Next Article in Journal
Multiscale Carbon Fibre–Carbon Nanotube Composites of Poly(Vinyl Chloride)—An Evaluation of Their Properties and Structure
Previous Article in Journal
[BMP]+[BF4]-Modified CsPbI1.2Br1.8 Solar Cells with Improved Efficiency and Suppressed Photoinduced Phase Segregation
Previous Article in Special Issue
Targeting Members of the Chemokine Family as a Novel Approach to Treating Neuropathic Pain
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Neuroinflammation of Microglial Regulation in Alzheimer’s Disease: Therapeutic Approaches

by
Haiyun Chen
1,†,
Yuhan Zeng
2,3,4,5,†,
Dan Wang
2,3,4,5,
Yichen Li
6,
Jieyu Xing
1,
Yuejia Zeng
1,
Zheng Liu
7,
Xinhua Zhou
8,* and
Hui Fan
2,3,4,5,*
1
College of Pharmacy, Clinical Pharmacy (School of Integrative Pharmacy), Guangdong Pharmaceutical University, Guangzhou 510006, China
2
Guangdong Metabolic Diseases Research Center of Integrated Chinese and Western Medicine, Guangzhou 510006, China
3
Guangdong TCM Key Laboratory for Metabolic Diseases, Guangzhou 510006, China
4
Key Laboratory of Glucolipid Metabolic Disorder, Ministry of Education of China, Guangzhou 510006, China
5
Key Unit of Modulating Liver to Treat Hyperlipemia SATCM, State Administration of Traditional Chinese Medicine, Guangzhou 510006, China
6
Guangdong Provincial Key Laboratory of Research and Development of Natural Drugs, School of Pharmacy, Guangdong Medical University, Zhanjiang 524023, China
7
School of Medicine, Foshan University, Foshan 528000, China
8
Guangzhou Eighth People’s Hospital, Guangzhou Medical University, Guangzhou 510000, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2024, 29(7), 1478; https://doi.org/10.3390/molecules29071478
Submission received: 4 February 2024 / Revised: 13 March 2024 / Accepted: 23 March 2024 / Published: 26 March 2024
(This article belongs to the Special Issue Developing Drug Strategies for the Neuroprotective Treatment)

Abstract

:
Alzheimer’s disease (AD) is a complex degenerative disease of the central nervous system that is clinically characterized by a progressive decline in memory and cognitive function. The pathogenesis of AD is intricate and not yet fully understood. Neuroinflammation, particularly microglial activation-mediated neuroinflammation, is believed to play a crucial role in increasing the risk, triggering the onset, and hastening the progression of AD. Modulating microglial activation and regulating microglial energy metabolic disorder are seen as promising strategies to intervene in AD. The application of anti-inflammatory drugs and the targeting of microglia for the prevention and treatment of AD has emerged as a new area of research interest. This article provides a comprehensive review of the role of neuroinflammation of microglial regulation in the development of AD, exploring the connection between microglial energy metabolic disorder, neuroinflammation, and AD development. Additionally, the advancements in anti-inflammatory and microglia-regulating therapies for AD are discussed.

1. Introduction

Alzheimer’s disease (AD) is a complex neurodegenerative order that primarily affects memory and cognitive function. It is the most common cause of dementia, accounting for 60–80% of cases. According to the World Alzheimer Report 2019, the prevalence of AD dementia is projected to rise from 50 million globally in 2019 to 152 million by 2050 due to the aging population [1]. This translates to a new dementia case every three seconds and an estimated annual cost of USD 1 trillion, which is expected to double by 2030. Given the increasing number of AD patients and the rising costs associated with their care, there is a pressing need to develop effective treatments for this condition.
The pathological features of AD are primarily characterized by the formation of senile plaques (SPs) through amyloid beta (Aβ) aggregation and neurofibrillary tangles (NFTs) due to tau protein hyperphosphorylation [2]. Research indicates that many neurodegenerative diseases stem from the accumulation of misfold proteins, such as Aβ aggregation and hyperphosphorylated tau in AD, as well as α-synuclein accumulation in Parkinson’s disease (PD) [3]. Targeting these misfolded proteins to slow down or delay the progression of neurodegenerative disorder has been a significant focus in drug development. Over 200 anti-AD drugs targeting Aβ and tau therapies have undergone clinical trials, but only a few have received approval from the Food and Drug Administration (FDA) with considerable controversy [4], highlighting the need for a better understanding of the pathological mechanism of AD. Recent studies have revealed the complexity of AD pathogenesis, involving factors like Aβ, tau, neuroinflammation, and the immune system, forming an intricate network that regulates AD pathology [5]. Among them, microglial activation can enhance Aβ clearance [6] and directly drive the spread of tau proteins in the Braak phase, and even Aβ-mediated tau spread is also dependent on microglial activation [7], indicating the pivotal role of microglial activation-mediated neuroinflammation in AD risk, onset, and advancement [8]. Such microglial activation is also closely linked to metabolic inflammation in the body, leading to a growing interest in the use of anti-inflammatory drugs for AD prevention and treatment [9]. Therefore, we present a review of the role of microglial activation-mediated neuroinflammation and microglial energy metabolic disorder in the pathological process of AD, as well as the potential therapeutic interventions through small molecule drug treatments.
As reported, Aβ deposition is key factor in the pathogenesis of AD, resulting from the cleavage of amyloid precursor protein (APP) by secretase enzymes, mainly including alpha-secretase, β-secretase, and gamma-secretase [10]. These enzymes generate soluble APPα, APPβ, Aβ, and APP intracellular domains [11,12]. While soluble peptides can be recycled intracellularly, Aβ peptides can be further cleaved into shorter peptides like Aβ40 and Aβ42 [13]. Aβ42 is particularly prone to forming beta-amyloid plaques (ABPs), considered critical in AD pathogenesis. ABPs may disrupt the signaling between neurons, leading to brain damage, memory loss, and recognition issues [14]. ABPs have also been reported to initiate an immune response that leads to neuroinflammation, which may damage the surrounding neurons [15]. Additionally, ABPs trigger an immune response causing neuroinflammation that harms surrounding neurons. ABPs are also found in cerebral amyloid angiopathy (CAA) associated with AD [16]. CAA can eventually result in the leakage or rupture of blood vessels located outside of the cell. Another neuropathological characteristic of AD is the presence of NFTs, which are aggregates of abnormal tau proteins found in the neuronal cytoplasm that form paired helical filaments [17]. Normally, tau proteins are situated on the surface of microtubules and contribute to their structural integrity. When the formation of ABPs in the extracellular region occurs, ABPs can transfer a phosphate group to the tau protein through the activation of the kinases pathway, resulting in the phosphorylated tau protein becoming isolated from the microtubule [18]. Such phosphorylated tau proteins can make microtubules incomplete, leading to a loss of their signaling function, cell death, or apoptosis [19]. Given the crucial roles of Aβ and tau in the pathogenesis of AD, over 200 anti-AD drugs targeting these proteins have been tested in clinical trials to slow down disease progression, although only a few have received FDA approval [2]. The Aβ and tau hypothesis of AD is under scrutiny, and the development of disease-modifying treatments for AD has proven to be challenging. In recent decades, neuroinflammation has emerged as a third major neuropathological feature in the brains of AD patients, alongside Aβ and NFTs, forming a complex network that influences AD pathology [20,21,22,23]. Increased pro-inflammatory cytokines, such as tumor necrosis factor α (TNF-α) and interleukin 6 (IL-6), have been found in the serum and brain tissue of individuals with AD when compared to controls [24]. The initial phases of AD are marked by persistent neuroinflammation. This early inflammation, as the disease progresses, contributes to and intensifies the production of Aβ and NFT, leading to neuronal toxicity and death [25,26,27]. Additionally, endogenous bioactive lipids including eicosanoids, specialized pro-catabolic lipid mediators, lysophospholipids, and endocannabinoids in the brain which regulate a multitude of cellular and molecular processes, are closely linked to chronic inflammation [28]. The disruption of these bioactive lipids significantly enhances the development and progression of neurodegenerative diseases, leading to chronic damage like AD [10]. Ongoing inflammation leads to the continuous release of different inflammatory cytokines, resulting in a pro-inflammatory response that outweighs the anti-inflammatory response, and ultimately harming neurons and causing various pathological changes in the body [29,30,31]. Epidemiological studies have shown that the delayed onset of AD in certain populations due to the use of anti-inflammatory drugs implies a potential role in modulating neuroinflammation [32], underscoring the significance of neuroinflammation in the development of neurodegenerative diseases.

2. Microglia’s Role in AD

2.1. Main Physiological Functions of Microglia

The central nervous system (CNS) is a complex network of neurons and glial cells, of which glial cells are composed of astrocytes, oligodendrocytes, and NG2 cells composed of macroglia and microglia. These cells provide crucial support for brain functions. Microglia, a type of macrophage, are innate immune cells in the CNS that originate from a single mesoderm but have multiple neuroectodermal lineages, distinguishing them from other brain cells. Microglia play key roles in various physiological processes such as brain development, activity, and plasticity [33]. As reported, microglia are crucial regulators of CNS development and homeostasis through neuron–microglial interactions, removal of cellular debris, secretion of trophic factors, and synaptic pruning and remodeling [34]. Additionally, microglia also contribute to memory and cognitive functions by modulation of neuronal numbers and neural networks [35]. Furthermore, they support neurogenesis and neuronal survival, promoting circuit plasticity, as well as neuronal survival by supporting trophic factors, suggesting neuron–glia interaction is important for network formation in the developing brain [36,37,38]. Importantly, microglia can also accumulate in the cerebellar region to eliminate the dying neurons and nonfunctional synapses with high levels of basal clearance activity [39]. As the primary innate immune cells in the CNS, microglia respond to harmful stimuli through recognition of pathogen-associated molecular patterns (PAMPs) and damage-associated molecular patterns (DAMPs) via receptors like pattern recognition receptors (PRRs) and viral receptors [40]. Concretely, PAMPs are typically shared structures of pathogens, which can be recognized by lipopolysaccharide (LPS) in bacterial cell walls and double-stranded RNA in viruses through PRRs including Toll-like receptors (TLRs) and NOD-like receptors (NLRs) activating immune cells and promoting inflammatory responses, whereas DAMPs are endogenous molecules released during cell damage or death, which can also be recognized by PRRs and trigger immune responses [41]. The recognition and immune responses to PAMPs and DAMPs play a crucial role in the immune system’s defense against external and internal threats [42,43]. In response to these insults, microglia can be activated and then secrete pro-inflammatory and anti-inflammatory factors, contributing to a double-edged sword role in CNS damage [44,45].

2.2. Microglial Activation-Mediated Neuroinflammation in AD

Under physiological conditions, microglia may appear to be in a ‘resting’ state, but they are actually extending their cell bodies to continually monitor potential brain damage [46,47]. In situations of pathological injury, microglia are rapidly activated and polarized, exhibiting processes such as proliferation, chemotaxis, phagocytosis, migration, and cytokine secretion. Depending on the severity of the damage, if the damage is mild, microglia quickly transition to anti-inflammatory ‘M2’ phenotypes, characterized by distant branching, minor cell body alterations, and release of anti-inflammatory factors. ‘M2’ microglia contribute to tissue repair, phagocytic activity enhancement, and neuroprotective effects. In cases of severe or prolonged damage, cells at the injury site emit signals to be phagocytized. Subsequently, microglia become fully activated, polarizing to pro-inflammatory ‘M1’ phenotypes or the toxic ‘M1’ state with round cell bodies and thickened protrusions [48,49,50]. These microglial cells mainly exert cytotoxic effects with the capability of efficiently clearing infected cells and debris from dying cells [50]. However, excessive activation of ‘M1’ microglia can result in neuronal function loss, damage, and degeneration, playing crucial roles in cerebrovascular and neurodegenerative diseases [49,51]. In the context of normal pathological damage, due to the diverse signals emitted by cells within the organism, microglial cells may polarize into either ‘M1’ phenotype or ‘M2’ phenotype states to protect the nervous system [50]. Overall, there is a strong controversy regarding this two-phenotype classification by simply defining their activated states since this method is not applicable for microglial cells in a complex brain environment. Commonly, over-activation of the ‘M1’ phenotype can harm neuronal function in AD [50,52,53]. It is well known that the anti-inflammatory phenotype microglia confer neuroprotective effects by increasing the expression of cytokines and proteins including the arginase 1 gene (Arg-1), transforming growth factor-β (TGF-β), and interleukin 10 (IL-10) involved in resolving inflammation, maintaining homeostasis, and promoting wound healing [54,55,56]. However, this initial anti-inflammatory response is self-limiting [57], since neuroinflammation occurring in neurodegenerative disease tends to be a chronic process, characterized by long-term stimuli-induced over-activation that can surpass the self-limiting capacity of the organism. Such over-activation leads to microglia continuously transforming into pro-inflammatory phenotypes and releasing significant quantities of pro-inflammatory factors such as TNF-α, interleukin 1β (IL-1β), and IL-6. These pro-inflammatory factors lead to neurotoxic responses such as oxidative stress and neuronal apoptosis, which are key triggers for the development of neurodegenerative diseases like AD [58]. This can also be used to explain the different phenotypes of activated microglia in the AD brain. Initially, microglia effectively remove excessive Aβ and tau through phagocytosis at the onset of neuroinflammation. However, prolonged microglial over-activation can lead to Aβ and tau aggregation, accelerating the progression of AD [59]. In particular, a large accumulation of Aβ can hinder microglial phagocytosis by reducing the activity of phagocytic receptors on microglia, increasing pro-inflammatory cytokines production and further Aβ accumulation. This forms a vicious positive feedback loop that accelerates neurodegeneration and neuronal death [60].
The modulation of microglia on tau involves the ability of microglia to remove the neurons containing tau protein, secrete tau protein, and then subsequently transmit tau protein to other neurons [61]. The secreted tau protein acts as “tau seeds”, leading to the accumulation of tau protein in the recipient cells [62,63,64]. However, the release of a large misfolded protein into the extracellular space may also lead to a role for microglia in clearance or processing of tau, since microglia are the obligate phagocytes of the brain. Microglia are capable of degrading tau protein, but their efficiency in this process is not as expected. The ineffective processing of tau protein by microglia could potentially contribute to the spread of tau pathology in vivo [65]. The continuous accumulation of tau in brain regions, along with the induction of phosphorylated tau protein by extracellular tau seeds, contributes to a vicious cycle. Additionally, neuroinflammation has been reported to exacerbate the pathological progression of tau by disrupting neural transport and inhibiting mitochondrial respiration [3]. It has been demonstrated that injurious stimuli such as LPS, prostaglandin E2, and tert-butyl hydroperoxide can trigger over-activation of microglia, leading to the promotion of Aβ aggregation [66]. Among them, LPS can also induce phosphorylation of tau protein in rTg4510 mice, resulting in the formation of NFTs [67], and this effect is associated with the activation of cyclin-dependent kinase 5 (CDK5) [68]. In vivo experiments utilizing gene silencing to induce microglial activation have revealed increased tau protein phosphorylation in the hippocampus and impaired synaptic integrity, highlighting the significant role of microglial activation-mediated neuroinflammation in the pathological progression of AD [69].

3. The Critical Role of Energy Metabolic Disorder of Microglia between Microglial Activation-Mediated Neuroinflammation and AD

The current evidence indicates that chronic peripheral inflammation could potentially initiate systemic inflammation, increasing production of pro-inflammatory cytokines and other mediators, which in turn trigger neuroinflammation in the diseased brain [70,71]. In instances of systemic metabolic inflammation, pro-inflammatory mediators can enter the brain, inducing microglial phagocytosis to engulf cell debris and damaged neurons by prompting microglia to swiftly alter their shape and extend their synapses [72]. The above morphological changes, phagocytosis, and translocation of microglial cells all require dynamic reorganization of the actin cytoskeleton, thus demanding a significant amount of adenosine triphosphate (ATP) support [73]. As the systemic metabolic inflammation-induced damage worsens, the insufficient ATP supply makes microglial cells transition from oxidative phosphorylation (OXPHOS) to frequent glycolysis to enhance ATP production [74]. As reported, in the context of neuroinflammation, pro-inflammatory microglial cells exhibit heightened glucose uptake and increased expression of glycolytic enzymes like hexokinase (HK), phosphofructokinase (PFK), and pyruvate kinase M2 (PKM2) [75]. When stimulated by LPS, microglial cells demonstrate elevated levels of glucose transporter 1 (GLUT1), HK2, and PFK1, indicating enhanced glycolytic capacity [76]. Conversely, blocking GLUT1 or inhibiting HK2 activity can diminish the production of pro-inflammatory cytokines, underscoring the critical role of glycolysis in driving the inflammatory response [77,78]. Recent studies have reported that the key glycolytic enzyme PKM2 also has a role in regulating pro-inflammatory microglial cell activation [78]. Nuclear PKM2 acts as a coactivator of the transcription factor STAT1, which promotes the expression of pro-inflammatory genes. Silencing PKM2 with TEPP-46 or inhibiting its nuclear translocation significantly reduces the inflammatory response in microglial cells [79]. Additionally, lactate, a product of glycolysis, has been found to enhance the release of pro-inflammatory factors by microglial cells [80] and induce histone lactylation, impacting macrophage polarization [81]. Studies have shown that pan-protein lysine lactylation and H3K18 lactylation are upregulated in aging microglial cells and hippocampal tissues of naturally aged mice and AD model mice [82], suggesting a crucial role of lactylation in glycolysis in the progression of AD. Conversely, sustained activation of microglial cells may increase the expression of genes related to glycolysis, promoting glycolysis and exacerbating neurodegenerative diseases like AD and PD [83,84,85]. Ruiyuan Pan’s research on AD revealed that excessive glucose uptake by microglial cells results in the production of high levels of lactate through glycolysis. This lactate induces lactylation epigenetic modifications on histones, which then regulate glycolytic genes like PKM2 to establish a glycolysis/H4K12 lactylation/PKM2 positive feedback loop. Such loop formation then promotes the build-up of significant inflammation, ultimately contributing to the pathology of AD [86]. Based on the crucial role of glycolysis in the activation and polarization of microglial cells during neuroinflammation, targeting related glycolytic enzymes or pathways during glycolysis period would be new therapeutic approaches to mitigate microglial activation-mediated neuroinflammation.
Current research suggested that the body undergoing frequent glycolysis decreased the microglial phagocytosis and migration activity for Aβ clearance [87]. The accumulation of Aβ could also impede glucose uptake, aerobic glycolysis, and ATP synthesis by regulating glucose transporter 4 (GLUT4) and phosphofructokinase. Studies have demonstrated a correlation between the glycolysis transition and the reduced ability of microglia to phagocytose Aβ [88]. As reported, the triggering receptor expressed on myeloid cells 2 (TREM2) found on microglia plays a vital role in Aβ phagocytosis [89,90], while the absence of TREM2 increased the accumulation of Aβ plaques [91]. Further analysis of mice lacking TREM2 reported that they displayed AD-like symptoms and exhibited reduced glycolysis and relative difficulty in clearing Aβ [92]. Besides the crucial role of TREM2 in altering microglial metabolism, inflammasomes, particularly the NLR family pyrin domain-containing 3 (NLRP3) inflammasome, are also implicated in AD pathology (Figure 1) [93]. The NLRP3 inflammasome serves as a sensor of the innate immune system and can be triggered by various factors, including Aβ and oxidative stress [94]. Activation of the NLRP3 inflammasome in microglia results in the release of pro-inflammatory cytokines like IL-1β and IL-18, leading to heightened neuroinflammation and neuronal damage [95]. The energy metabolic disorder of microglia is linked to the activation of the NLRP3 inflammasome, where the switch to glycolysis facilitates such inflammasome activation [96]. Blocking glycolysis in microglia has been demonstrated to inhibit Aβ-induced NLRP3 inflammasome activation, thereby reducing neuroinflammation and neurodegeneration [97]. Moreover, mitochondrial dysfunction characterized by increased ROS production in microglia can also trigger NLRP3 inflammasome activation, worsening AD-related pathology [98].
Collectively, the pro-inflammatory cytokines and other mediators produced by systemic inflammation could enter into the brain and induce microglial polarization and then weaken its phagocytosis for the misfolded proteins. The metabolic deficiency of ATP supply accelerates the glycolysis process during the above period of inflammatory response, which can further trigger the over-activation of microglial-mediated neuroinflammation, creating a negative feedback loop of ‘glucose hypometabolism-toxic protein accumulation-neurodegeneration’ [87,99], e.g., causing decreased memory and cognitive dysfunctions in AD patients, and may even lead to abnormal behaviors (Figure 1). Thus, targeting glycolytic metabolic pathways in microglia could potentially provide novel therapeutic approaches to mitigate the harmful effects of abnormal inflammatory microglia and regulate excessive inflammation in brain diseases.

4. Anti-Inflammatory Drugs for the Treatment of AD

Clearly, the success rate of anti-AD drug development is extremely low with 99% of candidate drugs discontinued due to lack of clinical benefits. Over 200 anti-AD drugs targeting Aβ and tau therapies have undergone clinical trials, but only two monoclonal antibodies targeting Aβ clearance have received approval from FDA. Aducanumab, one of the monoclonal antibodies targeting Aβ, has shown partial Aβ clearance and cognitive decline slowing effects, but is associated with vasogenic edema. Despite being deemed ineffective in Phase III in March 2019, it was ultimately FDA approved for treating mild AD, leading to significant debate [100]. Lecanemab, another monoclonal antibody against Aβ, targets soluble Aβ oligomers or mature fibrils, demonstrating promising outcomes in inhibiting Aβ and tau pathology with lower clearance activity and fewer side effects compared to aducanumab [101,102]. Consequently, based on the critical role of neuroinflammation in AD pathology, treating neural inflammation from its early stages has emerged as a widely accepted and promising treatment strategy. According to the data from the ClinicalTrials.gov registry, there are about 17% of drugs focusing on inflammation, and in Phase II, these drugs represent 20% of all drugs in development as of the index date of 1 January 2023. The number of anti-inflammatory drugs has been steadily increasing in recent years, with 20 in 2020 (16.5%), 19 in 2021 (15%), and 23 in 2022 (16%) [103]. This section explores investigational drugs that target neuroinflammation and modulate microglial polarization and energy metabolism. These drugs include cyclooxygenase inhibitors, kinase inhibitors, and drugs that regulate microglial polarization and energy metabolic disorder. Table 1 and Table 2 present a comprehensive list of anti-inflammatory small molecule drugs currently in clinical trials, while Table 3 presents the structural formulas of promising drugs in anti-AD research. These tables hope to offer valuable information for advancing drug development in this field.

4.1. Non-Steroidal Anti-Inflammatory Drugs (NSAIDs)

NSAIDs are currently the most commonly used antipyretic, analgesic, and anti-inflammatory drugs in clinic. The mechanisms of action of NSAIDs mainly inhibit cyclooxygenase 2 (COX-2) activity to reduce prostaglandin (PG) synthesis. Given the research progress on the relationship between systemic inflammation and microglial activation, NSAIDs also have drawn researchers’ widespread attention in AD treatment for their anti-neuroinflammatory activities, inhibitory effects on Aβ aggregation, and the modulation of γ-secretase [143,144,145]. CHF5074, a non-cyclooxygenase-inhibiting NSAID, has been shown to enhance cognitive performance and reduce intracerebral inflammation by promoting the expression of phagocytic and anti-inflammatory ‘M2’ phenotype markers in young Tg2576 mice [115]. Although CHF5074 (NCT01602393) has been shown to improve cognition and reduce brain inflammation in patients with mild cognitive impairment (MCI), its clinical trial progress has not been updated since the end of the Phase 2 clinical trial in 2014. Another example is azeliragon (NCT02916056), also known as TTP488, which is an inhibitor of the Receptor for Advanced Glycation End-products (RAGE). It is reported that RAGE can react with Aβ to form the RAGE–Aβ interaction, leading to a persistent neuroinflammatory response and the acceleration of AD degeneration [117]. The clinical data from Phase 2b trials of azeliragon suggested that it could slow cognitive decline in mild AD patients [146]. Recent investigations into the underlying mechanisms of azeliragon’s anti-AD properties have shown that it can improve NLRP3-associated inflammation, cell viability, apoptosis, and ROS production by mediating the JAK1/STAT3/NF-κB/IRF3 pathway [147]. Based on promising results of azeliragon from the Phase 2b study, a Phase 3 registration program (STEADFAST and STEADFAST Extension) is being conducted under a Special Protocol Assessment from the FDA in 2021. Although many epidemiological studies have shown that NSAIDs are effective in the treatment of AD, the relative risk of AD is reduced only in people taking long-term NSAIDs at least for more than 24 months [148]. Among them, only indomethacin has shown some slight efficacy in mild to moderate AD in numerous clinical trials of NSAIDs [5]. The protective effects of NSAIDs exhibited in the epidemiological statistics require more than 2 years’ administration of NSAIDs before the onset of AD, suggesting that NSAIDs must be taken very early and at doses that inhibit the advancing inflammation [149]. Other NSAIDs as anti-inflammatory agents under research are furosemide and dendrimers. As reported, furosemide can down-regulate COX-2 and iNOS protein expression to decrease the pro-inflammatory M1 phenotype and increase the anti-inflammatory M2 phenotype of microglia, respectively [150]. A series of low-generation dendrimers were designed and synthesized characterized by both targeting COX-2 inhibition and microglial activation suppression [151].

4.2. TLR4 Antagonists

TLR4 is an essential component of the natural immune system, playing an important role in intracellular signaling and neuronal death by regulating the TLR4/MAPKs/NF-κB pathway (Figure 1) to stimulate the production of pro-inflammatory mediators [152]. There is increasing evidence suggesting that the clearance of Aβ by microglia occurs on TLR4 signaling [153]. Studies on TLR4-mutant AD-model mice have shown that they accumulate more Aβ and exhibit greater cognitive deficits as compared to their TLR4-wild-type counterparts [154]. Conversely, TLR4-deficient/knockout AD-mice have demonstrated reduced Aβ deposits and neuroinflammatory factors, improving cognitive abilities [155]. Given the crucial role of TLR4 in AD pathogenesis, it is considered a promising therapeutic target for drug development. TLR4 antagonists such as TAK-242 have shown positive therapeutic effects in various models of peripheral diseases such as inflammatory bowel disease (IBD) by inhibiting the production of inflammatory mediators such as TNF-α, IL-6, and NO [156]. TAK-242, as a specific inhibitor of TLR4, has been found to enhance learning and memory abilities, decrease Aβ deposition, and protect neurons from apoptosis in APP/PS1 mice [118]. Additionally, naloxone has been shown to block TLR4 downstream signaling, leading to NO, TNF-α, and reactive oxygen species, through TLR4 signaling during neuroinflammation [120]. Another promising compound, Gx-50, extracted from Sichuan pepper, has demonstrated potent anti-inflammatory effects against Aβ-triggered microglial over-activation in AD mice by suppressing TLR4-mediated NF-κB/MAPK signaling [119]. Despite the fact that therapeutic targeting of TLR4 in animal models of inflammation or AD has yielded promising results, to date, there are few TLR4 antagonists reported in clinical trials.

4.3. p38 Mitogen-Activated Protein Kinase (p38 MAPK) Antagonists

Mitogen-activated protein kinase (MAPK) is a crucial protein family involved in various cellular processes such as cell proliferation, differentiation, inflammation, oxidative stress, and apoptosis in mammals. This family includes p38 MAPK kinase, extracellular signal-regulated kinase (ERK1/2), and c-Jun amino-terminal kinase (JNK)/stress-activated protein kinase (SAPK) [157]. Within the inflammatory signaling pathway, MAPK is primarily regulated by the upstream TLRs pathway, triggered by stimuli like LPS (Figure 1). This leads to the translocation of NF-κb, expression of p38 phosphorylation, and subsequent production of inflammatory factors like TNF-α and IL-6 [158]. Additionally, p38α MAPK kinase was reported to mediate impaired synaptic dysfunction in the hippocampus, causing memory deficits, and to be involved in Aβ production and tau pathology [159]. Based on the central role of p38 MAPK in chronic inflammation, many preclinical or clinical trials for the application of p38 MAPK inhibitors in inflammatory diseases were conducted. VX-745 is an inhibitor that selectively inhibits p38 MAPKα, which was found to be more permeable to the blood–brain barrier (BBB), significantly reducing IL-1β protein levels in the hippocampus and improving memory recognition in aging rats with 22–24 months of age [121]. In a small sample size of a short-term (12 months) Phase IIa clinical trial in early AD patients, VX745 showed good tolerability and adequate drug concentrations in the cerebrospinal fluid (CSF) and a tendency to improve episodic memory [121]. However, in a further long-term (24 months) Phase II clinical trial with a large sample size, VX745 did not show the ability to significantly improve episodic memory in patients with mild AD compared to the placebo group [159,160]. The results obtained from the Phase II clinical trials of VX745 suggest that higher doses and longer duration studies may be necessary to assess its effect on AD progression. Targeted protein degradation offers a unique advantage over gene knockout by selectively degrading proteins with pathological mutations or aberrant post-translational modifications. PRZ-18002, a protein degrader that binds to p38 MAPKα, has been reported to degrade phosphorylated p38 MAPK (p-p38) and mimetic mutants of p38 MAPK in a proteasome-dependent manner, alleviating microglial activation and Aβ deposition. This led to improved spatial memory and learning in a mouse model of AD [122]. Furthermore, natural products and gut microbiota metabolites have been found to inhibit the MAPK signaling pathway with significant anti-neuroinflammatory effects. Urolithin A (UA) and urolithin B (UB), isolated from metabolites of intestinal microorganisms, have shown to reduce NO levels and suppress mRNA levels of pro-inflammatory genes in LPS-treated microglia by inhibiting NF-κB, MAPKs (p38 and ERK1/2), and Akt signaling pathway activation [123].

4.4. Microglia Regulators

Cromolyn, a compound used in the clinical treatment of asthma, has been demonstrated to induce neuroprotective activation of microglia and reduce level of aggregation-prone Aβ [127]. Currently, cromolyn is undergoing a phase III trial as a potential agent for early AD treatment (NCT02547818). Furosemide also acted as a modulator of microglia, was found to rescue Aβ-induced neuroinflammation in microglia by regulating microglia polarization to promote anti-inflammatory phenotype transformation [150]. Resveratrol, a natural polyphenol, has the ability to modulate stress signals in microglia [125] leading to the attenuation of cognitive behavioral deficits [126]. Further investigation into the mechanism of action of resveratrol on microglia has revealed its ability to effectively reverse the LPS-induced polarization of microglia from a pro-inflammatory phenotype to an anti-inflammatory phenotype [124]. Research has indicated that PKM2 plays a crucial role in the regulation of energy metabolic disorder of microglia [86,159]. Additionally, TSG and CIAC001 have been identified as compounds with a targeted antagonistic effect on PKM2, demonstrating an anti-neuroinflammatory effect by restoring the PKM2 tetramer levels and inhibiting the nuclear translocation of PKM2 [124,128]. Rui-Yuan Pan discovered that sodium rutin treatment shifts the metabolic program from anaerobic glycolysis to mitochondrial OXPHOS and enhances microglial polarization towards anti-inflammatory phenotypes [130]. Benserazide has been shown to effectively inhibit PKM2, thereby blocking aerobic glycolysis and modulating OXPHOS [129]. While benserazide is a PKM2 inhibitor, currently used in combination with levodopa for PD treatment, its impact on energy metabolic disorder of microglia remains unexplored. Nevertheless, utilizing benserazide as a lead compound for developing novel PKM2 inhibitors to regulate energy metabolic disorder of microglia shows promise in AD treatment [129]. Currently, benzyl hydrazine is used in combination with levodopa for the treatment of Parkinson’s disease, but there is currently no research on how benserazide regulates energy metabolic disorder of microglia. However, we believe that using benserazide as a lead compound for research to develop novel PKM2 inhibitors and regulate energy metabolic disorder of microglia cells is one of the promising directions for treating AD. Rapamycin and salidroside, known mTOR inhibitors, have shown efficacy in AD treatment. However, there is a lack of research on the therapeutic effects of rapamycin in AD through the TREM2/mTOR pathway in regulating energy metabolic disorder of microglia [131,161,162]. Therefore, investigating the potential of rapamycin and salidroside as a research direction for modulating energy metabolic disorder of microglia in microglia is also promising.

4.5. Chinese Herbal Polysaccharides

In recent years, with the rapid development of isolation and purification techniques, more and more natural products, especially herbal polysaccharides, play an important role in neurodegenerative diseases such as AD due to their anti-neuroinflammatory and other neuroprotective activities [163]. For example, Ganoderma lucidum polysaccharides can not only promote neural re-generation to improve cognitive impairment in AD transgenic mice but also reduce Aβ levels and tau protein hyperphosphorylation and reduce ultrastructural damage to improve spatial memory in rats [134]. Astragalus polysaccharides can reduce astrocytic and microglial activation and exert biological activities such as antioxidant, anti-neuroinflammatory, and neuroprotective effects [135]. Dendrobium orchid polysaccharide can exert anti-neuroinflammatory effects by inhibiting oxidative stress and reducing the production of free radicals and pro-inflammatory factors such as TNF-α and IL-1β [136]. Furthermore, Astragalus polysaccharide can improve the learning memory ability of zebrafish, a model of AD, by upregulating N-cadherin protein levels and decreasing the phosphorylation of p38 [137]. Chinese medicine polysaccharides are mostly in preclinical studies due to their complex structure compared to small molecule drugs. Fortunately, with the conditional approval of GV-971 [133] by the China Food and Drug Administration in 2019 as a fucoidan for the treatment of mild to moderate AD, the unique advantages of Chinese medicine polysaccharides with mild therapeutic properties and low toxic side effects give them great potential for treatments of neurodegenerative diseases such as AD.

4.6. Other Anti-Inflammatory Drugs

In addition to the aforementioned anti-inflammatory drugs including the small molecule inhibitors and regulators, as well as Chinese medicine polysaccharides, there still exist some antibody biologics such as Anakinra, an IL-1β inhibitor [164], and Etanercept, a TNF-α inhibitor [165]. Most of these inhibitors are biomolecules, which are not only costly but have more difficulty to enter the CNS to exert their anti-neuroinflammation through the BBB, severely limiting their possibility in treating neurodegenerative diseases [166]. It is well known that a sharp decrease in acetylcholine occurs during the course of AD, a deficiency of which directly leads to cognitive deficits [167].
Acetylcholine, a neurotransmitter found in the cholinergic nervous system, is released by excited cholinergic neurons. Upon entering the synaptic cleft, acetylcholine can be hydrolyzed by acetylcholinesterase if not bound to receptors, leading to the termination of neurotransmission. Inhibiting acetylcholinesterase can result in the accumulation of acetylcholine, enhancing its action. The cholinergic hypothesis proposes that reduced synthesis of acetylcholine is a key factor in AD [167,168]. Therefore, inhibiting acetylcholinesterase can cause acetylcholine to accumulate, thereby prolonging and enhancing the action of acetylcholine. The cholinergic hypothesis suggests that the main cause of AD is the reduced synthesis of acetylcholine. Therefore, one potential therapeutic strategy for AD is to increase the cholinergic neurotransmitter levels in the brain by inhibiting the biological activity of acetylcholinesterase (AChE) [168]. Therefore, one potential therapeutic approach for AD involves increasing cholinergic neurotransmitter levels by inhibiting acetylcholinesterase activity. Drugs like tacrine, galantamine, donepezil, and cabalatin have been developed to target acetylcholinesterase and alleviate AD symptoms [169]. Due to tacrine’s severe hepatotoxicity, it has been withdrawn from the market [168]. Given its potent anti-AD activity, many new compounds were designed and synthesized by introducing pyrimidine, thiazole, and other heterocycles groups to the parent group of tacrine; one compound (No. 46) was able to inhibit neuroinflammation and aggregation of Aβ with low toxicity and IC50 value of 2 nM of AChE [138]. New compounds, such as compound No. 46, have been designed to inhibit neuroinflammation and Aβ aggregation with low toxicity and high AChE inhibition potency. Currently, acetylcholinesterase inhibitors like donepezil are being investigated in clinical trials, indicating that inhibiting AChE remains a potential strategy for AD treatment. (NCT04661280, NCT05345509, NCT04730635, NCT05525780).
Activation of the NLRP3 inflammasome play a key role in the pathogenesis of AD [45,170,171]. Several compounds, including MCC950, JC-124, CY-09, and inzomelid, have been recognized as inhibitors of NLRP3 with neuroprotective properties. These compounds effectively suppress NLRP3 activation and the production of IL-1β [140,141,142,172].

5. Discussion and Outlook

In summary, a large body of evidence suggests that microglial over-activation-mediated neuroinflammation and the energy metabolic disorder of microglia contribute significantly to the pathogenesis of AD, and research on anti-inflammatory drugs for the treatment of AD has also made some progress. In total, 25 anti-inflammatory candidate drugs, including biological preparations and small molecule compounds, have entered clinical trials, ranking second only to neurotransmitter receptor drugs as of 1 January 2023 [103]. Among them, masitinib and NE3107 have entered Phase III clinical trials, emphasizing that suppressing neuroinflammation has been considered as a potential therapeutic strategy for AD. Clearly, microglia are crucial mediators and effectors in the pathology of AD, but a slew of mysteries surrounding the interactions between microglia and AD remain unsolved. The early stage of microglial activation-mediated neurodegeneration has been linked to ‘M1’ phenotype-associated inflammatory factors such as TNF-α and IL-1β (Figure 1). In this stage, microglia are activated to endocytose pathological Aβ and tau but with cytotoxic effects, and the inefficiency of ATP supply accelerates the glycolysis process of microglia followed by continued microglial overactivation (Figure 1). In the later stage of microglial activation, the ‘M1’ phenotype fails to endocytose pathological Aβ and tau and their deposition return contributes to inflammatory activation and energy metabolism disorder of microglia. On the contrary, anti-inflammatory factors secreted by ‘M1’ phenotype microglia, such as IL-4 and IL-10, and inhibition of certain receptors, such as TREM2, PKM2, and NLRP3, aid in the restoration of learning and memory deficits in AD via various signaling pathways and mechanisms (Figure 1). The above double-edged sword mechanism of microglia regulation between ’protective immune effector‘ and ’harmful inducer-mediated neuroinflammation in AD‘ might constrain the development of anti-inflammatory drugs in AD treatment. Notably, long-term clinical use of NSAIDs has been epidemiologically shown to significantly reduce the risk of developing AD, suggesting the vital role of inflammation in AD pathology and the beneficial effect of anti-inflammatory agents for AD prevention and treatment. Although the short duration of clinical trials on NSAIDs for AD patients failed to exhibit direct evidence of clinical efficacy, as the research advances on the pathogenesis of microglia-mediated neuroinflammation in AD and given the continuous improvement of clinical trial design schemes considering the difference between the short and long duration, further progress will be made in the treatment of AD with anti-inflammatory drugs.

Author Contributions

H.C. and Y.Z. (Yuhan Zeng). reviewed the literature and drafted the original manuscript. D.W., Y.L., J.X. and Y.Z. (Yuejia Zeng). contributed to editing and expanding the manuscript. Z.L., X.Z. and H.F. revised the paper and approved the submitted manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported financially by the Natural Science Foundation of Guangdong Province (2022A1515010974), Science and Technology Program of Guangzhou (202201010387), the Scientific Research and Cultivation Foundation for young teachers of pharmacy of Guangdong Pharmaceutical University, the Scientific Research Projects (Characteristic Innovation) of General Universities in Guangdong Province (2022KTSCX059).

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Abbreviations

Aβ: Amyloid beta, AD: Alzheimer’s disease, APP: Amyloid protein precursor, ABPs: beta-amyloid plaques, AChE: acetylcholinesterase, CNS: Central nervous system, COX: Cyclooxygenase, DAMPs: Damage-associated molecular patterns, IL-1β: Interleukin 1 beta, IL-6: interleukin 6, LPS: Lipopolysaccharides, mTOR: Mammalian target of rapamycin, MAPK: mitogen-activated protein kinase, NLRP3: NLR: family pyrin domain containing 3, NSAIDs: Non-steroidal anti-inflammatory drugs, OXPHOS: oxidative phosphorylation, PAMPs: Pathogen-associated molecular patterns, PRRs: pattern recognition receptors, PKM2: pyruvate kinase M2, SPs: senile plaques, TREM2: triggering receptor expressed on myeloid cells 2, TGF-β: transforming growth factor-β, TLRs: Toll-like receptors, TNF-α: Tumor necrosis factor alpha, TLR4: Toll-like receptor 4, NFTs: neurofibrillary tangles.

References

  1. Lynch, C. World Alzheimer Report 2019: Attitudes to dementia, a global survey. Alzheimer’s Dement. 2020, 16, e038255. [Google Scholar] [CrossRef]
  2. Venegas, C.; Heneka, M.T. Inflammasome-mediated innate immunity in Alzheimer’s disease. FASEB J. 2019, 33, 13075–13084. [Google Scholar] [CrossRef] [PubMed]
  3. Yoshiyama, Y.; Higuchi, M.; Zhang, B.; Huang, S.-M.; Iwata, N.; Saido, T.C.; Maeda, J.; Suhara, T.; Trojanowski, J.Q.; Lee, V.M.-Y. Synapse loss and microglial activation precede tangles in a P301S tauopathy mouse model. Neuron 2007, 53, 337–351. [Google Scholar] [CrossRef]
  4. Walsh, S.; Merrick, R.; Milne, R.; Brayne, C. Aducanumab for Alzheimer’s disease? BMJ 2021, 374, n1682. [Google Scholar] [CrossRef] [PubMed]
  5. Lao, K.; Ji, N.; Zhang, X.; Qiao, W.; Tang, Z.; Gou, X. Drug development for Alzheimer’s disease: Review. J. Drug Target. 2019, 27, 164–173. [Google Scholar] [CrossRef]
  6. Wojcieszak, J.; Kuczyńska, K.; Zawilska, J.B. Role of Chemokines in the Development and Progression of Alzheimer’s Disease. J. Mol. Neurosci. 2022, 72, 1929–1951. [Google Scholar] [CrossRef]
  7. Pascoal, T.A.; Benedet, A.L.; Ashton, N.J.; Kang, M.S.; Therriault, J.; Chamoun, M.; Savard, M.; Lussier, F.Z.; Tissot, C.; Karikari, T.K.; et al. Microglial activation and tau propagate jointly across Braak stages. Nat. Med. 2021, 27, 1592–1599. [Google Scholar] [CrossRef]
  8. Ozben, T.; Ozben, S. Neuro-inflammation and anti-inflammatory treatment options for Alzheimer’s disease. Clin. Biochem. 2019, 72, 87–89. [Google Scholar] [CrossRef] [PubMed]
  9. Bornstein, S.R.; Voit-Bak, K.; Rosenthal, P.; Tselmin, S.; Julius, U.; Schatz, U.; Boehm, B.O.; Thuret, S.; Kempermann, G.; Reichmann, H.; et al. Extracorporeal apheresis therapy for Alzheimer disease—Targeting lipids, stress, and inflammation. Mol. Psychiatry 2020, 25, 275–282. [Google Scholar] [CrossRef]
  10. Wilkins, H.M.; Swerdlow, R.H. Amyloid precursor protein processing and bioenergetics. Brain Res. Bull. 2017, 133, 71–79. [Google Scholar] [CrossRef]
  11. Allinson, T.M.J.; Parkin, E.T.; Turner, A.J.; Hooper, N.M. ADAMs family members as amyloid precursor protein alpha-secretases. J. Neurosci. Res. 2003, 74, 342–352. [Google Scholar] [CrossRef]
  12. Cai, H.; Wang, Y.; McCarthy, D.; Wen, H.; Borchelt, D.R.; Price, D.L.; Wong, P.C. BACE1 is the major beta-secretase for generation of Abeta peptides by neurons. Nat. Neurosci. 2001, 4, 233–234. [Google Scholar] [CrossRef] [PubMed]
  13. Wolfe, M.S.; Miao, Y. Structure and mechanism of the γ-secretase intramembrane protease complex. Curr. Opin. Struct. Biol. 2022, 74, 102373. [Google Scholar] [CrossRef] [PubMed]
  14. Tolar, M.; Abushakra, S.; Sabbagh, M. The path forward in Alzheimer’s disease therapeutics: Reevaluating the amyloid cascade hypothesis. Alzheimer’s Dement. 2020, 16, 1553–1560. [Google Scholar] [CrossRef]
  15. Wang, Y.; Ulland, T.K.; Ulrich, J.D.; Song, W.; Tzaferis, J.A.; Hole, J.T.; Yuan, P.; Mahan, T.E.; Shi, Y.; Gilfillan, S.; et al. TREM2-mediated early microglial response limits diffusion and toxicity of amyloid plaques. J. Exp. Med. 2016, 213, 667–675. [Google Scholar] [CrossRef] [PubMed]
  16. Greenberg, S.M.; Bacskai, B.J.; Hernandez-Guillamon, M.; Pruzin, J.; Sperling, R.; van Veluw, S.J. Cerebral amyloid angiopathy and Alzheimer disease-one peptide, two pathways. Nat. Rev. Neurol. 2020, 16, 30–42. [Google Scholar] [CrossRef]
  17. Bloom, G.S. Amyloid-β and tau: The trigger and bullet in Alzheimer disease pathogenesis. JAMA Neurol. 2014, 71, 505–508. [Google Scholar] [CrossRef] [PubMed]
  18. Wang, Y.; Mandelkow, E. Tau in physiology and pathology. Nat. Rev. Neurosci. 2016, 17, 22–35. [Google Scholar] [CrossRef]
  19. Spires-Jones, T.L.; Attems, J.; Thal, D.R. Interactions of pathological proteins in neurodegenerative diseases. Acta Neuropathol. 2017, 134, 187–205. [Google Scholar] [CrossRef]
  20. Doens, D.; Fernández, P.L. Microglia receptors and their implications in the response to amyloid β for Alzheimer’s disease pathogenesis. J. Neuroinflamm. 2014, 11, 48. [Google Scholar] [CrossRef]
  21. Heneka, M.T.; Carson, M.J.; El Khoury, J.; Landreth, G.E.; Brosseron, F.; Feinstein, D.L.; Jacobs, A.H.; Wyss-Coray, T.; Vitorica, J.; Ransohoff, R.M.; et al. Neuroinflammation in Alzheimer’s disease. Lancet Neurol. 2015, 14, 388–405. [Google Scholar] [CrossRef] [PubMed]
  22. Calsolaro, V.; Edison, P. Neuroinflammation in Alzheimer’s disease: Current evidence and future directions. Alzheimer’s Dement. 2016, 12, 719–732. [Google Scholar] [CrossRef] [PubMed]
  23. Piirainen, S.; Youssef, A.; Song, C.; Kalueff, A.V.; Landreth, G.E.; Malm, T.; Tian, L. Psychosocial stress on neuroinflammation and cognitive dysfunctions in Alzheimer’s disease: The emerging role for microglia? Neurosci. Biobehav. Rev. 2017, 77, 148–164. [Google Scholar] [CrossRef]
  24. Fillit, H.; Ding, W.H.; Buee, L.; Kalman, J.; Altstiel, L.; Lawlor, B.; Wolf-Klein, G. Elevated circulating tumor necrosis factor levels in Alzheimer’s disease. Neurosci. Lett. 1991, 129, 318–320. [Google Scholar] [CrossRef]
  25. Bamberger, M.E.; Landreth, G.E. Microglial interaction with beta-amyloid: Implications for the pathogenesis of Alzheimer’s disease. Microsc. Res. Tech. 2001, 54, 59–70. [Google Scholar] [CrossRef] [PubMed]
  26. Kitazawa, M.; Yamasaki, T.R.; LaFerla, F.M. Microglia as a potential bridge between the amyloid beta-peptide and tau. Ann. N. Y. Acad. Sci. 2004, 1035, 85–103. [Google Scholar] [CrossRef] [PubMed]
  27. Garwood, C.J.; Pooler, A.M.; Atherton, J.; Hanger, D.P.; Noble, W. Astrocytes are important mediators of Aβ-induced neurotoxicity and tau phosphorylation in primary culture. Cell Death Dis. 2011, 2, e167. [Google Scholar] [CrossRef] [PubMed]
  28. Cristino, L.; Bisogno, T.; Di Marzo, V. Cannabinoids and the expanded endocannabinoid system in neurological disorders. Nat. Rev. Neurol. 2020, 16, 9–29. [Google Scholar] [CrossRef]
  29. Chiurchiù, V.; Tiberi, M.; Matteocci, A.; Fazio, F.; Siffeti, H.; Saracini, S.; Mercuri, N.B.; Sancesario, G. Lipidomics of Bioactive Lipids in Alzheimer’s and Parkinson’s Diseases: Where Are We? Int. J. Mol. Sci. 2022, 23, 6235. [Google Scholar] [CrossRef]
  30. Liu, P.; Wang, Y.; Sun, Y.; Peng, G. Neuroinflammation as a Potential Therapeutic Target in Alzheimer’s Disease. Clin. Interv. Aging 2022, 17, 665–674. [Google Scholar] [CrossRef]
  31. Wong-Guerra, M.; Calfio, C.; Maccioni, R.B.; Rojo, L.E. Revisiting the neuroinflammation hypothesis in Alzheimer’s disease: A focus on the druggability of current targets. Front. Pharmacol. 2023, 14, 1161850. [Google Scholar] [CrossRef] [PubMed]
  32. Scipioni, L.; Ciaramellano, F.; Carnicelli, V.; Leuti, A.; Lizzi, A.R.; De Dominicis, N.; Oddi, S.; Maccarrone, M. Microglial Endocannabinoid Signalling in AD. Cells 2022, 11, 1237. [Google Scholar] [CrossRef] [PubMed]
  33. Tay, T.L.; Carrier, M.; Tremblay, M.-È. Physiology of Microglia. Adv. Exp. Med. Biol. 2019, 1175, 129–148. [Google Scholar] [CrossRef] [PubMed]
  34. Tay, T.L.; Savage, J.C.; Hui, C.W.; Bisht, K.; Tremblay, M.-È. Microglia across the lifespan: From origin to function in brain development, plasticity and cognition. J. Physiol. 2017, 595, 1929–1945. [Google Scholar] [CrossRef] [PubMed]
  35. Weinhard, L.; d’Errico, P.; Leng Tay, T. Headmasters: Microglial regulation of learning and memory in health and disease. AIMS Mol. Sci. 2018, 5, 63–89. [Google Scholar] [CrossRef]
  36. Ueno, M.; Fujita, Y.; Tanaka, T.; Nakamura, Y.; Kikuta, J.; Ishii, M.; Yamashita, T. Layer V cortical neurons require microglial support for survival during postnatal development. Nat. Neurosci. 2013, 16, 543–551. [Google Scholar] [CrossRef] [PubMed]
  37. Arnò, B.; Grassivaro, F.; Rossi, C.; Bergamaschi, A.; Castiglioni, V.; Furlan, R.; Greter, M.; Favaro, R.; Comi, G.; Becher, B.; et al. Neural progenitor cells orchestrate microglia migration and positioning into the developing cortex. Nat. Commun. 2014, 5, 5611. [Google Scholar] [CrossRef] [PubMed]
  38. Ribeiro Xavier, A.L.; Kress, B.T.; Goldman, S.A.; Lacerda de Menezes, J.R.; Nedergaard, M. A Distinct Population of Microglia Supports Adult Neurogenesis in the Subventricular Zone. J. Neurosci. 2015, 35, 11848–11861. [Google Scholar] [CrossRef] [PubMed]
  39. Ayata, P.; Badimon, A.; Strasburger, H.J.; Duff, M.K.; Montgomery, S.E.; Loh, Y.-H.E.; Ebert, A.; Pimenova, A.A.; Ramirez, B.R.; Chan, A.T.; et al. Epigenetic regulation of brain region-specific microglia clearance activity. Nat. Neurosci. 2018, 21, 1049–1060. [Google Scholar] [CrossRef]
  40. Stephenson, J.; Nutma, E.; van der Valk, P.; Amor, S. Inflammation in CNS neurodegenerative diseases. Immunology 2018, 154, 204–219. [Google Scholar] [CrossRef]
  41. Zindel, J.; Kubes, P. DAMPs, PAMPs, and LAMPs in Immunity and Sterile Inflammation. Annu. Rev. Pathol. 2020, 15, 493–518. [Google Scholar] [CrossRef] [PubMed]
  42. Takeuchi, O.; Akira, S. Pattern recognition receptors and inflammation. Cell 2010, 140, 805–820. [Google Scholar] [CrossRef] [PubMed]
  43. Chen, G.Y.; Nuñez, G. Sterile inflammation: Sensing and reacting to damage. Nat. Rev. Immunol. 2010, 10, 826–837. [Google Scholar] [CrossRef] [PubMed]
  44. Kwon, H.S.; Koh, S.-H. Neuroinflammation in neurodegenerative disorders: The roles of microglia and astrocytes. Transl. Neurodegener. 2020, 9, 42. [Google Scholar] [CrossRef] [PubMed]
  45. Onyango, I.G.; Jauregui, G.V.; Čarná, M.; Bennett, J.P.; Stokin, G.B. Neuroinflammation in Alzheimer’s Disease. Biomedicines 2021, 9, 524. [Google Scholar] [CrossRef] [PubMed]
  46. Borst, K.; Dumas, A.A.; Prinz, M. Microglia: Immune and non-immune functions. Immunity 2021, 54, 2194–2208. [Google Scholar] [CrossRef] [PubMed]
  47. Harry, G.J. Microglia during development and aging. Pharmacol. Ther. 2013, 139, 313–326. [Google Scholar] [CrossRef] [PubMed]
  48. Umpierre, A.D.; Wu, L.-J. How microglia sense and regulate neuronal activity. Glia 2021, 69, 1637–1653. [Google Scholar] [CrossRef]
  49. Xiong, X.-Y.; Liu, L.; Yang, Q.-W. Functions and mechanisms of microglia/macrophages in neuroinflammation and neurogenesis after stroke. Prog. Neurobiol. 2016, 142, 23–44. [Google Scholar] [CrossRef]
  50. Wang, J.; He, W.; Zhang, J. A richer and more diverse future for microglia phenotypes. Heliyon 2023, 9, e14713. [Google Scholar] [CrossRef]
  51. Tang, Y.; Le, W. Differential Roles of M1 and M2 Microglia in Neurodegenerative Diseases. Mol. Neurobiol. 2016, 53, 1181–1194. [Google Scholar] [CrossRef] [PubMed]
  52. Ransohoff, R.M. A polarizing question: Do M1 and M2 microglia exist? Nat. Neurosci. 2016, 19, 987–991. [Google Scholar] [CrossRef] [PubMed]
  53. Bachiller, S.; Jiménez-Ferrer, I.; Paulus, A.; Yang, Y.; Swanberg, M.; Deierborg, T.; Boza-Serrano, A. Microglia in Neurological Diseases: A Road Map to Brain-Disease Dependent-Inflammatory Response. Front. Cell. Neurosci. 2018, 12, 488. [Google Scholar] [CrossRef] [PubMed]
  54. Jiang, T.; Zhang, Y.-D.; Chen, Q.; Gao, Q.; Zhu, X.-C.; Zhou, J.-S.; Shi, J.-Q.; Lu, H.; Tan, L.; Yu, J.-T. TREM2 modifies microglial phenotype and provides neuroprotection in P301S tau transgenic mice. Neuropharmacology 2016, 105, 196–206. [Google Scholar] [CrossRef] [PubMed]
  55. Colton, C.A. Heterogeneity of Microglial Activation in the Innate Immune Response in the Brain. J. Neuroimmune Pharmacol. 2009, 4, 399–418. [Google Scholar] [CrossRef] [PubMed]
  56. Gandy, S.; Heppner, F.L. Microglia as Dynamic and Essential Components of the Amyloid Hypothesis. Neuron 2013, 78, 575–577. [Google Scholar] [CrossRef] [PubMed]
  57. Chen, X.; Wang, S.; Cao, W. Mesenchymal stem cell-mediated immunomodulation in cell therapy of neurodegenerative diseases. Cell. Immunol. 2018, 326, 8–14. [Google Scholar] [CrossRef]
  58. Colonna, M.; Butovsky, O. Microglia Function in the Central Nervous System During Health and Neurodegeneration. Annu. Rev. Immunol. 2017, 35, 441–468. [Google Scholar] [CrossRef] [PubMed]
  59. Brown, M.R.; Radford, S.E.; Hewitt, E.W. Modulation of β-Amyloid Fibril Formation in Alzheimer’s Disease by Microglia and Infection. Front. Mol. Neurosci. 2020, 13, 609073. [Google Scholar] [CrossRef]
  60. Weekman, E.M.; Sudduth, T.L.; Abner, E.L.; Popa, G.J.; Mendenhall, M.D.; Brothers, H.M.; Braun, K.; Greenstein, A.; Wilcock, D.M. Transition from an M1 to a mixed neuroinflammatory phenotype increases amyloid deposition in APP/PS1 transgenic mice. J. Neuroinflamm. 2014, 11, 127. [Google Scholar] [CrossRef]
  61. Asai, H.; Ikezu, S.; Tsunoda, S.; Medalla, M.; Luebke, J.; Haydar, T.; Wolozin, B.; Butovsky, O.; Kügler, S.; Ikezu, T. Depletion of microglia and inhibition of exosome synthesis halt tau propagation. Nat. Neurosci. 2015, 18, 1584–1593. [Google Scholar] [CrossRef] [PubMed]
  62. DeVos, S.L.; Miller, R.L.; Schoch, K.M.; Holmes, B.B.; Kebodeaux, C.S.; Wegener, A.J.; Chen, G.; Shen, T.; Tran, H.; Nichols, B.; et al. Tau reduction prevents neuronal loss and reverses pathological tau deposition and seeding in mice with tauopathy. Sci. Transl. Med. 2017, 9, eaag0481. [Google Scholar] [CrossRef] [PubMed]
  63. Takeda, S.; Commins, C.; DeVos, S.L.; Nobuhara, C.K.; Wegmann, S.; Roe, A.D.; Costantino, I.; Fan, Z.; Nicholls, S.B.; Sherman, A.E.; et al. Seed-competent high-molecular-weight tau species accumulates in the cerebrospinal fluid of Alzheimer’s disease mouse model and human patients. Ann. Neurol. 2016, 80, 355–367. [Google Scholar] [CrossRef]
  64. Takeda, S.; Wegmann, S.; Cho, H.; DeVos, S.L.; Commins, C.; Roe, A.D.; Nicholls, S.B.; Carlson, G.A.; Pitstick, R.; Nobuhara, C.K.; et al. Neuronal uptake and propagation of a rare phosphorylated high-molecular-weight tau derived from Alzheimer’s disease brain. Nat. Commun. 2015, 6, 8490. [Google Scholar] [CrossRef] [PubMed]
  65. Hopp, S.C.; Lin, Y.; Oakley, D.; Roe, A.D.; DeVos, S.L.; Hanlon, D.; Hyman, B.T. The role of microglia in processing and spreading of bioactive tau seeds in Alzheimer’s disease. J. Neuroinflamm. 2018, 15, 269. [Google Scholar] [CrossRef] [PubMed]
  66. Pan, X.-D.; Zhu, Y.-G.; Lin, N.; Zhang, J.; Ye, Q.-Y.; Huang, H.-P.; Chen, X.-C. Microglial phagocytosis induced by fibrillar β-amyloid is attenuated by oligomeric β-amyloid: Implications for Alzheimer’s disease. Mol. Neurodegener. 2011, 6, 45. [Google Scholar] [CrossRef] [PubMed]
  67. Lee, D.C.; Rizer, J.; Selenica, M.-L.B.; Reid, P.; Kraft, C.; Johnson, A.; Blair, L.; Gordon, M.N.; Dickey, C.A.; Morgan, D. LPS- induced inflammation exacerbates phospho-tau pathology in rTg4510 mice. J. Neuroinflamm. 2010, 7, 56. [Google Scholar] [CrossRef]
  68. Kitazawa, M.; Oddo, S.; Yamasaki, T.R.; Green, K.N.; LaFerla, F.M. Lipopolysaccharide-induced inflammation exacerbates tau pathology by a cyclin-dependent kinase 5-mediated pathway in a transgenic model of Alzheimer’s disease. J. Neurosci. 2005, 25, 8843–8853. [Google Scholar] [CrossRef] [PubMed]
  69. Jain, S.; Singh, R.; Paliwal, S.; Sharma, S. Targeting Neuroinflammation as Disease Modifying Approach toAlzheimer’s Disease: Potential and Challenges. MRMC 2023, 23, 2097–2116. [Google Scholar] [CrossRef]
  70. Perry, V.H. The influence of systemic inflammation on inflammation in the brain: Implications for chronic neurodegenerative disease. Brain Behav. Immun. 2004, 18, 407–413. [Google Scholar] [CrossRef]
  71. Aktas, O.; Ullrich, O.; Infante-Duarte, C.; Nitsch, R.; Zipp, F. Neuronal damage in brain inflammation. Arch. Neurol. 2007, 64, 185–189. [Google Scholar] [CrossRef] [PubMed]
  72. Koizumi, S.; Shigemoto-Mogami, Y.; Nasu-Tada, K.; Shinozaki, Y.; Ohsawa, K.; Tsuda, M.; Joshi, B.V.; Jacobson, K.A.; Kohsaka, S.; Inoue, K. UDP acting at P2Y6 receptors is a mediator of microglial phagocytosis. Nature 2007, 446, 1091–1095. [Google Scholar] [CrossRef] [PubMed]
  73. Engl, E.; Attwell, D. Non-signalling energy use in the brain. J. Physiol. 2015, 593, 3417–3429. [Google Scholar] [CrossRef] [PubMed]
  74. Kelly, B.; O’Neill, L.A.J. Metabolic reprogramming in macrophages and dendritic cells in innate immunity. Cell Res. 2015, 25, 771–784. [Google Scholar] [CrossRef] [PubMed]
  75. Vander Heiden, M.G.; Cantley, L.C.; Thompson, C.B. Understanding the Warburg effect: The metabolic requirements of cell proliferation. Science 2009, 324, 1029–1033. [Google Scholar] [CrossRef] [PubMed]
  76. Van den Bossche, J.; Baardman, J.; Otto, N.A.; van der Velden, S.; Neele, A.E.; van den Berg, S.M.; Luque-Martin, R.; Chen, H.-J.; Boshuizen, M.C.S.; Ahmed, M.; et al. Mitochondrial Dysfunction Prevents Repolarization of Inflammatory Macrophages. Cell Rep. 2016, 17, 684–696. [Google Scholar] [CrossRef]
  77. Masoumi, M.; Mehrabzadeh, M.; Mahmoudzehi, S.; Mousavi, M.J.; Jamalzehi, S.; Sahebkar, A.; Karami, J. Role of glucose metabolism in aggressive phenotype of fibroblast-like synoviocytes: Latest evidence and therapeutic approaches in rheumatoid arthritis. Int. Immunopharmacol. 2020, 89, 107064. [Google Scholar] [CrossRef] [PubMed]
  78. Wang, Y.; Han, S.; Chen, J.; Sun, J.; Sun, X. PFKFB3 knockdown attenuates Amyloid β-Induced microglial activation and retinal pigment epithelium disorders in mice. Int. Immunopharmacol. 2023, 115, 109691. [Google Scholar] [CrossRef] [PubMed]
  79. Meiser, J.; Krämer, L.; Sapcariu, S.C.; Battello, N.; Ghelfi, J.; D’Herouel, A.F.; Skupin, A.; Hiller, K. Pro-inflammatory Macrophages Sustain Pyruvate Oxidation through Pyruvate Dehydrogenase for the Synthesis of Itaconate and to Enable Cytokine Expression. J. Biol. Chem. 2016, 291, 3932–3946. [Google Scholar] [CrossRef]
  80. Andersson, A.K.; Rönnbäck, L.; Hansson, E. Lactate induces tumour necrosis factor-alpha, interleukin-6 and interleukin-1beta release in microglial- and astroglial-enriched primary cultures. J. Neurochem. 2005, 93, 1327–1333. [Google Scholar] [CrossRef]
  81. Zhang, D.; Tang, Z.; Huang, H.; Zhou, G.; Cui, C.; Weng, Y.; Liu, W.; Kim, S.; Lee, S.; Perez-Neut, M.; et al. Metabolic regulation of gene expression by histone lactylation. Nature 2019, 574, 575–580. [Google Scholar] [CrossRef] [PubMed]
  82. Wei, L.; Yang, X.; Wang, J.; Wang, Z.; Wang, Q.; Ding, Y.; Yu, A. H3K18 lactylation of senescent microglia potentiates brain aging and Alzheimer’s disease through the NFκB signaling pathway. J. Neuroinflamm. 2023, 20, 208. [Google Scholar] [CrossRef] [PubMed]
  83. Baik, S.H.; Kang, S.; Lee, W.; Choi, H.; Chung, S.; Kim, J.I.; Mook-Jung, I. A Breakdown in Metabolic Reprogramming Causes Microglia Dysfunction in Alzheimer’s Disease. Cell Metab. 2019, 30, 493–507. [Google Scholar] [CrossRef] [PubMed]
  84. Li, Y.; Lu, B.; Sheng, L.; Zhu, Z.; Sun, H.; Zhou, Y.; Yang, Y.; Xue, D.; Chen, W.; Tian, X.; et al. Hexokinase 2-dependent hyperglycolysis driving microglial activation contributes to ischemic brain injury. J. Neurochem. 2018, 144, 186–200. [Google Scholar] [CrossRef] [PubMed]
  85. Qiao, H.; He, X.; Zhang, Q.; Yuan, H.; Wang, D.; Li, L.; Hui, Y.; Wu, Z.; Li, W.; Zhang, N. Alpha-synuclein induces microglial migration via PKM2-dependent glycolysis. Int. J. Biol. Macromol. 2019, 129, 601–607. [Google Scholar] [CrossRef] [PubMed]
  86. Pan, R.-Y.; He, L.; Zhang, J.; Liu, X.; Liao, Y.; Gao, J.; Liao, Y.; Yan, Y.; Li, Q.; Zhou, X.; et al. Positive feedback regulation of microglial glucose metabolism by histone H4 lysine 12 lactylation in Alzheimer’s disease. Cell Metab. 2022, 34, 634–648.e6. [Google Scholar] [CrossRef]
  87. Holland, R.; McIntosh, A.L.; Finucane, O.M.; Mela, V.; Rubio-Araiz, A.; Timmons, G.; McCarthy, S.A.; Gun’ko, Y.K.; Lynch, M.A. Inflammatory microglia are glycolytic and iron retentive and typify the microglia in APP/PS1 mice. Brain Behav. Immun. 2018, 68, 183–196. [Google Scholar] [CrossRef] [PubMed]
  88. Cai, Y.; Liu, J.; Wang, B.; Sun, M.; Yang, H. Microglia in the Neuroinflammatory Pathogenesis of Alzheimer’s Disease and Related Therapeutic Targets. Front. Immunol. 2022, 13, 856376. [Google Scholar] [CrossRef] [PubMed]
  89. Lambert, J.C.; Ibrahim-Verbaas, C.A.; Harold, D.; Naj, A.C.; Sims, R.; Bellenguez, C.; DeStafano, A.L.; Bis, J.C.; Beecham, G.W.; Grenier-Boley, B.; et al. Meta-analysis of 74,046 individuals identifies 11 new susceptibility loci for Alzheimer’s disease. Nat. Genet. 2013, 45, 1452–1458. [Google Scholar] [CrossRef]
  90. Piers, T.M.; Cosker, K.; Mallach, A.; Johnson, G.T.; Guerreiro, R.; Hardy, J.; Pocock, J.M. A locked immunometabolic switch underlies TREM2 R47H loss of function in human iPSC-derived microglia. FASEB J. 2020, 34, 2436–2450. [Google Scholar] [CrossRef]
  91. Butovsky, O.; Jedrychowski, M.P.; Moore, C.S.; Cialic, R.; Lanser, A.J.; Gabriely, G.; Koeglsperger, T.; Dake, B.; Wu, P.M.; Doykan, C.E.; et al. Identification of a unique TGF-β-dependent molecular and functional signature in microglia. Nat. Neurosci. 2014, 17, 131–143. [Google Scholar] [CrossRef] [PubMed]
  92. Harms, A.S.; Cao, S.; Rowse, A.L.; Thome, A.D.; Li, X.; Mangieri, L.R.; Cron, R.Q.; Shacka, J.J.; Raman, C.; Standaert, D.G. MHCII is required for α-synuclein-induced activation of microglia, CD4 T cell proliferation, and dopaminergic neurodegeneration. J. Neurosci. Off. J. Soc. Neurosci. 2013, 33, 9592–9600. [Google Scholar] [CrossRef] [PubMed]
  93. Miao, J.; Chen, L.; Pan, X.; Li, L.; Zhao, B.; Lan, J. Microglial Metabolic Reprogramming: Emerging Insights and Therapeutic Strategies in Neurodegenerative Diseases. Cell. Mol. Neurobiol. 2023, 43, 3191–3210. [Google Scholar] [CrossRef] [PubMed]
  94. Heneka, M.T.; Kummer, M.P.; Stutz, A.; Delekate, A.; Schwartz, S.; Vieira-Saecker, A.; Griep, A.; Axt, D.; Remus, A.; Tzeng, T.C.; et al. NLRP3 is activated in Alzheimer’s disease and contributes to pathology in APP/PS1 mice. Nature 2013, 493, 674–678. [Google Scholar] [CrossRef]
  95. Venegas, C.; Kumar, S.; Franklin, B.S.; Dierkes, T.; Brinkschulte, R.; Tejera, D.; Vieira-Saecker, A.; Schwartz, S.; Santarelli, F.; Kummer, M.P.; et al. Microglia-derived ASC specks cross-seed amyloid-β in Alzheimer’s disease. Nature 2017, 552, 355–361. [Google Scholar] [CrossRef] [PubMed]
  96. Mills, E.L.; Kelly, B.; Logan, A.; Costa, A.S.; Varma, M.; Bryant, C.E.; Tourlomousis, P.; Däbritz, J.H.M.; Gottlieb, E.; Latorre, I.; et al. Succinate Dehydrogenase Supports Metabolic Repurposing of Mitochondria to Drive Inflammatory Macrophages. Cell 2016, 167, 457–470. [Google Scholar] [CrossRef] [PubMed]
  97. Gomez Perdiguero, E.; Klapproth, K.; Schulz, C.; Busch, K.; Azzoni, E.; Crozet, L.; Garner, H.; Trouillet, C.; De Bruijn, M.F.; Geissmann, F.; et al. Tissue-resident macrophages originate from yolk-sac-derived erythro-myeloid progenitors. Nature 2015, 518, 547–551. [Google Scholar] [CrossRef]
  98. Nakahira, K.; Haspel, J.A.; Rathinam, V.A.K.; Lee, S.-J.; Dolinay, T.; Lam, H.C.; Englert, J.A.; Rabinovitch, M.; Cernadas, M.; Kim, H.P.; et al. Autophagy proteins regulate innate immune responses by inhibiting the release of mitochondrial DNA mediated by the NALP3 inflammasome. Nat. Immunol. 2011, 12, 222–230. [Google Scholar] [CrossRef]
  99. Cunnane, S.C.; Trushina, E.; Morland, C.; Prigione, A.; Casadesus, G.; Andrews, Z.B.; Beal, M.F.; Bergersen, L.H.; Brinton, R.D.; de la Monte, S.; et al. Brain energy rescue: An emerging therapeutic concept for neurodegenerative disorders of ageing. Nat. Rev. Drug Discov. 2020, 19, 609–633. [Google Scholar] [CrossRef]
  100. Tatulian, S.A. Challenges and hopes for Alzheimer’s disease. Drug Discov. Today 2022, 27, 1027–1043. [Google Scholar] [CrossRef]
  101. Elmaleh, D.R.; Farlow, M.R.; Conti, P.S.; Tompkins, R.G.; Kundakovic, L.; Tanzi, R.E. Developing Effective Alzheimer’s Disease Therapies: Clinical Experience and Future Directions. J. Alzheimer’s Dis. 2019, 71, 715–732. [Google Scholar] [CrossRef] [PubMed]
  102. Avgerinos, K.I.; Ferrucci, L.; Kapogiannis, D. Effects of monoclonal antibodies against amyloid-β on clinical and biomarker outcomes and adverse event risks: A systematic review and meta-analysis of phase III RCTs in Alzheimer’s disease. Ageing Res. Rev. 2021, 68, 101339. [Google Scholar] [CrossRef] [PubMed]
  103. Cummings, J.; Zhou, Y.; Lee, G.; Zhong, K.; Fonseca, J.; Cheng, F. Alzheimer’s disease drug development pipeline: 2023. Alzheimer’s Dement. Transl. Res. Clin. Interv. 2023, 9, e12385. [Google Scholar] [CrossRef] [PubMed]
  104. Ketabforoush, A.H.M.E.; Chegini, R.; Barati, S.; Tahmasebi, F.; Moghisseh, B.; Joghataei, M.T.; Faghihi, F.; Azedi, F. Masitinib: The promising actor in the next season of the Amyotrophic Lateral Sclerosis treatment series. Biomed. Pharmacother. 2023, 160, 114378. [Google Scholar] [CrossRef] [PubMed]
  105. Balzano, T.; Esteban-García, N.; Blesa, J. Neuroinflammation, immune response and α-synuclein pathology: How animal models are helping us to connect dots. Expert. Opin. Drug Discov. 2023, 18, 13–23. [Google Scholar] [CrossRef] [PubMed]
  106. Zhang, J.; Qi, F.; Dong, J.; Tan, Y.; Gao, L.; Liu, F. Application of Baricitinib in Dermatology. J. Inflamm. Res. 2022, 15, 1935–1941. [Google Scholar] [CrossRef] [PubMed]
  107. Zhang, P.; Kishimoto, Y.; Grammatikakis, I.; Gottimukkala, K.; Cutler, R.G.; Zhang, S.; Abdelmohsen, K.; Bohr, V.A.; Misra Sen, J.; Gorospe, M.; et al. Senolytic therapy alleviates Aβ-associated oligodendrocyte progenitor cell senescence and cognitive deficits in an Alzheimer’s disease model. Nat. Neurosci. 2019, 22, 719–728. [Google Scholar] [CrossRef]
  108. Querfurth, H.; Lee, H.-K. Mammalian/mechanistic target of rapamycin (mTOR) complexes in neurodegeneration. Mol. Neurodegener. 2021, 16, 44. [Google Scholar] [CrossRef]
  109. Kopp, K.O.; Greer, M.E.; Glotfelty, E.J.; Hsueh, S.-C.; Tweedie, D.; Kim, D.S.; Reale, M.; Vargesson, N.; Greig, N.H. A New Generation of IMiDs as Treatments for Neuroinflammatory and Neurodegenerative Disorders. Biomolecules 2023, 13, 747. [Google Scholar] [CrossRef]
  110. Agarwal, M.; Khan, S. Plasma Lipids as Biomarkers for Alzheimer’s Disease: A Systematic Review. Cureus 2020, 12, e12008. [Google Scholar] [CrossRef]
  111. Staal, R.G.W.; Weinstein, J.R.; Nattini, M.; Cajina, M.; Chandresana, G.; Möller, T. Senicapoc: Repurposing a Drug to Target Microglia KCa3.1 in Stroke. Neurochem. Res. 2017, 42, 2639–2645. [Google Scholar] [CrossRef] [PubMed]
  112. Devanand, D.P.; Andrews, H.; Kreisl, W.C.; Razlighi, Q.; Gershon, A.; Stern, Y.; Mintz, A.; Wisniewski, T.; Acosta, E.; Pollina, J.; et al. Antiviral therapy: Valacyclovir Treatment of Alzheimer’s Disease (VALAD) Trial: Protocol for a randomised, double-blind, placebo-controlled, treatment trial. BMJ Open 2020, 10, e032112. [Google Scholar] [CrossRef] [PubMed]
  113. Esser, N.; Paquot, N.; Scheen, A.J. Anti-inflammatory agents to treat or prevent type 2 diabetes, metabolic syndrome and cardiovascular disease. Expert. Opin. Investig. Drugs 2015, 24, 283–307. [Google Scholar] [CrossRef] [PubMed]
  114. Ramirez, P.; Zuniga, G.; Sun, W.; Beckmann, A.; Ochoa, E.; DeVos, S.L.; Hyman, B.; Chiu, G.; Roy, E.R.; Cao, W.; et al. Pathogenic tau accelerates aging-associated activation of transposable elements in the mouse central nervous system. Prog. Neurobiol. 2022, 208, 102181. [Google Scholar] [CrossRef] [PubMed]
  115. Porrini, V.; Lanzillotta, A.; Branca, C.; Benarese, M.; Parrella, E.; Lorenzini, L.; Calzà, L.; Flaibani, R.; Spano, P.F.; Imbimbo, B.P.; et al. CHF5074 (CSP-1103) induces microglia alternative activation in plaque-free Tg2576 mice and primary glial cultures exposed to beta-amyloid. Neuroscience 2015, 302, 112–120. [Google Scholar] [CrossRef] [PubMed]
  116. Dhapola, R.; Hota, S.S.; Sarma, P.; Bhattacharyya, A.; Medhi, B.; Reddy, D.H. Recent advances in molecular pathways and therapeutic implications targeting neuroinflammation for Alzheimer’s disease. Inflammopharmacology 2021, 29, 1669–1681. [Google Scholar] [CrossRef] [PubMed]
  117. Chen, Y.; Akirav, E.M.; Chen, W.; Henegariu, O.; Moser, B.; Desai, D.; Shen, J.M.; Webster, J.C.; Andrews, R.C.; Mjalli, A.M.; et al. RAGE ligation affects T cell activation and controls T cell differentiation. J. Immunol. 2008, 181, 4272–4278. [Google Scholar] [CrossRef] [PubMed]
  118. Cui, W.; Sun, C.; Ma, Y.; Wang, S.; Wang, X.; Zhang, Y. Inhibition of TLR4 Induces M2 Microglial Polarization and Provides Neuroprotection via the NLRP3 Inflammasome in Alzheimer’s Disease. Front. Neurosci. 2020, 14, 444. [Google Scholar] [CrossRef]
  119. Shi, S.; Liang, D.; Chen, Y.; Xie, Y.; Wang, Y.; Wang, L.; Wang, Z.; Qiao, Z. Gx-50 reduces β-amyloid-induced TNF-α, IL-1β, NO, and PGE2 expression and inhibits NF-κB signaling in a mouse model of Alzheimer’s disease. Eur. J. Immunol. 2016, 46, 665–676. [Google Scholar] [CrossRef]
  120. Wang, X.; Zhang, Y.; Peng, Y.; Hutchinson, M.R.; Rice, K.C.; Yin, H.; Watkins, L.R. Pharmacological characterization of the opioid inactive isomers (+)-naltrexone and (+)-naloxone as antagonists of toll-like receptor 4. Br. J. Pharmacol. 2016, 173, 856–869. [Google Scholar] [CrossRef]
  121. Scheltens, P.; Prins, N.; Lammertsma, A.; Yaqub, M.; Gouw, A.; Wink, A.M.; Chu, H.-M.; van Berckel, B.N.M.; Alam, J. An exploratory clinical study of p38α kinase inhibition in Alzheimer’s disease. Ann. Clin. Transl. Neurol. 2018, 5, 464–473. [Google Scholar] [CrossRef] [PubMed]
  122. Son, S.H.; Lee, N.-R.; Gee, M.S.; Song, C.W.; Lee, S.J.; Lee, S.-K.; Lee, Y.; Kim, H.J.; Lee, J.K.; Inn, K.-S.; et al. Chemical Knockdown of Phosphorylated p38 Mitogen-Activated Protein Kinase (MAPK) as a Novel Approach for the Treatment of Alzheimer’s Disease. ACS Cent. Sci. 2023, 9, 417–426. [Google Scholar] [CrossRef] [PubMed]
  123. Xu, J.; Yuan, C.; Wang, G.; Luo, J.; Ma, H.; Xu, L.; Mu, Y.; Li, Y.; Seeram, N.P.; Huang, X.; et al. Urolithins Attenuate LPS-Induced Neuroinflammation in BV2Microglia via MAPK, Akt, and NF-κB Signaling Pathways. J. Agric. Food Chem. 2018, 66, 571–580. [Google Scholar] [CrossRef] [PubMed]
  124. Wang, L.; Zhao, H.; Wang, L.; Tao, Y.; Du, G.; Guan, W.; Liu, J.; Brennan, C.; Ho, C.-T.; Li, S. Effects of Selected Resveratrol Analogues on Activation and Polarization of Lipopolysaccharide-Stimulated BV-2 Microglial Cells. J. Agric. Food Chem. 2020, 68, 3750–3757. [Google Scholar] [CrossRef] [PubMed]
  125. Carey, A.N.; Fisher, D.R.; Rimando, A.M.; Gomes, S.M.; Bielinski, D.F.; Shukitt-Hale, B. Stilbenes and anthocyanins reduce stress signaling in BV-2 mouse microglia. J. Agric. Food Chem. 2013, 61, 5979–5986. [Google Scholar] [CrossRef] [PubMed]
  126. Joseph, J.A.; Fisher, D.R.; Cheng, V.; Rimando, A.M.; Shukitt-Hale, B. Cellular and behavioral effects of stilbene resveratrol analogues: Implications for reducing the deleterious effects of aging. J. Agric. Food Chem. 2008, 56, 10544–10551. [Google Scholar] [CrossRef] [PubMed]
  127. Zhang, C.; Griciuc, A.; Hudry, E.; Wan, Y.; Quinti, L.; Ward, J.; Forte, A.M.; Shen, X.; Ran, C.; Elmaleh, D.R.; et al. Cromolyn Reduces Levels of the Alzheimer’s Disease-Associated Amyloid β-Protein by Promoting Microglial Phagocytosis. Sci. Rep. 2018, 8, 1144. [Google Scholar] [CrossRef] [PubMed]
  128. Jin, S.; Lin, C.; Wang, Y.; Wang, H.; Wen, X.; Xiao, P.; Li, X.; Peng, Y.; Sun, J.; Lu, Y.; et al. Cannabidiol Analogue CIAC001 for the Treatment of Morphine-Induced Addiction by Targeting PKM2. J. Med. Chem. 2023, 66, 11498–11516. [Google Scholar] [CrossRef]
  129. Zhou, Y.; Huang, Z.; Su, J.; Li, J.; Zhao, S.; Wu, L.; Zhang, J.; He, Y.; Zhang, G.; Tao, J.; et al. Benserazide is a novel inhibitor targeting PKM2 for melanoma treatment. Int. J. Cancer 2020, 147, 139–151. [Google Scholar] [CrossRef]
  130. Pan, R.-Y.; Ma, J.; Kong, X.-X.; Wang, X.-F.; Li, S.-S.; Qi, X.-L.; Yan, Y.-H.; Cheng, J.; Liu, Q.; Jin, W.; et al. Sodium rutin ameliorates Alzheimer’s disease-like pathology by enhancing microglial amyloid-β clearance. Sci. Adv. 2019, 5, eaau6328. [Google Scholar] [CrossRef]
  131. Wang, H.; Fu, J.; Xu, X.; Yang, Z.; Zhang, T. Rapamycin Activates Mitophagy and Alleviates Cognitive and Synaptic Plasticity Deficits in a Mouse Model of Alzheimer’s Disease. J. Gerontol. Ser. A 2021, 76, 1707–1713. [Google Scholar] [CrossRef] [PubMed]
  132. Zhang, N.; Nao, J.; Dong, X. Neuroprotective Mechanisms of Salidroside in Alzheimer’s Disease: A Systematic Review and Meta-analysis of Preclinical Studies. J. Agric. Food Chem. 2023, 71, 17597–17614. [Google Scholar] [CrossRef] [PubMed]
  133. Syed, Y.Y. Sodium Oligomannate: First Approval. Drugs 2020, 80, 441–444. [Google Scholar] [CrossRef] [PubMed]
  134. Cai, Q.; Li, Y.; Pei, G. Polysaccharides from Ganoderma lucidum attenuate microglia-mediated neuroinflammation and modulate microglial phagocytosis and behavioural response. J. Neuroinflamm. 2017, 14, 63. [Google Scholar] [CrossRef]
  135. Zheng, Y.; Ren, W.; Zhang, L.; Zhang, Y.; Liu, D.; Liu, Y. A Review of the Pharmacological Action of Astragalus Polysaccharide. Front. Pharmacol. 2020, 11, 349. [Google Scholar] [CrossRef] [PubMed]
  136. Guo, L.; Qi, J.; Du, D.; Liu, Y.; Jiang, X. Current advances of Dendrobium officinale polysaccharides in dermatology: A literature review. Pharm. Biol. 2020, 58, 664–673. [Google Scholar] [CrossRef]
  137. Xia, G.; Li, X.; Zhang, Z.; Jiang, Y. Effect of food processing on the antioxidant activity of flavones from Polygonatum odoratum (Mill.) Druce. Open Life Sci. 2021, 16, 92–101. [Google Scholar] [CrossRef]
  138. Javed, M.A.; Ashraf, N.; Saeed Jan, M.; Mahnashi, M.H.; Alqahtani, Y.S.; Alyami, B.A.; Alqarni, A.O.; Asiri, Y.I.; Ikram, M.; Sadiq, A.; et al. Structural Modification, In Vitro, In Vivo, Ex Vivo, and In Silico Exploration of Pyrimidine and Pyrrolidine Cores for Targeting Enzymes Associated with Neuroinflammation and Cholinergic Deficit in Alzheimer’s Disease. ACS Chem. Neurosci. 2021, 12, 4123–4143. [Google Scholar] [CrossRef] [PubMed]
  139. Li, J.; Zhuang, L.; Luo, X.; Liang, J.; Sun, E.; He, Y. Protection of MCC950 against Alzheimer’s disease via inhibiting neuronal pyroptosis in SAMP8 mice. Exp. Brain Res. 2020, 238, 2603–2614. [Google Scholar] [CrossRef]
  140. Yin, J.; Zhao, F.; Chojnacki, J.E.; Fulp, J.; Klein, W.L.; Zhang, S.; Zhu, X. NLRP3 Inflammasome Inhibitor Ameliorates Amyloid Pathology in a Mouse Model of Alzheimer’s Disease. Mol. Neurobiol. 2018, 55, 1977–1987. [Google Scholar] [CrossRef]
  141. Jiang, H.; He, H.; Chen, Y.; Huang, W.; Cheng, J.; Ye, J.; Wang, A.; Tao, J.; Wang, C.; Liu, Q.; et al. Identification of a selective and direct NLRP3 inhibitor to treat inflammatory disorders. J. Exp. Med. 2017, 214, 3219–3238. [Google Scholar] [CrossRef]
  142. Coll, R.C.; Schroder, K.; Pelegrín, P. NLRP3 and pyroptosis blockers for treating inflammatory diseases. Trends Pharmacol. Sci. 2022, 43, 653–668. [Google Scholar] [CrossRef] [PubMed]
  143. Luo, J.E.; Li, Y.-M. Turning the tide on Alzheimer’s disease: Modulation of γ-secretase. Cell Biosci. 2022, 12, 2. [Google Scholar] [CrossRef]
  144. Choi, S.-H.; Aid, S.; Caracciolo, L.; Minami, S.S.; Niikura, T.; Matsuoka, Y.; Turner, R.S.; Mattson, M.P.; Bosetti, F. Cyclooxygenase-1 inhibition reduces amyloid pathology and improves memory deficits in a mouse model of Alzheimer’s disease. J. Neurochem. 2013, 124, 59–68. [Google Scholar] [CrossRef]
  145. Ajmone-Cat, M.A.; Bernardo, A.; Greco, A.; Minghetti, L. Non-Steroidal Anti-Inflammatory Drugs and Brain Inflammation: Effects on Microglial Functions. Pharmaceuticals 2010, 3, 1949–1965. [Google Scholar] [CrossRef]
  146. Burstein, A.H.; Sabbagh, M.; Andrews, R.; Valcarce, C.; Dunn, I.; Altstiel, L. Development of Azeliragon, an Oral Small Molecule Antagonist of the Receptor for Advanced Glycation Endproducts, for the Potential Slowing of Loss of Cognition in Mild Alzheimer’s Disease. J. Prev. Alzheimer’s Dis. 2018, 5, 149–154. [Google Scholar] [CrossRef] [PubMed]
  147. Xue, J.; Jia, P.; Zhang, D.; Yao, Z. TTP488 ameliorates NLRP3-associated inflammation, viability, apoptosis, and ROS production in an Alzheimer’s disease cell model by mediating the JAK1/STAT3/NFκB/IRF3 pathway. Cell Biochem. Funct. 2021, 39, 555–561. [Google Scholar] [CrossRef] [PubMed]
  148. Breitner, J.C.S.; Haneuse, S.J.P.A.; Walker, R.; Dublin, S.; Crane, P.K.; Gray, S.L.; Larson, E.B. Risk of dementia and AD with prior exposure to NSAIDs in an elderly community-based cohort. Neurology 2009, 72, 1899–1905. [Google Scholar] [CrossRef]
  149. McGeer, P.L.; McGeer, E.G. NSAIDs and Alzheimer disease: Epidemiological, animal model and clinical studies. Neurobiol. Aging 2007, 28, 639–647. [Google Scholar] [CrossRef]
  150. Wang, Z.; Vilekar, P.; Huang, J.; Weaver, D.F. Furosemide as a Probe Molecule for the Treatment of Neuroinflammation in Alzheimer’s Disease. ACS Chem. Neurosci. 2020, 11, 4152–4168. [Google Scholar] [CrossRef]
  151. Neibert, K.; Gosein, V.; Sharma, A.; Khan, M.; Whitehead, M.A.; Maysinger, D.; Kakkar, A. “Click” dendrimers as anti-inflammatory agents: With insights into their binding from molecular modeling studies. Mol. Pharm. 2013, 10, 2502–2508. [Google Scholar] [CrossRef] [PubMed]
  152. García Bueno, B.; Caso, J.R.; Madrigal, J.L.M.; Leza, J.C. Innate immune receptor Toll-like receptor 4 signalling in neuropsychiatric diseases. Neurosci. Biobehav. Rev. 2016, 64, 134–147. [Google Scholar] [CrossRef] [PubMed]
  153. Go, M.; Kou, J.; Lim, J.-E.; Yang, J.; Fukuchi, K.-I. Microglial response to LPS increases in wild-type mice during aging but diminishes in an Alzheimer’s mouse model: Implication of TLR4 signaling in disease progression. Biochem. Biophys. Res. Commun. 2016, 479, 331–337. [Google Scholar] [CrossRef] [PubMed]
  154. Song, M.; Jin, J.; Lim, J.E.; Kou, J.; Pattanayak, A.; Rehman, J.A.; Kim, H.D.; Tahara, K.; Lalonde, R.; Fukuchi, K.I. TLR4 mutation reduces microglial activation, increases Aβ deposits and exacerbates cognitive deficits in a mouse model of Alzheimer’s disease. J. Neuroinflamm. 2011, 8, 92. [Google Scholar] [CrossRef] [PubMed]
  155. Zhong, Q.; Zou, Y.; Liu, H.; Chen, T.; Zheng, F.; Huang, Y.; Chen, C.; Zhang, Z. Toll-like receptor 4 deficiency ameliorates β2-microglobulin induced age-related cognition decline due to neuroinflammation in mice. Mol. Brain 2020, 13, 20. [Google Scholar] [CrossRef] [PubMed]
  156. Leitner, G.R.; Wenzel, T.J.; Marshall, N.; Gates, E.J.; Klegeris, A. Targeting toll-like receptor 4 to modulate neuroinflammation in central nervous system disorders. Expert. Opin. Ther. Targets 2019, 23, 865–882. [Google Scholar] [CrossRef] [PubMed]
  157. Irving, E.A.; Bamford, M. Role of mitogen- and stress-activated kinases in ischemic injury. J. Cereb. Blood Flow Metab. 2002, 22, 631–647. [Google Scholar] [CrossRef]
  158. Dong, N.; Li, X.; Xue, C.; Zhang, L.; Wang, C.; Xu, X.; Shan, A. Astragalus polysaccharides alleviates LPS-induced inflammation via the NF-κB/MAPK signaling pathway. J. Cell. Physiol. 2020, 235, 5525–5540. [Google Scholar] [CrossRef] [PubMed]
  159. Prins, N.D.; Harrison, J.E.; Chu, H.-M.; Blackburn, K.; Alam, J.J.; Scheltens, P.; REVERSE-SD Study Investigators. A phase 2 double-blind placebo-controlled 24-week treatment clinical study of the p38 alpha kinase inhibitor neflamapimod in mild Alzheimer’s disease. Alzheimer’s Res. Ther. 2021, 13, 106. [Google Scholar] [CrossRef]
  160. Johnson, E.C.B.; Dammer, E.B.; Duong, D.M.; Ping, L.; Zhou, M.; Yin, L.; Higginbotham, L.A.; Guajardo, A.; White, B.; Troncoso, J.C.; et al. Large-scale proteomic analysis of Alzheimer’s disease brain and cerebrospinal fluid reveals early changes in energy metabolism associated with microglia and astrocyte activation. Nat. Med. 2020, 26, 769–780. [Google Scholar] [CrossRef]
  161. Carosi, J.M.; Sargeant, T.J. Rapamycin and Alzheimer disease: A hypothesis for the effective use of rapamycin for treatment of neurodegenerative disease. Autophagy 2023, 19, 2386–2390. [Google Scholar] [CrossRef] [PubMed]
  162. Wang, H.; Li, Q.; Sun, S.; Chen, S. Neuroprotective Effects of Salidroside in a Mouse Model of Alzheimer’s Disease. Cell. Mol. Neurobiol. 2020, 40, 1133–1142. [Google Scholar] [CrossRef] [PubMed]
  163. Di Paolo, M.; Papi, L.; Gori, F.; Turillazzi, E. Natural Products in Neurodegenerative Diseases: A Great Promise but an Ethical Challenge. Int. J. Mol. Sci. 2019, 20, 5170. [Google Scholar] [CrossRef]
  164. Cvetkovic, R.S.; Keating, G. Anakinra. BioDrugs 2002, 16, 303–311; discussion 313–314. [Google Scholar] [CrossRef] [PubMed]
  165. Nair, B.; Raval, G.; Mehta, P. TNF-alpha inhibitor etanercept and hematologic malignancies: Report of a case and review of the literature. Am. J. Hematol. 2007, 82, 1022–1024. [Google Scholar] [CrossRef]
  166. Ardura-Fabregat, A.; Boddeke, E.W.G.M.; Boza-Serrano, A.; Brioschi, S.; Castro-Gomez, S.; Ceyzériat, K.; Dansokho, C.; Dierkes, T.; Gelders, G.; Heneka, M.T.; et al. Targeting Neuroinflammation to Treat Alzheimer’s Disease. CNS Drugs 2017, 31, 1057–1082. [Google Scholar] [CrossRef] [PubMed]
  167. Hampel, H.; Mesulam, M.M.; Cuello, A.C.; Farlow, M.R.; Giacobini, E.; Grossberg, G.T.; Khachaturian, A.S.; Vergallo, A.; Cavedo, E.; Snyder, P.J.; et al. The cholinergic system in the pathophysiology and treatment of Alzheimer’s disease. Brain J. Neurol. 2018, 141, 1917–1933. [Google Scholar] [CrossRef] [PubMed]
  168. Marucci, G.; Buccioni, M.; Ben, D.D.; Lambertucci, C.; Volpini, R.; Amenta, F. Efficacy of acetylcholinesterase inhibitors in Alzheimer’s disease. Neuropharmacology 2021, 190, 108352. [Google Scholar] [CrossRef]
  169. Akıncıoğlu, H.; Gülçin, İ. Potent Acetylcholinesterase Inhibitors: Potential Drugs for Alzheimer’s Disease. Mini Rev. Med. Chem. 2020, 20, 703–715. [Google Scholar] [CrossRef]
  170. Bai, H.; Zhang, Q. Activation of NLRP3 Inflammasome and Onset of Alzheimer’s Disease. Front. Immunol. 2021, 12, 701282. [Google Scholar] [CrossRef]
  171. Zhang, W.; Xiao, D.; Mao, Q.; Xia, H. Role of neuroinflammation in neurodegeneration development. Signal Transduct. Target. Ther. 2023, 8, 267. [Google Scholar] [CrossRef] [PubMed]
  172. Coll, R.C.; Robertson, A.A.B.; Chae, J.J.; Higgins, S.C.; Muñoz-Planillo, R.; Inserra, M.C.; Vetter, I.; Dungan, L.S.; Monks, B.G.; Stutz, A.; et al. A small-molecule inhibitor of the NLRP3 inflammasome for the treatment of inflammatory diseases. Nat. Med. 2015, 21, 248–255. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Energy metabolic disorder and inflammation of microglia in AD in response to therapy. When exposed to the AD environment, microglia polarizing to pro-inflammatory ‘M1’ phenotypes release many inflammatory cytokines. Microglia activate TLRs, which, in turn activate the TREM2/AKT/mTOR signaling pathway, leading to the upregulation of HIF-1α, GLUT, and PKM2. This results in reduced OXPHOS, leading to the production of reactive oxygen species (ROS) and inflammatory factors. Furthermore, the TREM2/AKT/mTOR and TLR4/NF-κB signaling pathways directly activate NLRP3, promoting inflammatory gene expression and further enhancing glycolysis. The drugs indicated in yellow text boxes act through target genes (in red) to induce anti-inflammatory responses and regulate microglial polarization and energy metabolic disorder, ultimately aiming to alleviate or treat AD.
Figure 1. Energy metabolic disorder and inflammation of microglia in AD in response to therapy. When exposed to the AD environment, microglia polarizing to pro-inflammatory ‘M1’ phenotypes release many inflammatory cytokines. Microglia activate TLRs, which, in turn activate the TREM2/AKT/mTOR signaling pathway, leading to the upregulation of HIF-1α, GLUT, and PKM2. This results in reduced OXPHOS, leading to the production of reactive oxygen species (ROS) and inflammatory factors. Furthermore, the TREM2/AKT/mTOR and TLR4/NF-κB signaling pathways directly activate NLRP3, promoting inflammatory gene expression and further enhancing glycolysis. The drugs indicated in yellow text boxes act through target genes (in red) to induce anti-inflammatory responses and regulate microglial polarization and energy metabolic disorder, ultimately aiming to alleviate or treat AD.
Molecules 29 01478 g001
Table 1. AD anti-inflammation drug in clinical development (https://classic.clinicaltrials.gov/ (access on 6 March 2024)).
Table 1. AD anti-inflammation drug in clinical development (https://classic.clinicaltrials.gov/ (access on 6 March 2024)).
AgentTypeCADRO TargetMechanism of ActionClinical Trial NCT#
Masitinib
(Phase 3)
small moleculeInflammationTyrosine kinase inhibitor exhibits neuroprotection via inhibition of mast cell and microglia/macrophage activityNCT05564169
NE3107
(Phase 3)
small moleculeInflammationBeta-androstenetriol with anti-inflammatory and insulin signaling effects via ERK 1 and 2NCT04669028
Baricitinib
(Phase 2)
small moleculeInflammationJanus kinase (JAK) inhibitorNCT05189106
Dasatinib
+
Quercetin
(Phase 2)
small moleculeInflammationDasatinib induces apoptosis in senescent cells to allow their removal; quercetin is a flavonoidNCT04063124
NCT04685590
NCT04785300
NCT05422885
L-Serine
(Phase 2)
small moleculeInflammationNaturally occurring dietary amino acid; inhibits toxic misfoldingNCT03062449
Lenalidomide
(Phase 2)
small moleculeInflammationImmunomodulatorNCT04032626
Montelukast
(Phase 2)
small moleculeInflammationLeukotriene receptor antagonist (LTRA); anti-inflammatory effectsNCT03402503
Senicapoc
(Phase 2)
small moleculeInflammationCalcium-activated potassium channel inhibitorNCT04804241
Valacyclovir
(Phase 2)
small moleculeInflammationAnti-viral against HSV-1 and −2; reduces vira-related “seeding” of ABP depositionNCT03282916
Salsalate
(Phase 1)
small moleculeInflammationNon-steroidal anti-inflammatory (NSAID)NCT03277573
Emtricitabine
(Phase 1)
small moleculeInflammationNucleoside reverse transcriptase inhibitor (NRTI)NCT04500847
Table 2. AD anti-inflammation drug mechanism (https://classic.clinicaltrials.gov/ (access on 6 March 2024)).
Table 2. AD anti-inflammation drug mechanism (https://classic.clinicaltrials.gov/ (access on 6 March 2024)).
Drug NameChemical StructuresCategoryCompound Effects
MasitinibMolecules 29 01478 i001Tyrosine kinase inhibitorDownregulated proinflammatory cytokines.
Induced neuroprotection [104].
NE3107Molecules 29 01478 i002NF-κB inhibitor Decreased activated microglia, Aβ [105].
BaricitinibMolecules 29 01478 i003JAK inhibitorBlocked intracellular delivery of cytokines via JAK-STAT [106].
Dasatinib
+
Quercetin
Molecules 29 01478 i004Tyrosine kinase inhibitorAlleviated neurodegeneration in AD [107].
Molecules 29 01478 i005PI3K/Akt inhibitor
L-SerineMolecules 29 01478 i006mTOR inhibitorAutophagic clearance of Aβ [108].
LenalidomideMolecules 29 01478 i007ImmunomodulatorDecreased the expression of TNFα, IL-6, IL-8.
Increased the expression of anti-inflammatory cytokines [109].
MontelukastMolecules 29 01478 i008Cytochrome P-450 Enzyme InducersIncreased expression of P450 enzymes [110].
SenicapocMolecules 29 01478 i009KCa3.1 inhibitorRegulated microglia polarization [111].
ValacyclovirMolecules 29 01478 i010Anti-virusReduced accumulation of Aβ and p-tau [112].
SalsalateMolecules 29 01478 i011NASIDInhibited inflammatory mediators [113].
EmtricitabineMolecules 29 01478 i012Anti-virusSuppressed neuroinflammation [114].
Table 3. Promising drugs for treating AD by inhibiting inflammation and regulating microglia polarization and energy metabolic disorder.
Table 3. Promising drugs for treating AD by inhibiting inflammation and regulating microglia polarization and energy metabolic disorder.
Drug NameChemical StructuresCategoryCompound Effects
CHF5074Molecules 29 01478 i013 Improved cognition. Reduced brain inflammation [115].
IndomethacinMolecules 29 01478 i014Cox-2
inhibitor
Inhibited inflammatory mediators released from microglia [116].
FurosemideMolecules 29 01478 i015 Inhibited protein expression of COX-2,iNOS.
AzeliragonMolecules 29 01478 i016Mediated the JAK1/STAT3/NF-κB/IRF3 pathway [117].
TAK242Molecules 29 01478 i017TLR4
antagonists
Reduced Aβ deposition [118].
Gx-50Molecules 29 01478 i018Suppressed of TLR4-mediated NF-κB/MAPK signaling [119].
NaloxoneMolecules 29 01478 i019Inhibited reactive oxygen species production [120].
VX745Molecules 29 01478 i020p38-MAPK
antagonists
Reduced IL-1β protein levels in the hippocampus [121].
PRZ-18002Molecules 29 01478 i021Degraded phosphorylated p38 MAPK (p-p38) [122].
Alleviated microglial activation and Aβ deposition [122].
UrolithinMolecules 29 01478 i022Reduced NO levels and suppressed pro-inflammatory genes [123].
TSGMolecules 29 01478 i023Microglia
regulators
Inhibited PKM2 to adjust microglia polarization [124].
ResveratrolMolecules 29 01478 i024Adjusted microglia polarization [124,125,126].
CromolynMolecules 29 01478 i025Induced neuroprotective microglial activation [127].
CIAC001Molecules 29 01478 i026PKM2
inhibitor
Ameliorated morphine-induced addiction through anti-neuroinflammation [128].
BenserazideMolecules 29 01478 i027Inhibited PKM2, thereby blocking aerobic glycolysis and modulating OXPHOS [129].
RutinMolecules 29 01478 i028Promoted a metabolic switch from anaerobic glycolysis to mitochondrial OXPHOS [130].
RapamycinMolecules 29 01478 i029mTOR
inhibitor
Activated mitophagy and alleviated cognitive impairment [131].
SalidrosideMolecules 29 01478 i030Inhibited Aβ deposit.
Anti-inflammation [132].
GV-971Molecules 29 01478 i031Chinese herbal polysaccharides Remodeled gut microbiota to inhibit AD progression [133].
Ganoderma lucidum polysaccharides/Reduced Aβ levels and tau protein hyperphosphorylation [134].
Astragalus polysaccharides/Reduced astrocytic and microglial activation [135].
Dendrobium orchid polysaccharide/Inhibited oxidative stress [136].
Astragalus polysaccharide/Decreased the phosphorylation of p38 MAPK [137].
TacrineMolecules 29 01478 i032AChE
inhibitor
Inhibited acetylcholinesterase activity [138].
The derivative of tacrineMolecules 29 01478 i033
DonepezilMolecules 29 01478 i034
DapansutrileMolecules 29 01478 i035NLRP3
Inhibitor
Inhibited the associated NLRP3 inflammatory response [120].
MC950Molecules 29 01478 i036Reduced Aβ deposition and associated neurotoxicity [139].
JC124Molecules 29 01478 i037Reduced CAA, microgliosis and oxidative stress [140].
CY-09Molecules 29 01478 i038Inhibited the associated NLRP3 inflammatory response [141].
InzomelidMolecules 29 01478 i039Inhibited the associated NLRP3 inflammatory response [142].
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Chen, H.; Zeng, Y.; Wang, D.; Li, Y.; Xing, J.; Zeng, Y.; Liu, Z.; Zhou, X.; Fan, H. Neuroinflammation of Microglial Regulation in Alzheimer’s Disease: Therapeutic Approaches. Molecules 2024, 29, 1478. https://doi.org/10.3390/molecules29071478

AMA Style

Chen H, Zeng Y, Wang D, Li Y, Xing J, Zeng Y, Liu Z, Zhou X, Fan H. Neuroinflammation of Microglial Regulation in Alzheimer’s Disease: Therapeutic Approaches. Molecules. 2024; 29(7):1478. https://doi.org/10.3390/molecules29071478

Chicago/Turabian Style

Chen, Haiyun, Yuhan Zeng, Dan Wang, Yichen Li, Jieyu Xing, Yuejia Zeng, Zheng Liu, Xinhua Zhou, and Hui Fan. 2024. "Neuroinflammation of Microglial Regulation in Alzheimer’s Disease: Therapeutic Approaches" Molecules 29, no. 7: 1478. https://doi.org/10.3390/molecules29071478

Article Metrics

Back to TopTop