Next Article in Journal
Addition of Natural Extracts with Antioxidant Function to Preserve the Quality of Meat Products
Next Article in Special Issue
Stereological Evidence of Non-Selective Hippocampal Neurodegeneration, IGF-1 Depletion, and Behavioral Deficit following Short Term Bilateral Adrenalectomy in Wistar Rats
Previous Article in Journal
High-Resolution Ribosome Profiling Reveals Gene-Specific Details of UGA Re-Coding in Selenoprotein Biosynthesis
Previous Article in Special Issue
Aberrant Ganglioside Functions to Underpin Dysregulated Myelination, Insulin Signalling, and Cytokine Expression: Is There a Link and a Room for Therapy?
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Diverse and Composite Roles of miRNA in Non-Neuronal Cells and Neuronal Synapses in Alzheimer’s Disease

1
Department of Biomedical Engineering, City University of Hong Kong, Hong Kong 999077, China
2
Hong Kong Centre for Cerebro-Cardiovascular Health Engineering (CoCHE), Hong Kong Science Park, Hong Kong 999077, China
3
Department of Mechanical and Automation Engineering, The Chinese University of Hong Kong, Hong Kong 999077, China
4
School of Biomedical Sciences, The Chinese University of Hong Kong, Hong Kong 999077, China
*
Author to whom correspondence should be addressed.
Biomolecules 2022, 12(10), 1505; https://doi.org/10.3390/biom12101505
Submission received: 30 August 2022 / Revised: 28 September 2022 / Accepted: 10 October 2022 / Published: 17 October 2022

Abstract

:
Neurons interact with astrocytes, microglia, and vascular cells. These interactions become unbalanced in disease states, resulting in damage to neurons and synapses, and contributing to cognitive impairment. Importantly, synaptic loss and synaptic dysfunction have been considered for years as a main pathological factor of cognitive impairment in Alzheimer’s disease (AD). Recently, miRNAs have emerged as essential regulators of physiological and pathological processes in the brain. Focusing on the role of miRNAs in regulating synaptic functions, as well as different cell types in the brain, offers opportunities for the early prevention, diagnosis, and potential treatment of AD-related cognitive impairment. Here, we review the recent research conducted on miRNAs regulating astrocytes, microglia, cerebrovasculature, and synaptic functions in the context of AD-related cognitive impairment. We also review potential miRNA-related biomarkers and therapeutics, as well as emerging imaging technologies relevant for AD research.

1. Introduction

MicroRNAs (miRNAs) are a class of non-coding RNAs (about 21 nucleotides) that play a crucial role in regulating gene expression in eukaryotes and are highly conserved [1]. miRNAs are transcribed from DNA sequences by RNA polymerase II (Pol II) into primary miRNAs (pri-miRNA) with local hairpin structures embedded with miRNA sequences, then precursor miRNAs, and eventually mature miRNAs [1]. The miRNA-induced silencing complex (miRISC) is a core component of the miRNA-mediated gene silencing process, and its primary function is to discriminate and repress miRNA-targeted mRNAs [2]. The centerpiece structure of miRISC is the GW182 protein and the Argonaute (Ago) protein loaded with mature miRNA [3]. GW182 has contributed to miRNA-mediated target-gene-expression silencing. The functional silencing of target gene expression by the Ago–miRNA interaction is disturbed when miRNAs are not entirely complementary to their target mRNAs [4]. GW182 binds to and silences target mRNAs independent of Ago to ensure translational repression, decapping, deadenylation, and mRNA degradation [3]. Interestingly, the strength of the GW182–Argonaute interaction is affected by whether the guide RNA is loaded on the Ago [5]. Multiple reports have revealed that miRNAs can alternatively target the miRNA-recognition elements (MREs) in protein-coding sequence (CDS) regions in an Ago-dependent but GW182-independent manner in order to suppress translation by inducing transient ribosomal arrest rather than mRNA destabilization, resulting in gene silencing [6]. A recent study on Drosophila suggests that the middle region domain of GW182 significantly contributes to circular RNA (circRNA) degradation. GW182 may regulate at least a subset of circRNA degradation in an Ago-slicer or P-body-independent manner [7]. In this review, we discuss the diverse roles of miRNA in non-neuronal cells and neuronal synapses in AD.

2. Role of miRNAs in Regulating Astrocytes and Microglia in AD

Astrocytes are the most abundant cell type in the brain [8]. The roles of astrocytes in AD depend on the neuronal environment. Reactive astrocytes are classified as A1 and A2. A1 astrocytes promote neuroinflammation by producing massive amyloid precursor protein-cleaving enzymes, β-secretase (BACE1) and γ-secretase, causing a high production of Aβ and secreting inflammatory factors, such as IFN γ, IL-1β, TNF-α, IL–6, and TGF-β [9] (Figure 1a). Conversely, A2 astrocytes can upregulate neurotrophic factors, such as brain-derived neurotrophic factor (BDNF), vascular endothelial growth factor (VEGF), and basic fibroblast growth factor (bFGF), contributing to neuroprotection and synaptic repair [10] (Figure 1b). Astrocytes remove cellular debris by releasing extracellular vesicles (EVs), a mechanism for the homeostatic regulation of the neuronal function [11,12]. EVs are membrane vesicles (MVs) produced by the germination of plasma membranes (ectosomes) or by the cytosolic action of multivesicular bodies (exosomes). Individual EVs carrying specific molecular messages in the form of RNA, proteins, and lipids play a role in transmitting signals between cells [13]. Astrocyte-derived small extracellular vesicles (sEVs) reduce hippocampal dendritic complexity by upregulating endogenous miR-26a-5p levels and potentially regulating dendritic spines and synaptic transmission as well [14]. Astrocyte-derived extracellular vesicles (ADEV) are a novel astrocyte–neuron communication mediator that has recently been discovered [11]. ADEVs are released in response to IL-1β (ADEV-IL-1β) and TNFα (ADEV-TNFα) enriched with miR-125a-5p and miR-16-5p, targeting NTKR3 and its downstream effector Bcl2 to regulate dendritic growth, dendritic complexity, and synaptic stability [15] (Figure 1c).
Microglia interact with neurons, astrocytes, and oligodendrocytes in the brain, and mediate synaptic dysfunction or loss in AD through multiple pathways. The activation of microglia is closely related to the activation of the complement system [16]. The activation of complement-mediated mechanisms (e.g., C1q and C3) has been demonstrated to directly mediate synaptic loss and dysfunction in AD mice [17,18] (Figure 1d). C1q and C3 levels are parallel to Aβ deposition in AD mice [19]. Rescuing cognitive impairment in AD induced by Aβ-mediated synaptic loss by reducing C1q and C3 levels is a potential therapeutic avenue [20,21,22]. Microglia destabilize its homeostasis to an inflammatory state and release a series of inflammatory factors, such as IL-1α, TNF, and IFN-1, leading to synaptic loss, neuronal defects, and alterations in learning and memory [23,24] (Figure 1e). AD-related risk genes expressed by the microglia (e.g., TREM2 and ApoE) are demonstrated to mediate synaptic loss [25]. The microglia are able to modulate the expression of dendritic spines and synapse-associated proteins by releasing EVs to deliver miR-146a-5p that are not originally present in the neurons, to alter the number of dendritic spines and synaptic function in hippocampal neurons by repressing the translation of presynaptic synaptotagmin1 (Syt1) and postsynaptic neuroligin1 (Nlg1) [26] (Figure 1f). miR-146a prevents synaptic loss in AD through the alteration of microglia polarization, a phagocytic phenotype that reduces the secretion of inflammatory factors and protects neurons in AD [27].

3. Role of miRNAs in Regulating the Cerebrovascular System in AD

The cerebrovascular system is composed of a complex community of cells that interact with each other to enable the proper functioning of material circulation, metabolism, and permeability properties within the brain, including endothelial cells, smooth muscle cells, pericytes, perivascular immune cells (e.g., microglia), and surrounding astrocytes (Figure 1g). It is essential for maintaining brain health. Cerebrovascular disease (CVD) and AD share several risk factors [28], such as APOE ε4 [29], midlife hypertension, and hypercholesterolemia [30,31]. Vascular risk factors are predictive of the progression of AD and the conversion of mild cognitive impairment (MCI) to AD [32,33]. One of the most prominent implications is the abnormalities of cerebral blood-flow autoregulation and impaired removal of metabolic waste and neurotoxins from the brain. The paravalvular interstices around the cerebral vessels provide a low-resistance pathway for the convective clearance of solutes in the brain, with major contributions from the lymphatic and glymphatic systems (water channels represented by AQP4) [34,35]. Interstitial fluid drains from the brain along with the perivascular space and basement membrane [36]. Mouse models of cerebral amyloid angiopathy display abnormal cerebral blood flow and a greater tendency for Aβ to be deposited within the basement membrane of blood vessels [37].
The cerebrovascular system governs cerebral blood flow, the delivery of nutrients to the brain, the filtering of the blood–brain barrier (BBB) for brain-entering substances, and angiogenesis. Emerging evidence indicates that these factors are closely related to AD [38,39,40]. A novel technique, vessel isolation and nuclei extraction for sequencing (VINE-seq), has been developed to investigate genome-wide association studies (GWASs) in the vascular system of cortical and hippocampal samples obtained from Alzheimer’s patients. Surprisingly, the results show that up to 30 of the top 45 genes linked to AD are expressed in the vascular system of the human brain [41]. Cells of the human brain vasculature predominantly include endothelial cells, smooth muscle cells, pericytes, astrocytes, macrophages, T cells, and perivascular and meningeal fibroblasts.
miRNAs intervene in cardiovascular diseases by participating in the pathogenic mechanism of the disease. A decrease in cerebral blood flow to the striatum, hippocampus, or thalamus in the deep part of the brain triggers inflammation resulting in organ damage, causing the impairment of cognitive and memory functions [42]. The deletion of miR-126 in endothelial cells decreases the cerebral blood flow in mice (Figure 1h) and activates the astrocytes and microglia, thereby increasing inflammation as well as triggering the impairment of water channels and the lymphatic system, lower synaptic plasticity, and dendritic spine density [43]. Changes in miRNA profiles were detected while improving cerebral blood flow in a stroke model with therapies, indicating the possible involvement of miRNAs in the mechanism of improving cerebrovascular and neurological functions [44]. The role of miRNAs in vascular cognitive impairment progression warrants investigation.
Cerebrovascular miRNA markers corresponding to post-transcriptional regulation can indicate the stage of AD pathogenesis [45]. In the total brain vessel segments of 3xTg-AD mice, let-7g, miR-1944 (in males), miR-133a, and miR-2140 (in females) were significantly downregulated upon the transition from the young stage (1–2 months old) to cognitive impairment (CI). miR-99a in males was downregulated during the period of CI to extracellular Aβ formation. From the pre-AD (3–5 months old) to AD (6–15 months old) phase, let-7d, let-7i, miR-23a, miR-34b-3p, miR-99a, miR-126-3p, miR-132, miR-150, miR-1151-5p, miR-181a (in males), and miR-150, miR-539 (in females) were detected as being markedly altered in expression [45]. miR-200b-3p, miR-200c-3p, and miR-205-5p considerably reduced low-density lipoprotein receptor-related protein 1 (LRP1) in microvascular endothelial cells (MVECs) [46]. LRP1 impaired Aβ clearance [47,48,49], allowing Aβ peptides but not Aβ plaques, soluble and insoluble Aβ40, and Aβ42 to accumulate in the cortex, triggering increased microglia and astrocytes to induce neuroinflammation and to impair spatial memory in mice (Figure 1i). This process was induced by copper [46]. A high expression of miR-126 and miR-145, specifically in the brain capillaries, elevated pericyte coverage and promoted endothelial cell activation, resulting in an increased clearance of Aβ oligomers through the perivascular route [50] (Figure 1j). miR-181a is suspected to be involved in hippocampal memory formation [51,52]. Amyloid precursor protein (APP) is an important factor impacting synaptogenesis and neuronal health, and its presence is bidirectional and not just harmful as a source of Aβ [53]. miRNA is essential in the homeostatic regulation of APP dynamics in vivo [54]. miR-181a negatively targets forkhead box protein O1 (FOXO1), decreasing amyloid plaque deposition in the hippocampus and cortex, mitigating pericyte loss and BBB disruption in APP/PS1 mice, and ameliorating cognitive impairment in mice [55]. As the most abundant miRNA in the brain tissue, miR-124 is involved in the pathological process of age-related ischemic encephalopathy [56,57,58,59]. Paradoxically, miR-124 demonstrates beneficial aspects in ischemic encephalopathy by regulating autophagy, neuroinflammation, oxidative stress, Aβ deposition, and tau protein hyperphosphorylation [59,60,61], while in AD, it may adversely affect synaptic plasticity via the miR-124/PTPN1 pathway, through neural differentiation by mediating apoptosis [62,63] (Table 1). A thorough understanding of the role of miR-124 in ischemic encephalopathy and AD could help develop new avenues for the exploration of ischemic encephalopathy and AD prevention or treatment therapies.

4. miRNAs Are Associated with AD-Related Synaptic Impairment

Dendrites and dendritic spines are functionally critical structures in cognitive processes, such as learning and memory, and dysfunction and morphological changes in dendritic spines are essential characteristics of AD [73]. Pre-AD-related cognitive impairment is correlated with the progressive loss of synapses and neurons [74]. Synaptic density is believed to be a better predictor of a person’s cognitive performance than the loss of gray-matter volume in the brain, and this correlation also extends to the mild and pre-AD stages [75]. Synapses interconnect neurons via electrical and chemical signaling (e.g., glutamate and GABA). Animal studies have indicated that dendritic spine morphological changes preceded synaptic loss [76], and showed an increased proportion of abnormal spine morphology. The shortening and widening of the neck of the dendritic spine alter the electrical compartmentalization of the spines, resulting in reduced postsynaptic potentials in the head of the spine, long-term potentiation (LTP) enhancement, and impaired spatial memory [77]. In recent years, a series of publications have demonstrated the regulation of miRNAs on synapses, including synaptic plasticity, synapse formation, synapse morphology, and synapse quantity. The overexpression of miR-34a in AD inhibits presynaptic and postsynaptic proteins, such as VAMP2, SYT1HCN, NR2A, GluR1, and SHANK3 [78]; it impairs rapid synaptic-vesicle endocytosis, fast calcium-triggered synaptic vesicle exocytosis, synaptic input integration, and synaptic plasticity [79]. A meta-analysis of the gene expression data showed that miR-129 was substantially upregulated in the hippocampus of scopolamine-induced AD rats and was concentrated in the synapses with a high expression. miR-129 has over 600 target genes in the hippocampus, four of which are considered to play a function in synaptic plasticity, LTP, and long-term depression (LTD): CAMK2D, CAMK4, PPP3CA, and PRKCB, which encode CaMKII, CaMKIV, CaN, and PKC, respectively [67]. This result echoes another meta-analysis predicting that miR-30a-5p targets CAMK2B, CAMK3, MAPK1, PPP3CB, and RSP6KA2 in the LTP in AD [80]. Here, we summarized several miRNAs related to synapses in AD that were exploited by bioinformatic tools or experimental phases in recent years (Table 1).

4.1. miRNAs Involved in Aβ-Mediated AD Synaptic Dysfunction

Aβ interacts directly or indirectly with various synaptic proteins to disrupt synaptic function, such as AMPAR, NMDAR, mGluR5, PRPC, RAGE, and ephrin type-B receptor 2 (EPHB2).
miR-30b: miR-30b, which is highly analogous to miR-30a, has the NF-κB p65 subunit, which binds to miR-30b and is regulated by upstream Aβ42 to promote its upregulation in the brains of AD patients and APP transgenic PSD mice. Nevertheless, the specific mechanism of how upstream Aβ42 regulates miR-30b is unclear, and the expression of the enzymes that produce and synthesize Aβ does not change. miR-30b overexpression impairs synaptic integrity and cognitive function. Its target genes are EPHB2, sirtuin1 (SIRT1), and glutamate ionotropic receptor AMPA-type subunit 2 (GluA2), which are essential for maintaining synaptic integrity, exhibiting the downregulated expression in AD. The knockdown of miR-30b in TG mice rescued dendritic spine density, GluA1, GluN2B, and PSD95 expression levels in the hippocampal neurons, presumably mediated by EPHB2 and SIRT1 [71].
miR-181a: miR-181a acts as a negative regulator of synaptic plasticity and a positive regulator of memory. miR-181a mediates Aβ-induced synaptotoxicity in the mouse hippocampus. miR-181a and its target protein, GluA2, are reduced and increased, respectively, during the process of learning. In this way, miR-181a effectively reversed Aβ oligomers-induced impairments in synaptic plasticity and memory deficits in 3xTg-AD mice. Yet, PSD95, SYP, and GluA1 presented no significant differences [51]. In accession to GluA2, miR-181a has been proven to target CREB1 and PRKAA1, which are transcription factors and proteins that function in learning and memory, respectively [51,52].
miR-132: the significance of miR-132 in neural development and the regulation of synaptic function has been widely studied [81]. miR-132 acts as a mediator of inflammation [82,83]. An enriched environment enhances hippocampal LTP by upregulating miR-132 and reducing histone deacetylase (HDAC) signaling, thereby counteracting the synaptotoxicity of Aβ and acting as a preventive factor against hippocampal damage. This implies that the accumulation of Aβ in the brain can be reduced by prolonged stimulation with novelty, which may prevent AD [68].
miR-34: the expression of miR-34a plays a predictive function in Aβ-mediated AD, as its expression is upregulated before the aggregation of Aβ [84]. The inhibition of miR-34c rescued Aβ-induced AD-related spine morphology and synaptic integrity and memory damage by targeting SYT1 through the ROS-JNK-p53 pathway. However, its upstream regulatory signaling pathways and downstream molecular mechanisms in AD are unclear [72].

4.2. Regulation of Synaptic Function by miRNA-Mediated Tauopathies

Tau is a microtubule-associated protein. In the human genome, it is derived from the selective splicing of microtubule-associated protein tau (MAPT) mRNA [85]. Tau maintains the assembly and structural maintenance of microtubule structures through the binding of microtubule structures in normal physiological conditions [86,87]. Tau is abundantly expressed in central nervous system neurons and weakly expressed in astrocytes and oligodendrocytes. Structural defects in tau proteins are believed to be one of the cardinal pathogeneses of AD [88,89].
Bilen et al. [90] first raised the prospective role of miRNAs in tauopathies in a Drosophila study. They observed that the loss of the protein R3D1/loquacious blocked miRNA processing, preventing miRNA maturation, and significantly increased tau-mediated neuronal death [90]. miRNAs regulate tau phosphorylation in two principal ways: the first is direct binding to the 3’UTR of the mRNAs of protein kinases and phosphatases. The second is indirect regulation by modulating the factors involved in activating protein kinases and phosphatases. Smith et al. showed that miR-124, miR-9, miR-132, and miR-137 caused the misplacement of tau exon 10, leading to an imbalance in the 4R:3R-tau ratio in neuronal cells and resulting in disease [91]. Phosphorylation is the most frequent post-transcriptional modification of tau, usually occurring in the serine (Ser), threonine (Thr), and tyrosine (Tyr) residues of the protein [92].
miR-132: miR-132 acts as a positive regulator during postnatal brain development. miR-132 directly targets the polypyrimidine tract-binding protein 2 (PTBP2) in the neurons to aberrantly splice tau exon 10, leading to an imbalance in the ratio of three-repeat (3R) and four-repeat (4R) tau isoforms [91]. miR-132 prevents microtubule dynamics and neurofibrillary tangle (NFT) accumulation to enhance LTP in the P301S tau transgenic mice by directly targeting the RNA-binding protein RNA Binding Fox-1 Homolog 1 (RBFOX1) to degrade tau mRNA, acetylation via EP300, cleavage via calpain 2 and caspases 3/7, and tau kinase glycogen synthase kinase 3β (GSK3β), which hyperphosphorylates tau [69] to disrupt microtubule dynamics and NFT accumulation [93]. In addition to this, miR-132 can target glycosyltransferase-like domain containing 1 (GTDC-1), indirectly inhibiting the tau kinase cyclin-dependent kinase 5 (CDK5) through the NOS1 signaling pathway [94].
miR-135a-5p: miR-135a-5p displays a Foxd3-mediated tau-dependent expression decline in AD mice. In MAPT knockout neurons, tau overexpression was able to induce a loss of miR-135a-5p expression, leading to reduced LTP and basal transmission in CA3-CA1 projections, as well as reduced hippocampal pyramidal neuronal dendritic complexity and spine maturation [64]. Rho-Associated Coiled-Coil Containing Protein Kinase 2 (ROCK2) is one of six potential major regulators of synaptic and axonal degeneration in vivo [95]. ROCK2 is the direct target of miR-135a-5p. The loss of miR-135a-5p expression results in aberrant ROCK2 activation and the subsequent deregulation of the cellular cytoskeleton via phosphorylation of ADD1 at Ser726. The interaction of ROCK2 with its downstream target ADD1 promotes synaptic dysfunction. [64]. Synaptic density, synaptic transmission, and hippocampal LTP are considerably impaired in ROCK2 knockout mice [96]. The overexpression of ROCK2 also impairs dendritic morphology and LTP and basal transmission [64].
miR-101b: previous studies have revealed that HDAC is closely associated with excessive Aβ production [68,97]. Liu et al. identified that HDAC2 is also involved in the hyperphosphorylation of tau [98]. Tau is one of the cytoplasmic substrates of HDAC. HDAC inhibitors, such as Scriptaid, M344, BML281, and SAHA, can increase the level of acetylated tau, leading to the activation of the tau pathology [99]. 5′-adenosine monophosphate (AMP)-activated protein kinase (AMPK) reduces Tau phosphorylation and improves brain function in C57 mice [100]. The AMPK activity is significantly reduced in aging C57 mice and 3 × Tg AD mice. It inhibits GSK3β activity in HEK293 tau cells transfected by the AMPK plasmid [100]. The ectopic expression of HDAC2 promotes hyperphosphorylation and the aggregation of tau, and scrambles the combination of the hepatocyte nuclear factor 4α (HNF-4A) transcription factor with the miR-101b promoter. This results in the upregulation of AMPK, the target protein of miR-101b, a loss of spine density, and memory deficits in AD mice [98].
miR-9: miR-9 has been reported to play a vital role in neurodevelopment, neurogenesis, neuronal morphology, and dendrite development [101,102]. SIRT1 is a histone deacetylase abundantly expressed in the brain and is closely associated with transcriptional silencing, aging, and inflammation [103]. SIRT1 serves as a target of miR-9-5p, which is also one of the critical enzymes for removing acetyl groups from tau. Its deficiency causes synaptic loss and premature death in tauP301S transgenic mice. In contrast, the stereotaxic delivery of the adeno-associated virus that encodes SIRT1 into the hippocampus reduces acetylated Lys174 (K174) of tau [70]. UBE4B, an E3/E4 ubiquitin ligase, functions as a target gene for miR-9a, and its overexpression enhances the ubiquitination and degradation of tau. In a Drosophila tau overexpression model, UBE4B alleviated eye neurodegeneration and synaptic bouton defects [104]. However, the relationship between UEB4B and synaptic function is unclear.
miR-138: in the previous subsection, we briefly described the significance of SIRT1 in synapse and neural development. In addition to miR-9, miR-138 targets SIRT1 to regulate tau pathology. Interestingly, the in vitro overexpression of miR-138 in cultured cells was associated with increased Aβ production and tau phosphorylation, whereas in vivo experiments also showed an increased production of soluble Aβ42 and significantly reduced tau expression in the hippocampus and frontal cortex, but there was no significant effect on phosphorylation (PHF1, AT8, and S422 phosphorylation epitopes) [105]. The effects of miR-138 on the synapses were primarily manifested by increased postsynaptic marker PSD95, and decreased presynaptic marker synaptophysin and synaptic density [65,105]. In terms of cognition, the upregulation of miR-138 induced alterations in learning and memory, as well as anxiety-like behavior in mice [65].

5. Mitochondria as a Potential Therapeutic Target for AD-Related Cognitive Disorders

Neuronal energetics is related to neurotransmission. Most of the energy employed by neurons is used for synaptic signaling [106]. Hippocampal pyramidal neurons have the highest energy demand of all neurons in the brain [107]. Therefore, when the metabolic and energetic demands of hippocampal neurons are not satisfied, they are more at risk for cognitive impairment. In the early stages of AD development and aging, cellular metabolism becomes disturbed, causing a disturbance in the dynamic balance between glutamatergic and GABAergic signals released by the neurons and synapses [108]. By providing neurons with energy chemicals, such as adenosine triphosphate (ATP) [109], nicotinamide adenine dinucleotide (NAD+) [110], creatine [111], or the mitochondrial potassium channel-opener diazoxide [112], they can effectively protect the neurons from metabolic and Aβ toxicity damage and are expected to improve cognitive impairment (Figure 1k).
The mitochondria provide 90% of cellular energy as ATP. A study of brain tissue obtained from AD patients found region-specific changes in the synaptic mitochondria in AD [113]. Mitochondrial transport to the presynaptic terminal or synaptic mitochondrial dynamics may be altered in AD with abnormal mitochondrial morphology and the presence of multivesicular bodies in the synapse [113]. miR-455-3p dysregulation in the neuronal mitochondria is closely related to synaptic activity [114]. Mitochondrial miRNAs are emerging targets in AD and synaptic research [115]. Mitochondrial miRNAs have the potential for the development of biomarkers and targeted therapies for AD. However, research regarding the effects of mitochondrial miRNAs on specific mitochondrial functions is limited.

6. miRNAs as Potential Therapeutics or Biomarkers of AD

The FDA approved the first miRNA-based therapy for hereditary transthyretin amyloidosis in 2018 [116]. Biotechnology companies, such as Miragen, Synlogic, and Regulus therapeutics, are focused on developing miRNA-related medication. Nevertheless, there are no FDA-approved miRNA therapies for AD in current clinical use, and it is still in the preclinical phase.
The application of miRNA therapies in AD remains behind other disease areas, and more preclinical work is in demand. How to penetrate the BBB to target the brain to deliver miRNA supplements or inhibitors is a pivotal challenge that needs to be addressed. miRNA medications are selected for in situ injection instead of intravenous administration in cancer therapies in order to strengthen the target specificity and minimize side effects [117]. Nonetheless, the greatest obstacle to oligonucleotide-based miRNA-targeted treatments for AD is, at present, navigating the medicine to cross the BBB; only fat-soluble small molecules below 400 daltons can pass the BBB. Researchers have shown in rats that antimiR-134 can be selectively transfected into striatal neurons for the treatment of Parkinson’s disease by intracerebroventricular injection of the magnetic particle NeuroMag® complex plasmid into the lateral ventricles of the pars striatum, and the expression of miR-134 decreased by 0.35 fold [118].
Individual miRNAs can target multiple mRNAs, although their effects might be contradictory [119], whereas multiple miRNAs can also simultaneously target one mRNA [120]. It has been proposed that MG-6267, a dual inhibitor of AChE and miR-15b biogenesis, is able to address the multifactorial properties of AD as a therapeutic strategy. Preliminary predictions suggest that MG-6267 has a high CNS penetration [121].
In human patients, miRNA cannot be extracted from brain tissues as in animals, and samples are mainly derived from either peripheral blood or cerebrospinal fluid, especially in the exosomes. A clinical study predicted miRNA biomarkers in the peripheral blood serum exosomes of AD patients by high-throughput next-generation sequencing (NGS) and conducted random forest analyses with patients’ clinical, medical, and cognitive assessments and amyloid neuroimaging with positron emission tomography. A total of 16 miRNAs (hsa-miR-101-3p, hsa-miR-106a-5p, hsa-miR-106b-5p, hsa-miR-1306-5p, hsa-miR-143-3p, hsa-miR-15a-5p, hsa-miR-15b-3p, hsa-miR-18b-5p, hsa-miR-20a-5p, hsa-miR-30a-5p, hsa-miR-335-5p, hsa-miR-342-3p, hsa-miR-361-5p, hsa-miR-424-5p, hsa-miR-582-5p, and hsa-miR-93-5p) were obtained that were responsive to patients’ AD progression and had potential as peripheral screening tools [122].

7. Emerging Imaging Techniques for AD Research and Potential Applications in Diagnosis

7.1. Two-Photon Microscopy

Using two-photon imaging techniques, researchers have revealed in vivo in APP/PS1 mice that deficits in spine maintenance and plasticity occur in specific brain regions [123,124,125]. Reactive astrocytes can be observed in the brains of patients with mild cognitive impairment in the early stages of AD, as well as in APP/PS1 mice, and precede the appearance of amyloid plaques [126]. In vitro astrocyte studies failed to fulfill the demands of exploring the relationship between astrocytes, neurons, and the vascular system. Employing genetically encoded calcium indicators (GECIs), such as GCaMP, provided a stable and specific fluorescent indicator for studying calcium signaling and synaptic function in astrocytes [127]. Fluorescent probes for Aβ proteins [128], molecular probes for p-tau proteins [129], metal-ion fluorescent probes [130], two-photon monoamine oxidase (MAO) probes [131], reactive astrocyte probes [126], and miRNA probes [132] have been developed for two-photon imaging. Fluorescent probes not only serve as a localization tool for biomarkers, but also serve as potential targeted therapeutics based on the pharmaceutical molecules they carry. For example, curcumin has anti-inflammatory and antioxidant properties. Curcumin two-photon probes capable of targeting Aβ is being considered as a theranostic agent due to its imaging and potential therapeutic functions [133]. Although there are still technical challenges to be overcome, including low-fluorescence quantum yields, an inability to detect soluble Aβ, and difficulty in penetrating the BBB, this finding provides evidence for visualizing the effects of targeted therapy [133]. Another limitation at present is that two-photon in vivo imaging is limited to a 450 μm depth from the brain surface and requires replacing part of the skull with a cranial window to track fluorescently labeled neurons [134]. The synaptic imaging of the hippocampus requires more invasive micro-endoscopy or removing the overlying cortical tissue [135].

7.2. Optical Coherence Tomography (OCT)

Optical coherence tomography is an imaging technique that uses low-coherence light to capture two- and three-dimensional images from within the biological tissue with micron-level resolution. In the field of medicine, it is commonly applied in ophthalmology, cardiology, endovascular applications, and oncology. In the field of cardiology, researchers have used OCT to scan patients’ coronary arteries as well as to measure circulating miRNA levels from patients’ serum samples, with the finding that a decrease in miR-24 is strongly associated with coronary neointimal hyperplasia [136,137]. A study in the atherosclerotic rabbit model found that miR-let-7b regulates macrophages by targeting the TGF-βR1 signaling pathway, thereby affecting plaque stability [138]. Aβ causes vascular damage in the brain and OCT measurements show that the thickness of the retinal neuron and vascular layers of the eye are significantly thinner in AD patients [139,140]. Aβ brain deposition was positively correlated with vascular density in all regions of the retina [141]. One possible mechanism is the retrograde deformation of the optic nerve. Aβ pathological features visible in the subcortical visual centers could support this claim [142]. However, the combined approach of OCT and miRNA profiling in AD is rarely reported and possesses great potential for development.

7.3. Surface Plasmon Resonance (SPR)

Surface plasmon resonance can track biomolecular interactions in natural situations without any markers. The interaction processes between various biomolecules, such as peptides, proteins, oligonucleotides, and oligosaccharides, as well as viruses, bacteria, cells, and small molecule compounds, can be measured in real-time, in situ, and dynamically when passing through the sensor chip. The principle is that a biomolecule target is bonded to the biosensor surface, and then a solution containing an analyte that interacts with the target molecule is injected into and flows over the biosensor surface. The bonding between the biomolecules causes the refractive index to increase in the same proportion, and changes in the reaction between the biomolecules are observed. Song et al. [143] reported a DNA-assembled advanced plasma structure (DAPA)-based plasma biosensor capable of the selective single-nucleotide detection of exosomal miRNA at the attomolar level for AD diagnosis with an accuracy of 97.83%. The enhanced SPR biosensor enabled the high-throughput and ultra-sensitive screening of miRNAs and the simultaneous detection of multiple targets, which has great potential for future clinical diagnostics [144].

8. Outstanding Challenges for the Future

The multifaceted modulation of synapses by various miRNAs contributes to AD-related cognitive impairment. This review summarized key examples of miRNAs identified in recent years as upstream or downstream regulators of Aβ and tau proteins that modulate synaptic composition (synaptic vesicles and neurotransmitters), synaptic morphology, and synaptic function. Nevertheless, the research to date on miRNAs in AD has unresolved problems. Does the change in miRNA levels directly or indirectly contribute to synaptosome and synaptic function alterations and the toxicity induced by Aβ and tau proteins? Is the modulation of synapses by miRNAs the sole contributor to AD-related cognitive impairment? If not, how do we discover and verify the presence of other factors involved as well? Can the outcomes of miRNA expression-level abnormalities in situ at specific sites of pathogenesis be justified for commonly available humoral biomarkers (e.g., CSF and peripheral blood)? In response to the above questions, we need to consider whether the methodology for exploring somatic biomarkers is compatible with investigating the role of miRNAs in disease mechanisms. There has been considerable preclinical, basic research conducted on miRNAs in AD, but translational applications are still lacking.

9. Conclusions

In the brain, various cell types interact and functionally depend on each other in a tightly regulated manner. In particular, healthy vascular cells ensure sufficient cerebral blood flow and the release of trophic factors to nourish neurons. Aged vascular cells stimulate astrocytes to secrete inflammatory factors and damage synapses and neurons in pathological situations, such as AD. Perivascular immune cells (microglia and macrophages) and BBB play a monitoring role to investigate the molecules entering the brain. We presented a review of how miRNAs play an essential and role in the cellular interaction previously described, and how miRNA dysregulation affects AD-related cognitive impairment.

Author Contributions

Conceptualization, J.P.K.I.; writing—original draft preparation and editing, X.L., J.P.K.I. and S.-C.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work is partially supported by the Lo Kwee-Seong Biomedical Research Fund (J.I.), Faculty Innovation Awards (FIA2020/A/04) from the Faculty of Medicine, CUHK (J.I.), and the Hong Kong Research Grants Council General Research Fund (14117221; J.I.), Area of Excellence Scheme (AoE/M-604/16; J.I.), Theme-based Research Scheme (T13-605/18-W; J.I.).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors would like to thank Ip lab members for their helpful discussions.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. O’Brien, J.; Hayder, H.; Zayed, Y.; Peng, C. Overview of MicroRNA Biogenesis, Mechanisms of Actions, and Circulation. Front. Endocrinol. 2018, 9, 402. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Hutvágner, G.; Zamore, P.D. A MicroRNA in a Multiple-Turnover RNAi Enzyme Complex. Science 2002, 297, 2056–2060. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Ding, L.; Han, M. GW182 Family Proteins Are Crucial for MicroRNA-Mediated Gene Silencing. Trends Cell Biol. 2007, 17, 411–416. [Google Scholar] [CrossRef] [PubMed]
  4. Bartel, D.P. MicroRNAs: Target Recognition and Regulatory Functions. Cell 2009, 136, 215–233. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Elkayam, E.; Faehnle, C.R.; Morales, M.; Sun, J.; Li, H.; Joshua-Tor, L. Multivalent Recruitment of Human Argonaute by GW182. Mol. Cell 2017, 67, 646–658.e3. [Google Scholar] [CrossRef] [PubMed]
  6. Zhang, K.; Zhang, X.; Cai, Z.; Zhou, J.; Cao, R.; Zhao, Y.; Chen, Z.; Wang, D.; Ruan, W.; Zhao, Q.; et al. A Novel Class of MicroRNA-Recognition Elements That Function Only within Open Reading Frames. Nat. Struct. Mol. Biol. 2018, 25, 1019–1027. [Google Scholar] [CrossRef] [PubMed]
  7. Jia, R.; Xiao, M.S.; Li, Z.; Shan, G.; Huang, C. Defining an Evolutionarily Conserved Role of GW182 in Circular RNA Degradation. Cell Discov. 2019, 5, 45. [Google Scholar] [CrossRef] [Green Version]
  8. Oksanen, M.; Lehtonen, S.; Jaronen, M.; Goldsteins, G.; Hämäläinen, R.H.; Koistinaho, J. Astrocyte Alterations in Neurodegenerative Pathologies and Their Modeling in Human Induced Pluripotent Stem Cell Platforms. Cell. Mol. Life Sci. 2019, 76, 2739–2760. [Google Scholar] [CrossRef] [Green Version]
  9. Frost, G.R.; Li, Y.M. The Role of Astrocytes in Amyloid Production and Alzheimer’s Disease. Open Biol. 2017, 7, 170228. [Google Scholar] [CrossRef] [Green Version]
  10. Chiareli, R.A.; Carvalho, G.A.; Marques, B.L.; Mota, L.S.; Oliveira-Lima, O.C.; Gomes, R.M.; Birbrair, A.; Gomez, R.S.; Simão, F.; Klempin, F.; et al. The Role of Astrocytes in the Neurorepair Process. Front. Cell Dev. Biol. 2021, 9, 1304. [Google Scholar] [CrossRef]
  11. Gosselin, R.-D.; Meylan, P.; Decosterd, I. Extracellular Microvesicles from Astrocytes Contain Functional Glutamate Transporters: Regulation by Protein Kinase C and Cell Activation. Front. Cell. Neurosci. 2013, 7, 251. [Google Scholar] [CrossRef] [Green Version]
  12. Hajj, G.N.M.; Arantes, C.P.; Dias, M.V.S.; Roffé, M.; Costa-Silva, B.; Lopes, M.H.; Porto-Carreiro, I.; Rabachini, T.; Lima, F.R.; Beraldo, F.H.; et al. The Unconventional Secretion of Stress-Inducible Protein 1 by a Heterogeneous Population of Extracellular Vesicles. Cell. Mol. Life Sci. 2013, 70, 3211–3227. [Google Scholar] [CrossRef]
  13. Cocucci, E.; Meldolesi, J. Ectosomes and Exosomes: Shedding the Confusion between Extracellular Vesicles. Trends Cell Biol. 2015, 25, 364–372. [Google Scholar] [CrossRef]
  14. Luarte, A.; Henzi, R.; Fernández, A.; Gaete, D.; Cisternas, P.; Pizarro, M.; Batiz, L.F.; Villalobos, I.; Masalleras, M.; Vergara, R.; et al. Astrocyte-Derived Small Extracellular Vesicles Regulate Dendritic Complexity through MiR-26a-5p Activity. Cells 2020, 9, 930. [Google Scholar] [CrossRef] [Green Version]
  15. Chaudhuri, A.D.; Dastgheyb, R.M.; Yoo, S.W.; Trout, A.; Talbot, C.C.; Hao, H.; Witwer, K.W.; Haughey, N.J. TNFα and IL-1β Modify the MiRNA Cargo of Astrocyte Shed Extracellular Vesicles to Regulate Neurotrophic Signaling in Neurons. Cell Death Dis. 2018, 9, 1–18. [Google Scholar] [CrossRef] [Green Version]
  16. Eikelenboom, P.; Veerhuis, R. The Role of Complement and Activated Microglia in the Pathogenesis of Alzheimer’s Disease. Neurobiol. Aging 1996, 17, 673–680. [Google Scholar] [CrossRef]
  17. Leng, F.; Edison, P. Neuroinflammation and Microglial Activation in Alzheimer Disease: Where Do We Go from Here? Nat. Rev. Neurol. 2020, 17, 157–172. [Google Scholar] [CrossRef]
  18. Schafer, D.P.; Lehrman, E.K.; Kautzman, A.G.; Koyama, R.; Mardinly, A.R.; Yamasaki, R.; Ransohoff, R.M.; Greenberg, M.E.; Barres, B.A.; Stevens, B. Microglia Sculpt Postnatal Neural Circuits in an Activity and Complement-Dependent Manner. Neuron 2012, 74, 691–705. [Google Scholar] [CrossRef] [Green Version]
  19. Reichwald, J.; Danner, S.; Wiederhold, K.H.; Staufenbiel, M. Expression of Complement System Components during Aging and Amyloid Deposition in APP Transgenic Mice. J. Neuroinflamm. 2009, 6, 35. [Google Scholar] [CrossRef] [Green Version]
  20. Peng, L.; Bestard-Lorigados, I.; Song, W. The Synapse as a Treatment Avenue for Alzheimer’s Disease. Mol. Psychiatry 2022, 27, 2940–2949. [Google Scholar] [CrossRef]
  21. Ye, Y.; Xu, H.; Su, X.; He, X. Role of MicroRNA in Governing Synaptic Plasticity. Neural Plast. 2016, 2016, 4959523. [Google Scholar] [CrossRef] [Green Version]
  22. Shi, Q.; Chowdhury, S.; Ma, R.; Le, K.X.; Hong, S.; Caldarone, B.J.; Stevens, B.; Lemere, C.A. Complement C3 Deficiency Protects against Neurodegeneration in Aged Plaque-Rich APP/PS1 Mice. Sci. Transl. Med. 2017, 9, eaaf6295. [Google Scholar] [CrossRef] [Green Version]
  23. Liddelow, S.A.; Guttenplan, K.A.; Clarke, L.E.; Bennett, F.C.; Bohlen, C.J.; Schirmer, L.; Bennett, M.L.; Münch, A.E.; Chung, W.S.; Peterson, T.C.; et al. Neurotoxic Reactive Astrocytes Are Induced by Activated Microglia. Nature 2017, 541, 481–487. [Google Scholar] [CrossRef] [Green Version]
  24. Lall, D.; Lorenzini, I.; Mota, T.A.; Bell, S.; Mahan, T.E.; Ulrich, J.D.; Davtyan, H.; Rexach, J.E.; Muhammad, A.K.M.G.; Shelest, O.; et al. C9orf72 Deficiency Promotes Microglial-Mediated Synaptic Loss in Aging and Amyloid Accumulation. Neuron 2021, 109, 2275–2291.e8. [Google Scholar] [CrossRef]
  25. Shi, Y.; Holtzman, D.M. Interplay between Innate Immunity and Alzheimer Disease: APOE and TREM2 in the Spotlight. Nat. Rev. Immunol. 2018, 18, 759–772. [Google Scholar] [CrossRef]
  26. Prada, I.; Gabrielli, M.; Turola, E.; Iorio, A.; D’Arrigo, G.; Parolisi, R.; de Luca, M.; Pacifici, M.; Bastoni, M.; Lombardi, M.; et al. Glia-to-Neuron Transfer of MiRNAs via Extracellular Vesicles: A New Mechanism Underlying Inflammation-Induced Synaptic Alterations. Acta Neuropathol. 2018, 135, 529–550. [Google Scholar] [CrossRef] [Green Version]
  27. Liang, C.; Zou, T.; Zhang, M.; Fan, W.; Zhang, T.; Jiang, Y.; Cai, Y.; Chen, F.; Chen, X.; Sun, Y.; et al. MicroRNA-146a Switches Microglial Phenotypes to Resist the Pathological Processes and Cognitive Degradation of Alzheimer’s Disease. Theranostics 2021, 11, 4103. [Google Scholar] [CrossRef]
  28. Love, S.; Miners, J.S. Cerebrovascular Disease in Ageing and Alzheimer’s Disease. Acta Neuropathol. 2015, 131, 645–658. [Google Scholar] [CrossRef] [Green Version]
  29. Schilling, S.; DeStefano, A.L.; Sachdev, P.S.; Choi, S.H.; Mather, K.A.; DeCarli, C.D.; Wen, W.; Høgh, P.; Raz, N.; Au, R.; et al. APOE Genotype and MRI Markers of Cerebrovascular Disease. Neurology 2013, 81, 292–300. [Google Scholar] [CrossRef] [Green Version]
  30. Meng, X.F.; Yu, J.T.; Wang, H.F.; Tan, M.S.; Wang, C.; Tan, C.C.; Tan, L. Midlife Vascular Risk Factors and the Risk of Alzheimer’s Disease: A Systematic Review and Meta-Analysis. J. Alzheimer’s Dis. 2014, 42, 1295–1310. [Google Scholar] [CrossRef]
  31. Chuang, Y.F.; An, Y.; Bilgel, M.; Wong, D.F.; Troncoso, J.C.; O’Brien, R.J.; Breitner, J.C.; Ferruci, L.; Resnick, S.M.; Thambisetty, M. Midlife Adiposity Predicts Earlier Onset of Alzheimer’s Dementia, Neuropathology and Presymptomatic Cerebral Amyloid Accumulation. Mol. Psychiatry 2015, 21, 910–915. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Li, J.; Wang, Y.J.; Zhang, M.; Xu, Z.Q.; Gao, C.Y.; Fang, C.Q.; Yan, J.C.; Zhou, H.D. Vascular Risk Factors Promote Conversion from Mild Cognitive Impairment to Alzheimer Disease. Neurology 2011, 76, 1485–1491. [Google Scholar] [CrossRef] [PubMed]
  33. De Bruijn, R.F.A.G.; Ikram, M.A. Cardiovascular Risk Factors and Future Risk of Alzheimer’s Disease. BMC Med. 2014, 12, 130. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Mentis, A.F.A.; Dardiotis, E.; Chrousos, G.P. Apolipoprotein E4 and Meningeal Lymphatics in Alzheimer Disease: A Conceptual Framework. Mol. Psychiatry 2020, 26, 1075–1097. [Google Scholar] [CrossRef]
  35. Rasmussen, M.K.; Mestre, H.; Nedergaard, M. The Glymphatic Pathway in Neurological Disorders. Lancet Neurol. 2018, 17, 1016–1024. [Google Scholar] [CrossRef] [Green Version]
  36. Zhang, E.T.; Richards, H.K.; Kida, S.; Weller, R.O. Directional and Compartmentalised Drainage of Interstitial Fluid and Cerebrospinal Fluid from the Rat Brain. Acta Neuropathol. 1992, 83, 233–239. [Google Scholar] [CrossRef]
  37. Weller, R.O.; Massey, A.; Newman, T.A.; Hutchings, M.; Kuo, Y.M.; Roher, A.E. Cerebral Amyloid Angiopathy: Amyloid β Accumulates in Putative Interstitial Fluid Drainage Pathways in Alzheimer’s Disease. Am. J. Pathol. 1998, 153, 725–733. [Google Scholar] [CrossRef]
  38. Gate, D.; Saligrama, N.; Leventhal, O.; Yang, A.C.; Unger, M.S.; Middeldorp, J.; Chen, K.; Lehallier, B.; Channappa, D.; de Los Santos, M.B.; et al. Clonally Expanded CD8 T Cells Patrol the Cerebrospinal Fluid in Alzheimer’s Disease. Nature 2020, 577, 399–404. [Google Scholar] [CrossRef]
  39. Iadecola, C.; Gottesman, R.F. Cerebrovascular Alterations in Alzheimer Disease Incidental or Pathogenic? Circ. Res. 2018, 123, 406–408. [Google Scholar] [CrossRef]
  40. Safaiyan, S.; Kannaiyan, N.; Snaidero, N.; Brioschi, S.; Biber, K.; Yona, S.; Edinger, A.L.; Jung, S.; Rossner, M.J.; Simons, M. Age-Related Myelin Degradation Burdens the Clearance Function of Microglia during Aging. Nat. Neurosci. 2016, 19, 995–998. [Google Scholar] [CrossRef]
  41. Yang, A.C.; Vest, R.T.; Kern, F.; Lee, D.P.; Agam, M.; Maat, C.A.; Losada, P.M.; Chen, M.B.; Schaum, N.; Khoury, N.; et al. A Human Brain Vascular Atlas Reveals Diverse Mediators of Alzheimer’s Risk. Nature 2022, 603, 885–892. [Google Scholar] [CrossRef]
  42. Gustavsson, A.M.; van Westen, D.; Stomrud, E.; Engström, G.; Nägga, K.; Hansson, O. Midlife Atherosclerosis and Development of Alzheimer or Vascular Dementia. Ann. Neurol. 2020, 87, 52–62. [Google Scholar] [CrossRef] [Green Version]
  43. Yu, P.; Venkat, P.; Chopp, M.; Zacharek, A.; Shen, Y.; Ning, R.; Liang, L.; Li, W.; Zhang, L.; Landschoot-Ward, J.; et al. Role of MicroRNA-126 in Vascular Cognitive Impairment in Mice. J. Cereb. Blood Flow Metab. 2019, 39, 2497–2511. [Google Scholar] [CrossRef]
  44. Zheng, H.Z.; Jiang, W.; Zhao, X.F.; Du, J.; Liu, P.G.; Chang, L.D.; Li, W.B.; Hu, H.T.; Shi, X.M. Electroacupuncture Induces Acute Changes in Cerebral Cortical MiRNA Profile, Improves Cerebral Blood Flow and Alleviates Neurological Deficits in a Rat Model of Stroke. Neural Regen. Res. 2016, 11, 1940. [Google Scholar] [CrossRef]
  45. Chum, P.P.; Hakim, M.A.; Behringer, E.J. Cerebrovascular MicroRNA Expression Profile During Early Development of Alzheimer’s Disease in a Mouse Model. J. Alzheimer’s Dis. 2022, 85, 91–113. [Google Scholar] [CrossRef]
  46. Hsu, H.W.; Rodriguez-Ortiz, C.J.; Lim, S.L.; Zumkehr, J.; Kilian, J.G.; Vidal, J.; Kitazawa, M. Copper-Induced Upregulation of MicroRNAs Directs the Suppression of Endothelial LRP1 in Alzheimer’s Disease Model. Toxicol. Sci. 2019, 170, 144–156. [Google Scholar] [CrossRef]
  47. Storck, S.E.; Meister, S.; Nahrath, J.; Meißner, J.N.; Schubert, N.; di Spiezio, A.; Baches, S.; Vandenbroucke, R.E.; Bouter, Y.; Prikulis, I.; et al. Endothelial LRP1 Transports Amyloid-Β1–42 across the Blood–brain barrier. J. Clin. Investig. 2016, 126, 123–136. [Google Scholar] [CrossRef] [Green Version]
  48. Deane, R.; Wu, Z.; Zlokovic, B.V. RAGE (Yin) versus LRP (Yang) Balance Regulates Alzheimer Amyloid β-Peptide Clearance through Transport across the Blood–brain barrier. Stroke 2004, 35, 2628–2631. [Google Scholar] [CrossRef] [Green Version]
  49. Bell, R.D.; Sagare, A.P.; Friedman, A.E.; Bedi, G.S.; Holtzman, D.M.; Deane, R.; Zlokovic, B.V. Transport Pathways for Clearance of Human Alzheimer’s Amyloid β-Peptide and Apolipoproteins E and J in the Mouse Central Nervous System. J. Cereb. Blood Flow Metab. 2007, 27, 909–918. [Google Scholar] [CrossRef] [Green Version]
  50. Fu, L.; Jiang, G.; Weng, H.; Dick, G.M.; Chang, Y.; Kassab, G.S. Cerebrovascular MiRNAs Correlate with the Clearance of Aβ through Perivascular Route in Younger 3xTg-AD Mice. Brain Pathol. 2020, 30, 92–105. [Google Scholar] [CrossRef]
  51. Rodriguez-Ortiz, C.J.; Prieto, G.A.; Martini, A.C.; Forner, S.; Trujillo-Estrada, L.; LaFerla, F.M.; Baglietto-Vargas, D.; Cotman, C.W.; Kitazawa, M. MiR-181a Negatively Modulates Synaptic Plasticity in Hippocampal Cultures and Its Inhibition Rescues Memory Deficits in a Mouse Model of Alzheimer’s Disease. Aging Cell 2020, 19, e13118. [Google Scholar] [CrossRef] [Green Version]
  52. Xu, X.F.; Jing, X.; Ma, H.X.; Yuan, R.R.; Dong, Q.; Dong, J.L.; Han, X.F.; Chen, Z.Y.; Li, X.Z.; Wang, Y. MiR-181a Participates in Contextual Fear Memory Formation Via Activating MTOR Signaling Pathway. Cerebral Cortex 2018, 28, 3309–3321. [Google Scholar] [CrossRef] [Green Version]
  53. Southam, K.A.; Stennard, F.; Pavez, C.; Small, D.H. Knockout of Amyloid β Protein Precursor (APP) Expression Alters Synaptogenesis, Neurite Branching and Axonal Morphology of Hippocampal Neurons. Neurochem. Res. 2019, 44, 1346–1355. [Google Scholar] [CrossRef]
  54. Wang, R.; Chopra, N.; Nho, K.; Maloney, B.; Obukhov, A.G.; Nelson, P.T.; Counts, S.E.; Lahiri, D.K. Human MicroRNA (MiR-20b-5p) Modulates Alzheimer’s Disease Pathways and Neuronal Function, and a Specific Polymorphism Close to the MIR20B Gene Influences Alzheimer’s Biomarkers. Mol. Psychiatry 2022, 27, 1256–1273. [Google Scholar] [CrossRef]
  55. Wu, Q.; Yuan, X.; Bai, J.; Han, R.; Li, Z.; Zhang, H.; Xiu, R. MicroRNA-181a Protects against Pericyte Apoptosis via Directly Targeting FOXO1: Implication for Ameliorated Cognitive Deficits in APP/PS1 Mice. Aging 2019, 11, 6120. [Google Scholar] [CrossRef]
  56. Liu, X.; Feng, Z.; Du, L.; Huang, Y.; Ge, J.; Deng, Y.; Mei, Z. The Potential Role of MicroRNA-124 in Cerebral Ischemia Injury. Int. J. Mol. Sci. 2019, 21, 120. [Google Scholar] [CrossRef] [Green Version]
  57. Sun, M.; Hou, X.; Ren, G.; Zhang, Y.; Cheng, H. Dynamic Changes in MiR-124 Levels in Patients with Acute Cerebral Infarction. Int. J. Neurosci. 2019, 129, 649–653. [Google Scholar] [CrossRef]
  58. Chen, S.-H.; Sun, H.; Zhang, Y.M.; Xu, H.; Yang, Y.; Wang, F.M. Effects of Acupuncture at Baihui (GV 20) and Zusanli (ST 36) on Peripheral Serum Expression of MicroRNA 124, Laminin and Integrin Β1 in Rats with Cerebral Ischemia Reperfusion Injury. Chin. J. Integr. Med. 2015, 22, 49–55. [Google Scholar] [CrossRef]
  59. Doeppner, T.R.; Doehring, M.; Bretschneider, E.; Zechariah, A.; Kaltwasser, B.; Müller, B.; Koch, J.C.; Bähr, M.; Hermann, D.M.; Michel, U. MicroRNA-124 Protects against Focal Cerebral Ischemia via Mechanisms Involving Usp14-Dependent REST Degradation. Acta Neuropathol. 2013, 126, 251–265. [Google Scholar] [CrossRef]
  60. Saraiva, C.; Talhada, D.; Rai, A.; Ferreira, R.; Ferreira, L.; Bernardino, L.; Ruscher, K. MicroRNA-124-Loaded Nanoparticles Increase Survival and Neuronal Differentiation of Neural Stem Cells in Vitro but Do Not Contribute to Stroke Outcome in Vivo. PLoS ONE 2018, 13, e0193609. [Google Scholar] [CrossRef]
  61. Zhu, H.; Wang, J.; Shao, Y.; Wan, D. Catalpol May Improve Axonal Growth via Regulating MiR-124 Regulated PI3K/AKT/MTOR Pathway in Neurons after Ischemia. Ann. Transl. Med. 2019, 7, 306. [Google Scholar] [CrossRef] [PubMed]
  62. Wang, X.; Liu, D.; Huang, H.Z.; Wang, Z.H.; Hou, T.Y.; Yang, X.; Pang, P.; Wei, N.; Zhou, Y.F.; Dupras, M.J.; et al. A Novel MicroRNA-124/PTPN1 Signal Pathway Mediates Synaptic and Memory Deficits in Alzheimer’s Disease. Biol. Psychiatry 2018, 83, 395–405. [Google Scholar] [CrossRef] [PubMed]
  63. Zhao, Y.N.; Li, W.F.; Li, F.; Zhang, Z.; Dai, Y.D.; Xu, A.L.; Qi, C.; Gao, J.M.; Gao, J. Resveratrol Improves Learning and Memory in Normally Aged Mice through MicroRNA-CREB Pathway. Biochem. Biophys. Res. Commun. 2013, 435, 597–602. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Zheng, K.; Hu, F.; Zhou, Y.; Zhang, J.; Zheng, J.; Lai, C.; Xiong, W.; Cui, K.; Hu, Y.Z.; Han, Z.T.; et al. MiR-135a-5p Mediates Memory and Synaptic Impairments via the Rock2/Adducin1 Signaling Pathway in a Mouse Model of Alzheimer’s Disease. Nat. Commun. 2021, 12, 1903. [Google Scholar] [CrossRef]
  65. Lu, Y.; Tan, L.; Wang, X. Circular HDAC9/MicroRNA-138/Sirtuin-1 Pathway Mediates Synaptic and Amyloid Precursor Protein Processing Deficits in Alzheimer’s Disease. Neurosci. Bull. 2019, 35, 877–888. [Google Scholar] [CrossRef]
  66. Wingo, T.S.; Yang, J.; Fan, W.; Min Canon, S.; Gerasimov, E.S.; Lori, A.; Logsdon, B.; Yao, B.; Seyfried, N.T.; Lah, J.J.; et al. Brain MicroRNAs Associated with Late-Life Depressive Symptoms Are Additionally, Associated with Cognitive Trajectory and Dementia. NPJ Genom. Med. 2020, 5, 6. [Google Scholar] [CrossRef] [Green Version]
  67. Hosseinian, S.; Arefian, E.; Rakhsh-Khorshid, H.; Eivani, M.; Rezayof, A.; Pezeshk, H.; Marashi, S.A. A Meta-Analysis of Gene Expression Data Highlights Synaptic Dysfunction in the Hippocampus of Brains with Alzheimer’s Disease. Sci. Rep. 2020, 10, 8384. [Google Scholar] [CrossRef]
  68. Wei, Z.; Meng, X.; el Fatimy, R.; Sun, B.; Mai, D.; Zhang, J.; Arora, R.; Zeng, A.; Xu, P.; Qu, S.; et al. Environmental Enrichment Prevents Aβ Oligomer-Induced Synaptic Dysfunction through Mirna-132 and Hdac3 Signaling Pathways. Neurobiol. Dis 2020, 134, 104617. [Google Scholar] [CrossRef]
  69. Shi, L.; Zhang, R.; Li, T.; Han, X.; Yuan, N.; Jiang, L.; Zhou, H.; Xu, S. Decreased MiR-132 Plays a Crucial Role in Diabetic Encephalopathy by Regulating the GSK-3β/Tau Pathway. Aging 2021, 13, 4590. [Google Scholar] [CrossRef]
  70. Min, S.W.; Sohn, P.D.; Li, Y.; Devidze, N.; Johnson, J.R.; Krogan, N.J.; Masliah, E.; Mok, S.A.; Gestwicki, J.E.; Gan, L. SIRT1 Deacetylates Tau and Reduces Pathogenic Tau Spread in a Mouse Model of Tauopathy. J. Neurosci. 2018, 38, 3680–3688. [Google Scholar] [CrossRef] [Green Version]
  71. Song, Y.; Hu, M.; Zhang, J.; Teng, Z.Q.; Chen, C. A Novel Mechanism of Synaptic and Cognitive Impairments Mediated via MicroRNA-30b in Alzheimer’s Disease. EBioMedicine 2019, 39, 409–421. [Google Scholar] [CrossRef] [Green Version]
  72. Shi, Z.; Zhang, K.; Zhou, H.; Jiang, L.; Xie, B.; Wang, R.; Xia, W.; Yin, Y.; Gao, Z.; Cui, D.; et al. Increased MiR-34c Mediates Synaptic Deficits by Targeting Synaptotagmin 1 through ROS-JNK-P53 Pathway in Alzheimer’s Disease. Aging Cell 2020, 19, e13125. [Google Scholar] [CrossRef] [Green Version]
  73. Penzes, P.; Cahill, M.E.; Jones, K.A.; Vanleeuwen, J.E.; Woolfrey, K.M. Dendritic Spine Pathology in Neuropsychiatric Disorders. Nat. Neurosci. 2011, 14, 285–293. [Google Scholar] [CrossRef] [Green Version]
  74. Scheff, S.W.; Price, D.A.; Schmitt, F.A.; Dekosky, S.T.; Mufson, E.J. Synaptic Alterations in CA1 in Mild Alzheimer Disease and Mild Cognitive Impairment. Neurology 2007, 68, 1501–1508. [Google Scholar] [CrossRef]
  75. Mecca, A.P.; O’Dell, R.S.; Sharp, E.S.; Banks, E.R.; Bartlett, H.H.; Zhao, W.; Lipior, S.; Diepenbrock, N.G.; Chen, M.K.; Naganawa, M.; et al. Synaptic Density and Cognitive Performance in Alzheimer’s Disease: A PET Imaging Study with [11C]UCB-J. Alzheimer’s Dement. 2021, 1–10. [Google Scholar] [CrossRef]
  76. Harris, K.M.; Jensen, F.E.; Tsao, B. Three-Dimensional Structure of Dendritic Spines and Synapses in Rat Hippocampus (CA1) at Postnatal Day 15 and Adult Ages: Implications for the Maturation of Synaptic Physiology and Long-Term Potentiation [Published Erratum Appears in J Neurosci 1992 Aug;12(8):Following Table of Contents]. J. Neurosci. 1992, 12, 2685–2705. [Google Scholar] [CrossRef] [Green Version]
  77. Androuin, A.; Potier, B.; Nägerl, U.V.; Cattaert, D.; Danglot, L.; Thierry, M.; Youssef, I.; Triller, A.; Duyckaerts, C.; el Hachimi, K.H.; et al. Evidence for Altered Dendritic Spine Compartmentalization in Alzheimer’s Disease and Functional Effects in a Mouse Model. Acta Neuropathol. 2018, 135, 839–854. [Google Scholar] [CrossRef]
  78. Sarkar, S.; Jun, S.; Rellick, S.; Quintana, D.D.; Cavendish, J.Z.; Simpkins, J.W. Expression of MicroRNA-34a in Alzheimer’s Disease Brain Targets Genes Linked to Synaptic Plasticity, Energy Metabolism, and Resting State Network Activity. Brain Res. 2016, 1646, 139–151. [Google Scholar] [CrossRef] [Green Version]
  79. Bazrgar, M.; Khodabakhsh, P.; Prudencio, M.; Mohagheghi, F.; Ahmadiani, A. The Role of MicroRNA-34 Family in Alzheimer’s Disease: A Potential Molecular Link between Neurodegeneration and Metabolic Disorders. Pharmacol. Res. 2021, 172, 105805. [Google Scholar] [CrossRef]
  80. Moradifard, S.; Hoseinbeyki, M.; Ganji, S.M.; Minuchehr, Z. Analysis of MicroRNA and Gene Expression Profiles in Alzheimer’s Disease: A Meta-Analysis Approach. Sci. Rep. 2018, 8, 4767. [Google Scholar] [CrossRef]
  81. Salta, E.; Sierksma, A.; Eynden, E.V.; de Strooper, B. MiR-132 Loss de-Represses ITPKB and Aggravates Amyloid and TAU Pathology in Alzheimer’s Brain. EMBO Mol. Med. 2016, 8, 1005–1018. [Google Scholar] [CrossRef]
  82. Shaked, I.; Meerson, A.; Wolf, Y.; Avni, R.; Greenberg, D.; Gilboa-Geffen, A.; Soreq, H. MicroRNA-132 Potentiates Cholinergic Anti-Inflammatory Signaling by Targeting Acetylcholinesterase. Immunity 2009, 31, 965–973. [Google Scholar] [CrossRef] [Green Version]
  83. O’Neill, L.A.J. Boosting the Brain’s Ability to Block Inflammation via MicroRNA-132. Immunity 2009, 31, 854–855. [Google Scholar] [CrossRef] [Green Version]
  84. Xu, Y.; Chen, P.; Wang, X.; Yao, J.; Zhuang, S. MiR-34a Deficiency in APP/PS1 Mice Promotes Cognitive Function by Increasing Synaptic Plasticity via AMPA and NMDA Receptors. Neurosci. Lett. 2018, 670, 94–104. [Google Scholar] [CrossRef]
  85. Ghetti, B.; Oblak, A.L.; Boeve, B.F.; Johnson, K.A.; Dickerson, B.C.; Goedert, M. Invited Review: Frontotemporal Dementia Caused by Microtubule-Associated Protein Tau Gene (MAPT) Mutations: A Chameleon for Neuropathology and Neuroimaging. Neuropathol. Appl. Neurobiol. 2015, 41, 24–46. [Google Scholar] [CrossRef] [Green Version]
  86. Guo, T.; Noble, W.; Hanger, D.P. Roles of Tau Protein in Health and Disease. Acta Neuropathol. 2017, 133, 665–704. [Google Scholar] [CrossRef] [Green Version]
  87. Tapia-Rojas, C.; Cabezas-Opazo, F.; Deaton, C.A.; Vergara, E.H.; Johnson, G.V.W.; Quintanilla, R.A. It’s All about Tau. Prog. Neurobiol. 2019, 175, 54–76. [Google Scholar] [CrossRef]
  88. Shin, R.W.; Iwaki, T.; Kitamoto, T.; Sato, Y.; Tateishi, J. Massive Accumulation of Modified Tau and Severe Depletion of Normal Tau Characterize the Cerebral Cortex and White Matter of Alzheimer’s Disease. Demonstration Using the Hydrated Autoclaving Method. Am. J. Pathol. 1992, 140, 937. [Google Scholar]
  89. Dugger, B.N.; Dickson, D.W. Pathology of Neurodegenerative Diseases. Cold Spring Harb. Perspect. Biol. 2017, 9, a028035. [Google Scholar] [CrossRef] [Green Version]
  90. Bilen, J.; Liu, N.; Burnett, B.G.; Pittman, R.N.; Bonini, N.M. MicroRNA Pathways Modulate Polyglutamine-Induced Neurodegeneration. Mol. Cell 2006, 24, 157–163. [Google Scholar] [CrossRef]
  91. Smith, P.Y.; Delay, C.; Girard, J.; Papon, M.A.L.; Planel, E.; Sergeant, N.; Buée, L.; Hébert, S.S. MicroRNA-132 Loss Is Associated with Tau Exon 10 Inclusion in Progressive Supranuclear Palsy. Hum. Mol. Genet. 2011, 20, 4016–4024. [Google Scholar] [CrossRef] [PubMed]
  92. Bazrgar, M.; Khodabakhsh, P.; Mohagheghi, F.; Prudencio, M.; Ahmadiani, A. Brain MicroRNAs Dysregulation: Implication for Missplicing and Abnormal Post-Translational Modifications of Tau Protein in Alzheimer’s Disease and Related Tauopathies. Pharmacol. Res. 2020, 155, 104729. [Google Scholar] [CrossRef] [PubMed]
  93. El Fatimy, R.; Li, S.; Chen, Z.; Mushannen, T.; Gongala, S.; Wei, Z.; Balu, D.T.; Rabinovsky, R.; Cantlon, A.; Elkhal, A.; et al. MicroRNA-132 Provides Neuroprotection for Tauopathies via Multiple Signaling Pathways. Acta Neuropathol. 2018, 136, 537–555. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Zhang, M.; Bian, Z. Alzheimer’s Disease and MicroRNA-132: A Widespread Pathological Factor and Potential Therapeutic Target. Front. Neurosci. 2021, 15, 687973. [Google Scholar] [CrossRef]
  95. Wishart, T.M.; Rooney, T.M.; Lamont, D.J.; Wright, A.K.; Morton, A.J.; Jackson, M.; Freeman, M.R.; Gillingwater, T.H. Combining Comparative Proteomics and Molecular Genetics Uncovers Regulators of Synaptic and Axonal Stability and Degeneration In Vivo. PLoS Genet. 2012, 8, e1002936. [Google Scholar] [CrossRef]
  96. Zhou, Z.; Meng, Y.; Asrar, S.; Todorovski, Z.; Jia, Z. A Critical Role of Rho-Kinase ROCK2 in the Regulation of Spine and Synaptic Function. Neuropharmacology 2009, 56, 81–89. [Google Scholar] [CrossRef]
  97. Zhang, J.; Liu, Y.; Wang, S.; Que, R.; Zhao, W.; An, L. Exploration of the Molecular Mechanism for Lipoprotein Lipase Expression Variations in SH-SY5Y Cells Exposed to Different Doses of Amyloid-Beta Protein. Front. Aging Neurosci. 2020, 12, 132. [Google Scholar] [CrossRef]
  98. Liu, D.; Tang, H.; Li, X.Y.; Deng, M.F.; Wei, N.; Wang, X.; Zhou, Y.F.; Wang, D.Q.; Fu, P.; Wang, J.Z.; et al. Targeting the HDAC2/HNF-4A/MiR-101b/AMPK Pathway Rescues Tauopathy and Dendritic Abnormalities in Alzheimer’s Disease. Mol. Ther. 2017, 25, 752–764. [Google Scholar] [CrossRef] [Green Version]
  99. Jeong, H.; Shin, S.; Lee, J.S.; Lee, S.H.; Baik, J.H.; Lim, S.; Kim, Y.K. Pan-HDAC Inhibitors Promote Tau Aggregation by Increasing the Level of Acetylated Tau. Int. J. Mol. Sci. 2019, 20, 4283. [Google Scholar] [CrossRef] [Green Version]
  100. Wang, L.; Li, N.; Shi, F.X.; Xu, W.Q.; Cao, Y.; Lei, Y.; Wang, J.Z.; Tian, Q.; Zhou, X.W. Upregulation of AMPK Ameliorates Alzheimer’s Disease-Like Tau Pathology and Memory Impairment. Mol. Neurobiol. 2020, 57, 3349–3361. [Google Scholar] [CrossRef]
  101. Radhakrishnan, B.; Alwin Prem Anand, A. Role of MiRNA-9 in Brain Development. J. Exp. Neurosci. 2016, 10, 101–120. [Google Scholar] [CrossRef]
  102. Giusti, S.A.; Vogl, A.M.; Brockmann, M.M.; Vercelli, C.A.; Rein, M.L.; Trümbach, D.; Wurst, W.; Cazalla, D.; Stein, V.; Deussing, J.M.; et al. MicroRNA-9 Controls Dendritic Development by Targeting REST. Elife 2014, 3, e02755. [Google Scholar] [CrossRef]
  103. Elibol, B.; Kilic, U. High Levels of SIRT1 Expression as a Protective Mechanism against Disease-Related Conditions. Front. Endocrinol. 2018, 9, 614. [Google Scholar] [CrossRef]
  104. Subramanian, M.; Hyeon, S.J.; Das, T.; Suh, Y.S.; Kim, Y.K.; Lee, J.S.; Song, E.J.; Ryu, H.; Yu, K. UBE4B, a MicroRNA-9 Target Gene, Promotes Autophagy-Mediated Tau Degradation. Nat. Commun. 2021, 12, 1–15. [Google Scholar] [CrossRef]
  105. Boscher, E.; Goupil, C.; Petry, S.; Keraudren, R.; Loiselle, A.; Planel, E.; Hébert, S.S. MicroRNA-138 Overexpression Alters Aβ42 Levels and Behavior in Wildtype Mice. Front. Neurosci. 2021, 14, 1445. [Google Scholar] [CrossRef]
  106. Kann, O.; Kovács, R. Mitochondria and Neuronal Activity. Am. J. Physiol. Cell Physiol. 2007, 292, 641–657. [Google Scholar] [CrossRef] [Green Version]
  107. LaManna, J.C.; Harik, S.I. Regional Comparisons of Brain Glucose Influx. Brain Res. 1985, 326, 299–305. [Google Scholar] [CrossRef]
  108. Kapogiannis, D.; Mattson, M.P. Disrupted Energy Metabolism and Neuronal Circuit Dysfunction in Cognitive Impairment and Alzheimer’s Disease. Lancet Neurol. 2011, 10, 187–198. [Google Scholar] [CrossRef] [Green Version]
  109. Goldberg, J.; Currais, A.; Prior, M.; Fischer, W.; Chiruta, C.; Ratliff, E.; Daugherty, D.; Dargusch, R.; Finley, K.; Esparza-Moltó, P.B.; et al. The Mitochondrial ATP Synthase Is a Shared Drug Target for Aging and Dementia. Aging Cell 2018, 17, e12715. [Google Scholar] [CrossRef]
  110. Fang, E.F. Mitophagy and NAD+ Inhibit Alzheimer Disease. Autophagy 2019, 15, 1112–1114. [Google Scholar] [CrossRef] [Green Version]
  111. Marques, E.P.; Wyse, A.T.S. Creatine as a Neuroprotector: An Actor That Can Play Many Parts. Neurotox. Res. 2019, 36, 411–423. [Google Scholar] [CrossRef]
  112. Guan, L.; Ji, Y.Q.; Liu, J.; Kong, M.; Sun, Z.W.; Shen, X.Q.; Ren, C.; Yu, G.P.; Ba, M.W. Diazoxide Induces Endoplasmic Reticulum Stress-Related Neuroprotection Mediated by P38 MAPK against Aβ25-35 Insults. Eur. Rev. Med. Pharmacol. Sci. 2018, 22, 6133–6138. [Google Scholar] [CrossRef]
  113. Pickett, E.K.; Rose, J.; McCrory, C.; McKenzie, C.A.; King, D.; Smith, C.; Gillingwater, T.H.; Henstridge, C.M.; Spires-Jones, T.L. Region-Specific Depletion of Synaptic Mitochondria in the Brains of Patients with Alzheimer’s Disease. Acta Neuropathol. 2018, 136, 747–757. [Google Scholar] [CrossRef] [Green Version]
  114. Kumar, S.; Hemachandra Reddy, P. A New Discovery of MicroRNA-455-3p in Alzheimer’s Disease. J. Alzheimer’s Dis. 2019, 72, S117–S130. [Google Scholar] [CrossRef]
  115. Gowda, P.; Reddy, P.H.; Kumar, S. Deregulated Mitochondrial MicroRNAs in Alzheimer’s Disease: Focus on Synapse and Mitochondria. Ageing Res. Rev. 2022, 73, 101529. [Google Scholar] [CrossRef]
  116. Adams, D.; Gonzalez-Duarte, A.; O’Riordan, W.D.; Yang, C.-C.; Ueda, M.; Kristen, A.V.; Tournev, I.; Schmidt, H.H.; Coelho, T.; Berk, J.L.; et al. Patisiran, an RNAi Therapeutic, for Hereditary Transthyretin Amyloidosis. N. Engl. J. Med. 2018, 379, 11–21. [Google Scholar] [CrossRef]
  117. Chen, Y.; Gao, D.Y.; Huang, L. In Vivo Delivery of MiRNAs for Cancer Therapy: Challenges and Strategies. Adv. Drug Deliv. Rev. 2015, 81, 128–141. [Google Scholar] [CrossRef] [Green Version]
  118. De Almeida, S.S.T.; Horst, C.H.; Soto-Sánchez, C.; Fernandez, E.; de Almeida, R.T. Delivery of MiRNA-Targeted Oligonucleotides in the Rat Striatum by Magnetofection with Neuromag®. Molecules 2018, 23, 1825. [Google Scholar] [CrossRef] [Green Version]
  119. Lim, L.P.; Lau, N.C.; Garrett-Engele, P.; Grimson, A.; Schelter, J.M.; Castle, J.; Bartel, D.P.; Linsley, P.S.; Johnson, J.M. Microarray Analysis Shows That Some MicroRNAs Downregulate Large Numbers of Target MRNAs. Nature 2005, 433, 769–773. [Google Scholar] [CrossRef]
  120. Liufu, Z.; Zhao, Y.; Guo, L.; Miao, G.; Xiao, J.; Lyu, Y.; Chen, Y.; Shi, S.; Tang, T.; Wu, C.I. Redundant and Incoherent Regulations of Multiple Phenotypes Suggest MicroRNAs’ Role in Stability Control. Genome Res. 2017, 27, 1665–1673. [Google Scholar] [CrossRef] [Green Version]
  121. Gabr, M.T.; Brogi, S. MicroRNA-Based Multitarget Approach for Alzheimer’s Disease: Discovery of the First-In-Class Dual Inhibitor of Acetylcholinesterase and MicroRNA-15b Biogenesis. J. Med. Chem. 2020, 63, 9695–9704. [Google Scholar] [CrossRef] [PubMed]
  122. Cheng, L.; Doecke, J.D.; Sharples, R.A.; Villemagne, V.L.; Fowler, C.J.; Rembach, A.; Martins, R.N.; Rowe, C.C.; Macaulay, S.L.; Masters, C.L.; et al. Prognostic Serum MiRNA Biomarkers Associated with Alzheimer’s Disease Shows Concordance with Neuropsychological and Neuroimaging Assessment. Mol. Psychiatry 2014, 20, 1188–1196. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Wu, W.; Du, S.; Shi, W.; Liu, Y.; Hu, Y.; Xie, Z.; Yao, X.; Liu, Z.; Ma, W.; Xu, L.; et al. Inhibition of Rac1-Dependent Forgetting Alleviates Memory Deficits in Animal Models of Alzheimer’s Disease. Protein Cell 2019, 10, 745–759. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Borin, M.; Saraceno, C.; Catania, M.; Lorenzetto, E.; Pontelli, V.; Paterlini, A.; Fostinelli, S.; Avesani, A.; di Fede, G.; Zanusso, G.; et al. Rac1 Activation Links Tau Hyperphosphorylation and Aß Dysmetabolism in Alzheimer’s Disease. Acta Neuropathol. Commun. 2018, 6, 1–17. [Google Scholar] [CrossRef] [Green Version]
  125. Noguchi, J.; Nagaoka, A.; Watanabe, S.; Ellis-Davies, G.C.R.; Kitamura, K.; Kano, M.; Matsuzaki, M.; Kasai, H. In Vivo Two-Photon Uncaging of Glutamate Revealing the Structure–Function Relationships of Dendritic Spines in the Neocortex of Adult Mice. J. Physiol. 2011, 589, 2447–2457. [Google Scholar] [CrossRef]
  126. Jo, S.; Yarishkin, O.; Hwang, Y.J.; Chun, Y.E.; Park, M.; Woo, D.H.; Bae, J.Y.; Kim, T.; Lee, J.; Chun, H.; et al. GABA from Reactive Astrocytes Impairs Memory in Mouse Models of Alzheimer’s Disease. Nat. Med. 2014, 20, 886–896. [Google Scholar] [CrossRef]
  127. Corkrum, M.; Covelo, A.; Lines, J.; Bellocchio, L.; Pisansky, M.; Loke, K.; Quintana, R.; Rothwell, P.E.; Lujan, R.; Marsicano, G.; et al. Dopamine-Evoked Synaptic Regulation in the Nucleus Accumbens Requires Astrocyte Activity. Neuron 2020, 105, 1036–1047.e5. [Google Scholar] [CrossRef] [Green Version]
  128. Heo, C.H.; Kim, K.H.; Kim, H.J.; Baik, S.H.; Song, H.; Kim, Y.S.; Lee, J.; Mook-Jung, I.; Kim, H.M. A Two- Photon Fluorescent Probe for Amyloid-β Plaques in Living Mice. Chem. Commun. 2013, 49, 1303–1305. [Google Scholar] [CrossRef]
  129. Verwilst, P.; Kim, H.R.; Seo, J.; Sohn, N.W.; Cha, S.Y.; Kim, Y.; Maeng, S.; Shin, J.W.; Kwak, J.H.; Kang, C.; et al. Rational Design of in Vivo Tau Tangle-Selective Near-Infrared Fluorophores: Expanding the BODIPY Universe. J. Am. Chem. Soc. 2017, 139, 13393–13403. [Google Scholar] [CrossRef]
  130. Savelieff, M.G.; Detoma, A.S.; Derrick, J.S.; Lim, M.H. The Ongoing Search for Small Molecules to Study Metal-Associated Amyloid-β Species in Alzheimers Disease. Acc. Chem. Res. 2014, 47, 2475–2482. [Google Scholar] [CrossRef]
  131. Kim, D.; Baik, S.H.; Kang, S.; Cho, S.W.; Bae, J.; Cha, M.Y.; Sailor, M.J.; Mook-Jung, I.; Ahn, K.H. Close Correlation of Monoamine Oxidase Activity with Progress of Alzheimer’s Disease in Mice, Observed by in Vivo Two-Photon Imaging. ACS Cent. Sci. 2016, 2, 967–975. [Google Scholar] [CrossRef]
  132. Wu, W.; Zheng, T.; Tian, Y. An Enzyme-Free Amplification Strategy Based on Two-Photon Fluorescent Carbon Dots for Monitoring MiR-9 in Live Neurons and Brain Tissues of Alzheimer’s Disease Mice. Chem. Commun. 2020, 56, 8083–8086. [Google Scholar] [CrossRef]
  133. Zhang, X.; Tian, Y.; Yuan, P.; Li, Y.; Yaseen, M.A.; Grutzendler, J.; Moore, A.; Ran, C. A Bifunctional Curcumin Analogue for Two-Photon Imaging and Inhibiting Crosslinking of Amyloid Beta in Alzheimer’s Disease. Chem. Commun. 2014, 50, 11550–11553. [Google Scholar] [CrossRef] [Green Version]
  134. Takasaki, K.; Abbasi-Asl, R.; Waters, J. Superficial Bound of the Depth Limit of Two-Photon Imaging in Mouse Brain. eNeuro 2020, 7, ENEURO.0255-19.2019. [Google Scholar] [CrossRef] [Green Version]
  135. Gu, L.; Kleiber, S.; Schmid, L.; Nebeling, F.; Chamoun, M.; Steffen, J.; Wagner, J.; Fuhrmann, M. Long-Term In Vivo Imaging of Dendritic Spines in the Hippocampus Reveals Structural Plasticity. J. Neurosci. 2014, 34, 13948–13953. [Google Scholar] [CrossRef] [Green Version]
  136. Hong, S.J.; Choi, S.C.; Cho, J.Y.; Joo, H.J.; Park, J.H.; Yu, C.W.; Lim, D.S. Pioglitazone Increases Circulating MicroRNA-24 with Decrease in Coronary Neointimal Hyperplasia in Type 2 Diabetic Patients—Optical Coherence Tomography Analysis. Circ. J. 2015, 79, CJ-14-0964. [Google Scholar] [CrossRef] [Green Version]
  137. Koga, S.; Ikeda, S.; Yoshida, T.; Nakata, T.; Takeno, M.; Koide, Y.; Kawano, H.; Maemura, K.; Georges, J.L.; Hanssen, M.; et al. Decrease in Circulating MicroRNA-24 Is Associated with Increase in Neointimal Hyperplasia: Optical Coherence Tomography Analysis. Eur. Heart J. 2013, 34, P5438. [Google Scholar] [CrossRef] [Green Version]
  138. Liu, X.; Sun, Y.; Wu, J.; Tian, J.; Zheng, Y.; Wang, P.; Yu, M.; Zhang, S.; Wang, M.; Dong, H.; et al. GW27-E0063 Application of Optical Coherent Tomography to Evaluate the Stability of Atherosclerotic Plaque in Macrophages Regulated by MicroRNA Let-7b. J. Am. Coll. Cardiol. 2016, 68, C3. [Google Scholar] [CrossRef]
  139. O’Bryhim, B.E.; Lin, J.B.; van Stavern, G.P.; Apte, R.S. OCT Angiography Findings in Preclinical Alzheimer’s Disease: 3-Year Follow-Up. Ophthalmology 2021, 128, 1489–1491. [Google Scholar] [CrossRef]
  140. Chan, V.T.T.; Sun, Z.; Tang, S.; Chen, L.J.; Wong, A.; Tham, C.C.; Wong, T.Y.; Chen, C.; Ikram, M.K.; Whitson, H.E.; et al. Spectral-Domain OCT Measurements in Alzheimer’s Disease: A Systematic Review and Meta-Analysis. Ophthalmology 2019, 126, 497–510. [Google Scholar] [CrossRef]
  141. Van de Kreeke, J.A.; Nguyen, H.T.; Konijnenberg, E.; Tomassen, J.; den Braber, A.; ten Kate, M.; Yaqub, M.; van Berckel, B.; Lammertsma, A.A.; Boomsma, D.I.; et al. Optical Coherence Tomography Angiography in Preclinical Alzheimer’s Disease. Br. J. Ophthalmol. 2020, 104, 157–161. [Google Scholar] [CrossRef]
  142. Leuba, G.; Saini, K. Pathology of Subcortical Visual Centres in Relation to Cortical Degeneration in Alzheimer’s Disease. Neuropathol. Appl. Neurobiol. 1995, 21, 410–422. [Google Scholar] [CrossRef]
  143. Song, S.; Lee, J.U.; Jeon, M.J.; Kim, S.; Sim, S.J. Detection of Multiplex Exosomal MiRNAs for Clinically Accurate Diagnosis of Alzheimer’s Disease Using Label-Free Plasmonic Biosensor Based on DNA-Assembled Advanced Plasmonic Architecture. Biosens. Bioelectron. 2022, 199, 113864. [Google Scholar] [CrossRef]
  144. Jebelli, A.; Oroojalian, F.; Fathi, F.; Mokhtarzadeh, A.; Guardia, M. de la Recent Advances in Surface Plasmon Resonance Biosensors for MicroRNAs Detection. Biosens. Bioelectron. 2020, 169, 112599. [Google Scholar] [CrossRef]
Figure 1. Interconnection of synapses and the cerebrovascular system. (a) A1 astrocytes cause a high production of Aβ by producing inflammatory factors, including IFN γ, IL-1β, TNF-α, IL-6, TGF-β, and massive amyloid precursor protein-cleaving enzymes, such as β-secretase (BACE1) and γ-secretase. (b) A2 astrocytes release neurotrophic factors, such BDNF, VEGF, and bFGF, which contribute to synaptic repair and neuroprotection. (c) Astrocytes secrete inflammatory factors, such as IL-1β and TNFα, to stimulate the production of ADEVs, carrying miR-125a-p and miR-16-5p to target NTKR3 to regulate synaptic stability. (d) Microglia alter the synaptic phenotypes in the hippocampal neuron by releasing a series of complement factors (e.g., C1q and C3), (e) inflammatory factors (e.g., IL-1α, TNF, and type-I IFN), and (f) extracellular vesicles (EVs) to deliver miR-146a-5p to target SYT1 and NLG1 to regulate synaptic density. (g) The cerebrovascular system is composed of endothelial cells, smooth muscle cells (SMCs), pericytes, perivascular immune cells (e.g., microglia), and surrounding astrocytes. Abnormal cerebral blood flow leads to a greater tendency for Aβ to be deposited within the basement membrane of blood vessels. (h) miR-126 in the endothelial cells participate in the regulation of cerebral blood flow. (i) miR-200b-3p, miR-200c-3p, and miR-205-5p considerably reduced low-density LRP1 in microvascular endothelial cells impairing Aβ clearance. (j) miR-126 and miR-145, specifically in brain capillaries, elevate pericyte coverage as well as promote endothelial cell activation. (k) Providing neurons with mitochondria-related energy chemicals, such as adenosine triphosphate (ATP), nicotinamide adenine dinucleotide (NAD+), creatine, or diazoxide are efficacious treatments that can effectively protect neurons from metabolic and Aβ-toxicity damage.
Figure 1. Interconnection of synapses and the cerebrovascular system. (a) A1 astrocytes cause a high production of Aβ by producing inflammatory factors, including IFN γ, IL-1β, TNF-α, IL-6, TGF-β, and massive amyloid precursor protein-cleaving enzymes, such as β-secretase (BACE1) and γ-secretase. (b) A2 astrocytes release neurotrophic factors, such BDNF, VEGF, and bFGF, which contribute to synaptic repair and neuroprotection. (c) Astrocytes secrete inflammatory factors, such as IL-1β and TNFα, to stimulate the production of ADEVs, carrying miR-125a-p and miR-16-5p to target NTKR3 to regulate synaptic stability. (d) Microglia alter the synaptic phenotypes in the hippocampal neuron by releasing a series of complement factors (e.g., C1q and C3), (e) inflammatory factors (e.g., IL-1α, TNF, and type-I IFN), and (f) extracellular vesicles (EVs) to deliver miR-146a-5p to target SYT1 and NLG1 to regulate synaptic density. (g) The cerebrovascular system is composed of endothelial cells, smooth muscle cells (SMCs), pericytes, perivascular immune cells (e.g., microglia), and surrounding astrocytes. Abnormal cerebral blood flow leads to a greater tendency for Aβ to be deposited within the basement membrane of blood vessels. (h) miR-126 in the endothelial cells participate in the regulation of cerebral blood flow. (i) miR-200b-3p, miR-200c-3p, and miR-205-5p considerably reduced low-density LRP1 in microvascular endothelial cells impairing Aβ clearance. (j) miR-126 and miR-145, specifically in brain capillaries, elevate pericyte coverage as well as promote endothelial cell activation. (k) Providing neurons with mitochondria-related energy chemicals, such as adenosine triphosphate (ATP), nicotinamide adenine dinucleotide (NAD+), creatine, or diazoxide are efficacious treatments that can effectively protect neurons from metabolic and Aβ-toxicity damage.
Biomolecules 12 01505 g001
Table 1. Summary of studies implying miRNA in synapse-associated phenotypes.
Table 1. Summary of studies implying miRNA in synapse-associated phenotypes.
PhenotypemiRNATarget GeneReferences
Synaptic densitymiR-135a-5pROCK2[64]
miR-138SIRT1[65]
Synaptic transmissionmiR-26a-5p-[14]
miR-135a-5pROCK2[64]
miR-484-[66]
Synaptic plasticitymiR-484-[66]
miR-129CAMK2D[67]
CAMK4
PPP3CA
PRKCB
miR-181aGluA2[51]
miR-132HDAC[68]
RBFOX1[69]
miR-135a-5pROCK2[64]
miR-124PTPN1[62]
SynaptotoxicitymiR-181aGluA2[51]
miR-132HDAC[68]
Synaptic lossmiR-146aNKD2[27]
miR-9-5pSIRT1[70]
Synaptic integritymiR-30bEPHB2[71]
SIRT1
GluA2
miR-34cSYT1[72]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, X.; Chen, S.-C.; Ip, J.P.K. Diverse and Composite Roles of miRNA in Non-Neuronal Cells and Neuronal Synapses in Alzheimer’s Disease. Biomolecules 2022, 12, 1505. https://doi.org/10.3390/biom12101505

AMA Style

Li X, Chen S-C, Ip JPK. Diverse and Composite Roles of miRNA in Non-Neuronal Cells and Neuronal Synapses in Alzheimer’s Disease. Biomolecules. 2022; 12(10):1505. https://doi.org/10.3390/biom12101505

Chicago/Turabian Style

Li, Xinrong, Shih-Chi Chen, and Jacque Pak Kan Ip. 2022. "Diverse and Composite Roles of miRNA in Non-Neuronal Cells and Neuronal Synapses in Alzheimer’s Disease" Biomolecules 12, no. 10: 1505. https://doi.org/10.3390/biom12101505

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop