, Gueant JL, et al. Gestational methyl donor deficiency alters key proteins involved in neurosteroidogenesis in the olfactory bulbs of newborn female rats and is associated with impaired olfactory performance. Br J Nutr. 2014;111:1021–31.

    Article  Google Scholar 

  268. Wang F. The protective effect and mechanism of folic acid on early-life exposure 5-Aza induced mental disorder in adult mice. Graduate Thesis. Xuzhou, China: Xuzhou Medical University; 2013. p. 35–51.

  269. Adwan L, Zawia NH. Epigenetics: a novel therapeutic approach for the treatment of Alzheimer’s disease. Pharmacol Ther. 2013;139:41–50.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  270. Delay C, Mandemakers W, Hébert SS. MicroRNAs in Alzheimer’s disease. Neurobiol Dis. 2012;46:285–90.

    Article  CAS  PubMed  Google Scholar 

  271. Gaudet AD, Fonken LK, Watkins LR, Nelson RJ, Popovich PG. MicroRNAs: Roles in regulating neuroinflammation. Neuroscientist. 2018;24:221–45.

    Article  CAS  PubMed  Google Scholar 

  272. Martinez B, Peplow PV. MicroRNAs as diagnostic and therapeutic tools for Alzheimer’s disease: advances and limitations. Neural Regen Res. 2019;14:242–55.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  273. Walgrave H, Zhou L, De Strooper B, Salta E. The promise of microRNA-based therapies in Alzheimer’s disease: challenges and perspectives. Mol Neurodegener. 2021;16:76.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  274. Fyfe I. MicroRNAs - diagnostic markers in Parkinson disease? Nat Rev Neurol. 2020;16:65.

    Article  PubMed  Google Scholar 

  275. Goh SY, Chao YX, Dheen ST, Tan EK, Tay SS. Role of microRNAs in Parkinson's disease. Int J Mol Sci. 2019;20:5649.

  276. Cai M, Chai S, Xiong T, Wei J, Mao W, Zhu Y, et al. Aberrant expression of circulating microRNA leads to the dysregulation of alpha-synuclein and other pathogenic genes in Parkinson’s disease. Front Cell Dev Biol. 2021;9:695007.

  277. Walgrave H, Balusu S, Snoeck S, Vanden Eynden E, Craessaerts K, Thrupp N, et al. Restoring miR-132 expression rescues adult hippocampal neurogenesis and memory deficits in Alzheimer’s disease. Cell Stem Cell. 2021;28(1805–21): e8.

    Google Scholar 

  278. Dong J, Liu Y, Zhan Z, Wang X. MicroRNA-132 is associated with the cognition improvement following voluntary exercise in SAMP8 mice. Brain Res Bull. 2018;140:80–7.

    Article  CAS  PubMed  Google Scholar 

  279. Xu L, Zheng YL, Yin X, Xu SJ, Tian D, Zhang CY, et al. Excessive treadmill training enhances brain-specific microRNA-34a in the mouse hippocampus. Front Mol Neurosci. 2020;13:7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  280. Wang Z, van Praag H. Exercise and the Brain: Neurogenesis, Synaptic Plasticity, Spine Density, and Angiogenesis. In: Boecker H, Hillman C, Scheef L, Strüder H, editors. Functional Neuroimaging in Exercise and Sport Sciences. New York: Springer; 2012. p. 237–47.

  281. Shi Y, Zhao X, Hsieh J, Wichterle H, Impey S, Banerjee S, et al. MicroRNA regulation of neural stem cells and neurogenesis. J Neurosci. 2010;30:14931–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  282. Domańska-Senderowska D, Laguette M-J, Jegier A, Cięszczyk P, September A, Brzeziańska-Lasota E. MicroRNA profile and adaptive response to exercise training: a review. Int J Sports Med. 2019;40:227–35.

    Article  PubMed  Google Scholar 

  283. Da Silva FC, Rode MP, Vietta GG, Iop RDR, Creczynski-Pasa TB, Martin AS, et al. Expression levels of specific microRNAs are increased after exercise and are associated with cognitive improvement in Parkinson's disease. Mol Med Rep. 2021;24:618.

  284. Fernandes J, Vieira AS, Kramer-Soares JC, Da Silva EA, Lee KS, Lopes-Cendes I, et al. Hippocampal microRNA-mRNA regulatory network is affected by physical exercise. Biochim Biophys Acta Gen Subj. 2018;1862:1711–20.

    Article  CAS  PubMed  Google Scholar 

  285. Dong XJ, Chen JJ, Xue LL, Al-Hawwas M. Treadmill training improves cognitive function by increasing IGF2 targeted downregulation of miRNA-483. Ibrain. 2022;8:264–75.

    Article  PubMed  PubMed Central  Google Scholar 

  286. Improta-Caria AC, Nonaka CKV, Cavalcante BRR, De Sousa RAL, Aras Júnior R, Souza BSdF. Modulation of microRNAs as a potential molecular mechanism involved in the beneficial actions of physical exercise in Alzheimer disease. Int J Mol Sci. 2020;21:4977.

  287. Esteves M, Serra-Almeida C, Saraiva C, Bernardino L. New insights into the regulatory roles of microRNAs in adult neurogenesis. Curr Opin Pharmacol. 2020;50:38–45.

    Article  CAS  PubMed  Google Scholar 

  288. Bao TH, Miao W, Han JH, Yin M, Yan Y, Wang WW, et al. Spontaneous running wheel improves cognitive functions of mouse associated with miRNA expressional alteration in hippocampus following traumatic brain injury. J Mol Neurosci. 2014;54:622–9.

    Article  CAS  PubMed  Google Scholar 

  289. Miao W, Bao TH, Han JH, Yin M, Yan Y, Wang WW, et al. Voluntary exercise prior to traumatic brain injury alters miRNA expression in the injured mouse cerebral cortex. Braz J Med Biol Res. 2015;48:433–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  290. De Sousa RAL, Improta-Caria AC. Regulation of microRNAs in Alzheimer´s disease, type 2 diabetes, and aerobic exercise training. Metab Brain Dis. 2022;37:559–80.

    Article  PubMed  Google Scholar 

  291. Liu W, Li L, Liu S, Wang Z, Kuang H, Xia Y, et al. MicroRNA expression profiling screen miR-3557/324-targeted CaMK/mTOR in the rat striatum of Parkinson’s disease in regular aerobic exercise. Biomed Res Int. 2019;2019:1–12.

    Google Scholar 

  292. Juzwik CA, Zhang Y, Paradis-Isler N, Sylvester A, Amar-Zifkin A, et al. microRNA dysregulation in neurodegenerative diseases: A systematic review. Prog Neurobiol. 2019;182.

    Article  CAS  PubMed  Google Scholar 

  293. Cui JG, Li YY, Zhao Y, Bhattacharjee S, Lukiw WJ. Differential regulation of interleukin-1 receptor-associated kinase-1 (IRAK-1) and IRAK-2 by microRNA-146a and NF-kappaB in stressed human astroglial cells and in Alzheimer disease. J Biol Chem. 2010;285:38951–60.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  294. Taganov KD, Boldin MP, Chang KJ, Baltimore D. NF-kappaB-dependent induction of microRNA miR-146, an inhibitor targeted to signaling proteins of innate immune responses. Proc Natl Acad Sci U S A. 2006;103:12481–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  295. Fan B, Chopp M, Zhang ZG, Liu XS. Emerging roles of microRNAs as biomarkers and therapeutic targets for diabetic neuropathy. Front Neurol. 2020;11:558758.

    Article  PubMed  PubMed Central  Google Scholar 

  296. Martin NA, Hyrlov KH, Elkjaer ML, Thygesen EK, Wlodarczyk A, Elbaek KJ, et al. Absence of miRNA-146a differentially alters microglia function and proteome. Front Immunol. 2020;11:1110.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  297. Soplinska A, Zareba L, Wicik Z, Eyileten C, Jakubik D, Siller-Matula JM, et al. MicroRNAs as biomarkers of systemic changes in response to endurance exercise—a comprehensive review. Diagnostics. 2020;10:813.

  298. Alipour MR, Yousefzade N, Bavil FM, Naderi R, Ghiasi R. Swimming impacts on pancreatic inflammatory cytokines, miR-146a and NF-кB expression levels in type-2 diabetic rats. Curr Diabetes Rev. 2020;16:889–94.

    Article  CAS  PubMed  Google Scholar 

  299. Baggish AL, Hale A, Weiner RB, Lewis GD, Systrom D, Wang F, et al. Dynamic regulation of circulating microRNA during acute exhaustive exercise and sustained aerobic exercise training. J Physiol. 2011;589:3983–94.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  300. Mai H, Fan W, Wang Y, Cai Y, Li X, Chen F, et al. Intranasal administration of miR-146a agomir rescued the pathological process and cognitive impairment in an AD mouse model. Mol Ther Nucleic Acids. 2019;18:681–95.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  301. Xiao WZ, Lu AQ, Liu XW, Li Z, Zi Y, Wang ZW. Role of miRNA-146 in proliferation and differentiation of mouse neural stem cells. Biosci Rep. 2015;35:e00245.

  302. Sung PS, Lin PY, Liu CH, Su HC, Tsai KJ. Neuroinflammation and neurogenesis in alzheimer's disease and potential therapeutic approaches. Int J Mol Sci. 2020;21:701.

  303. Ryan SM, Nolan YM. Neuroinflammation negatively affects adult hippocampal neurogenesis and cognition: can exercise compensate? Neurosci Biobehav Rev. 2016;61:121–31.

    Article  CAS  PubMed  Google Scholar 

  304. Guedes JR, Custodia CM, Silva RJ, de Almeida LP, Pedroso de Lima MC, Cardoso AL. Early miR-155 upregulation contributes to neuroinflammation in Alzheimer’s disease triple transgenic mouse model. Hum Mol Genet. 2014;23:6286–301.

    Article  CAS  PubMed  Google Scholar 

  305. Lopez MS, Morris-Blanco KC, Ly N, Maves C, Dempsey RJ, Vemuganti R. MicroRNA miR-21 decreases post-stroke brain damage in rodents. Transl Stroke Res. 2021;13:483–93.

    Article  PubMed  Google Scholar 

  306. Zhao J, Zhou Y, Guo M, Yue D, Chen C, Liang G, et al. MicroRNA-7: expression and function in brain physiological and pathological processes. Cell Biosci. 2020;10:77.

  307. Lei Y, Jin X, Sun M, Ji Z. miR-129-5p ameliorates ischemic brain injury by binding to SIAH1 and activating the mTOR signaling pathway. J Mol Neurosci. 2021;71:1761–71.

    Article  CAS  PubMed  Google Scholar 

  308. Liu A-H, Wu Y-T, Wang Y-P. MicroRNA-129-5p inhibits the development of autoimmune encephalomyelitis-related epilepsy by targeting HMGB1 through the TLR4/NF-kB signaling pathway. Brain Res Bull. 2017;132:139–49.

    Article  CAS  PubMed  Google Scholar 

  309. Yang C, Wang H, Li C, Niu H, Luo S, Guo X. Association between clusterin concentration and dementia: a systematic review and meta-analysis. Metab Brain Dis. 2019;34:129–40.

    Article  PubMed  Google Scholar 

  310. Prikrylova Vranova H, Henykova E, Mares J, Kaiserova M, Mensikova K, Vastik M, et al. Clusterin CSF levels in differential diagnosis of neurodegenerative disorders. J Neurol Sci. 2016;361:117–21.

    Article  CAS  PubMed  Google Scholar 

  311. Lenzi C, Ramazzina I, Russo I, Filippini A, Bettuzzi S, Rizzi F. The Down-Regulation of Clusterin Expression Enhances the alphaSynuclein Aggregation Process. Int J Mol Sci. 2020;21:7181.

  312. Palihati N, Tang Y, Yin Y, Yu D, Liu G, Quan Z, et al. Clusterin is a potential therapeutic target in Alzheimer's disease. Mol Neurobiol. 2023. https://doi.org/10.1007/s12035-023-03801-1.

  313. Narayan P, Orte A, Clarke RW, Bolognesi B, Hook S, Ganzinger KA, et al. The extracellular chaperone clusterin sequesters oligomeric forms of the amyloid-beta(1–40) peptide. Nat Struct Mol Biol. 2011;19:79–83.

    Article  PubMed  PubMed Central  Google Scholar 

  314. Wojtas AM, Carlomagno Y, Sens JP, Kang SS, Jensen TD, Kurti A, et al. Clusterin ameliorates tau pathology in vivo by inhibiting fibril formation. Acta Neuropathol Commun. 2020;8:210.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  315. Cordero-Llana O, Scott SA, Maslen SL, Anderson JM, Boyle J, Chowhdury RR, et al. Clusterin secreted by astrocytes enhances neuronal differentiation from human neural precursor cells. Cell Death Differ. 2011;18:907–13.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  316. Oh SB, Kim MS, Park S, Son H, Kim SY, Kim MS, et al. Clusterin contributes to early stage of Alzheimer’s disease pathogenesis. Brain Pathol. 2019;29:217–31.

    Article  CAS  PubMed  Google Scholar 

  317. Sahay A, Hen R. Adult hippocampal neurogenesis in depression. Nat Neurosci. 2007;10:1110–5.

    Article  CAS  PubMed  Google Scholar 

  318. Vaidya V, Vadodaria K, Jha S. Neurotransmitter regulation of adult neurogenesis: putative therapeutic targets. CNS Neurol Disord Drug Targets. 2007;6:358–74.

    Article  CAS  PubMed  Google Scholar 

  319. Kumar AM, Sevush S, Kumar M, Ruiz J, Eisdorfer C. Peripheral serotonin in Alzheimer’s disease. Neuropsychobiology. 1995;32:9–12.

    Article  CAS  PubMed  Google Scholar 

  320. Middlemiss DN, Bowen DM, Palmer AM. Serotonin Neurones and Receptors in Alzheimer’s Disease. In: Briley M, Kato A, Weber M, editors. New Concepts in Alzheimer’s Disease. London: Macmillan Education; 1986. p. 89–102.

  321. Mayeux R, Stern Y, Sano M, Williams JB, Cote LJ. The relationship of serotonin to depression in Parkinson’s disease. Mov Disord. 1988;3:237–44.

    Article  CAS  PubMed  Google Scholar 

  322. Tong Q, Zhang L, Yuan Y, Jiang S, Zhang R, Xu Q, et al. Reduced plasma serotonin and 5-hydroxyindoleacetic acid levels in Parkinson’s disease are associated with nonmotor symptoms. Parkinsonism Relat Disord. 2015;21:882–7.

    Article  PubMed  Google Scholar 

  323. Alenina N, Klempin F. The role of serotonin in adult hippocampal neurogenesis. Behav Brain Res. 2015;277:49–57.

    Article  CAS  PubMed  Google Scholar 

  324. Klempin F, Beis D, Mosienko V, Kempermann G, Bader M, Alenina N. Serotonin is required for exercise-induced adult hippocampal neurogenesis. J Neurosci. 2013;33:8270–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  325. Kondo M, Nakamura Y, Ishida Y, Shimada S. The 5-HT3 receptor is essential for exercise-induced hippocampal neurogenesis and antidepressant effects. Mol Psychiatry. 2015;20:1428–37.

    Article  CAS  PubMed  Google Scholar 

  326. Lauritzen HP, Brandauer J, Schjerling P, Koh HJ, Treebak JT, Hirshman MF, et al. Contraction and AICAR stimulate IL-6 vesicle depletion from skeletal muscle fibers in vivo. Diabetes. 2013;62:3081–92.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  327. Ouchi N, Shibata R, Walsh K. AMP-activated protein kinase signaling stimulates VEGF expression and angiogenesis in skeletal muscle. Circ Res. 2005;96:838–46.

    Article  CAS  PubMed  Google Scholar 

  328. He G, Zhang YW, Lee JH, Zeng SX, Wang YV, Luo Z, et al. AMP-activated protein kinase induces p53 by phosphorylating MDMX and inhibiting its activity. Mol Cell Biol. 2014;34:148–57.

    Article  PubMed  PubMed Central  Google Scholar 

  329. Kobilo T, Guerrieri D, Zhang Y, Collica SC, Becker KG, van Praag H. AMPK agonist AICAR improves cognition and motor coordination in young and aged mice. Learn Mem. 2014;21:119–26.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  330. Li H, Wu J, Zhu L, Sha L, Yang S, Wei J, et al. Insulin degrading enzyme contributes to the pathology in a mixed model of Type 2 diabetes and Alzheimer’s disease: possible mechanisms of IDE in T2D and AD. Biosci Rep. 2018;38:BSR20170862.

  331. Du L-L, Chai D-M, Zhao L-N, Li X-H, Zhang F-C, Zhang H-B, et al. AMPK activation ameliorates Alzheimer’s disease-like pathology and spatial memory impairment in a streptozotocin-induced Alzheimer’s disease model in rats. J Alzheimers Dis. 2015;43:775–84.

    Article  CAS  PubMed  Google Scholar 

  332. Ayasolla KR, Giri S, Singh AK, Singh I. 5-aminoimidazole-4-carboxamide-1-beta-4-ribofuranoside (AICAR) attenuates the expression of LPS- and Aβ peptide-induced inflammatory mediators in astroglia. J Neuroinflammation. 2005;2:21.

  333. Cheng JT, Huang CC, Liu IM, Tzeng TF, Chang CJ. Novel mechanism for plasma glucose-lowering action of metformin in streptozotocin-induced diabetic rats. Diabetes. 2006;55:819–25.

    Article  CAS  PubMed  Google Scholar 

  334. Matsui Y, Hirasawa Y, Sugiura T, Toyoshi T, Kyuki K, Ito M. Metformin reduces body weight gain and improves glucose intolerance in high-fat diet-fed C57BL/6J mice. Biol Pharm Bull. 2010;33:963–70.

    Article  CAS  PubMed  Google Scholar 

  335. Wang J, Gallagher D, DeVito LM, Cancino GI, Tsui D, He L, et al. Metformin activates an atypical PKC-CBP pathway to promote neurogenesis and enhance spatial memory formation. Cell Stem Cell. 2012;11:23–35.

    Article  CAS  PubMed  Google Scholar 

  336. Jin Q, Cheng J, Liu Y, Wu J, Wang X, Wei S, et al. Improvement of functional recovery by chronic metformin treatment is associated with enhanced alternative activation of microglia/macrophages and increased angiogenesis and neurogenesis following experimental stroke. Brain Behav Immun. 2014;40:131–42.

    Article  CAS  PubMed  Google Scholar 

  337. Campbell JM, Stephenson MD, de Courten B, Chapman I, Bellman SM, Aromataris E. Metformin use associated with reduced risk of dementia in patients with diabetes: a systematic review and meta-analysis. J Alzheimers Dis. 2018;65:1225–36.

    Article  PubMed  PubMed Central  Google Scholar 

  338. Sluggett JK, Koponen M, Bell JS, Taipale H, Tanskanen A, Tiihonen J, et al. Metformin and risk of Alzheimer's disease among community-dwelling people with diabetes: a national case-control study. J Clin Endocrinol Metab. 2020;105:dgz234.

  339. Hsu CC, Wahlqvist ML, Lee MS, Tsai HN. Incidence of dementia is increased in type 2 diabetes and reduced by the use of sulfonylureas and metformin. J Alzheimers Dis. 2011;24:485–93.

    Article  CAS  PubMed  Google Scholar 

  340. Imfeld P, Bodmer M, Jick SS, Meier CR. Metformin, other antidiabetic drugs, and risk of Alzheimer’s disease: a population-based case-control study. J Am Geriatr Soc. 2012;60:916–21.

    Article  PubMed  Google Scholar 

  341. Ng TP, Feng L, Yap KB, Lee TS, Tan CH, Winblad B. Long-term metformin usage and cognitive function among older adults with diabetes. J Alzheimers Dis. 2014;41:61–8.

    Article  CAS  PubMed  Google Scholar 

  342. Foretz M, Guigas B, Viollet B. Metformin: update on mechanisms of action and repurposing potential. Nat Rev Endocrinol. 2023:19:460–76.

  343. Liao W, Xu J, Li B, Ruan Y, Li T, Liu J. Deciphering the roles of metformin in Alzheimer’s disease: a snapshot. Front Pharmacol. 2021;12:728315.

  344. Sanati M, Aminyavari S, Afshari AR, Sahebkar A. Mechanistic insight into the role of metformin in Alzheimer's disease. Life Sci. 2022;291:120299.

  345. Khezri MR, Yousefi K, Mahboubi N, Hodaei D, Ghasemnejad-Berenji M. Metformin in Alzheimer’s disease: An overview of potential mechanisms, preclinical and clinical findings. Biochem Pharmacol. 2022;197:114945.

  346. Iwashita A, Muramatsu Y, Yamazaki T, Muramoto M, Kita Y, Yamazaki S, et al. Neuroprotective efficacy of the peroxisome proliferator-activated receptor delta-selective agonists in vitro and in vivo. J Pharmacol Exp Ther. 2007;320:1087–96.

    Article  CAS  PubMed  Google Scholar 

  347. Defaux A, Zurich MG, Braissant O, Honegger P, Monnet-Tschudi F. Effects of the PPAR-beta agonist GW501516 in an in vitro model of brain inflammation and antibody-induced demyelination. J Neuroinflammation. 2009;6:15.

    Article  PubMed  PubMed Central  Google Scholar 

  348. Kobilo T, Yuan C, van Praag H. Endurance factors improve hippocampal neurogenesis and spatial memory in mice. Learn Mem. 2011;18:103–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  349. Marangos PJ, Loftus T, Wiesner J, Lowe T, Rossi E, Browne CE, et al. Adenosinergic modulation of homocysteine-induced seizures in mice. Epilepsia. 1990;31:239–46.

    Article  CAS  PubMed  Google Scholar 

  350. Wu J, Puppala D, Feng X, Monetti M, Lapworth AL, Geoghegan KF. Chemoproteomic analysis of intertissue and interspecies isoform diversity of AMP-activated protein kinase (AMPK). J Biol Chem. 2013;288:35904–12.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  351. Moore EM, Mander AG, Ames D, Kotowicz MA, Carne RP, Brodaty H, et al. Increased risk of cognitive impairment in patients with diabetes is associated with metformin. Diabetes Care. 2013;36:2981–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  352. Cardarine (GW-501516): Dosage, Side Effects & Before And After Pictures.  https://insidebodybuilding.com/cardarine/. Access Date: 29 Apr 2024.

  353. Bianchi VE, Locatelli V. GW501516 (Cardarine): pharmacological and clinical effects. Genet Mol Med. 2023;5:1–12.

    Google Scholar 

  354. Patil SP, Jain PD, Ghumatkar PJ, Tambe R, Sathaye S. Neuroprotective effect of metformin in MPTP-induced Parkinson’s disease in mice. Neuroscience. 2014;277:747–54.

    Article  CAS  PubMed  Google Scholar 

  355. Park JH, Burgess JD, Faroqi AH, DeMeo NN, Fiesel FC, Springer W, et al. Alpha-synuclein-induced mitochondrial dysfunction is mediated via a sirtuin 3-dependent pathway. Mol Neurodegener. 2020;15:5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  356. Dulovic M, Jovanovic M, Xilouri M, Stefanis L, Harhaji-Trajkovic L, Kravic-Stevovic T, et al. The protective role of AMP-activated protein kinase in alpha-synuclein neurotoxicity in vitro. Neurobiol Dis. 2014;63:1–11.

    Article  CAS  PubMed  Google Scholar 

  357. Nigam SM, Xu S, Kritikou JS, Marosi K, Brodin L, Mattson MP. Exercise and BDNF reduce Abeta production by enhancing alpha-secretase processing of APP. J Neurochem. 2017;142:286–96.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  358. Kopec BM, Zhao L, Rosa-Molinar E, Siahaan TJ. Non-invasive brain delivery and efficacy of BDNF in APP/PS1 transgenic mice as a model of Alzheimer's disease. Med Res Arch. 2020;8:2043.

  359. Real CC, Ferreira AF, Chaves-Kirsten GP, Torrao AS, Pires RS, Britto LR. BDNF receptor blockade hinders the beneficial effects of exercise in a rat model of Parkinson’s disease. Neuroscience. 2013;237:118–29.

    Article  CAS  PubMed  Google Scholar 

  360. Hansen KF, Karelina K, Sakamoto K, Wayman GA, Impey S, Obrietan K. miRNA-132: a dynamic regulator of cognitive capacity. Brain Struct Funct. 2013;218:817–31.

    Article  PubMed  Google Scholar 

  361. Feng M-G, Liu C-F, Chen L, Feng W-B, Liu M, Hai H, et al. MiR-21 attenuates apoptosis-triggered by amyloid-β via modulating PDCD4/ PI3K/AKT/GSK-3β pathway in SH-SY5Y cells. Biomed Pharmacother. 2018;101:1003–7.

    Article  CAS  PubMed  Google Scholar 

  362. Li Z, Chen Q, Liu J, Du Y. Physical exercise ameliorates the cognitive function and attenuates the neuroinflammation of Alzheimer’s disease via miR-129–5p. Dement Geriatr Cogn Disord. 2020;49:163–9.

    Article  PubMed  Google Scholar 

  363. Long JM, Ray B, Lahiri DK. MicroRNA-339-5p down-regulates protein expression of β-site amyloid precursor protein-cleaving enzyme 1 (BACE1) in human primary brain cultures and is reduced in brain tissue specimens of Alzheimer disease subjects. J Biol Chem. 2014;289:5184–98.

    Article  CAS  PubMed  Google Scholar 

  364. Kou X, Chen D, Chen N. The regulation of microRNAs in Alzheimer’s disease. Front Neurol. 2020;11:288.

    Article  PubMed  PubMed Central  Google Scholar 

  365. Martinez B, Peplow PV. Altered microRNA expression in animal models of Huntington’s disease and potential therapeutic strategies. Neural Regen Res. 2021;16:2159–69.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  366. Tung CW, Huang PY, Chan SC, Cheng PH, Yang SH. The regulatory roles of microRNAs toward pathogenesis and treatments in Huntington’s disease. J Biomed Sci. 2021;28:59.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  367. Fuso A, Seminara L, Cavallaro RA, D’Anselmi F, Scarpa S. S-adenosylmethionine/homocysteine cycle alterations modify DNA methylation status with consequent deregulation of PS1 and BACE and beta-amyloid production. Mol Cell Neurosci. 2005;28:195–204.

    Article  CAS  PubMed  Google Scholar 

  368. Chen F, Swartzlander DB, Ghosh A, Fryer JD, Wang B, Zheng H. Clusterin secreted from astrocyte promotes excitatory synaptic transmission and ameliorates Alzheimer’s disease neuropathology. Mol Neurodegener. 2021;16:5.

    Article  PubMed  PubMed Central  Google Scholar 

  369. Clark A, Mach N. Exercise-induced stress behavior, gut-microbiota-brain axis and diet: a systematic review for athletes. J Int Soc Sports Nutr. 2016;13:43.

    Article  PubMed  PubMed Central  Google Scholar 

  370. Carey RA, Montag D. Exploring the relationship between gut microbiota and exercise: short-chain fatty acids and their role in metabolism. BMJ Open Sport Exerc Med. 2021;7:e000930.

  371. Giovannini MG, Lana D, Traini C, Vannucchi MG. The microbiota-gut-brain axis and alzheimer disease. from dysbiosis to neurodegeneration: focus on the central nervous system glial cells. J Clin Med. 2021;10:2358.

  372. Wang Q, Luo Y, Ray Chaudhuri K, Reynolds R, Tan EK, Pettersson S. The role of gut dysbiosis in Parkinson’s disease: mechanistic insights and therapeutic options. Brain. 2021;144:2571–93.

    Article  PubMed  Google Scholar 

  373. Aho VTE, Houser MC, Pereira PAB, Chang J, Rudi K, Paulin L, et al. Relationships of gut microbiota, short-chain fatty acids, inflammation, and the gut barrier in Parkinson’s disease. Mol Neurodegener. 2021;16:6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  374. Elfil M, Kamel S, Kandil M, Koo BB, Schaefer SM. Implications of the gut microbiome in Parkinson’s disease. Mov Disord. 2020;35:921–33.

    Article  CAS  PubMed  Google Scholar 

  375. Rhutik SP, Sanjoli GV, Wasiyoddin TQ, Harshwardhan JT, Priya SM, Milind JU. The gut microbiome in Huntington disease: A review. GSC Biol Pharm Sci. 2021;15:317–26.

    Article  Google Scholar 

  376. Zhao Y, Pang Q, Liu M, Pan J, Xiang B, Huang T, et al. Treadmill exercise promotes neurogenesis in ischemic rat brains via Caveolin-1/VEGF signaling pathways. Neurochem Res. 2017;42:389–97.

    Article  PubMed  Google Scholar 

  377. Fabel K, Fabel K, Tam B, Kaufer D, Baiker A, Simmons N, et al. VEGF is necessary for exercise-induced adult hippocampal neurogenesis. Eur J Neurosci. 2003;18:2803–12.

    Article  PubMed  Google Scholar 

  378. Pang Q, Zhang H, Chen Z, Wu Y, Bai M, Liu Y, et al. Role of caveolin-1/vascular endothelial growth factor pathway in basic fibroblast growth factor-induced angiogenesis and neurogenesis after treadmill training following focal cerebral ischemia in rats. Brain Res. 2017;1663:9–19.

    Article  CAS  PubMed  Google Scholar 

  379. Correia AS, Cardoso A, Vale N. Oxidative stress in depression: the link with the stress response, neuroinflammation, serotonin, neurogenesis and synaptic plasticity. Antioxidants. 2023;12:470.

  380. Kim DD, Barr AM, Honer WG, Procyshyn RM. Exercise-induced hippocampal neurogenesis: 5-HT(3) receptor antagonism by antipsychotics as a potential limiting factor in Schizophrenia. Mol Psychiatry. 2018;23:2252–3.

    Article  PubMed  Google Scholar 

  381. Rogers J, Chen F, Stanic D, Farzana F, Li S, Zeleznikow-Johnston AM, et al. Paradoxical effects of exercise on hippocampal plasticity and cognition in mice with a heterozygous null mutation in the serotonin transporter gene. Br J Pharmacol. 2019;176:3279–96.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  382. Pilz GA, Bottes S, Betizeau M, Jorg DJ, Carta S, Simons BD, et al. Live imaging of neurogenesis in the adult mouse hippocampus. Science. 2018;359:658–62.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  383. Urban N, Blomfield IM, Guillemot F. Quiescence of adult mammalian neural stem cells: a highly regulated rest. Neuron. 2019;104:834–48.

    Article  CAS  PubMed  Google Scholar 

  384. Knobloch M, Jessberger S. Metabolism and neurogenesis. Curr Opin Neurobiol. 2017;42:45–52.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

None.

Funding

Not applicable.

Author information

Authors and Affiliations

Authors

Contributions

RZ: conceptualisation and design of the study; RZ: acquisition and analysis of data, preparation of figures; RZ: writing – original draft, writing – review and editing.

Corresponding author

Correspondence to Renqing Zhao.

Ethics declarations

Ethics approval and Consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The author declares no competing interests.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zhao, R. Can exercise benefits be harnessed with drugs? A new way to combat neurodegenerative diseases by boosting neurogenesis. Transl Neurodegener 13, 36 (2024). https://doi.org/10.1186/s40035-024-00428-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s40035-024-00428-7

Keywords