Dersleri yüzünden oldukça stresli bir ruh haline sikiş hikayeleri bürünüp özel matematik dersinden önce rahatlayabilmek için amatör pornolar kendisini yatak odasına kapatan genç adam telefonundan porno resimleri açtığı porno filmini keyifle seyir ederek yatağını mobil porno okşar ruh dinlendirici olduğunu iddia ettikleri özel sex resim bir masaj salonunda çalışan genç masör hem sağlık hem de huzur sikiş için gelip masaj yaptıracak olan kadını gördüğünde porn nutku tutulur tüm gün boyu seksi lezbiyenleri sikiş dikizleyerek onları en savunmasız anlarında fotoğraflayan azılı erkek lavaboya geçerek fotoğraflara bakıp koca yarağını keyifle okşamaya başlar
Reach Us +44-3308186230

GET THE APP

Cellular and Molecular Biology - Progress on the Long Non-coding RNAs Involved in Alzheimer’s Disease
1165-158X

Cellular and Molecular Biology
Open Access

Our Group organises 3000+ Global Conferenceseries Events every year across USA, Europe & Asia with support from 1000 more scientific Societies and Publishes 700+ Open Access Journals which contains over 50000 eminent personalities, reputed scientists as editorial board members.

Open Access Journals gaining more Readers and Citations
700 Journals and 15,000,000 Readers Each Journal is getting 25,000+ Readers

This Readership is 10 times more when compared to other Subscription Journals (Source: Google Analytics)
  • Research Article   
  • Cell Mol Biol , Vol 70(1)

Progress on the Long Non-coding RNAs Involved in Alzheimer’s Disease

Aodeng Qimuge1, Su Dun2, Bai Ying3, Yang Xingxin1 and Wuhan Qimuge1*
1Experimental center, Inner Mongolia Traditional Chinese & Mongolian Medical Research Institute, Hohhot, China
2Department of Research and Development Center, HUA Cloud Intelligent Healthcare Co. Ltd, Shenzhen, China
3Department of Mongolian Medicine, Inner Mongolia Medical University, Hohhot, China
*Corresponding Author: Wuhan Qimuge, Experimental center, Inner Mongolia Traditional Chinese & Mongolian Medical Research Institute, Hohhot, China, Email: 939839367@qq.com

Received: 17-Jan-2024 / Manuscript No. CMB-24-125155 / Editor assigned: 19-Jan-2024 / PreQC No. CMB-24-125155 / Reviewed: 05-Feb-2024 / QC No. CMB-24-125155 (QC) / Revised: 12-Feb-2024 / Manuscript No. CMB-24-125155 / Published Date: 19-Feb-2024

Abstract

Alzheimer’s Disease (AD) is terminal and is considered the most common neurodegenerative disease characterized by progressive memory loss and cognitive impairment, accounting for 60% of all dementia cases. AD has been widely studied, but its pathological mechanism is still unclear. Noncoding RNAs (ncRNAs) are a class of RNAs that do not encode proteins. Increasing evidence has indicated that long noncoding RNAs (lncRNAs) play essential roles in protein coding processes, biological activities and various diseases, including AD. This study aims to highlight the emerging roles of lncRNAs in AD by summarizing current studies about lncRNAs as potential biomarkers, the roles of lncRNAs involved in Aβ plaque formation, tau hyperphosphorylation, microglia and astrocyte activation, NLRP3 inflammation and oxidative stress. Finally, we tried to elucidate the possible mechanisms by which lncRNAs function and provide new ideas for the diagnosis and treatment of AD.

Keywords: Long noncoding RNA; Alzheimer’s disease; Biomarker; Biological roles

Introduction

The Telomere-to-telomere (T2T) Consortium contains a complete 3.055 billion base pair (bp) sequence of a human genome called T2TCHM13, which includes gapless assemblies for all chromosomes except Y. To date, T2T-CHM13 is the newest and most complete sequence and was published in “Science” in 2022. However, T2T-ChM13 still does not represent the entire human genome. It was reported that 63,494 of the presently annotated human genes and 233,615 transcripts. Among them, 19,969 genes (86,245 transcripts) were predicted to be proteincoding genes, while the rest consisted of introns, regulatory sequences, noncoding RNAs, and unknown and uncertain protein-coding genes [1]. There are many noncoding sequences, including various ncRNAs, introns and other sequences. In recent decades, protein-coding mRNAs and microRNAs have been widely studied, while reports about lncRNAs are still relatively rare.

In recent years, increasing evidence has proven the significant roles of lncRNAs in cellular processes and various diseases, such as cancer, the cardiovascular system rheumatoid arthritis ischemia and Alzheimer’s Disease (AD) [2-10]. For instance, the lncRNA XIST is upregulated in colon cancer tissues and cell lines, and its knockdown can inhibit tumor growth. Further study investigated whether the lncRNA X-inactive Specific Transcript (XIST) promotes colon cancer tumor growth by targeting miR-34a and the Wnt/β-catenin pathway [11]. Exosomal lncRNA Psoriasis Susceptibility-related RNA Gene Induced by Stress (PRINS) in peripheral blood serum in Monoclonal Gammopathy of Undetermined Significance (MGUS) patients has potential as a biomarker with a sensitivity and specificity of 84.9% and 83.3%, respectively, when compared to that in healthy donors [12]. Lnc- SOX6-1 expression was elevated in pediatric Acute Myeloid Leukemia (AML) patients and in various AML cells. This finding correlates with the outcomes of treatments and a worse risk of stratification [13]. The lncRNA SNHG20 has also been shown to play a vital role in Epithelial Ovarian Cancer (EOC) since it is highly expressed in EOC tissues of patients, and its knockdown can inhibit EOC cell proliferation, migration and invasion. Therefore, the lncRNA Small Nucleolar RNA Host Gene 20 (SNHG20) has potential as a diagnostic marker and treatment target [14].

lncRNA-HOTAIR affects ovarian cancer susceptibility in an Iranian population [15]. lncRNA H19 in the Iranian population was reported to be an important factor that increases the risk of breast cancer [16].

lncRNA MEG3 is a tumor suppressor. It was reported to be a predictive and prognostic marker for patients with breast cancer, while lncRNA-NEF has been found to be involved in intrahepatic cholangiocarcinoma (IHCC), a liver cancer, as a tumor suppressor [17]. Intrahepatic Cholangiocarcinoma (IHCC) patients with high plasma levels of lncRNA-Neighboring Enhancer of FoxA2 (NEF) had significantly better survival conditions than IHCC patients with low plasma levels [18]. The lncRNAs TINCR and CARLo-7 participate in bladder cancer and play a critical role in mediating bladder cancer through the Wnt/β- catenin and JAK2/STAT3 signaling pathways [19,20].

Noncoding RNAs (ncRNAs) are a class of genes that do not encode proteins. ncRNAs longer than 200 nucleotides (nt) are called long noncoding RNAs (lncRNAs), while ncRNAs shorter than 200 nt are known as small or short ncRNAs. Noncoding RNAs reported in the literature are classified as lncRNAs, microRNAs (miRNAs), transfer RNAs (tRNAs), ribosomal RNAs (rRNAs), small nuclear RNAs (snRNAs), piwi-associated RNAs (piRNAs), endogenous short-interfering RNAs (siRNAs) and circular RNAs (circRNAs) [10]. However, the type and number of human ncRNAs currently annotated by the HUGO Gene Nomenclature Committee (HGNC, www.genenames.org) are long noncoding RNAs (4514), microRNAs (1914), transfer RNAs (587), small nuclear RNAs (568), ribosomal RNAs (60), small nuclear RNAs (50), small NF90 (IRF3) (50)-associated RNAs, Y RNAs (4) and vault RNAs (4). The HUGO is the only group with the authority to approve symbols for human genes. Long noncoding RNAs consist of most noncoding RNAs [21]. miRNAs have been widely studied as a group of small ncRNAs. In contrast, scientific research on lncRNAs is emerging. Here, we focused mainly on the recent emerging roles of lncRNAs in Alzheimer’s disease.

Long noncoding RNAs (lncRNAs)

lncRNAs are a group of highly heterogeneous regulatory ncRNAs lacking a significant open reading frame. More than 50,000 lncRNAs are estimated to exist in the human genome. The majority of lncRNAs have a similar but different structure than mRNAs. In terms of similarity, they have a similar biological origin and are also transcribed by RNA polymerase II, 5-capped, spliced and polyadenylated. The difference is that lncRNAs are shorter and have fewer weakly encoded but longer exons, typically lower expression and poor conservation of primary sequences [22]. Another distinguishing characteristic of lncRNAs is strong tissue-specific expression patterns. Brain tissues are more conservative in the expression of lncRNAs than other tissues, and it is estimated that approximately 40% of lncRNAs are expressed specifically in brain tissues. Hence, lncRNAs are potentially involved in many important physiological processes in the nervous system, such as neuronal differentiation, neurite outgrowth, synaptogenesis and synaptic plasticity [23,24]. lncRNAs include linear lncRNAs and circular RNAs (circRNAs). Some studies have investigated circular long noncoding RNAs that also play vital roles in AD, such as ciRS-7, ciRS-7, circNF1-419, circHDAC9, circ_0000950 and circAβ [25].

lncRNAs have been implicated in the pathogenesis of several neurodegenerative diseases, including Alzheimer’s, Parkinson’s, amyotrophic lateral sclerosis and Huntington’s diseases [26-31]. For example, the lncRNA NONMMUG014387 promotes Schwann Cells (SCs) after peripheral nerve injury [32]. The lncRNA GAS5 affects neuron death and neuronal function in ischemic stroke. The apoptosis of neurons can be inhibited, and neuronal injury can be improved by inhibiting DNMT3B-dependent MAP4K4 methylation when lncRNA GAS5 is silenced [33]. Additionally, lncRNA SPRY4 was found to be dysregulated in Non-small Cell Lung Cancer (NSCLC) cells. NSCLC patients with low levels of lncRNA SPRY4 Intronic Transcript 1 (SPRY4- IT1) might have relatively shorter survival times. Hence, lncRNA SPRY4 also has biomarker potential [34]. In addition, lncRNA MEG3 was also found to be involved in nerve injury and injured nerve regeneration in rats with sciatic nerve defects through regulating the proliferation and migration of SCs [35]. A comprehensive profile of lncRNAs in a transgenic mouse model of Alzheimer’s disease was generated, and a total of 4622 lncRNAs were analyzed. Among them, 205 lncRNAs were significantly different. MAPT-AS1, a lncRNA, may play an important role in breast cancer [36].

The lncRNA linc00657 was found to induce neuropathic pain by mediating the miR-136/ZEB1 axis in a rat model. The reduction in linc00657 can inhibit neuroinflammation in CCI rats by targeting the expression of cyclooxygenase-2, tumor necrosis factor-α and interleukin-1β, and these outcomes were reversed by miR-136 inhibitors [37]. lncRNA embryonic stem cells expressed 1 (Lncenc1) is also a lncRNA involved in neuropathic pain. Lncenc1 interacts with EZH2 and downregulates BAI1 gene expression to affect neuropathic pain in mouse microglia [38]. Recently, downregulation of lncRNA ZEB1-AS1 was found in peripheral blood mononuclear cells of Sporadic amyotrophic lateral sclerosis (SALS) patients. ZEB1-AS1 was implicated in the pathology of neuronal differentiation [39]. Here, we emphasize linear long noncoding RNAs involved in AD.

Alzheimer’s disease

Alzheimer’s Disease (AD) is the most common neurodegenerative disease characterized by progressive memory loss and cognitive impairment. It is a disease that threatens human health worldwide. However, research into the mechanisms of AD is lagging far behind expectations [40]. Although the exact causes of the disease are not understood, it is believed that the main pathological features of AD are the dysregulation of two proteins in brain tissue, extracellular protein amyloid β (Aβ) deposition and intracellular type 2 microtubuleassociated protein (tau) aggregation or insoluble Neurofibrillary Tangle (NFT) formation [41]. The plaques caused by Aβ aggregation play a major role in cognitive decline and cell death by disrupting cellular communication and causing microglial activation and inflammation [42]. In addition to Aβ deposition, dysregulation of the protein tau also leads to AD. Hyperphosphorylated tau protein aggregates to form NFTs, which block synaptic transmission and are thought to be closely linked to cognitive decline in AD [43].

While Aβ plaques and NFTs in neurons are key features of AD, neuroinflammation has emerged as another major cause. In the pathogenesis of AD, aggregates of Aβ and NFTs trigger an inflammatory response in the nervous system, termed neuroinflammation, mediated by central Nervous System (CNS) resident glial cells such as microglia and astrocytes, inflammasomes, and inflammatory mediators including cytokines, chemokines and reactive oxygen species. Conversely, neuroinflammation leads to the deterioration of Aβ plaques and tau hyperphosphorylation, which further potentiate AD progression and contribute to AD pathology [44-46].

Many studies have shown that neuroinflammation and oxidative stress are inextricably linked and lead to most neurodegenerative pathologies, including Alzheimer’s disease, amyotrophic lateral sclerosis, Huntington’s disease and Parkinson’s disease. Inflammatory cells secrete excess oxygen free radicals, leading to oxidative stress, and some Reactive Oxygen Species (ROS) and Reactive Nitrogen Species (RNS) can further promote intracellular signaling cascades to increase the secretion of proinflammatory factors, ultimately leading to neuroinflammation. Thus, neuroinflammation and oxidative stress can be mutually stimulating, especially in the disease state [47].

Dementia is diagnosed by core clinical criteria, including the appearance of interference with the ability to work or function; decline from previous levels of functioning and performing; not explained by delirium or major psychiatric disorder; and cognitive impairment is detected and diagnosed, which is related to impaired abilities in daily activities. The possible AD dementia diagnosis meets the cognitive deficits for AD, and etiologically mixed presentation meets all AD dementia clinical criteria. Currently, well-known biomarkers for judging the pathological process of AD are Aβ Positron Emission Tomography (PET) of the brain, the Aβ42/Aβ40 ratio in Cerebrospinal Fluid (CSF), low Cerebrospinal Fluid (CSF) Aβ42, increased CFS tau, including total and phosphorylated tau, and decreased Fluorodeoxy Glucose (FDG) uptake on PET in the brain [48].

For the diagnosis of AD, cognitive testing is the most common way to check for AD symptoms. If the cognitive test has difficulty determining the disease, Magnetic Resonance Imaging (MRI) and other tests mentioned above will be performed. However, there are some disadvantages. For instance, PET and MRI are not easy to widely use due to their high cost. Additionally, CSF collection requires an invasive lumbar puncture procedure with general anesthetics. It causes side effects of headache [49]. Therefore, investigating other biomarkers in blood, such as the plasma Aβ42/Aβ40 ratio, tau and Neurofilament Light (NFL), has become a topic of interest in recent years. The aims for biomarkers such as other proteins, miRNAs, metabolites and exosomes have been reported [50]. In this mini-review, we mainly summarize the comprehensive role of various lncRNAs involved in AD, including Aβ deposition, tau protein hyperphosphorylation and neuroinflammationrelated pathological processes, such as microglia and astrocyte activation, the NLRP3 inflammasome and oxidative stress.

lncRNAs as potential biomarkers or therapeutic targets in Alzheimer’s Disease (AD)

lncRNAs have been investigated to have great potential as diagnostic biomarkers and therapeutic targets in AD. lncRNAs might be a new focus for AD biomarker studies, which are critical for earlystage screening and timely prevention of AD. One of the most studied lncRNAs is BACE1-AS. The expression level of lncRNA BACE1-AS in whole plasma samples was lower in pre-AD subgroup patients and higher in full-AD patients than in healthy people. Additionally, the level of lncRNA BACE1-AS in plasma can accurately reflect pre-AD, full-AD and healthy people with good sensitivity and specificity [30,51].

lncRNA 51A in plasma, an antisense transcript of the SORL1 gene SORL1-AS (51A), has been implicated as a potential biomarker. The expression of lncRNA 51A can cause aggregation in AD patients in both the brain and plasma. The plasma 51A level has a negative correlation with Mini-Mental State Examination (MMSE) scores in AD patients [52,53]. The results from a study enrolled 120 AD patients, 120 Parkinson’s Disease (PD) patients and 120 controls and showed that lncRNA MALAT1 and its target are potential biomarkers for AD management involving FOXQ1, PTGS2 and CDK5 [54]. The lncRNA EBF3-AS promotes neuronal apoptosis in AD model mice and affects EBF3 expression, indicating that EBF3-AS has the potential as a new therapeutic target for the treatment of AD [55].

lncRNA NEAT1, reported to be a novel target for Alzheimer’s disease progression via the miR-124/BACE1 axis as mentioned above, can Mediate Microtubule (MT) stabilization by the FZD3/GSK3β/ptau pathway in SH-SY5Y cells and APP/PS1 mice. Specifically, lncRNA NEAT1 mediates the phosphorylation of tau protein through FZD3 [54,56].

lncRNAs involved in Aβ plaque formation

Aβ plays a critical role in the pathophysiology of AD by causing the death of neurons due to its insoluble and neurotoxic characteristics. Aβ usually has two isoforms, Aβ40 and Aβ42, which are generally released from amyloid precursor protein (APP) by β-secretase 1 (BACE1) and γ-secretase. The Aβ42 peptide is more toxic and has faster growth of Aβ42 aggregates. The expression of BACE1 was upregulated in the brain enzymes of patients with AD [57]. It should be noted that BACE1 transcripts are upregulated by the lncRNA BACE1-AS, which is transcribed from the opposite strand of the BACE1 gene and shares complementary sequences, and its expression is mediated in brain and vascular homeostasis [30]. In other words, lncRNA BACE1-AS expression could increase BACE1 mRNA, which in turn increases the production of Aβ [58]. Therefore, lncRNABACE1-AS plays an important role in Aplaque formation.

In addition, lncRNA X-inactive Specific Transcript (XIST) is a functional lncRNA. Yue et al., found that lncRNA XIST expression was upregulated in vivo and in vitro in AD models. The molecular mechanism of lncRNA XIST in AD models showed that silencing of lncRNA XIST downregulated Aβ protein-related BACE1 through miR- 124/BACE1 signaling pathways [59]. Similarly, lncRNA MAGI2-AS3 has been acknowledged as an important factor for Aβ deposition. Studies have shown direct interactions between MAGI2-AS3 and miR-374b- 5p and between BACE1 and miR-374b-5p. Specifically, when SH-SY5Y and BV2 cells were stimulated with Aβ25-35, MAGI2-AS3 expression increased, while miR-374b-5p expression decreased. Reduction of MAGI2-AS3 and overexpression of miR-374b-5p can cause the same effect: enhancement of neuronal survival and impairment of neuroinflammation in AD cell models. These effects can be reversed by inhibiting miR-374b-5p. Furthermore, in AD patients, MAGI2-AS3 and miR-374b-5p blood levels were also negatively correlated. lncRNA MAGI2-AS3 can interact with miR-374b-5p, indicating its critical role in attenuating neurotoxicity and neuroinflammation [60]. On the other hand, lncRNAs 17A, 51A, and NDM29 were reported to increase Aβ formation and/or the Aβx-42/Aβx-40 ratio directly or indirectly, although they may play other roles [61].

lncRNAs involved in tau hyperphosphorylation

AD is also characterized by the hyperphosphorylation of tau protein or Neurofibrillary Tangles (NFTs) in cells. Activated kinases add phosphate groups to tau proteins, which leads to their hyperphosphorylation and NFT formation [62]. These steps are regulated by multiple kinases, primarily Glycogen Synthase Kinase 3 (GSK3β) and Cyclin-Dependent Kinase 5 (CDK5) [63]. lncRNAs, such as the lncRNAs linc-00507, RP11-543N12.1 and SOX21-AS1, participate in tau hyperphosphorylation. The level of linc00507 was significantly increased in both animal and cell models of AD. Linc00507 sponges miR-181c-5p to regulate the Microtubule-associated Protein Tau (MAPT) or Tau-Tubulin Kinase-1 (TTBK1) pathway, which activates the P25/P35/GSK3β signaling pathway. Consequently, linc-00507 mediates tau protein hyperphosphorylation [64].

lncRNAs RP11-543N12.1 and SOX21-AS1 are highly expressed in SH-SY5Y cells following Aβ25-35 treatment and act as a lncRNA/ miRNA axis to promote phosphorylation of tau protein and cell apoptosis during AD progression. lncRNA RP11543N12.1 inhibits proliferation and enhances apoptosis by targeting miR3243p, indicating that RP11543N12.1 and miR3243p may be potential biomarkers and therapeutic targets in AD. lncRNA SOX21-AS1 has been found to sponge miR-107 in SH-SY5Y and SK-N-SH cells. The upregulation of tau protein phosphorylation and apoptosis and the viability decrease can be inhibited by miR-107 in Aβ1-42-treated SH-SY5Y cells. However, lncRNA SOX21-AS1 can reverse this effect of miR-107 [65,66]. The lncRNA NEAT1, reported to be a novel target for Alzheimer’s disease progression via the miR-124/BACE1 axis above, can mediate Microtubule (MT) stabilization by the FZD3/GSK3β/p-tau pathway in SH-SY5Y cells and APP/PS1 mice [54, 56].

lncRNAs involved in Microglia (MG) and astrocyte activation

Neuroinflammation is largely mediated by glial cells, including microglia and astrocytes. Activated microglia and astrocytes have been implicated as key players in the progression of AD [67]. Microglia are the major immune cells of the CNS. In response to stimuli, microglia are activated into two phenotypes, M1 and M2. The M1 type is a neurotoxic phenotype, secreting pro-inflammatory cytokines (TNF-α, IL-16, etc.) and chemokines (CCL2, IL-18, etc.) that promote neuroinflammation and induce neuronal damage and oxidative stress, thereby triggering synaptic loss and accelerating the disease process, the M2 type is a neuroprotective phenotype, releasing arginase 1, IGF-1 and Fzd1, which are involved in neuroprotection and tissue repair. Astrocytes are the most abundant glial cells in the brain, and similar to microglia, which have two phenotypes, neurotoxic and neuroprotective activation states (A1/A2), astrocytes in Alzheimer’s disease have both protective and inflammatory effects on neurons [68].

Some lncRNAs have been found to regulate AD progression by modulating the ratio of neurotoxic and neuroprotective activation states of microglia and astrocytes. The expression of lncRNA MALAT1 was significantly reduced in AD cells, animal models and human AD brain tissues. This occurred possibly through interacting with miRNAs miRs-200a, -26a and -26b, which are generally increased in AD [69]. In addition, lncRNA MALAT1 was dysregulated in Acute Spinal Cord Injury (ASCI) rat models, and its knockdown reduced ASCI possibly through inhibiting the inflammatory response of microglia [70].

Likewise, in AD rats, the expression of lncRNA MEG3 was downregulated. Conversely, upregulation of the lncRNA MEG3 improved cognitive impairment and neuronal damage and prevented astrocyte activation in hippocampal tissues in AD rats by impeding the PI3K/Akt signaling pathway [71]. In addition, several lncRNAs XIST and NEAT1 act as sponges of miR-124, which is specifically expressed in microglia, and switch microglia to an anti-inflammatory state by imbalancing the M1/M2 ratio in CNS macrophages, so these lncRNAs may also be involved in the activation states of microglia [10]. The lncRNA LEF1-AS1 can attenuate apoptosis and improve the viability of LPS-treated Spinal Cord Injury (SCI) microglia, while the lncRNA SNHG14 promotes microglial activation by regulating miR-145-5p/ PLA2G4A in cerebral infarction in a Middle Cerebral Artery Occlusion (MCAO) mouse model and microglia [72,73]. Similarly, lncRNA SNHG5 can promote astrocyte and microglial viability [74]. In addition, LPS-induced microglial neurotoxicity can be decreased by silencing lncRNA nostrill [75]. lncRNA-Cox2 induced from astrocyte-derived Extracellular Vesicles (EVs) involves the impairment of microglial phagocytic activity in mice.

Importantly, lncRNA-11496 was found to contribute to microglial apoptosis in chronic T. gondii-infected mice by targeting Mef2c, which is related to neuronal differentiation, survival, and synapse formation, by binding to Histone Deacetylase 2 (HDAC2) [76]. On the other hand, lncRNA-11496 positively regulated the expression of its target. The overexpression of lncRNA uc.80 improves depression in rats by promoting M2 polarization of microglia [77]. These findings indicate that lncRNA- might be used as a therapeutic target for treating T. gondii-induced mental and behavioral disorders in the brain. Further study on the mechanism is needed. The increase in lincRNACox2 expression led to a reduction in microglial phagocytic activity, which was restored by inhibiting the expression of lincRNA-Cox2 via siRNA interference [78]. LincRNA-EPS can reduce Cerebral Ischemia Reperfusion (CIR) injury, inflammation and oxidative stress by maintaining High Temperature Requirement Factor A1 Htra1 stability through recruiting HNRNPL [79]. lncRNA AK148321 also reduces neuroinflammation in LPS-stimulated BV2 microglia via the miR- 1199-5p/HSA5 axis [80].

lncRNAs involved in the NLRP3 inflammasome

The formation of inflammation is a key aspect of inflammatory responses. Nucleotide-binding Oligomerization domain-like Receptor containing Pyrin Domain 3 (NLRP3) is one of the most studied inflammasomes and is present in microglia and astrocytes in the CNS. Studies have shown that the NLRP3 inflammasome plays an important role in neurodegenerative diseases, particularly Alzheimer’s disease. The NLRP3 inflammasome is a multiprotein complex containing NLRP3, ASC and procaspase-1. Activation of the NLRP3 inflammasome leads to the release of caspase-1, which mediates the production of the proinflammatory cytokines IL-1β and IL-18 and induces several downstream inflammatory factors, thereby initiating inflammatory responses. These inflammatory mediators are closely linked to Aβ deposition and tau aggregation, contributing to the pathogenesis of AD [81].

The lncRNA SNHG14 expression is higher in the cytoplasm than in the nucleus of astrocytes, suggesting that the lncRNA SNHG14 may downregulate gene expression by sponging specific miRNAs.

Bioinformatics analysis predicted a binding site between miR-223-3p and SNHG14 as well as between miR-223-3p and NLRP3. The direct interactions between miR-223-3p and SNHG14, as well as miR-223-3p and NLRP34, were further confirmed by using a luciferase reporter assay. Altogether, it could be concluded that SNHG14 might sponge miR-223- 3p and thus regulate NLRP3 expression at the posttranscriptional level in the AD study [82].

The lncRNA NONRATT004344.2 is a lncRNA that is highly expressed in the brain and is also known as lncRNA 4344. Researchers found that lncRNA 4344 regulates the activated NLRP3 inflammasome in microglia, which contributes to brain function decline and neurodegenerative diseases. The results of the animal experiments showed that lncRNA 4344 sponges miR-138-5p to upregulate the expression of the inflammasome NLRP3. Conclusively, the lncRNA-4344/miR-138-5p/ NLRP3 axis plays an essential role in regulating neuroinflammation and cognitive decline by acting as a sponge of miR-138-5p to downregulate NLRP3 expression [83].

lncRNAsinvolved in oxidative stress

Oxidative stress plays an important role in the pathogenesis of AD. Oxidants and oxidative products can regulate APP and the associated secretase. This increases the expression of APP, leading to the aggregation of Aβ. In turn, irreversibly deposited Aβ disrupts the activity of N-Methyl-D-Aspartate NMDA receptors to promote the production of oxygen free radicals and excessive calcium influx into neurons, thereby inducing neuronal damage. Oxidative stress also directly interacts with the tau protein kinases GSK-3 and CDK5 to promote their hyperphosphorylation. Similar to Aβ protein, free p-tau protein aggregates to form NTFs, leading to significant oxidative stress and neurological impairment [84].

In Aβ25-35-stimulated PC12 cells, silencing of lncRNA H19 and BDNF-AS resulted in significantly increased Superoxide Dismutase (SOD) and Catalase (CAT) expression and decreased Malondialdehyde (MDA) expression, leading to a reversal of the Aβ25-35-induced oxidative stress response, suggesting that lncRNA H19 and BDNF-AS may be involved in AD development and progression via oxidative stress [85]. lncRNA SOX21-AS1 plays an important role in tau hyperphosphorylation and cell apoptosis during AD progression. In another study, scientists also found that silencing the lncRNA SOX21- AS1 alleviated oxidative stress and inhibited apoptosis in hippocampal neurons of AD mice by upregulating FZD3/5 via the Wnt signaling pathway. In other words, the lncRNA SOX21-AS1 can also regulate oxidative stress to participate in cell cycle distribution and apoptosis during AD pathology [86].

Other long noncoding RNAs involved in AD

There are many other very important lncRNAs involved in AD, but their specific biological roles and mechanisms still need to be further investigated. The expression of the long noncoding RNA MALAT1 was previously found to inhibit Cyclic AMP Response Element-binding Protein (CREB) dephosphorylation, a transcription factor that regulates cell proliferation, by PP2A. Furthermore, lower levels of MALAT1 were measured in the CSF of AD patients than in the CSF of control patients. In addition, dysregulated MALAT1 was also found in neurodegenerative and neuro-oncological disorders, indicating an important role of lncRNA MALAT1 in neurodegenerative diseases, including AD [87]. A lncRNA activated by TGF-beta, called lncRNA-ATB, was elevated in the CSF and serum of patients with AD. Prohibiting lncRNA-ATB can protect PC12 cells from neurotoxicity induced by Aβ25-35 through regulating the miR-200/ZNF217 axis [88]. In addition, lncRNAs RP11-462G22.1 and PCA3 were also upregulated in the CSF exosomes of both PD and AD patients. Some lncRNAs, such as lncRNA BC200 and NATRad18, were discovered as early as 2007. The lncRNA BC200 found in the AD patient brain may play a critical role in maintaining long-term synaptic plasticity during AD progression by regulating local protein synthesis [89]. NAT-Rad18 expressed in rat cortical neurons was previously shown to be involved in apoptosis [90] (Figure 1).

cellullar-biology-plague-125155

Figure 1: LncRNAs as potential biomarkers or therapeutic targets in Alzheimer’s Disease (AD).

A systematic study on the lncRNA profile investigated a total of 14 upregulated lncRNAs and 20 downregulated lncRNAs in Peripheral Blood Mononuclear Cells (PBMCs) of AD patients compared with those in the control groups (fold change ≥ 2.0, p<0.05). Among them, lncRNAs TTC39C-AS1, lnc-AL445989.1-2, LINC01420, lnc-CSTB-1 and LOC401557 may play critical roles in AD pathogenesis. Notably, the most significant one is the antisense transcript of the TTC39CAS1 gene, which is closely related to neurogenic atrophy [91,92]. In addition, hippocampal lncRNA profiles in AD rats were reported. A total of 315 lncRNAs were found to be significantly dysregulated in the AD model rats (≥ 2.0-fold, p<0.05). The expression profiles of lncRNAs in the hippocampus of intranasal LPS-induced AD model mice were determined by microarray analysis. A total of 395 lncRNAs were differentially expressed in AD rat and mouse models compared with controls (>2.0-fold, p<0.05) [93,94].

Other lncRNAs reported in AD patients, mice, rat or cells are BDNF-AS, upregulated in peripheral blood of AD patients; ANRIL, knockdown of it reduces Aβ-induced apoptosis, inflammation and autophagy in PC12 cells; HOTAIR, downregulated in the brain by exercise in rat model; N336694, upregulated in the brain of AD model mice; PART1, increased by Aβ in endothelial cells; SNHG1, increased by A25-35 in SH-SY5Y cells; SOX2-OT, lincRNA-p21, upregulated in AD model mice; Rpph1, upregulated in the cortex of AD model mice [95-97]. There are still deregulated lncRNAs, such as TTC39C-AS1, LOC401557, CH507-513H4.4, CH507-513H4.6, CH507-513H4.3, RP11-462G22.1, and PCA3, in AD patients [52].

Conclusion

Over the past two decades, the role of miRNAs has been comprehensively studied. lncRNAs account for the majority of ncRNAs and have been found to be involved in various biological processes and diseases. lncRNA XIST is involved in colon cancer, lncRNA PRINS in Monoclonal Gammopathies (MGUS), lnc-SOX6-1 in Acute Myeloid Leukemia (AML), lncRNA SNHG20 in Epithelial Ovarian Cancer (EOC), lncRNA-HOTAIR in ovarian cancer, lncRNA H19, lncRNA MEG3 and lncRNA MAPT-AS1 in breast cancer, lncRNA-NEF in liver cancer, lncRNAs TINCR and lncRNAs CARLo-7 in bladder cancer, and lncRNA SPRY4 in lung cancer. There are certainly more unknown lncRNAs that contribute to cancers.

In this review, we focused on summarizing lncRNAs and their critical potential roles in key pathological features of AD, such as Aβ plaques, tau hyperphosphorylation and neuroinflammation, as well as their interactions. As described above, these lncRNAs directly or indirectly regulate vital genes or proteins that may play an important role in the development of AD. Therefore, studying these lncRNAs and their disease-related regulatory mechanisms may provide new ideas for identifying effective treatments for AD. Seventy-five percent of AD patients worldwide have not been diagnosed, mostly in underdeveloped countries. Further study on lncRNAs as biomarkers or intervention targets may provide new insights into diagnosing and treating AD.

Notably, some lncRNA studies have only been carried out in the brains of laboratory animals and or cell models, but these models have not been able to accurately recapitulate the various features of AD. Therefore, there may be significant difficulties in translating results from AD models to humans. In-depth studies investigating more lncRNAs that have essential biological roles in the development of AD and studying their mechanism are still the major topic in the future. Furthermore, research on lncRNA biomarkers and lncRNA-targeted drugs is also very important. However, the application of regulatory lncRNAs in AD still has many challenges to overcome. For example, due to insufficient research, unexpected risks and inappropriate pathological effects may occur when lncRNA-targeted drugs are used clinically. Nevertheless, we believe that with continued innovation in experimental methods and techniques, the application of safe and effective lncRNA-associated biomarkers or drugs for the treatment of AD is possible.

References

  1. Nurk S, Koren S, Rhie A, Rautiainen M, Bzikadze AV, et al. (2022) The complete sequence of a human genome. Science 376: 44-53.

    [Crossref][Google Scholar]

  2. Saw PE, Xu X, Chen J, Song EW (2021) Non-coding RNAs: The new central dogma of cancer biology. Sci China Life Sci 64:22-50.

    [Crossref][Google Scholar][Pubmed]

  3. Yan H, Bu P (2021) Non-coding RNAin cancer. Essays Biochem 65:625-639.

    [Crossref][Google Scholar][Pubmed]

  4. Fasolo F, Di Gregoli K, Maegdefessel L, Johnson JL (2019) Non-coding RNAs in cardiovascular cell biology and atherosclerosis. Cardiovasc Res 115:1732-1756 .

    [Crossref][Google Scholar][Pubmed]

  5. Wang J, Yan S, Yang J, Lu H, Xu D, et al. (2020) Non-coding RNAs in rheumatoid arthritis: From bench to bedside. Front Immunol 10:3129.

    [Crossref][Google Scholar][Pubmed]

  6. Wang Y, Luo Y, Yao Y, Ji Y, Feng L, et al. (2020) Silencing the lncRNA Maclpil in pro-inflammatory macrophages attenuates acute experimental ischemic stroke via LCP1 in mice. J Cereb Blood Flow Metab 40:747-759.

    [Crossref][Google Scholar][Pubmed]

  7. Jia H, Ma H, Li Z, Chen F, Fang B, et al. (2019) Downregulation of LncRNATUG1 inhibited TLR4 signaling pathway-mediated inflammatory damage after spinal cord ischemia reperfusion in rats via suppressing TRIL expression. J Neuropathol Exp Neurol 78:268-282.

    [Crossref][Google Scholar][Pubmed]

  8. Wang J, Cao B, Zhao H, Gao Y, Luo Y, et al. (2019) Long noncoding RNA H19 prevents neurogenesis in ischemic stroke through p53/Notch1 pathway. Brain Res Bull 150:111-117.

    [Crossref][Google Scholar][Pubmed]

  9. Ghafouri-Fard S, Shoorei H, Taheri M (2020) Non-coding RNAsparticipate in the ischemia-reperfusion injury. Biomed Pharmacother 129:110419.

    [Crossref][Google Scholar][Pubmed]

  10. Liu Y, Cheng X, Li H, Hui S, Zhang Z, et al. (2022) Non-coding rnas as novel regulators of neuroinflammation in alzheimer’s disease. Front Immunol 13:908076.

    [Crossref][Google Scholar][Pubmed]

  11. Sun N, Zhang G, Liu Y. (2018) Long non-coding RNA XIST sponges miR-34a to promotes colon cancer progression via Wnt/β-catenin signaling pathway. Gene 665:141-148.

    [Crossref][Google Scholar][Pubmed]

  12. Sedlarikova L, Bollova B, Radova L, Brozova L, Jarkovsky J, et al. (2018) Circulating exosomal long noncoding RNA PRINS—First findings in monoclonal gammopathies. Hematol Oncol 36:786-791.

    [Crossref][Google Scholar][Pubmed]

  13. Guan X, Wen X, Xiao J, An X, Yu J, et al. (2019) Lnc‐SOX6‐1 upregulation correlates with poor risk stratification and worse treatment outcomes, and promotes cell proliferation while inhibits apoptosis in pediatric acute myeloid leukemia. Int J Lab Hematol 41:234-241.

    [Crossref][Google Scholar][Pubmed]

  14. Wang D, Dai J, Hou S,Qian Y (2019) LncRNA SNHG20 predicts a poor prognosis and promotes cell progression in epithelial ovarian cancer. Biosci Rep 39:BSR20182186.

    [Crossref][Google Scholar][Pubmed]

  15. Saeedi N, Ghorbian S (2020) Analysis of clinical important of LncRNA-HOTAIR gene variations and ovarian cancer susceptibility. Mol Biol Rep 47:7421-7427.

    [Crossref][Google Scholar][Pubmed]

  16. Safari MR, Rezaei FM, Dehghan A, Noroozi R, Taheri M, et al. (2019) Genomic variants within the long non-coding RNA H19 confer risk of breast cancer in Iranian population. Gene 701:121-124.

    [Crossref][Google Scholar][Pubmed]

  17. Bayarmaa B, Wu Z, Peng J, Wang Y, Xu S, et al. (2019) Association of LncRNA MEG3 polymorphisms with efficacy of neoadjuvant chemotherapy in breast cancer. BMC cancer 19:1-8.

    [Crossref][Google Scholar][Pubmed]

  18. Liang Z, Zhu B, Meng D, Shen X, Li X, et al. (2019) Down-regulation of lncRNA-NEF indicates poor prognosis in intrahepatic cholangiocarcinoma. Biosci Rep 39:BSR20181573.

    [Crossref][Google Scholar][Pubmed]

  19. Huang H, Fan X, Zhang X, Xie Y, Ji (2020) LncRNA CARLo-7facilitates proliferation, migration, invasion, and EMT of bladder cancer cells by regulating Wnt/β-catenin and JAK2/STAT3 signaling pathways. Transl Androl Urol 9:2251.

    [Crossref][Google Scholar][Pubmed]

  20. Xu C, Liu M, Jia D, Tao T, Hao D (2021) lncRNA TINCR SNPsand expression levels are associated with bladder cancer susceptibility. Genet Test Mol Biomarkers 25:31-41.

    [Crossref][Google Scholar][Pubmed]

  21. Seal RL, Chen LL, Griffiths‐Jones S, Lowe TM, Mathews MB, et al. (2020) A guide to naming human non‐coding RNA genes. EMBOJ39:e103777.

    [Crossref][Google Scholar][Pubmed]

  22. Stojic L (2020) Tuning the Expression of Long Noncoding RNA Loci with CRISPR Interference. RNA-Chromatin Interactions: Methods Mol Biol 1-6.

    [Crossref][Google Scholar][Pubmed]

  23. Zhou S, Yu X, Wang M, Meng Y, Song D, et al. (2021) Long non-coding RNAsin pathogenesis of neurodegenerative diseases. Front Cell Dev Biol 9:719247.

    [Crossref][Google Scholar][Pubmed]

  24. Srinivas T, Mathias C, Oliveira-Mateos C, Guil S (2023) Roles for lncRNAs in brain development and pathogenesis: emerging therapeutic opportunities. Mol Ther

    [Crossref][Google Scholar][Pubmed]

  25. Ghafouri-Fard S, Safari M, Taheri M, Samadian M (2022) Expression of linear and circular lncRNAs in Alzheimer’s disease. J Mol Neurosci 72:187-200.

    [Crossref][Google Scholar][Pubmed]

  26. Zhang H, Liu X, Liu Y, Liu J, Gong X, et al. (2022) Crosstalk between regulatory non-coding RNAs and oxidative stress in Parkinson’s disease. Front Aging Neurosci 14:975248.

    [Crossref][Google Scholar][Pubmed]

  27. Ruffo P, De Amicis F, Giardina E, Conforti FL. (2023) Long-noncoding RNAsas epigenetic regulators in neurodegenerative diseases. Neural Regen Res 18:1243.

    [Crossref][Google Scholar][Pubmed]

  28. Tan X, Liu Y, Liu Y, Zhang T, Cong S (2021) Dysregulation of long non‐coding RNAs and their mechanisms in Huntington's disease. J Neurosci Res 9:2074-2090.

    [Crossref][Google Scholar][Pubmed]

  29. Sivagurunathan N, Ambatt AT, Calivarathan L. (2022) Role of Long Non-coding RNAs in the Pathogenesis of Alzheimer’s and Parkinson’s Diseases. Curr Aging Sci 15:84-96.

    [Crossref][Google Scholar][Pubmed]

  30. Fotuhi SN, Khalaj-Kondori M, Hoseinpour Feizi MA, Talebi M (2019) Long non-coding RNA BACE1-ASmay serve as an Alzheimer’s disease blood-based biomarker. J Mol Neurosci 69:351-359.

    [Crossref][Google Scholar][Pubmed]

  31. Wang D, Gao H, Li Y, Jiang S, Yong Y, et al. (2020) Genome-scale expression pattern of long non-coding RNAs in Chinese Uyghur patients with Parkinson’s disease. Med Sci Monit 26:e925888-1.

    [Crossref][Google Scholar][Pubmed]

  32. Pan B, Shi ZJ, Yan JY, Li JH, Feng SQ (2017) Long non-coding RNA NONMMUG014387 promotes Schwann cell proliferation after peripheral nerve injury. Neural Regen Res 12:2084.

    [Crossref][Google Scholar][Pubmed]

  33. Deng Y, Chen D, Gao F, Lv H, Zhang G, et al. (2020) Silencing of long non-coding RNA GAS5 suppresses neuron cell apoptosis and nerve injury in ischemic stroke through inhibiting DNMT3B-dependent MAP4K4 methylation. Transl Stroke Res 950-966.

    [Crossref][Google Scholar][Pubmed]

  34. Sun M, Liu XH, Lu KH, Nie FQ, Xia R, et al. (2014) EZH2-mediated epigenetic suppression of long noncoding RNA SPRY4-IT1 promotes NSCLC cell proliferation and metastasis by affecting the epithelial–mesenchymal transition. Cell Death Dis 5:e1298.

    [Crossref][Google Scholar][Pubmed]

  35. Ma Y, Zhai D, Zhang W, Zhang H, Dong L, et al. (2020) Down‐regulation of long non‐coding RNA MEG3 promotes Schwann cell proliferation and migration and repairs sciatic nerve injury in rats. J Cell Mol Med 24:7460-7469.

    [Crossref][Google Scholar][Pubmed]

  36. Wang D, Li J, Cai F, Xu Z, Li L, et al. (2019) Overexpression of MAPT-AS1 is associated with better patient survival in breast cancer. Biochem Cell Biol 97:158-164.

    [Crossref][Google Scholar][Pubmed]

  37. Shen F, Zheng H, Zhou L, Li W, Zhang Y et al. (2019) LINC00657 expedites neuropathic pain development by modulating miR‐136/ZEB1 axis in a rat model. J Cell Biochem 120:1000-1010

    [Crossref][Google Scholar][Pubmed]

  38. Zhang Z, Sun X, Zhao G, Ma Y, Zeng G. (2021) LncRNA embryonic stem cells expressed 1 (Lncenc1) is identified as a novel regulator in neuropathic pain by interacting with EZH2 and downregulating the expression of Bai1 in mouse microglia. Exp Cell Res 399:112435.

    [Crossref][Google Scholar][Pubmed]

  39. Rey F, Maghraby E, Messa L, Esposito L, Barzaghini B, et al. (2023) Identification of a novel pathway in sporadic Amyotrophic Lateral Sclerosis mediated by the long non-coding RNA ZEB1-AS1. Neurobiol Dis 178:106030.

    [Crossref][Google Scholar][Pubmed]

  40. Breijyeh Z, Karaman R (2020) Comprehensive review on Alzheimer’s disease: Causes and treatment. Molecules 25:5789.

    [Crossref][Google Scholar][Pubmed]

  41. Busche MA, Hyman BT (2020) Synergy between amyloid-β and tau in Alzheimer’s disease. Nat Neurosci 23:1183-1193.

    [Crossref][Google Scholar][Pubmed]

  42. Moujalled D, Strasser A, Liddell JR (2021) Molecular mechanisms of cell death in neurological diseases. Cell Death Differ 28(7):2029-2044.

    [Crossref][Google Scholar][Pubmed]

  43. Tiwari S, Atluri V, Kaushik A, Yndart A, Nair M (2019) Alzheimer’s disease: pathogenesis, diagnostics, and therapeutics. Int J Nanomedicine19:5541-5554.

    [Crossref][Google Scholar][Pubmed]

  44. Rajesh Y, Kanneganti TD (2022) Innate immune cell death in neuroinflammation and Alzheimer’s disease. Cells 11:1885.

    [Crossref][Google Scholar][Pubmed]

  45. Cai Y, Liu J, Wang B, Sun M, Yang H (2022) Microglia in the neuroinflammatory pathogenesis of Alzheimer’s disease and related therapeutic targets. Front Immunol 13:856376.13.

    [Crossref][Google Scholar][Pubmed]

  46. Whitson HE, Colton C, El Khoury J, Gate D, Goate A, et al. (2022) Infection and inflammation: New perspectives on Alzheimer's disease. Brain Behav Immun Health 22:100462.

    [Crossref][Google Scholar][Pubmed]

  47. Teleanu DM, Niculescu AG, Lungu II, Radu CI, Vladâcenco O, et al. (2022) An overview of oxidative stress, neuroinflammation, and neurodegenerative diseases. Int J Mol Sci 23(11):5938.

    [Crossref][Google Scholar][Pubmed]

  48. Jack Jr CR, Albert MS, Knopman DS, McKhann GM, Sperling RA, et al. (2011) Introduction to the recommendations from the National Institute on Aging-Alzheimer's Association workgroups on diagnostic guidelines for Alzheimer's disease. Alzheimers Dement 7:257-262.

    [Crossref][Google Scholar][Pubmed]

  49. Swarbrick S, Wragg N, Ghosh S, Stolzing A (2019) Systematic review of miRNA as biomarkers in Alzheimer’s disease. Mol Neurobiol 56:6156-6167.

    [Crossref][Google Scholar][Pubmed]

  50. Zetterberg H, Burnham SC (2019) Blood-based molecular biomarkers for Alzheimer’s disease. Mol Brain 12:1-7.

    [Crossref][Google Scholar][Pubmed]

  51. Feng L, Liao YT, He JC, Xie CL, Chen SY, et al. (2018) Plasma long non-coding RNA BACE1 as a novel biomarker for diagnosis of Alzheimer disease. BMC Neurol 18:1-8.

    [Crossref][Google Scholar][Pubmed]

  52. Garofalo M, Pandini C, Sproviero D, Pansarasa O, Cereda C, et al. (2021) Advances with long non-coding rnas in alzheimer’s disease as peripheral biomarker Genes 12:1124.

    [Crossref][Google Scholar][Pubmed]

  53. Deng Y, Xiao L, Li W, Tian M, Feng X, et al. (2017) Plasma long noncoding RNA 51A as a stable biomarker of Alzheimer's disease. Int J Clin Exp Pathol 10(4):4694-4699.

    [Google Scholar]

  54. Zhuang J, Cai P, Chen Z, Yang Q, Chen X, et al (2020). Long noncoding RNA MALAT1 and its target microRNA-125b are potential biomarkers for Alzheimer’s disease management via interactions with FOXQ1, PTGS2 and CDK5. Am J Transl Res 12(9):5940.

    [Google Scholar][Pubmed]

  55. Gu C, Chen C, Wu R, Dong T, Hu X, et al. (2018) Long noncodingRNA EBF3-AS promotes neuron apoptosis in Alzheimer's disease. DNA Cell Biol 37:220-226.

    [Crossref][Google Scholar][Pubmed]

  56. Zhao Y, Wang Z, Mao Y, Li B, Zhu Y, et al. (2020) NEAT1 regulates microtubule stabilization via FZD3/GSK3β/P-tau pathway in SH-SY5Ycells and APP/PS1 mice. Aging (Albany NY) 12:23233-23250.

    [Crossref][Google Scholar][Pubmed]

  57. Zeng T, Ni H, Yu Y, Zhang M, Wu M, et al. (2019) BACE1-ASprevents BACE1 mRNA degradation through the sequestration of BACE1-targeting miRNAs. J Chem Neuroanat 98:87-96.

    [Crossref][Google Scholar][Pubmed]

  58. Sayad A, Najafi S, Hussen BM, Abdullah ST, Movahedpour A, et al. (2022) The emerging roles of the beta-secretase BACE1 and the long non-coding RNA BACE1-AS in human diseases: A focus on neurodegenerative diseases and cancer. Front Aging Neurosci 14:853180.

    [Crossref][Google Scholar][Pubmed]

  59. Yue D, Guanqun G, Jingxin L, Sen S, Shuang L, et al. (2020) Silencing of long noncoding RNA XISTattenuated Alzheimer's disease-related BACE1 alteration through miR-124. Cell Biol Int 44:630-636.

    [Crossref][Google Scholar][Pubmed]

  60. Zhang J, Wang R (2021) Deregulated lncRNA MAGI2-AS3 in Alzheimer's disease attenuates amyloid-beta induced neurotoxicity and neuroinflammation by sponging miR-374b-5p. Exp Gerontol 144:111180.

    [Crossref][Google Scholar][Pubmed]

  61. Luo Q, Chen Y (2016) Long noncoding RNAs and Alzheimer's disease. Clin Interv Aging 11:867-872.

    [Crossref][Google Scholar][Pubmed]

  62. Roda AR, Serra-Mir G, Montoliu-Gaya L, Tiessler L, Villegas S (2022) Amyloid-beta peptide and tau protein crosstalk in Alzheimer’s disease. Neural Regen Res 17:1666-1674.

    [Crossref][Google Scholar][Pubmed]

  63. Hisanaga SI, Krishnankutty A, Kimura T (2022) In vivo analysis of the phosphorylation of tau and the tau protein kinases Cdk5-p35 and GSK3beta by using Phos-tag SDS-PAGE. J Proteomics 262:104591.

    [Crossref][Google Scholar][Pubmed]

  64. Yan Y, Yan H, Teng Y, Wang Q, Yang P, et al. (2020) Long non-coding RNA 00507/miRNA-181c-5p/TTBK1/MAPT axis regulates tau hyperphosphorylation in Alzheimer's disease. J Gene Med 22:e3268.

    [Crossref][Google Scholar][Pubmed]

  65. Cai M, Wang YW, Xu SH, Qiao S, Shu QF, et al. (2018) Regulatory effects of the long non‑coding RNA RP11‑543N12.1 and microRNA‑324‑3p axis on the neuronal apoptosis induced by the inflammatory reactions of microglia. Int J Mol Med 42:1741-1755.

    [Crossref][Google Scholar][Pubmed]

  66. Xu W, Li K, Fan Q, Zong B, Han L. (2020) Knockdown of long non-coding RNA SOX21-AS1 attenuates amyloid-beta-induced neuronal damage by sponging miR-107. Biosci Rep 40:BSR20194295.

    [Crossref][Google Scholar][Pubmed]

  67. Merighi S, Nigro M, Travagli A, Gessi S (2022) Microglia and Alzheimer's disease. Int J Mol Sci 23:12990.

    [Crossref][Google Scholar][Pubmed]

  68. Jiang H, Zhang Y, Yue J, Shi Y, Xiao B, et al. (2022) Non-coding RNAs: The neuroinflammatory regulators in neurodegenerative diseases. Front Neurol 13:929290.

    [Crossref][Google Scholar][Pubmed]

  69. Chanda K, Jana NR, Mukhopadhyay D (2022) Long non-coding RNA MALAT1 protects against Abeta(1-42) induced toxicity by regulating the expression of receptor tyrosine kinase EPHA2 via quenching miR-200a/26a/26b in Alzheimer's disease. Life Sci 302:120652.

    [Crossref][Google Scholar][Pubmed]

  70. Zhou HJ, Wang LQ, Wang DB, Yu JB, Zhu Y, et al. (2018) Long noncoding RNA MALAT1 contributes to inflammatory response of microglia following spinal cord injury via the modulation of a miR-199b/IKKbeta/NF-kappaB signaling pathway. Am J Physiol Cell Physiol 315:C52-C61.

    [Crossref][Google Scholar][Pubmed]

  71. Yi J, Chen B, Yao X, Lei Y, Ou F, et al. (2019) Upregulation of the lncRNA MEG3 improves cognitive impairment, alleviates neuronal damage, and inhibits activation of astrocytes in hippocampus tissues in Alzheimer's disease through inactivating the PI3K/Akt signaling pathway. J Cell Biochem 120:18053-18065.

    [Crossref][Google Scholar][Pubmed]

  72. Cui SY, Zhang W, Cui ZM, Yi H, Xu DW, et al. (2021) Knockdown of long non-coding RNA LEF1-AS1attenuates apoptosis and inflammatory injury of microglia cells following spinal cord injury. J Orthop Surg Res 16:6.

    [Crossref][Google Scholar][Pubmed]

  73. Qi X, Shao M, Sun H, Shen Y, Meng D, et al. (2017) Long non-coding RNA SNHG14 promotes microglia activation by regulating miR-145-5p/PLA2G4A in cerebral infarction. Neuroscience 348:98-106.

    [Crossref][Google Scholar][Pubmed]

  74. Jiang ZS, Zhang JR (2018) LncRNA SNHG5 enhances astrocytes and microglia viability via upregulating KLF4in spinal cord injury. Int J Biol Macromol 120:66-72.

    [Crossref][Google Scholar][Pubmed]

  75. Mathy NW, Burleigh O, Kochvar A, Whiteford ER, Behrens M, et al. (2021) A novel long intergenic non-coding RNA, nostrill, regulates iNOS gene transcription and neurotoxicity in microglia. J Neuroinflammation 18:16.

    [Crossref][Google Scholar][Pubmed]

  76. Sun X, Wang T, Wang Y, Ai K, Pan G, et al (2020) Downregulation of lncRNA-11496 in the brain contributes to microglia apoptosis via regulation of Mef2c in chronic T. gondii infection mice. Front Mol Neurosci 13:77.

    [Crossref][Google Scholar][Pubmed]

  77. Gu XH, Xu LJ, Zheng LL, Yang YJ, Tang ZY, et al (2020) Long non‐coding RNA uc. 80‐overexpression promotes M2 polarization of microglias to ameliorate depression in rats. IUBMB life. 72:2194-203.

    [Crossref][Google Scholar][Pubmed]

  78. Hu G, Liao K, Niu F, Yang L, Dallon BW, et al (2018) Astrocyte EV-induced lincRNA-Cox2 regulates microglial phagocytosis: Implications for morphine-mediated neurodegeneration. Mol Ther Nucleic Acids 13:450-63.

    [Crossref][Google Scholar][Pubmed]

  79. Guo H, Guo X, Jiang S (2022) Long non-coding RNA lincRNA-erythroid prosurvival (EPS) alleviates cerebral ischemia/reperfusion injury by maintaining high-temperature requirement protein A1 (Htra1) stability through recruiting heterogeneous nuclear ribonucleoprotein L (HNRNPL). Bioengineered 13:12248-60.

    [Crossref][Google Scholar][Pubmed]

  80. Gao S, Cheng QC, Hu YG, Tan ZZ, Chen L, et al (2021) LncRNA AK148321 alleviates neuroinflammation in LPS-stimulated BV2 microglial cell through regulating microRNA-1199-5p/HSPA5 axis. Life Sci 266:118863.

    [Crossref][Google Scholar][Pubmed]

  81. Barczuk J, Siwecka N, Lusa W, Rozpędek-Kamińska W, Kucharska E, et al. (2022) Targeting NLRP3-mediated Neuroinflammation in Alzheimer’s disease treatment. Int J Mol Sci 23:8979.

    [Crossref][Google Scholar][Pubmed]

  82. Duan R, Wang SY, Wei B, Deng Y, Fu XX, et al. (2021) Angiotensin-(1–7) analogue AVE0991 modulates astrocyte-mediated neuroinflammation via lncRNA SNHG14/miR-223-3p/NLRP3 pathway and offers neuroprotection in a transgenic mouse model of Alzheimer's disease. J Inflamm Res 18:7007-19.

    [Crossref][Google Scholar][Pubmed]

  83. Feng X, Zhan F, Luo D, Hu J, Wei G, et al. (2021) LncRNA 4344 promotes NLRP3-related neuroinflammation and cognitive impairment by targeting miR‐138-5p. Brain Behav Immun 98:283-98.

    [Crossref][Google Scholar][Pubmed]

  84. Bai R, Guo J, Ye XY, Xie Y, Xie T (2022) Oxidative stress: The core pathogenesis and mechanism of Alzheimer’s disease. Ageing Res Rev 77:101619.

    [Crossref][Google Scholar][Pubmed]

  85. Xu X, Zhang Y (2022) Regulation of oxidative stress by long non-coding RNAs in central nervous system disorders. Front Mol Neurosci 15:931704.

    [Crossref][Google Scholar][Pubmed]

  86. Zhang L, Fang Y, Cheng X, Lian YJ, Xu HL (2019) Silencing of long noncoding RNA SOX21-AS1 relieves neuronal oxidative stress injury in mice with Alzheimer’s disease by upregulating FZD3/5 via the Wnt signaling pathway. Mol Neurobiol 56:3522-37.

    [Crossref][Google Scholar][Pubmed]

  87. Yao J, Wang XQ, Li YJ, Shan K, Yang H, et al. (2016) Long non‐coding RNA MALAT 1 regulates retinal neurodegeneration through CREB signaling. EMBO Mol Med 8:346-62.

    [Crossref][Google Scholar][Pubmed]

  88. Wang J, Zhou T, Wang T, Wang B (2018) Suppression of lncRNA-ATB prevents amyloid-β-induced neurotoxicity in PC12 cells via regulating miR-200/ZNF217axis. Biomed Pharmacother 108:707-15.

    [Crossref][Google Scholar][Pubmed]

  89. Mus E, Hof PR, Tiedge H (2007) Dendritic BC200 RNA in aging and in Alzheimer's disease. Proc Natl Acad Sci U S A 104:10679-84.

    [Crossref][Google Scholar][Pubmed]

  90. Parenti R, Paratore S, Torrisi A, Cavallaro S (2007) A natural antisense transcript against Rad18, specifically expressed in neurons and upregulated during β‐amyloid‐induced apoptosis. Eur J Neurosci 26:2444-57.

    [Crossref][Google Scholar][Pubmed]

  91. Kurt S, Tomatir AG, Tokgun PE, Oncel C (2020) Altered expression of long non-coding RNAs in peripheral blood mononuclear cells of patients with Alzheimer’s disease. Molecular neurobiology 57:5352-5361.

    [Crossref][Google Scholar][Pubmed]

  92. Gui Y, Liu H, Zhang L, Lv W, Hu X (2015) Altered microRNA profiles in cerebrospinal fluid exosome in Parkinson disease and Alzheimer disease. Oncotarget 6:37043.

    [Crossref][Google Scholar][Pubmed]

  93. Aquino R, de Concini V, Dhenain M, Lam S, Gosset D, et al. (2023) Intrahippocampal Inoculation of Aβ1–42 Peptide in Rat as a Model of Alzheimer’s Disease Identified MicroRNA-146a-5p as Blood Marker with Anti-Inflammatory Function in Astrocyte Cells. Cells 12:694.

    [Crossref][Google Scholar][Pubmed]

  94. Tang L, Liu L, Li G, Jiang P, Wang Y, et al. (2019) Expression profiles of long noncoding RNAs in intranasal LPS-mediated Alzheimer’s disease model in mice. Biomed Res Int

    [Crossref][Google Scholar][Pubmed]

  95. Canseco-Rodriguez A, Masola V, Aliperti V, Meseguer-Beltran M, Donizetti A, et al. (2022) Long Non-Coding RNAs, Extracellular Vesicles and Inflammation in Alzheimer’s Disease. Int J Mol Sci 23:13171.

    [Crossref][Google Scholar][Pubmed]

  96. Chen Y, Li Z, Chen X, Zhang S (2021) Long non-coding RNAs: From disease code to drug role. Acta Pharm Sin B 11:340-354.

    [Crossref][Google Scholar][Pubmed]

  97. Chen ZY, Zhang Y (2022) Animal models of Alzheimer’s disease: Applications, evaluation, and perspectives. Zool Res 43:1026-1040.

    [Crossref][Google Scholar][Pubmed]

Citation: Aodeng Q, Su D, Bai Y, Yang X.X, Wuhan Q (2024) Progress on the Long Non-coding RNAs involved in Alzheimer’s Disease. Cell Mol Biol S1:001.

Copyright: © 2024 Aodeng Qimuge et al. This is an open-access article distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited.

Top