Download PDF
Review  |  Open Access  |  4 Sep 2024

The role of N6-methyladenosine modification in neurodegenerative diseases

Views: 153 |  Downloads: 17 |  Cited:  0
Ageing Neur Dis 2024;4:15.
10.20517/and.2024.14 |  © The Author(s) 2024.
Author Information
Article Notes
Cite This Article

Abstract

In recent years, transcriptomics has emerged as a key focus in neuroscience research, as transcriptome modifications play a significant role in influencing various biological processes. N6-methyladenosine (m6A) represents a dynamic and reversible form of mRNA modification prevalent in eukaryotes. This modification is involved in virtually every critical stage of RNA metabolism, including mRNA stability, transcription, translation, splicing, nuclear export, and decay, thereby playing a pivotal role in normal brain development. Accumulating evidence suggests that m6A modification plays a substantial role in various neurodegenerative diseases, such as Alzheimer’s disease (AD), Parkinson’s disease (PD), amyotrophic lateral sclerosis (ALS), and Huntington’s disease (HD), while abnormal m6A modification can lead to neurodevelopmental disorders. This review summarizes the relationship between m6A modification and various neurodegenerative diseases and elucidates its potential pathogenic mechanisms at the molecular level.

Keywords

N6 methyladenosine modification (m6A), Alzheimer’s disease, Parkinson’s disease, amyotrophic lateral sclerosis, frontotemporal dementia, Huntington’s disease

INTRODUCTION

Chemical modifications in RNA are considered important mechanisms for regulating gene expression and protein translation. RNA methylation modifications include N6 methyladenosine (m6A), N1 methyladenosine (m1A), N6, 2-O-dimethyladenosine (m6Am), 5-methylcytosine (m5C), 5-hydroxymethylcytosine (5hmC), and 7-methylguanine (m7G), among which m6A is the most common and abundant eukaryotic methylation modification type in RNA[1-3].

M6A methylation modification is a reversible process that involves the participation of methyltransferases (Writers), demethylases (Erasers), and methylated reading proteins (Readers)[4]. Methyltransferases, which include METTL3, METTL14, WTAP, and KIAA1429, catalyze the addition of a methyl group to the sixth nitrogen of adenosine during the methylation process[5]. Demethylases, such as FTO and ALKBH5, can remove this methylation[6]. At the same time, methylated RNA base sites require specific enzymes for recognition, known as methylated reading proteins (Readers). These include YTH domain proteins (e.g., YTHDF1/2/3 and YTHDC1/2), nuclear heterogeneous ribonucleoprotein (hnRNP), and IGF2BP7[7] [Figure 1].

The role of N6-methyladenosine modification in neurodegenerative diseases

Figure 1. Three kinds of enzymes involved in m6A methylation modification. m6A: N6-methyladenosine.

Neurodegenerative disorders, encompassing Alzheimer’s disease (AD), Parkinson’s disease (PD), amyotrophic lateral sclerosis (ALS), frontotemporal dementia (FTD), and Huntington’s disease (HD), represent a spectrum of neurological conditions characterized by the relentless degeneration of neurons within the central nervous system (CNS) or the peripheral nervous system (PNS)[8,9]. Recent studies have illuminated a pivotal role for m6A RNA methylation modification in the intricate tapestry of brain development. Notably, m6A-associated proteins exhibit ubiquitous expression and profound enrichment within neuronal tissue, intricately modulating fundamental processes such as memory formation and consolidation, adult neurogenesis, axonal regeneration, and cerebellar maturation[10].

Indeed, the exploration of m6A RNA modification stands as a forefront and highly sought-after research avenue within the vast landscape of life sciences. While the spotlight has predominantly shone on its implications in cancer research, encompassing diverse malignancies like hematological tumors, uterine cancer, and breast cancer[11-13], a notable research void persists in the realm of neurodegenerative diseases. This gap underscores the urgency to delve deeper into the role of RNA methylation modification, particularly m6A, in these complex and devastating neurological conditions. In this context, we review the pivotal roles of m6A modification in select neurodegenerative disorders, aiming to stimulate research and enhance understanding of its intricate interplay with neurodegeneration.

AD

AD is a neurodegenerative disorder associated with advancing age, representing the most prevalent underlying cause of dementia[14]. In the United States, about 10% of people aged 65 and above suffer from AD. The ubiquitous neuropathological hallmarks of AD encompass synaptic and neuronal dysregulation, manifested by the presence of intracellular neurofibrillary tangles, coupled with elevated concentrations of neurotoxic amyloid proteins and the extracellular accumulation of neuropathic plaques[15,16]. Multiple studies have demonstrated that dysregulation of RNA methylation is associated with AD, and m6A is the predominant modification type found in eukaryotic RNA[17].

APP/PS1 and 5XFAD mice have been widely used in the study of the pathogenesis and therapeutic effectiveness of AD[18]. Quantitative analysis of RNA m6A methylation levels in APP/PS1 transgenic mice of AD showed an increase in m6A methylation levels in the cortex and hippocampus. Meanwhile, in the APP mouse model, the expression of m6A methyltransferase METTL3 increased while the expression of m6A demethylase FTO decreased[19]. However, in the 6-month-aged 5XFAD mouse model, an increase in FTO expression and a decrease in METTL3 levels were observed through RNA sequencing and proteomic analysis, which is consistent with a reduction in m6A modification levels in these mice[20]. Interestingly, in human AD samples, METTL3 is downregulated in the soluble fraction of the hippocampus but upregulated in the insoluble fraction, mirroring the accumulation of insoluble Tau observed in AD patients[21,22].

AD patients have a significant amount of β-amyloid protein, known as Aβ, in their brains. This protein can lead to the formation of senile plaques and the apoptosis of nerve cells, which are the hallmarks of AD[23]. Recent investigations have uncovered that a scarcity of methyltransferase such as METTL3 in macrophages derived from monocytes can enhance cognitive performance in amyloid-beta-induced AD mouse models. Specifically, the absence of METTL3 triggers the downregulation of α-Tubulin Acetyltransferase 1 (ATAT1), resulting in decreased levels of α-tubulin acetylation. As a consequence, this promotes the migration of monocyte-derived macrophages and facilitates the clearance of Amyloid-beta (Aβ), ultimately mitigating the symptoms of AD[24].

Tau protein hyperphosphorylation is one of the most important causes of AD, which can lead to neurofibrillary tangles, neuronal death, and functional impairment[25]. In a fruit fly model, it was found that the absence of m6A can enhance the toxicity of Tau in the fruit fly AD model, while the absence of METTL3, METTL14, and YTHDF can result in profound motor impairments in fruit flies[26]. Lysine-specific demethylase 1A (KDM1A) plays a protective role in the hippocampus and cortex of mice, preventing neurodegeneration and cognitive decline. However, in its absence, neurodegeneration and cognitive impairment are observed[27]. Recent research indicates that KDM1A promotes the expression of METTL3 through the upregulation of the m6A-dependent pathway. This, in turn, enhances the autophagic clearance of phosphorylated Tau (p-Tau), offering protection against neurodegenerative lesions in AD[28].

Synaptic degeneration arises during the initial phases of AD and is intimately linked to the decline in cognitive abilities[29]. In the AD mouse model with m6A methyltransferase METTL3 knockdown, m6A modification was reduced in hippocampal neurons, resulting in notable memory impairments in mice, accompanied by widespread synaptic loss and neuronal cell death, as well as oxidative stress and abnormal cell cycle events[21]. Circular RNA (circRNA) is abundant in neural tissue, and its abnormal expression precedes AD symptoms, which is strongly associated with the severity of clinical dementia[30,31]. CircRIMS2, mediated by METTL3-dependent m6A modification, exhibited marked upregulation in 4-month-old APP/PS1 mice. Elevated expression of circRIMS2 led to deficits in synaptic function and memory capacity in 4-month-old C57BL/6 mice, while silencing circRIMS2 significantly rescued synaptic dysfunction in AD mice[32].

The modification of m6A demonstrates temporal and spatial differences during neural development and aging processes. Specifically, in the brains of mice with AD, m6A functions in a distinct spatial and temporal pattern[33]. The m6A level is highest in the brains of 2-week-aged and 52-week-aged (elderly) mice, while the lowest level is detected in the brains of 4-week-aged and 6-week-aged (adolescent) mice[26]. Investigations have demonstrated that m6A assumes a pivotal function in modulating neuronal maturation and exerts a more profound influence during both the initial and terminal phases of development[33].

In essence, investigations utilizing AD mouse models reveal intricate dynamics in m6A methyltransferases and demethylases, intimately linked to disease progression. Specifically, METTL3 depletion in macrophages bolsters Aβ clearance, whereas its aberrant regulation in neurons disrupts synaptic function and memory. Notably, Tau hyperphosphorylation, a process modulated by m6A-associated pathways, contributes significantly to AD pathogenesis. Moreover, circRNA dysregulation, a consequence of METTL3-mediated m6A modifications, precedes AD symptoms and adversely affects synaptic function. The m6A modification displays distinct temporal and spatial patterns during neural development and aging, underscoring its pivotal role in orchestrating neuronal maturation and AD progression.

PD

PD stands as one of the most prevalent age-related neurodegenerative disorders, primarily targeting dopaminergic neurons[34]. The pathological hallmarks of this disease include neural inclusion bodies in the form of Lewis bodies and Lewy neurites, as well as cell loss in the substantia nigra and other brain regions[35,36]. Recent research endeavors have unveiled a compelling association between multiple m6A-methylated proteins and PD, hinting at a potential role of m6A regulatory factors in the intricate pathogenesis of this disorder[37].

1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP) is commonly used to construct PD animal models. This neurotoxin exerts its neurotoxicity by causing a series of injuries, ultimately leading to damage to dopaminergic neurons in the substantia nigra compacta and striatum[38]. In the PD mouse model constructed by MPTP, the m6A modification level decreased and the demethylase FTO significantly increased in the tissues of PD mice. Furthermore, the striatal expression of m6A methyltransferases, METTL3 and RBM15, decreased in PD mice, whereas the expression of the m6A demethylase ALKBH5 increased in the substantia nigra. In parallel, compared to control group mice, the expression level of methylated reading proteins YTHDF1 was downregulated in MPTP-treated mice[39]. Prior research underscores the pivotal role of YTHDF1 in the CNS, as it has been implicated in inducing axonal regeneration and regulating axonal guidance[40,41]. These findings suggest that disruptions in m6A modification pathways may contribute to the pathological processes underlying PD.

Alpha-synuclein (α-Syn), a highly representative biomarker of PD, has been instrumental in the diagnosis of this condition[42]. The abnormal accumulation of α-Syn plays a pivotal role in the onset and progression of PD, impacting various pathological processes including mitochondrial dysfunction, oxidative stress, neuroinflammation, and autophagy[43]. Analysis shows that METTL14 is associated with plasma α-Syn, which is closely related to the severity of PD disease. Further investigation suggests that METTL14 interacts with the m6A motif within the coding sequence (CDS) region of α-Syn, thereby modulating the m6A modification of α-Syn mRNA through an m6A-YTHDF2-dependent mechanism[44]. Notably, in peripheral blood mononuclear cells of PD patients, the levels of both m6A and its methyltransferase METTL14 are significantly reduced. This reduction leads to impaired m6A modification of α-Syn mRNA, ultimately contributing to the abnormal accumulation of α-Syn and exacerbating PD pathology[45,46].

Glutaredoxin (GLRX), a small protein featuring a single active site cysteine pair, is paramount in sustaining the intracellular reducing environment and safeguarding cells against oxidative stress. Moreover, it plays a pivotal role in the pathogenesis of numerous neurodegenerative disorders, including PD[47]. Specifically, GLRX1 is instrumental in modulating the apoptotic signals in dopaminergic neurons, and its absence exacerbates the demise of these cells in PD patients[48]. NRF1, a vital transcriptional regulatory factor, governs the expression of proteasome genes in neurons. Notably, its expression levels are diminished in PD rat models[49]. It has been reported that NRF1 can alleviate motor impairment and dopamine neuronal deterioration, triggered by MPTP in PD mice, by elevating the transcription level of METTL3 and enhancing the m6A modification process of GLRX mRNA[50].

In general, PD is a neurodegenerative condition primarily impacting dopaminergic neurons, marked by the presence of Lewis bodies and cell loss, and has recently been implicated in the context of m6A-methylated proteins. MPTP-induced models of PD exhibit perturbations in m6A modifications and alterations in the expression of associated enzymes, consequentially impacting dopaminergic neurons. α-Syn, an important biomarker for PD, undergoes abnormal accumulation and interacts with METTL14, modulating its m6A modification status, which is notably diminished in PD patients. GLRX plays a pivotal role in maintaining neuronal health, and NRF1, by regulating METTL3, fortifies the m6A modification of GLRX mRNA, thereby safeguarding neurons. This intricate interplay of m6A modifications underscores the complexity of PD pathogenesis.

ALS/FTD

ALS is a neurodegenerative disease that progresses rapidly and has a high mortality rate[51]. FTD belongs to the second largest family of hereditary cognitive impairment, mainly affecting personality, social behavior, and language function[52]. Up to 50% of ALS patients may experience varying degrees of cognitive impairment, with some developing into FTD, and patients with advanced FTD may also exhibit clinical manifestations of ALS. Moreover, ALS and FTD can occur simultaneously in the same family or even in the same patient[53]. TDP-43 positive aggregates were found in neurons of the vast majority of ALS and FTD patients, and the same FUS positive aggregates were also observed in neurons of ALS patients with FUS mutations and some FTD patients[54,55]. In recent years, it has been discovered that mutations in certain genes can lead to the phenotype of both ALS and FTD, and can even manifest as both ALS and FTD phenotypes simultaneously[56]. Therefore, the overlap in clinical manifestations, pathology, and genetics closely links ALS and FTD, two seemingly dissimilar diseases.

A hexanucleotide repeat expansion (HRE) of a GGGGCC repeat within the first intron of C9orf72 (GGGGCC) is the most prevalent genetic factor of ALS and FTD[57,58]. Based on post-mortem tissue study in patients with C9orf72-ALS/FTD, a conspicuous reduction in the expression levels of methyltransferases METTL3 and METTL14 was discerned, accompanied by a marked downregulation of m6A levels. This finding is in harmony with the observed downregulation of m6A methylation in spinal cord neurons (iPSNs) derived from induced pluripotent stem cells (iPSCs) sourced from ALS/FTD patients[54,59]. Intriguingly, in iPSNs differentiated from iPSCs, the strategic modulation of either methyltransferase or demethylase aimed at reinstating the m6A level exhibits the potential to mitigate the disease-related phenotypes associated with ALS/FTD[59].

The nuclear clearance and cytoplasmic localization errors of TDP-43 are pathological markers of ALS, FTD, and related neurodegenerative diseases, collectively referred to as TDP-43 proteinopathy[60,61]. This mislocalization of TDP-43 from the nucleus to the cytoplasm, along with its clearance from the nucleus, is intimately correlated with RNA splicing errors, translational impairments, and widespread RNA instability in ALS[62]. Furthermore, studies have shown that the upregulation of mRNA expression levels encoding ribosomal proteins and oxidative phosphorylase is associated with TDP-43 deposition in neurons lacking the RNA methyltransferase METTL14[63]. Intriguingly, research has uncovered that not only do most TDP-43 substrates bear m6A modifications, but also YTHDF2 exacerbates TDP-43-related toxicity in both rodent and human neuronal models[64].

Human alphaherpesvirus 1 (HSV-1) commonly establishes a latent infection within the trigeminal nerve, which can persist lifelong and has garnered attention as a potential contributor to the development of ALS and FTD[65]. Recent research has illuminated the mechanisms underlying HSV-1’s interaction with cellular RNA methylation pathways. Specifically, during the early stages of infection, HSV-1 enhances the expression of m6A methyltransferases, such as METTL3 and METTL14, as well as m6A reader proteins, including YTHDF1, YTHDF2, and YTHDF3. Subsequently, HSV-1 inhibits the expression of m6A demethylases like FTO and ALKBH5, thereby promoting viral replication by modulating the RNA methylation landscape[66].

The formation of inclusion bodies composed of dipeptide repeat proteins (DPRs), particularly the positively charged and arginine-rich DPR protein poly(GR), is recognized as a pathological hallmark of ALS and FTD[67]. Poly(GR) exhibits considerable toxicity in clinical models of ALS/FTD, manifesting through nucleolar damage, inhibition of protein synthesis, disruption of ribosomal RNA processing and biogenesis, interactions with RNA-binding proteins, and alterations in the liquid-liquid phase separation of membraneless organelles[68]. Remarkably, studies have demonstrated that mRNA modified with m6A modification and the m6A-binding YTHDF proteins not only co-localize with poly(GR) inclusions in the brains of ALS/FTD mouse models and FTD patients but also contribute to the formation of these inclusions. By incorporating RNA into the inclusions, m6A modifications and YTHDF proteins may interrupt essential interactions between poly(GR) and proteins like G3BP1 or YTHDF1, ultimately enhancing poly(GR) aggregation[69].

Translocated in liposarcoma/fused in sarcoma (TLS/FUS) is a nuclear RNA/DNA binding protein that, when mutated, can lead to the development of ALS of the spinal cord. Mutations in the C-terminal nuclear localization signal (NLS) of TLS/FUS have been specifically linked to the onset and severity of ALS symptoms[70,71]. These mutations disrupt the proper localization of TLS/FUS within the cell, causing it to accumulate in the cytoplasm, where it forms cytotoxic aggregates through a process known as liquid-liquid phase separation (LLPS)[72,73]. Notably, TLS/FUS has a preference for binding to mRNA fragments that are modified with m6A modification. However, mutations in the NLS of TLS/FUS can reduce its specificity for binding to these modified RNA fragments. Furthermore, the interplay between TLS/FUS and m6A-modified mRNA segments has been demonstrated to hinder the assembly of cytotoxic agglomerates via LLPS, while the dispersion of these modified RNA fragments throughout cytoplasmic TLS/FUS foci conspicuously augments cellular viability[74]. This observation implies that this specific interaction may serve as a safeguard against the onset and progression of ALS, underscoring its protective function in the context of this neurodegenerative disorder.

ALS and FTD exhibit significant clinical, pathological, and genetic intersections, with TDP-43 mislocalization and alterations in m6A RNA methylation standing as pivotal factors in their pathogenesis. Mutations in the C9orf72 gene, resulting in the formation of HREs, contribute to the development of both ALS and FTD, impacting methyltransferases METTL3/14 and modulating m6A levels. Additionally, HSV-1 infection dynamically modulates m6A pathways, enhancing viral replication strategies. In the context of ALS and FTD, DPR proteins such as poly(GR) form characteristic inclusions that co-localize with m6A-modified RNA and YTHDF proteins, underscoring their involvement in disease mechanisms. Meanwhile, mutations in TLS/FUS genes, prevalent in ALS, disrupt proper nuclear localization, leading to the formation of cytotoxic aggregates. Intriguingly, these mutations’ interaction with m6A-modified RNA appears to hinder aggregate formation, suggesting a potential protective mechanism against disease progression.

HD

HD is an autosomal dominant neurodegenerative disease, primarily due to a genetic alteration in the HTT gene situated on chromosome 4, leading to the production of a mutated Huntington protein[75,76]. HD patients typically develop symptoms between the ages of 30 and 40, mainly characterized by involuntary dance-like movements, cognitive impairment, and mental abnormalities. The course of the disease ranges from 10 to 20 years, ultimately leading to the patient’s death[77]. So far, there is no specific drug that can prevent the pathological process of HD, mainly due to the unclear pathological mechanism of the disease[78,79].

Hdh knockout mice are one of the commonly used animal models for studying HD, and HdhQ111 knockout mice are a group of them with 111 HD polyglutamine fragment residues[80]. High methylation of m6A in synapse-related genes was observed in the hippocampal transcriptome of Hdh+/Q111 mice. Research has uncovered that m6A modification in the hippocampus of individuals with HD is abnormally regulated in a manner that is influenced by experiences. This aberrant regulation results in the demethylation of crucial synaptic tissue components, potentially underpinning the cognitive impairments observed in HD[81]. Consistent with these data, knockdown of m6A demethylase FTO in HD mice can improve spatial and recognition memory[82].

In the realm of amplifying CAG repeats, alongside the production of the full-length (FL) HTT mRNA isoform, two smaller transcripts arise due to incomplete splicing. These transcripts, designated HTT1a, encompass exon 1 and intron 1 sequences in their 5’ region, a phenomenon attributed to abnormal polyadenylation at an intrinsic polyA site within intron 1[83,84]. Notably, HTT1a encodes not only the well-known highly pathogenic HTT exon 1 protein, which is prone to aggregation, but also contributes to the formation of mRNA nuclear clusters that exhibit resilience against HTT antisense oligonucleotides (ASOs) therapy[85,86]. It is imperative to mention that pharmacological interventions aimed at inhibiting the m6A methyltransferase METTL3 or specifically demethylating HTT intron 1 have been found to significantly reduce the transcription level of HTT1a in HD cells[85]. This suggests that m6A methylation within intron 1 may be influenced by CAG amplification and plays a pivotal role in the production of aberrantly spliced HTT1a transcripts[87].

The pathological hallmark of TDP-43 mislocalization from the nucleus to the cytoplasm has emerged as a consistent feature in patients and mouse models of ALS, FTD, and HD[88,89]. Intriguingly, recent advancements in research have illuminated that TDP-43 deficiency can instigate somatic trinucleotide repeat expansion, exacerbating the neuropathological manifestations observed in HD. Specifically, targeted knockdown of endogenous TDP-43 in the striatum of HD knock-in mice has been shown to accelerate the expansion of CAG repeats[90,91]. Furthermore, recent investigations have highlighted TDP-43 dysfunction and aberrant m6A modification as potential contributors to incorrect splicing in HD, thereby modulating the expression of striatal genes associated with HD[88]. Notably, the reduced binding of TDP-43 is evident within genes that define the unique characteristics of striatal HD. Analogously, the m6A deposition levels within these same striatal genes are also diminished under HD conditions[92]. In the context of HD, the heightened expression of TDP-43 implicates a reliance on m6A modification for TDP-43 binding. The observed decline in m6A deposition within abnormal HD-associated genes results in a consequential reduction in TDP-43 binding and stability, revealing a novel mechanistic interplay between TDP-43 binding sites and m6A deposition sites[93]. This finding underscores a novel mechanistic correlation between TDP-43 binding sites and m6A deposition sites within the context of HD, suggesting that m6A methylation may serve as a crucial determinant for TDP-43’s binding capacity.

In essence, the relationship between HD and m6A methylation is multifaceted and plays a pivotal role in the pathogenesis of HD. Notably, animal models, particularly those with HdhQ111 gene knockouts, exhibit abnormal m6A methylation patterns in synaptic genes, which subsequently affect cognitive functioning. The generation of HTT1a transcripts, intimately linked to the pathogenesis of HD, is governed by m6A methylation occurring specifically within intron 1. Furthermore, the mislocalization of TDP-43 and its intricate interaction with m6A modifications significantly influence striatal gene expression in HD, uncovering a novel and potentially crucial interplay between TDP-43 binding and m6A deposition that may underlie the fundamental mechanisms of HD pathology.

CONCLUSION

RNA m6A methylation is highly prevalent in the mammalian brain and constitutes a crucial surface transcriptome modification. m6A modulates the function of target genes by influencing mRNA translation, splicing, degradation, and nuclear export, thereby exerting a broad impact on various neurodegenerative diseases, such as AD, PD, ALS/FTD, and HD [Table 1]. Numerous studies have highlighted the pivotal role of m6A in regulating cerebral functions. The absence of m6A methyltransferases in Drosophila insects leads to severe mobility deficits, characterized by disorientation, sluggish walking pace, and decreased activity, all of which can be traced back to impaired neuronal function[94]. Similarly, mice with dysfunctional m6A pathways exhibit a multitude of cerebral abnormalities, ranging from reduced brain size and compromised spatial learning and memory abilities to defects in synaptic transmission, long-term potentiation, and axon regeneration[95]. This emphasizes the significance of the m6A-modulated epitranscriptome in neuronal functioning, yet research on its influence on human neurological or neurodegenerative disorders remains scarce.

Table 1

The molecular mechanism between m6A modification and neurodegenerative diseases

DiseasesPathological featuresBiomarkerRelated enzymesm6A modification levelRef.
ADβ-amyloid proteinATAT1METTL3Elevated methylation levels[24]
Hyperphosphorylation of tau proteinKDM1AMETTL3
METTL14
YTHDF
Reduced methylation level[28]
Synaptic degenerationCircRIMS2METTL3Reduced methylation level[32]
PDDopaminergic neuron damageMPTPMETTL3
RBM15
ALKBH5
FTO
Reduced methylation level[39]
The abnormal accumulation of α-Synα-SynMETTL14
YTHDF
Reduced methylation level[44]
Oxidative stressGLRX1METTL3Reduced methylation level[48]
ALS/FTDDeposition of TDP-43 in neuronsTDP-43METTL14
YTHDF2
Reduced methylation level[63]
Human alphaherpesvirus-1 infectionHSV-1METTL3
METTL14
FTO
ALBKH5
Elevated methylation levels[66]
Inclusion body composed of DPR proteinsPoly(GR)G3BP1
YTHDF1
Elevated methylation levels[69]
Cytotoxic aggregates through LLPSTLS/FUSYTHDF2Elevated methylation levels[74]
HDHighly pathogenic HTT exon1 protein that is prone to aggregationHTT1aMETTL3Elevated methylation levels[87]
Accelerated expansion of CAG repeat caused by TDP-43 deficiencyTDP-43METTL3Reduced methylation level[90,93]

In the context of AD, the multifaceted role of m6A modification emerges as a pivotal regulatory mechanism. It intricately modulates macrophage functionality, enhancing the clearance of Amyloid-beta (Aβ) deposits, while simultaneously influencing Tau protein phosphorylation dynamics. In PD, the role of m6A modification revolves around its influence on key molecular players, particularly α-Syn and GLRX. Moreover, in ALS and FTD cases caused by the repetitive expansion of C9orf72, m6A modification and its associated proteins play complex roles in modulating the pathogenesis of ALS/FTD through their interactions with TDP-43, DPRs like poly(GR), and TLS/FUS. Lastly, in HD, heightened m6A modification of synaptic-related genes modulates HTT gene expression and splicing, further influencing the cytoplasmic localization of TDP-43 protein [Figure 2]. By elucidating these intricate methylation patterns, we can delve deeper into the fundamental mechanisms underlying neurodegeneration, thereby potentially forging a path for the development of targeted therapeutic interventions. Furthermore, the pivotal proteins intricately involved in m6A modification are anticipated to emerge as promising molecular targets for the diagnosis, treatment, and drug discovery endeavors aimed at combating neurodegenerative diseases.

The role of N6-methyladenosine modification in neurodegenerative diseases

Figure 2. Dynamic changes of m6A modification in four neurodegenerative diseases and pathological features associated with this methylation. m6A: N6-methyladenosine.

The intricate interplay between “Writers”, “Erasers”, and “Readers” in the m6A methylation modification process further complicates its pathogenic mechanisms. For instance, METTL3 serves as a pivotal component of the m6A methyltransferase “Writers” complex, demonstrating robust methyltransferase activity. However, there have also been reports suggesting that METTL3 functions as an m6A “reader” in certain contexts. Research indicates that METTL3 directly promotes the translation of several m6A-modified mRNAs by interacting with the translation initiation machinery, independently of its downstream m6A “reader” activity[96]. Such findings underscore the need for a nuanced understanding of the dynamic interplay between these epigenetic regulators and their intricate contributions to neurodegenerative diseases.

Moreover, variations in experimental conditions and the utilization of diverse animal models contribute to discrepancies in research outcomes. A case in point is the study utilizing APP/PS1 mice as models, where contrasting results were obtained. Specifically, while an elevation of m6A methylation levels in the cortex and hippocampus was observed in 9-month-old APP/PS1 mice[19], a study involving 6-month-old APP/PS1 mice reported a downregulation of METTL3 expression, an upregulation of FTO expression, and a consequential decrease in m6A methylation levels[26]. Indeed, the observed disparities in research outcomes likely stem from the intricate compensatory mechanisms that regulate m6A modification during neurodevelopment and aging. These processes, characterized by temporal and spatial differences, contribute to a dynamic landscape of epigenetic modifications. Therefore, a comprehensive experimental strategy that encompasses various developmental stages across a broad spectrum of neurodegenerative diseases is paramount. Such an approach will enable researchers to track the nuanced changes in RNA m6A methylation modification over time and across different disease models. By doing so, we can gain a deeper understanding of the role of m6A in neurodegeneration and how it is affected by disease progression.

Absolutely, the current focus on mouse models in research on m6A methylation modification and neurodegenerative diseases, while valuable, has its limitations. The pathological complexity and age-dependence of neurodegenerative diseases mean that transgenic rodent models often fail to fully recapitulate the selective neurodegenerative changes observed in human brains. This underscores the importance of exploring larger animal models, such as monkeys and pigs, which exhibit greater structural and functional similarities to humans[97,98]. Large animal models offer unique advantages in studying neurodegenerative diseases, including more accurate representation of disease progression, pathology, and response to therapeutic interventions. As such, extending research on the role of m6A methylation modification in neurodegenerative diseases to large-scale animal models has the potential to provide deeper insights into the mechanisms underlying these disorders. With the continued development and maturation of large animal neurodegenerative disease models, we can anticipate a more comprehensive and nuanced understanding of the function of m6A methylation modification in these diseases. This, in turn, will pave the way for the development of more effective and targeted therapeutic strategies aimed at addressing the challenges posed by neurodegenerative disorders.

DECLARATIONS

Authors’ contributions

Wrote the review paper: Song J

Revised manuscript, conceived and designed experiments: Yan S

Both authors read and approved the final manuscript.

Availability of data and materials

Not applicable.

Financial support and sponsorship

This work was supported by the National Key Research and Development Program of China (2021YFA0805300), The National Natural Science Foundation of China (82171244). Sincerely acknowledges the support of the K. C. Wong Education Foundation.

Conflicts of interest

Yan S is a Guest Editor of the journal Ageing and Neurodegenerative Diseases, while the other author declared that there are no conflicts of interest.

Ethical approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Copyright

© The Author(s) 2024.

REFERENCES

1. Li J, Xie H, Ying Y, et al. YTHDF2 mediates the mRNA degradation of the tumor suppressors to induce AKT phosphorylation in N6-methyladenosine-dependent way in prostate cancer. Mol Cancer 2020;19:152.

2. Huang W, Chen TQ, Fang K, Zeng ZC, Ye H, Chen YQ. N6-methyladenosine methyltransferases: functions, regulation, and clinical potential. J Hematol Oncol 2021;14:117.

3. Wang T, Kong S, Tao M, Ju S. The potential role of RNA N6-methyladenosine in cancer progression. Mol Cancer 2020;19:88.

4. Fang Z, Mei W, Qu C, et al. Role of m6A writers, erasers and readers in cancer. Exp Hematol Oncol 2022;11:45.

5. Sendinc E, Shi Y. RNA m6A methylation across the transcriptome. Mol Cell 2023;83:428-41.

6. Feng H, Yuan X, Wu S, et al. Effects of writers, erasers and readers within miRNA-related m6A modification in cancers. Cell Prolif 2023;56:e13340.

7. Zhang N, Ding C, Zuo Y, Peng Y, Zuo L. N6-methyladenosine and neurological diseases. Mol Neurobiol 2022;59:1925-37.

8. Temple S. Advancing cell therapy for neurodegenerative diseases. Cell Stem Cell 2023;30:512-29.

9. Kritsilis M, V Rizou S, Koutsoudaki PN, Evangelou K, Gorgoulis VG, Papadopoulos D. Ageing, cellular senescence and neurodegenerative disease. Int J Mol Sci 2018;19:2937.

10. Zhang R, Zhang Y, Guo F, Li S, Cui H. RNA N6-methyladenosine modifications and its roles in Alzheimer’s disease. Front Cell Neurosci 2022;16:820378.

11. Sun T, Wu R, Ming L. The role of m6A RNA methylation in cancer. Biomed Pharmacother 2019;112:108613.

12. Deng X, Su R, Weng H, Huang H, Li Z, Chen J. RNA N6-methyladenosine modification in cancers: current status and perspectives. Cell Res 2018;28:507-17.

13. Dong S, Wu Y, Liu Y, Weng H, Huang H. N6-methyladenosine steers RNA metabolism and regulation in cancer. Cancer Commun 2021;41:538-59.

14. Rostagno AA. Pathogenesis of Alzheimer’s disease. Int J Mol Sci 2022;24:107.

15. Scheltens P, De Strooper B, Kivipelto M, et al. Alzheimer’s disease. Lancet 2021;397:1577-90.

16. Breijyeh Z, Karaman R. Comprehensive review on Alzheimer’s disease: causes and treatment. Molecules 2020;25:5789.

17. Lv J, Xing L, Zhong X, Li K, Liu M, Du K. Role of N6-methyladenosine modification in central nervous system diseases and related therapeutic agents. Biomed Pharmacother 2023;162:114583.

18. Izco M, Martínez P, Corrales A, et al. Changes in the brain and plasma Aβ peptide levels with age and its relationship with cognitive impairment in the APPswe/PS1dE9 mouse model of Alzheimer’s disease. Neuroscience 2014;263:269-79.

19. Han M, Liu Z, Xu Y, et al. Abnormality of m6A mRNA methylation is involved in Alzheimer’s disease. Front Neurosci 2020;14:98.

20. Bai B, Wang X, Li Y, et al. Deep multilayer brain proteomics identifies molecular networks in Alzheimer’s disease progression. Neuron 2020;105:975-91.e7.

21. Zhao F, Xu Y, Gao S, et al. METTL3-dependent RNA m6A dysregulation contributes to neurodegeneration in Alzheimer’s disease through aberrant cell cycle events. Mol Neurodegener 2021;16:70.

22. Huang H, Camats-Perna J, Medeiros R, Anggono V, Widagdo J. Altered expression of the m6A methyltransferase METTL3 in Alzheimer’s disease. eNeuro 2020;7:ENEURO.0125-20.2020.

23. Sandberg A, Luheshi LM, Söllvander S, et al. Stabilization of neurotoxic Alzheimer amyloid-beta oligomers by protein engineering. Proc Natl Acad Sci U S A 2010;107:15595-600.

24. Yin H, Ju Z, Zheng M, et al. Loss of the m6A methyltransferase METTL3 in monocyte-derived macrophages ameliorates Alzheimer’s disease pathology in mice. PLoS Biol 2023;21:e3002017.

25. Wegmann S, Biernat J, Mandelkow E. A current view on Tau protein phosphorylation in Alzheimer’s disease. Curr Opin Neurobiol 2021;69:131-8.

26. Shafik AM, Zhang F, Guo Z, et al. N6-methyladenosine dynamics in neurodevelopment and aging, and its potential role in Alzheimer’s disease. Genome Biol 2021;22:17.

27. Christopher MA, Myrick DA, Barwick BG, et al. LSD1 protects against hippocampal and cortical neurodegeneration. Nat Commun 2017;8:805.

28. Tang Z, Cao J, Yao J, et al. KDM1A-mediated upregulation of METTL3 ameliorates Alzheimer’s disease via enhancing autophagic clearance of p-Tau through m6A-dependent regulation of STUB1. Free Radic Biol Med 2023;195:343-58.

29. Crimins JL, Pooler A, Polydoro M, Luebke JI, Spires-Jones TL. The intersection of amyloid β and tau in glutamatergic synaptic dysfunction and collapse in Alzheimer’s disease. Ageing Res Rev 2013;12:757-63.

30. You X, Vlatkovic I, Babic A, et al. Neural circular RNAs are derived from synaptic genes and regulated by development and plasticity. Nat Neurosci 2015;18:603-10.

31. Puri S, Hu J, Sun Z, et al. Identification of circRNAs linked to Alzheimer’s disease and related dementias. Alzheimers Dement 2023;19:3389-405.

32. Wang X, Xie J, Tan L, et al. N6-methyladenosine-modified circRIMS2 mediates synaptic and memory impairments by activating GluN2B ubiquitination in Alzheimer’s disease. Transl Neurodegener 2023;12:53.

33. Ma C, Chang M, Lv H, et al. RNA m6A methylation participates in regulation of postnatal development of the mouse cerebellum. Genome Biol 2018;19:68.

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

35. Tolosa E, Garrido A, Scholz SW, Poewe W. Challenges in the diagnosis of Parkinson's disease. Lancet Neurol 2021;20:385-97.

36. Weintraub D, Aarsland D, Chaudhuri KR, et al. The neuropsychiatry of Parkinson’s disease: advances and challenges. Lancet Neurol 2022;21:89-102.

37. Mathoux J, Henshall DC, Brennan GP. Regulatory mechanisms of the RNA modification m6A and significance in brain function in health and disease. Front Cell Neurosci 2021;15:671932.

38. Mustapha M, Mat Taib CN. MPTP-induced mouse model of Parkinson’s disease: a promising direction of therapeutic strategies. Bosn J Basic Med Sci 2021;21:422-33.

39. Yu Z, Huang L, Xia Y, et al. Analysis of m6A modification regulators in the substantia nigra and striatum of MPTP-induced Parkinson’s disease mice. Neurosci Lett 2022;791:136907.

40. Zhuang M, Li X, Zhu J, et al. The m6A reader YTHDF1 regulates axon guidance through translational control of Robo3.1 expression. Nucleic Acids Res 2019;47:4765-77.

41. Shi H, Zhang X, Weng YL, et al. m6A facilitates hippocampus-dependent learning and memory through YTHDF1. Nature 2018;563:249-53.

42. Blommer J, Pitcher T, Mustapic M, et al. Extracellular vesicle biomarkers for cognitive impairment in Parkinson’s disease. Brain 2023;146:195-208.

43. Lei Q, Wu T, Wu J, et al. Roles of α-synuclein in gastrointestinal microbiome dysbiosis-related Parkinson’s disease progression (Review). Mol Med Rep 2021;24:734.

44. He H, Zhang Q, Liao J, et al. METTL14 is decreased and regulates m6A modification of α-synuclein in Parkinson’s disease. J Neurochem 2023;166:609-22.

45. Wang X, Lu Z, Gomez A, et al. N6-methyladenosine-dependent regulation of messenger RNA stability. Nature 2014;505:117-20.

46. Wang X, Zhao BS, Roundtree IA, et al. N6-methyladenosine modulates messenger RNA translation efficiency. Cell 2015;161:1388-99.

47. Arodin L, Lamparter H, Karlsson H, et al. Alteration of thioredoxin and glutaredoxin in the progression of Alzheimer’s disease. J Alzheimers Dis 2014;39:787-97.

48. Miller O, Mieyal JJ. Critical roles of glutaredoxin in brain cells-implications for Parkinson’s disease. Antioxid Redox Signal 2019;30:1352-68.

49. Ferreira AFF, Binda KH, Singulani MP, et al. Physical exercise protects against mitochondria alterations in the 6-hidroxydopamine rat model of Parkinson’s disease. Behav Brain Res 2020;387:112607.

50. Gong X, Huang M, Chen L. NRF1 mitigates motor dysfunction and dopamine neuron degeneration in mice with Parkinson’s disease by promoting GLRX m6A methylation through upregulation of METTL3 transcription. CNS Neurosci Ther 2024;30:e14441.

51. Al-Chalabi A, Hardiman O. The epidemiology of ALS: a conspiracy of genes, environment and time. Nat Rev Neurol 2013;9:617-28.

52. Balendra R, Isaacs AM. C9orf72-mediated ALS and FTD: multiple pathways to disease. Nat Rev Neurol 2018;14:544-58.

53. Rascovsky K, Hodges JR, Knopman D, et al. Sensitivity of revised diagnostic criteria for the behavioural variant of frontotemporal dementia. Brain 2011;134:2456-77.

54. Chew J, Gendron TF, Prudencio M, et al. Neurodegeneration. C9ORF72 repeat expansions in mice cause TDP-43 pathology, neuronal loss, and behavioral deficits. Science 2015;348:1151-4.

55. Chatterjee M, Özdemir S, Fritz C, et al. Plasma extracellular vesicle tau and TDP-43 as diagnostic biomarkers in FTD and ALS. Nat Med 2024;30:1771-83.

56. Guerreiro R, Brás J, Hardy J. SnapShot: genetics of ALS and FTD. Cell 2015;160:798.e1.

57. DeJesus-Hernandez M, Mackenzie IR, Boeve BF, et al. Expanded GGGGCC hexanucleotide repeat in noncoding region of C9ORF72 causes chromosome 9p-linked FTD and ALS. Neuron 2011;72:245-56.

58. Todd TW, Shao W, Zhang YJ, Petrucelli L. The endolysosomal pathway and ALS/FTD. Trends Neurosci 2023;46:1025-41.

59. Li Y, Dou X, Liu J, et al. Globally reduced N6-methyladenosine (m6A) in C9ORF72-ALS/FTD dysregulates RNA metabolism and contributes to neurodegeneration. Nat Neurosci 2023;26:1328-38.

60. Neumann M. Molecular neuropathology of TDP-43 proteinopathies. Int J Mol Sci 2009;10:232-46.

61. Tziortzouda P, Van Den Bosch L, Hirth F. Triad of TDP43 control in neurodegeneration: autoregulation, localization and aggregation. Nat Rev Neurosci 2021;22:197-208.

62. Melamed Z, López-Erauskin J, Baughn MW, et al. Premature polyadenylation-mediated loss of stathmin-2 is a hallmark of TDP-43-dependent neurodegeneration. Nat Neurosci 2019;22:180-90.

63. Koranda JL, Dore L, Shi H, et al. Mettl14 is essential for epitranscriptomic regulation of striatal function and learning. Neuron 2018;99:283-92.e5.

64. McMillan M, Gomez N, Hsieh C, et al. RNA methylation influences TDP43 binding and disease pathogenesis in models of amyotrophic lateral sclerosis and frontotemporal dementia. Mol Cell 2023;83:219-36.e7.

65. Wan Q, Song D, Li H, He ML. Stress proteins: the biological functions in virus infection, present and challenges for target-based antiviral drug development. Signal Transduct Target Ther 2020;5:125.

66. Feng Z, Zhou F, Tan M, et al. Targeting m6A modification inhibits herpes virus 1 infection. Genes Dis 2022;9:1114-28.

67. Kwon I, Xiang S, Kato M, et al. Poly-dipeptides encoded by the C9orf72 repeats bind nucleoli, impede RNA biogenesis, and kill cells. Science 2014;345:1139-45.

68. Zhang YJ, Gendron TF, Ebbert MTW, et al. Poly(GR) impairs protein translation and stress granule dynamics in C9orf72-associated frontotemporal dementia and amyotrophic lateral sclerosis. Nat Med 2018;24:1136-42.

69. Park J, Wu Y, Shao W, et al. Poly(GR) interacts with key stress granule factors promoting its assembly into cytoplasmic inclusions. Cell Rep 2023;42:112822.

70. Chen L. FUS mutation is probably the most common pathogenic gene for JALS, especially sporadic JALS. Rev Neurol 2021;177:333-40.

71. Aksoy YA, Deng W, Stoddart J, et al. “STRESSED OUT”: the role of FUS and TDP-43 in amyotrophic lateral sclerosis. Int J Biochem Cell Biol 2020;126:105821.

72. Coady TH, Manley JL. ALS mutations in TLS/FUS disrupt target gene expression. Genes Dev 2015;29:1696-706.

73. Alberti S, Gladfelter A, Mittag T. Considerations and challenges in studying liquid-liquid phase separation and biomolecular condensates. Cell 2019;176:419-34.

74. Yoneda R, Ueda N, Kurokawa R. m6A modified short RNA fragments inhibit cytoplasmic TLS/FUS aggregation induced by hyperosmotic stress. Int J Mol Sci 2021;22:11014.

75. Fields E, Vaughan E, Tripu D, et al. Gene targeting techniques for Huntington’s disease. Ageing Res Rev 2021;70:101385.

76. Duan W, Urani E, Mattson MP. The potential of gene editing for Huntington’s disease. Trends Neurosci 2023;46:365-76.

77. Bates GP, Dorsey R, Gusella JF, et al. Huntington disease. Nat Rev Dis Primers 2015;1:15005.

78. Tabrizi SJ, Ghosh R, Leavitt BR. Huntingtin lowering strategies for disease modification in Huntington’s disease. Neuron 2019;102:899.

79. Saudou F, Humbert S. The biology of Huntingtin. Neuron 2016;89:910-26.

80. Agostoni E, Michelazzi S, Maurutto M, et al. Effects of Pin1 loss in HdhQ111 knock-in mice. Front Cell Neurosci 2016;10:110.

81. Merkurjev D, Hong WT, Iida K, et al. Synaptic N6-methyladenosine (m6A) epitranscriptome reveals functional partitioning of localized transcripts. Nat Neurosci 2018;21:1004-14.

82. Pupak A, Singh A, Sancho-Balsells A, et al. Altered m6A RNA methylation contributes to hippocampal memory deficits in Huntington’s disease mice. Cell Mol Life Sci 2022;79:416.

83. Neueder A, Dumas AA, Benjamin AC, Bates GP. Regulatory mechanisms of incomplete huntingtin mRNA splicing. Nat Commun 2018;9:3955.

84. Sathasivam K, Neueder A, Gipson TA, et al. Aberrant splicing of HTT generates the pathogenic exon 1 protein in Huntington disease. Proc Natl Acad Sci U S A 2013;110:2366-70.

85. Fienko S, Landles C, Sathasivam K, et al. Alternative processing of human HTT mRNA with implications for Huntington’s disease therapeutics. Brain 2022;145:4409-24.

86. Ly S, Didiot MC, Ferguson CM, et al. Mutant huntingtin messenger RNA forms neuronal nuclear clusters in rodent and human brains. Brain Commun 2022;4:fcac248.

87. Pupak A, Navarro IR, Sathasivam K, et al. m6A RNA modification of mHtt intron 1 regulates the generation of Htt1a in Huntington’s Disease. 2023. Available from: https://www.biorxiv.org/content/10.1101/2023.11.10.566530v1. [Last accessed on 4 Sep 2024].

88. Tada M, Coon EA, Osmand AP, et al. Coexistence of Huntington’s disease and amyotrophic lateral sclerosis: a clinicopathologic study. Acta Neuropathol 2012;124:749-60.

89. Sanchez II, Nguyen TB, England WE, et al. Huntington’s disease mice and human brain tissue exhibit increased G3BP1 granules and TDP43 mislocalization. J Clin Invest 2021;131:140723.

90. Bai D, Zhu L, Jia Q, et al. Loss of TDP-43 promotes somatic CAG repeat expansion in Huntington’s disease knock-in mice. Prog Neurobiol 2023;227:102484.

91. Dewan R, Chia R, Ding J, et al; American Genome Center (TAGC); FALS Sequencing Consortium; Genomics England Research Consortium; International ALS/FTD Genomics Consortium (iAFGC); International FTD Genetics Consortium (IFGC); International LBD Genomics Consortium (iLBDGC); NYGC ALS Consortium; PROSPECT Consortium. Pathogenic Huntingtin repeat expansions in patients with frontotemporal dementia and amyotrophic lateral sclerosis. Neuron 2021;109:448-60.e4.

92. Liu N, Dai Q, Zheng G, He C, Parisien M, Pan T. N6-methyladenosine-dependent RNA structural switches regulate RNA-protein interactions. Nature 2015;518:560-4.

93. Nguyen TB, Miramontes R, Chillon-Marinas C, et al. Aberrant splicing in Huntington’s disease via disrupted TDP-43 activity accompanied by altered m6A RNA modification. 2023. Available from: https://www.biorxiv.org/content/10.1101/2023.10.31.565004v1. [Last accessed on 4 Sep 2024].

94. Lence T, Akhtar J, Bayer M, et al. m6A modulates neuronal functions and sex determination in Drosophila. Nature 2016;540:242-7.

95. Livneh I, Moshitch-Moshkovitz S, Amariglio N, Rechavi G, Dominissini D. The m6A epitranscriptome: transcriptome plasticity in brain development and function. Nat Rev Neurosci 2020;21:36-51.

96. Lin S, Choe J, Du P, Triboulet R, Gregory RI. The m6A methyltransferase METTL3 promotes translation in human cancer cells. Mol Cell 2016;62:335-45.

97. Yin P, Li S, Li XJ, Yang W. New pathogenic insights from large animal models of neurodegenerative diseases. Protein Cell 2022;13:707-20.

98. Yan S, Li S, Li XJ. Use of large animal models to investigate Huntington’s diseases. Cell Regen 2019;8:9-11.

Cite This Article

Review
Open Access
The role of N6-methyladenosine modification in neurodegenerative diseases
Junzhu Song, Sen Yan

How to Cite

Song, J.; Yan S. The role of N6-methyladenosine modification in neurodegenerative diseases. Ageing. Neur. Dis. 2024, 4, 15. http://dx.doi.org/10.20517/and.2024.14

Download Citation

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. Simply select your manager software from the list below and click on download.

Export Citation File:

Type of Import

Tips on Downloading Citation

This feature enables you to download the bibliographic information (also called citation data, header data, or metadata) for the articles on our site.

Citation Manager File Format

Use the radio buttons to choose how to format the bibliographic data you're harvesting. Several citation manager formats are available, including EndNote and BibTex.

Type of Import

If you have citation management software installed on your computer your Web browser should be able to import metadata directly into your reference database.

Direct Import: When the Direct Import option is selected (the default state), a dialogue box will give you the option to Save or Open the downloaded citation data. Choosing Open will either launch your citation manager or give you a choice of applications with which to use the metadata. The Save option saves the file locally for later use.

Indirect Import: When the Indirect Import option is selected, the metadata is displayed and may be copied and pasted as needed.

About This Article

Special Issue

© The Author(s) 2024. Open Access This article is licensed under a Creative Commons Attribution 4.0 International License (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, sharing, adaptation, distribution and reproduction in any medium or format, for any purpose, even commercially, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.

Data & Comments

Data

Views
153
Downloads
17
Citations
0
Comments
0
0

Comments

Comments must be written in English. Spam, offensive content, impersonation, and private information will not be permitted. If any comment is reported and identified as inappropriate content by OAE staff, the comment will be removed without notice. If you have any queries or need any help, please contact us at support@oaepublish.com.

0
Download PDF
Share This Article
Scan the QR code for reading!
See Updates
Contents
Figures
Related
Ageing and Neurodegenerative Diseases
ISSN 2769-5301 (Online)

Portico

All published articles will be preserved here permanently:

https://www.portico.org/publishers/oae/

Portico

All published articles will be preserved here permanently:

https://www.portico.org/publishers/oae/