Next Article in Journal
Elevated Production of Mitochondrial Reactive Oxygen Species via Hyperthermia Enhanced Cytotoxic Effect of Doxorubicin in Human Breast Cancer Cell Lines MDA-MB-453 and MCF-7
Next Article in Special Issue
Mitochondrial Protein Network: From Biogenesis to Bioenergetics in Health and Disease
Previous Article in Journal
Three-Dimensional Cell Culture Models to Investigate Oral Carcinogenesis: A Scoping Review
Previous Article in Special Issue
Skeletal Phenotypes Due to Abnormalities in Mitochondrial Protein Homeostasis and Import
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Molecular Dysfunctions of Mitochondria-Associated Membranes (MAMs) in Alzheimer’s Disease

by
Fanny Eysert
1,†,
Paula Fernanda Kinoshita
1,2,†,
Arnaud Mary
1,
Loan Vaillant-Beuchot
1,
Frédéric Checler
1 and
Mounia Chami
1,*
1
Laboratory of Excellence DistAlz, INSERM, CNRS, Institut of Molecular and Cellular Pharmacology, Université Côte d’Azur, Sophia-Antipolis, 06560 Valbonne, France
2
Instituto de Ciências Biomédicas, Department of Pharmacology, Universidade de São Paulo, São Paulo 05508-900, Brazil
*
Author to whom correspondence should be addressed.
Share first author position.
Int. J. Mol. Sci. 2020, 21(24), 9521; https://doi.org/10.3390/ijms21249521
Submission received: 10 November 2020 / Revised: 4 December 2020 / Accepted: 8 December 2020 / Published: 14 December 2020

Abstract

:
Alzheimer’s disease (AD) is a multifactorial neurodegenerative pathology characterized by a progressive decline of cognitive functions. Alteration of various signaling cascades affecting distinct subcellular compartment functions and their communication likely contribute to AD progression. Among others, the alteration of the physical association between the endoplasmic reticulum (ER) and mitochondria, also referred as mitochondria-associated membranes (MAMs), impacts various cellular housekeeping functions such as phospholipids-, glucose-, cholesterol-, and fatty-acid-metabolism, as well as calcium signaling, which are all altered in AD. Our review describes the physical and functional proteome crosstalk between the ER and mitochondria and highlights the contribution of distinct molecular components of MAMs to mitochondrial and ER dysfunctions in AD progression. We also discuss potential strategies targeting MAMs to improve mitochondria and ER functions in AD.

1. Introduction

Alzheimer’s disease (AD) is a neurodegenerative disease characterized by two major histological hallmarks: (1) the neurofibrillary tangles (NFTs) corresponding to intracellular aggregates of abnormally hyperphosphorylated Tau protein (pTau) and, (2) senile plaques that are mainly composed of extracellular aggregates of β amyloid peptide (Aβ) derived from the sequential cleavage of its precursor, the amyloid precursor protein (APP), by β-secretase and γ-secretase enzymes [1,2]. Late-onset AD (LOAD) cases are considered to be sporadic (SAD) accounting for the majority of AD cases. For a long time, apolipoprotein E (APOE) has been described as the sole and main genetic risk factor for LOAD forms [3]. Recently, a genome wide-association study (GWAS) identified several new AD genetic risk factors [4,5,6]. Dominantly inherited familial AD (FAD) accounts for less than 1% of the cases and can be caused by mutations in APP, presenilin 1 (PS1), or presenilin 2 (PS2) genes [7,8]. Overall, AD is now considered a complex pathology, the mechanistic defects of which remain unclear. The amyloid cascade is the most widely accepted hypothesis in AD and proposes Aβ accumulation as the etiological trigger of the pathology. This has been supported by genetic evidence demonstrating that FAD mutations in genes coding for APP or its processing enzymes (PS1 and PS2) constituting the catalytic components of the γ-secretase enzymatic complex [9] lead to increased production and secretion of Aβ peptides [9]. The latter accumulate in the form of neurotoxic Aβ oligomers (Aβo) and are thought to trigger several stress responses in neurons leading to the onset of neurofibrillary degeneration [9]. Several therapeutic approaches were developed with the aim to reduce the accumulation of Aβ through active and passive immunizations. However, the failure of these trials to rescue cognitive declines or even stabilize them [10] casts some doubts about the amyloid cascade hypothesis. Notably, the sole contribution of Aβ to AD pathogenesis is challenged, since APP processing yields several fragments besides Aβ [2]. In physiological conditions, 90% of mature APP is cleaved by α-secretase at the plasma membrane, producing the secreted soluble α-APP (sAPPα) and the membrane-anchored APP-C-terminal (APP-CTF)α fragment (C83). The latter is further cleaved by γ-secretase producing p3 and AICD (APP intracellular domain) peptides. On the other hand, 10% of mature APP is cleaved by β-secretase (BACE1) following its internalization towards the endosomal/lysosomal pathway, producing the soluble β-APP (sAPPβ) released in the extracellular environment, and the membrane-anchored APP-CTFβ fragment (C99), which is further cleaved by γ-secretase producing AICD and Aβ peptides [2]. C99 is also cleaved by α-secretase to produce C83 [11]. Furthermore, other APP-derived fragments produced by other enzymes (such as η- and δ- secretases) were recently described [12,13,14], likely contributing to AD pathophysiology [12,13,14,15,16,17,18].
Both the endoplasmic reticulum (ER) and mitochondria are vital organelles of the cell. Rough and smooth ER participate in protein synthesis, folding, and transport. The ER is also involved in other fundamental cellular functions such as lipids- and carbohydrates- metabolism and is the major calcium (Ca²+) storage pool [19]. On the other hand, mitochondria act as the powerhouse of the cell by generating energy through adenosine triphosphate (ATP) production [20]. Besides, mitochondria play a major role in buffering the Ca²+ flux from the ER, but also in lipid- and amino acid- metabolism, beneficial and harmful reactive oxygen species (ROS) production, and apoptosis [20]. It is now well established that ER and mitochondria functions are highly interconnected. They physically interact to form specific microdomains called mitochondria-associated ER membranes (MAMs), where the outer mitochondrial membrane (OMM) is close to the ER in the order of 10–100 nanometers [21,22,23]. The maintenance of stable contact sites between the ER and mitochondria provides a platform for bidirectional crosstalk. Not surprisingly, MAMs control intracellular elementary events such as Ca²+ homeostasis, metabolic flow, protein transport [24], mitochondrial lipids production, phospholipid biosynthesis, mitochondrial fusion and fission, and global events such as autophagy and apoptosis [25] (Figure 1). In addition, MAMs are a hot-spot for the transfer of stress signals from the ER to mitochondria, particularly under ER stress conditions activating the unfolded protein response (UPR) [26] (Figure 1). As a result, one can argue that the alteration of ER and mitochondria communication may influence and disrupt these functions, thus leading to the development of several pathologies and vice-versa.
ER stress and mitochondrial dysfunctions occur early in AD, likely contributing to disease progression and irreversible neuronal death [27,28,29,30]. Numerous papers reported that several pathogenic paradigms associated with AD are closely linked to MAMs structure and function alterations. This review aims at delineating the contribution of several structural and functional partners of MAMs (i.e., Mitofusins (MFNs), Protein Kinase-like Endoplasmic Reticulum Kinase (PERK), the serine/threonine-protein kinase/endoribonuclease inositol-requiring enzyme 1α (IRE1α), and the sarco/endoplasmic reticulum Ca2+ ATPase 1 (SERCA) truncated isoform (S1T)), and MAMs complexes (i.e., IP3R-Grp75-VDAC, PACS2-PSS1, BAP31-FIS1-DRP1, VAPB-PTPIP51, and Sig-1R-Grp78) (Figure 2) in pathophysiological processes of AD. We also elucidate how AD associated proteins (APP and its derived fragments, PSs, Tau, and APOE) may impact MAMs structure and function (Figure 3). Finally, we highlight the potential relevance of MAMs molecular components as therapeutic targets to prevent the development of AD.

2. Structural and Functional Partners of MAMs Are Involved in AD

2.1. Mitofusins: MFNs

Mitofusins (MFN1 and MFN2) belongs to GTPases family and are located in the OMM. Both MFNs are involved in mitochondrial fusion process by forming homotypic and heterotypic interactions with the inner mitochondria membrane’s (IMM) OPA1 protein [31,32]. Besides, several studies proposed that MFN2 modulates mitochondrial membrane potential (ΔΨm) through the regulation of nuclear-encoded subunits of OXPHOS complexes I, II, III, and V [33]. MFN2 acts as a regulator of apoptosis in a cell type- and age-specific manner [34,35], and is linked to mitophagy [36]. The importance of MFN2 in organelles juxtaposition and the communication function of MAMs has been also proven through the regulation of mitochondria Ca2+ uptake [37]. MFN2 was also shown to be involved in dendritic outgrowth, spine formation in mice models, and to prevent neurodegeneration in the cerebellum through fusion mechanism [38].
Several studies reported that MFN2 is associated with AD [39] (Figure 2 and Table 1). MFN2 gene polymorphisms were reported in LOAD [40]. Moreover, the mitochondrial fusion and fission balance is altered in AD, correlating with enhanced or reduced MFNs expression [41]. A reduction of both MFN1 and MFN2 expression was reported in human AD brains [41,42]. Accordingly, a reduction of MFN2 expression was also reported in N2a cells expressing the APP familial Swedish mutation (APPswe) that overproduces Aβo [43]. On the contrary, an overexpression of MFN2 was observed in human cybrid cells in which mitochondria from mild cognitive impairment (MCI) subjects were incorporated into neuronal cells depleted of endogenous mitochondrial DNA [44]. These discrapencies may point out a complex regulation of MFN2 in AD pathogenesis. In fact, directly linked to APP processing in AD, MFN2 was shown to control γ-secretase activity and Aβ production by modulating ER and mitochondrial membrane contacts, without affecting BACE1 and neprilysin expression [45]. Moreover, PS2-linked FAD mutants promote ER-mitochondria coupling in a MFN2-dependent manner (Table 1) [46]. Altogether, these studies highlight a central role of MFN2 in AD pathogenesis through the regulation of MAMs tethering, mitochondrial structure and function, and cell death but also APP processing.
Figure 2. Schema showing the localization and alterations of MAMs molecular components in AD, including upregulation (↑) or down-regulation (↓) of MAMs proteins expression, MAMs proteins interactions (⟷), and genes polymorphism.
Figure 2. Schema showing the localization and alterations of MAMs molecular components in AD, including upregulation (↑) or down-regulation (↓) of MAMs proteins expression, MAMs proteins interactions (⟷), and genes polymorphism.
Ijms 21 09521 g002

2.2. IP3R-Grp75-VDAC

The inositol 1,4,5-trisphosphate receptors (IP3Rs), releasing Ca2+ from the ER to the cytosol [47], are enriched in ER-mitochondria contact sites playing a key role in cellular differentiation, survival, and apoptosis [48,49]. Ca2+ uptake by mitochondria occurs through VDAC, an OMM protein which is a permeable Ca2+ channel, together with the mitochondrial Ca2+ uniporter (MCU) complex [50] located on the IMM (Figure 1). In MAMs, a functional complex between VDAC (isoform 1) and IP3R (isoform 3) is formed through a chaperone called glucose-regulated protein 75 (Grp75, a member of the heat shock protein 70 family) [51]. The IP3R-Grp75-VDAC complex regulates Ca²+ transfer from the ER to mitochondria, suggesting its involvement in Ca²+ homeostasis deregulation largely described in AD [51]. While there is only one study documenting a reduction of Grp75 expression in the temporal and parietal cortex of AD post-mortem brains [52], several studies demonstrated alterations of VDAC expression in AD (Figure 2 and Table 1). VDAC1 expression increases progressively according to AD stages in the cortex of post-mortem brains, in aged APP mice, and also in neuroblastoma cells treated with Aβo [53]. Accordingly, another study reported that Aβ treatment of primary hippocampal neurons increased VDAC1 and IP3R expression, resulting in enhanced contacts between ER and mitochondria [54]. VDAC1 also interacts with Aβ and pTau, likely contributing to mitochondrial dysfunction [55]. In turn, Aβ triggers VDAC dephosphorylation in lipid rafts in AD brains correlating with cell death [56]. Inversely, reduced expression of VDAC1 decreased the mRNA levels of APP, Tau, PS1 and PS2, and also BACE1 [57].
Altearations of IP3Rs expressions and activity have been largely explored in AD and were demonstrated to be linked to both Aβ and PS1 and PS2 FAD mutations. Fibroblasts from FAD patients exhibited an enhanced response to bombesin- and bradykinin-mediated IP3 generation in comparison to control cells [58]. These data were further confirmed in cortical and hippocampal neurons from 3xTgAD mice [59]. Aβo are also involved in IP3R overactivation by increasing IP3 production and triggering cytotoxicity [60]. Importantly, the reduction of IP3R1-mediated Ca2+ release in the cortical and hippocampal neurons of PS1M146V and 3xTgAD mice reduces ryanodine receptor (RyR)-mediated Ca2+ release, restores cAMP response element-binding (CREB)-dependent gene expression, rescues long-term potentiation (LTP) to normal levels, attenuates Aβ accumulation and pTau and reverses memory deficits [61]. Another study also showed a physical interaction of IP3R with PS1 (M146L) and PS2 (N141I) mutants stimulating IP3R gating activity [62] (Figure 2 and Table 1). Of note, our laboratory and others reported a major role of enhanced RyR expression and activity in AD [63,64,65,66,67,68]. Nevertheless, dedicated studies are still needed to demonstrate if the interaction between IP3R and FAD PSs mutants occurs preferentially in MAMs and to unravel the physiopathological contribution of RyRs in MAMs dysfunction related to AD.

2.3. PACS2-PSS1

Phosphofurin acidic cluster sorting protein 2 (PACS2) is a multifunctional sorting protein that interacts with several cargo proteins and regulates their location in MAMs [69]. PACS2 controls apoptosis induction, through the translocation of Bim (Bcl-2-like protein 11) to lysosomes [70], or Bid (BH3 interacting domain) to mitochondria [71]. The precise localization of PACS2 in MAMs is still uncertain. However, it has been demonstrated that PACS2 controls ER-mitochondrial apposition through its tethering action and is also involved in Ca2+ transfer from ER to mitochondria through IP3R localization [71]. Accordingly, PACS2 deletion leads to decreased ER-mitochondria contacts and triggers mitochondria fragmentation [71] and apoptosis [72].
PSS1 (phosphatidylserine synthase-1) is synthesized and specifically located in MAMs, supporting the direct transfer of lipids between ER and mitochondria [73]. PACS2 and PSS1 are functionally linked, since the overexpression of PACS2 raises the levels of PSS1 in MAMs [71].
PACS2 and PSS1 regulations in AD were reported in a study by Hedskog et al. showing enhanced PACS2 and PSS1 protein levels in AD transgenic mice expressing Swe/LDN FAD mutations [54]. More precisely, PACS2 protein expression was enhanced in the hippocampus of 2 month-old aged mice before Aβ plaque formation, and in the cortex and the cerebellum at 6 months. Meanwhile, PSS1 protein expression was only enhanced in the cerebellum starting at 6 months. Importantly, the authors also revealed enhanced PSS1 and PACS2 protein levels in human SAD brains (Figure 2 and Table 1). Enhanced PSS1 and PACS2 protein expressions were not corroborated by an increase in PSS1 and PACS2 mRNAs levels, suggesting that the regulation of PSS1 and PACS2 expressions in AD occurs at the protein level only. The authors also reported unchanged phospholipid metabolism in both mice and AD brains, thus questioning the functional consequences related to enhanced PSS1 and PACS-2 protein expression. Nevertheless, ablation of PSS2 triggers degeneration of astrocytes and neuronal hippocampal primary cultures. In conclusion, this unique study unravels the potential contribution of an additional MAMs molecular component in AD [54].

2.4. BAP31-FIS1-DRP1

Another physical tether of MAMs has been described as the interaction between the resident integral ER protein B cell receptor–associated protein 31 (BAP31) and the OMM mitochondrial FIS1 protein involved in the initiation of mitochondrial fission process [74,75]. FIS1 interacts with dynamin-related protein1 (DRP1) after its recruitment at mitochondrial fission sites by the OMM receptors (Mff, Mid49 and Mid51) in the MAMs complex, including BAP31 [76]. Intriguingly, a recent study reported that the human FIS1 protein can cause mitochondrial fragmentation in the absence of DRP1 and Dyn2 GTPases through its ability to bind to MFN1, MFN2, and OPA1, also inhibiting the mitochondrial fusion [77]. Shen et al. also demonstrated that mutations in the FIS1 gene do not cause defects in mitochondrial fission in response to fission-inductor chemical treatments but trigger persisting large aggregates of autophagosomes [76]. Accordingly, FIS1 KO MEFs cells display a strongly impaired mitophagy [78]. Iwasawa et al. showed that FIS1 conveys an apoptosis signal from the mitochondria to the ER through its interaction with BAP31 at the ER and its cleavage into the pro-apoptotic p20 [74]. The latter triggers a Ca2+ transfer from the ER to the mitochondria leading to ΔΨm depolarization, hence priming cell death induction [74,75]. BAP31 is also essential for mitochondrial homeostasis as it stimulates the formation of the mitochondrial complex I [79].
Several studies reported that FIS1, BAP31, and DRP1 may contribute individually to AD physiopathology (Figure 2 and Table 1). It was first reported that the mRNA and protein levels of FIS1 and DRP1 are increased in the frontal cortex of patients with early, definite, and severe AD [42]. Joshi et al., reported later that Aβ42-treated neurons and fibroblasts from SAD or FAD patients feature an increased interaction between DRP1 and FIS1, which causes mitochondrial dysfunctions attested by increased mitochondrial fragmentation, oxidative stress, decreased ΔΨm and ATP levels, and increases in the activity of pro-apoptotic enzymes and Cyt c release [80]. Most importantly, inhibiting the DRP1-FIS1 interaction corrects these mitochondrial impairments, reduces in the 5xFAD mice AD model (that overexpress human APP and PS1 transgenes with a total of five AD-linked mutations) cognitive defects, Aβ40 and Aβ42 levels, and oxidative stress, and increases ATP levels [80]. In parallel, BAP31 knockdown in the hippocampus and the cerebral cortex of PS1 M146V mice results in an increased BACE1 protein level, exacerbating C99 accumulation and Aβ plaque formation [81]. Other studies showed that pTau and Aβ monomers and oligomers can also directly interact with Drp1 in AD late stages, likely contributing to abnormal mitochondrial dynamics [42,82]. Moreover, Aβ also triggers nitric oxide production, likely contributing to Drp1 S-nitrosylation-mediated mitochondrial fragmentation and neuronal damage [83] (Figure 2 and Table 1). Together, these studies support the involvement of the BAP31-FIS1-DRP1 complex in AD pathology.

2.5. VAPB-PTPIP51

The OMM protein, tyrosine phosphatase-interacting protein-51 (PTPIP51), was identified as a binding partner for the resident ER protein vesicle-associated membrane protein-associated protein-B (VAPB) [84]. As a physical tether of MAMs, VAPB has several functions, including vesicle trafficking and UPR [84]. It was then demonstrated that the loss of either VAPB or PTPIP51 perturbs the uptake of Ca2+ by mitochondria, decreasing ATP production [85] and stimulating the formation of autophagosomes [86]. Additionally, PTPIP51 can interact with the oxysterol-binding protein (OSBP)-related proteins ORP5 and ORP8 in MAMs [87], likely contributing to phosphatidylserine (PS) transport between ER and mitochondria.
VAPB P56S mutation is associated with dominantly inherited familial forms of type-8 amyotrophic lateral sclerosis (ALS) [88]. Moreover, TAR DNA binding protein 43 (TDP-43) and fused in sarcoma (FUS), two ALS/frontotemporal dementia (FTD)-associated proteins, were reported to activate the glycogen synthase kinase-3β (GSK-3β), which inhibits VAPB-PTPIP51 binding, ultimately resulting in disrupted Ca2+ homeostasis and reduced ATP production [85,89]. Strikingly, it has recently been observed that AD patients with granulovacuolar degeneration bodies (GVB) feature lower VAPB level [90] (Figure 2 and Table 1). Interestingly, AD is a pathology where TDP-43 and FUS accumulate in stress granules and GVB. Furthermore, GSK-3β is also overactivated in AD [91]. A recent study reports a reduction of VAPB and PTPIP51 interaction in the pyramidal cortex of patients with early/mid dementia, and a reduction of the expression of both VAPB and PTPIP51 in the cortex of AD patients at late-stages [92]. Together, these studies also suggest the involvement of VAPB-PTPIP51 physical/or functional alterations in AD pathophysiology.

2.6. Sig-1R-Grp78

Sigma non-opioid intracellular 1-receptor 1 (Sig-1R) is a chaperone protein present in ER lipid rafts that interacts and modulates different proteins in ER and plasma membrane [93]. Sig-1R is essential for MAMs stability [94], and is implicated in lipid synthesis and trafficking [95].
Grp78 (or BiP), a member of the heat shock protein 70 family, is a chaperone and an ER stress sensor [96]. Importantly, Grp78 is present in MAMs, where it folds the steroidogenic acute regulatory protein (StAR) transporting cholesterol for delivery to the OMM [97]. Under ER stress conditions, Grp78 dissociates from and activates three branches of the UPR: activating transcription factor 6 (ATF6), IRE1α, and PERK. The UPR generates an adaptive or a pro-apoptotic response according to the stress duration and/or intensity [98]. In aging, the pro-apoptotic response becomes more common than the adaptive pathway, probably due to a failure in chaperone systems [98]. Interestingly, the impairment in UPR, apoptosis, and accumulation of misfolded proteins are common features of AD and other neurodegenerative diseases [99]. Desipte this, there is still no consensus on the role of Grp78 in AD. While APP/PS1 mice and Aβ25-35 treated neurons presented increased Grp78 expression (reviewed in Reference [100]), another study in AD post-mortem brains with PS1 mutations showed a reduction in Grp78 mRNA levels [101]. Recently, a study demonstrated that the 5xFAD mice had no changes in Grp78 expression and other ER stress-related proteins [102]. These discrepancies could be linked to the pleiotropic functions regulated by and regulating Grp78 expression (Figure 2 and Table 1).
Sig-1R forms a complex with Grp78 in MAMs under resting conditions [103]. During ER stress, Sig-1R also dissociates from Grp78 and forms a complex with IP3R through ankyrin B to increase Ca2+ signaling from ER to the mitochondria, thereby enhancing ATP synthesis [104]. Besides, Sig-1R stimulates phospholipase C (PLC) that increases IP3 levels, resulting also in Ca2+ release from the ER [105]. The activation of Sig-1R is considered neuroprotective since Sig-1R silencing in hippocampal neurons results in Cyt c release, caspase-3 activation, and a deficit in spine formation. Remarkably, this mechanism is reversed by superoxide dismutase activation, showing that the reduction of Sig-1R increases the levels of ROS [106].
Interestingly, Sig-1R has been associated with a variety of neurodegenerative diseases, including AD [107,108]. Reduced Sig-1R binding sites were reported in AD post-mortem brains [109]. It has also been proposed that a polymorphism in Sig-1R (TT-P) correlates with the risk of developing AD [110]. This finding has been challenged in another study demonstrating that another Sig-1R polymorphism (Q2P) may worsen clinical outcomes only in association with the gentic risk factor APOEε(epsilon)4 [111]. The protective effect of Sig-1R has also been observed in two AD mice models (3xTg-AD and MacGill-R-Thy1-APP), where the treatment with a Sig-1R agonist, named AF710B (a concomitant agonist of the muscarinic 1 receptor), have been shown to restore cognitive impairment, and to reduce amyloid pathology and neuroinflammation [112,113]. Accordingly, the silencing of Sig-1R or application of the Sig-1R antagonist (NE-100) aggravated the pathological status in Aβ25-35 treated mice (i.e., learning deficits, reduced BDNF levels, and enhanced expression of Bax) [114]. The treatment of Aβ25-35 injected mice with ANAVEX2-73, another agonist of Sig-1R, decreases pTau, Bax/Bcl-2 ratio, lipid peroxidation, C99 accumulation, and Cyt c release [115,116], while decreasing Sig-1R expression enhances Tau hyperphosphorylation and reduces dendritic spines formation [117].
Overall, these studies converge to demonstrate the neuroprotective effect of Sig-1R agonists, acting through a diversity of Sig-1R-mediated signaling functions that are generally pro-survival and anti-apoptotic.

2.7. Protein Kinase-Like Endoplasmic Reticulum Kinase: PERK

Several studies reported that the activation of the PERK pathway attenuates protein synthesis until the unfolded protein accumulation is removed [98]. Interestingly, PERK is enriched at MAMs [118], where it facilitates the tethering of the ER to mitochondria and sensitizes cells to apoptosis [118,119]. PERK KO cells show reduced MAMs and disturbed cytosolic Ca2+ signaling and also protect cells from ROS-induced mitochondrial dysfunction [118]. A recent study demonstrates that PERK localized in MAMs phosphorylates mitochondrial E3 ubiquitin ligases and decreases the formation of mitochondria-associated ER membranes under ER stress condition [120]. UPR occurs relatively early in AD, and PERK was shown to be activated in human AD-derived brains [121,122]. The dysregulated PERK pathway has been recapitulated in AD mice models that develop β-amyloidosis [123,124,125], Tau-mediated neurodegeneration [126,127], and in SAD genetic risk factor APOEε4 mice models [128] (Figure 2 and Table 1). Several studies have provided mechanistic insights into the direct relationship between PERK overactivation and several paradigms occurring in AD. The phosphorylation of eIF2α, occurring down-stream of PERK activation, has been shown to increase BACE1 expression, thereby enhancing the production of Aβ in neurons [129], and genetic reduction of PERK reduces BACE1 protein level and Aβ production in the 5xFAD mouse model [123]. Accordingly, the selective ablation of PERK improved the synaptic plasticity and spatial memory in mice harboring APP and PS1 mutations [124], consistent with the requirement for active protein translation in memory consolidation [130]. Another study demonstrated that a reduction in eIF2α phosphorylation enhances the late phase of LTP and memory in mice [131]. Furthermore, the local expression of ATF4 (a transcription factor activated downstream of PERK and eIF2α ER stress responsive pathway) in axons induces axonal damage through a cell-non-autonomous mechanism that propagates between neurons [132]. These data suggest that sustained PERK-eIF2α-ATF4 activation worsens AD pathogenesis, and that PERK localization in MAMs may act as a new pathogenic route contributing to AD development.
Table 1. MAMs molecular component alterations in AD. These alterations include reduced (↓), increased (↑), unchanged expressions (=), or interaction (⟷) between proteins observed in AD study models and human-derived samples (SAD brains and SAD/ FAD fibroblasts); KO: knock-out; KI: knock-in; “OE”: overexpressing; MCI: mild cognitive impairment; GVB: granulovacuolar degeneration bodies; Aβo: Aβ oligomers.
Table 1. MAMs molecular component alterations in AD. These alterations include reduced (↓), increased (↑), unchanged expressions (=), or interaction (⟷) between proteins observed in AD study models and human-derived samples (SAD brains and SAD/ FAD fibroblasts); KO: knock-out; KI: knock-in; “OE”: overexpressing; MCI: mild cognitive impairment; GVB: granulovacuolar degeneration bodies; Aβo: Aβ oligomers.
ProteinsAlterationsStudy ModelsReferences
MFNsPolymorphism AD patients[40]
↓MFN2 & MFN1 (proteins and mRNA)AD brains (hippocampus & frontal cortex) & N2a “OE” APPswe[41,42,43]
↑MFN2 MCI cytoplasmic hybrid (cybrid) cells [44]
↑MFN2⟷PS2 (FAD), ↑MAMsMFN2 KO MEFs
FAD PS2 (N141I) Mice
[46]
IP3R-Grp75-VDAC↑VDAC
Aβ⟷VDAC
pTau⟷VDAC
Tg2576 and J20 mice
APP, APP/PS1, 3xTgAD mice & AD brains (cortical tissues)
[53]
[55]
↑VDAC, ↑IP3R activity, ↑MAMsAβ-treated hippocampal neurons[54]
VDAC dephosphorylationAD brains[56]
↓Grp75AD brains (temporal and parietal cortex)[52]
FAD PS1/2⟷IP3RSf9 cells[62]
PACS2-PSS1↑PACS2,
↑PSS1
APPswe/LDN mice
AD brains cortex
[54]
BAP31-FIS1-DRP1↑Fis1, ↑Drp1 AD brains (frontal cortex) [42]
↑Drp1⟷Fis42-treated neurons, SAD/FAD fibroblasts, N2a “OE” APPswe, & 5xTgAD mice[80]
pTau⟷Drp1
↓Drp1 mito localization
AD brains, APP, APP/PS1 & 3xTgAD mice
Drosophila “OE” Tau R406W
[55]
[133]
Aβ⟷Drp1AD brains (frontal cortex)[42]
VAPB-PTPIP51↓VAPB, ↓PTPIP51
↓VAPB⟷PTPIP51
AD brains (subicular with GVB neurons)
AD brains (cortex)
[90]
[92]
Sig-1R-Grp78↓Sig-1R
Sig-1R polymorphism
AD brains
AD patients
[109]
[110,111]
↑Grp78 Streptozotocin model, APP/PS1 mice & Aβ25-35-treated rats [134,135,136]
↓Grp78 mRNAAD brains (FAD PS1)[101]
=Grp78 expression5xFAD mice[102]
PERK↑PERK pathwayAD brains
5xFAD mice
APP/PS1, APP(SL)/PS1 KI & rTg4510 mice
APOE4 mice
[121,122,123,124,125,126]
[123]
[124,125,126,127]
[128,129]
IRE1α↑↓IRE1α pathway, ↑↓XBP1sAD brains & 5xFAD mice
42o-treated SH-SY5Y & CHO “OE” APPLDN
[137]
[138]
S1T↑S1TAD brains, SHSY-5Y “OE”APPswe or treated with Aβo[139]

2.8. Active Inositol-Requiring Transmembrane Kinase and Endonuclease Alpha: IRE1α

The UPR IRE1α branch increases protein folding trough the induction of chaperones and enhances the removal of misfolded proteins [98]. Active IRE1α processes the mRNA encoding X-box-binding protein 1 spliced isoform (XBP1s), a transcription factor that upregulates genes encoding mediators of protein folding, ERAD (ER-associated degradation), organelle biogenesis, and protein quality control [98]. A fraction of IRE1α is located at MAMs through Sig-1R stabilization responding to mitochondrial ROS [103,140]. A recent study demonstrated that the mitochondrial ubiquitin ligase (MITOL/MARCH5) inhibits ER stress-induced apoptosis through ubiquitylation and degradation of IRE1α at MAMs [141]. IRE1α at MAMs can also induce cell death via mitochondrial Ca2+ overload [142]. Another study linked IRE1α to mitochondrial dysfunction and apoptosis through the phosphorylation of ASK1 and JNK [143,144].
The activation of the IRE1α branch is implicated in AD pathogenesis, mainly through the activation of XBP1s [121,122] (Figure 2 and Table 1). Enhanced or reduced XBP1 mRNA splicing was reported in AD human brains [145]. XBP1s activation occurs upon Aβo treatment lowering BACE1 protein level [138] and activating Kalirin-7 (Kal7), a protein that controls synaptic plasticity [146]. This suggest that XBP1s acts as a protective mechanism in AD. Accordingly, the overexpression of XBP1s protects against Aβ toxicity [147] and pathological Tau [148]. In addition, a genomic screen identified a cluster of AD genes as possible direct targets of XBP1s, including APP, and components of the α-secretase, and proteins involved in APP trafficking and maturation [149]. Finally, in both Drosophila and mammalian cell culture models of AD, XBP1 overexpression also down-regulates RyR3 expression, which in turn prevents cytosolic Ca2+ overload [147]. These last data corroborate studies demonstrating that pharmacological and genetic approaches reducing RyR-mediated exacerbated ER Ca2+ leak provide protective effects and rescue several AD pathogenic paradigms in cellular and mice AD models [64,65,66].

2.9. The Truncated Variant of the Sarco-Endoplasmic Reticulum Ca2+-ATPase 1: S1T

S1T is among molecular partners at the ER-mitochondria interface regulating inter-organelle Ca2+ signaling and cell death [150,151,152]. We reported the induction of S1T expression under pathological ER stress conditions occurring through the PERK-eIF2α-ATF4-CHOP (C/EBP homologous protein) pathway. On the one hand, S1T induction amplifies ER stress response, and on the other hand, it enhances ER Ca2+ leak, increases the number of ER-mitochondria contact sites, and inhibits mitochondrial movements. This leads to increased Ca2+ transfer to mitochondria, thus activating the mitochondrial apoptotic pathway [152]. Interestingly, we recently reported a molecular interplay between S1T-dependent ER Ca2+ leak, UPR, and APP processing, likely contributing to AD pathogenesis [139] (Figure 2 and Table 1). We demonstrated that S1T expression is increased in SAD brains and correlates with Aβ and ER stress chaperone protein levels (Grp78 and Calreticulin). Increased S1T expression is induced in an Aβ-dependent manner and enhances in return the production of APP-CTFs and Aβ through specific increases of BACE1 expression and activity. Interestingly, the induction of S1T expression is also linked to neuroinflammation [139]. Other studies are ongoing to unravel the contribution of S1T-mediated ER Ca2+ leak to MAMs dysfunctions, synaptic plasticity alterations, and cognitive deficits in AD.

3. The MAMs Hypothesis in the Pathophysiological Process of Alzheimer’s Disease

In this chapter, we focus on the role of main AD players in mitochondrial and ER dysfunctions and their link with MAMs alterations in AD.

3.1. Apolipoprotein E: APOE

APOE is a glycoprotein mostly expressed in astrocytes, microglia, and neurons under stress conditions, and in the choroid plexus in the CNS [153,154]. APOE plays a relevant role in lipid transport, in particular cholesterol and cholesterol esters. This occurs via APOE binding to a low-density protein (LDL) receptor, such as LDLR and LDLR-related protein 1 (LDLR-1) [155]. APOE is essential for brain functions since APOE KO mice present synaptic loss and cognition dysfunction. The rescue of APOE in the periphery can improve learning and memory impairments, but without affecting synaptic loss [156]. APOE is a highly polymorphic gene encoding for three major APOE epsilon (ε) isoforms (APOEε2, APOEε3, and APOEε4). As stated above, APOE is a genetic risk factor for LOAD forms [3,4]. The elevated risk of AD correlates with the number of APOEε4 alleles [157]. While APOEε2 confers a decreased risk of AD in comparison with the APOEε3 allele, a single APOε4 allele increases the risk of AD four-fold compared to the more common APOEε3/APOEε3 genotype [157]. The presence of two APOEε4/APOEε4 alleles increases the risk of AD by approximately twelve-fold [158]. APOE polymorphism causes alterations in APOE structure and function that results in a different binding to lipids [157]. Notably, there is a relationship between APOE polymorphism and enhanced Aβ pathology [159].
Mitochondrial dynamics are differently impacted by APOE isoforms. APOEε4 knock-in (KI) mice exhibited higher expression of MFN1 and lower expression of DRP1 in the hippocampus in comparison to APOEε3-expressing mice. The decrease in DRP1 was further confirmed in AD patients carrying two APOEε4 alleles [160]. The expression of COX1 and translocase of outer mitochondrial membrane 40 homolog (TOMM40) were also higher in APOEε4-expressing mice, which supports the idea of increased mitochondrial fusion [160]. The APOEε4 isoform is also associated with altered mitophagy, characterized by reduced expression of cleaved PINK1 and higher levels of Parkin and SQSTM1/p62 in the hippocampus of APOEε4-expressing mice. This latter presented elongated mitochondria with a smaller cristae density compared to APOEε3-expressing mice hippocampus, demonstrating the contribution of APOEε4 to mitochondrial structure alteration [160]. In addition, a proteomic study showed that N2a cells expressing APOEε4 had impaired respiration and increased glycolysis in basal conditions and presented less reserve capacity to generate ATP, decreased NAD+/NADH ratio, and increased ROS and the mitochondrial Ca2+ load under a stress condition in comparison to APOEε3-expressing cells [161]. Importantly, APOEε4-expressing cells also manifest increased levels of VDAC1, IP3R, BAP31, FIS1, and MFN2 and a decrease in Grp75 expression, showing that APOEε4 modulates several molecular components of MAMs [161]. As shown in Figure 1, one main function of MAMs is phospholipid synthesis, specially phosphatidylserine (PS), which is transported from the ER to mitochondria, where it is decarboxylated and transformed in phosphatidylethanolamine (PE) that is transported to the ER [162]. Human fibroblasts and mice hippocampal neurons subjected to conditioned media (CM) obtained from astrocytes isolated from APOEε3 and APOEε4 KI mice showed an increase in PS and PE that was larger with APOEε4 CM than with APOEε3 CM [163]. However, this difference was abolished in MFN2 KO MEFs, suggesting that APOEε4-mediated phospholipids synthesis is strictly dependent on the increased interaction between ER and mitochondria [163]. The same study also reported that APOEε4 CM enhances cholesterol oleate level compared to APOEε3 CM and that this effect is abolished in MFN2 KO MEFs. Accordingly, fibroblasts treated with APOEε4 CM presented increased cholesteryl ester synthesis and lipid droplets as compared to APOEε3 CM treated fibroblasts [163]. This study was the first to unravel a relevant role of the APOEε4 allele in controlling MAMs activity, likely contributing to LOAD development.

3.2. Presenilins

Both PS1 and PS2 are localized in the ER, Golgi, and golgian network, the plasma membrane, endosomes, lysosomes, nuclear membrane, and mitochondria [164]. Being catalytic subunits of the γ-secretase complexe, PSs are well known to be involved in APP metabolism, leading to the production of different fragments and defining the length of Aβ peptide. Currently, more than 320 mutations of PS1 and 63 mutations of PS2 have been shown to be involved in AD or other dementias (Alzforum: https://www.alzforum.org/mutations). Many of these mutations are responsible for the increased Aβ42/Aβ40 ratio. Of note, Aβ42 peptides are more toxic and prone to aggregation and accumulation in plaques than Aβ40 peptides. In addition, PSs participate in several functions closely linked to MAMs. First, it was described that PSs participate in autophagy by favoring the fusion of autophagosomes with lysosomes [164]. Neely et al. reported enhanced autophagosomes and LC3-II levels in fibroblasts isolated from PS1 or PS2 KO mice arguing for a defective lysosomal degradation of the autophagosome [165]. Moreover, Száraz and colleagues suggested that the alteration in autophagy involves Ca2+ homeostasis dysregulation associated with ER stress response, particularly to Grp78 enrichment in MAMs [166]. Interestingly, PSs are also involved in ER Ca2+ regulation [167], forming a passive ER Ca²+ leak channel [168], directly interacting with SERCA2b [169], or directly or indirectly impacting IP3R- and RyR-mediated ER Ca2+ release [68,170]. These data support the idea that FAD PSs dysregulate both ER Ca²+ store filling and emptying.
More recently, PS1 and PS2 have been reported to be enriched in MAMs [171], with PS2 being directly involved in the interaction between ER and mitochondria [46,172]. Filadi and colleagues showed that PS2 can directly interact with and requires MFN2 expression to modulate the coupling between ER and mitochondria [46]. The authors also showed that two familial mutants of PS2 (T122R and N141I), known to be involved in Ca2+ release during atypical dementia and Aβ42 production during AD, are more efficient in ER-mitochondria tethering than wild type (WT) PS2, increasing mitochondrial Ca2+ uptake upon ER Ca2+ release [46,173]. Furthermore, SH-SY5Y neuroblastoma cells expressing these PS2 mutants and FAD-PS2 patient-derived fibroblasts showed a blockade in the late phase of autophagy [174], as demonstrated by an increase of LC3-II and SQSTM1/p62 levels, and alteration of the RAB7-dependent autophagosome-lysosome fusion step as a consequence of the alteration of the Ca2+ content of intracellular stores and cytosolic Ca2+ signaling [174]. These studies demonstrate that the reinforcement of MAMs linked to FAD-PS2 mutants contributes to several mitochondrial dysfunctions reported in AD, in particular the disruption of subcellular Ca2+ handling and the regulation of autophagy [46,173,174].
Although FAD PS1 mutants do not seem to be directly involved in the ER-mitochondria tethering, several data show that PS1 FAD mutations alter several functions linked to MAMs. Firstly, several FAD PS1 mutants (M146V, L166P, A246E, E273A, G384A, and P436Q) were shown to exert a dominant-negative effect on ER Ca²+ leak and/ or to exacerbate IP3R- and RyR-mediated ER Ca2+ release and alter neuronal Ca²+ signaling [67,68,168,170,175]. Moreover, the expression of PS1-E280A FAD mutation in Purkinje cells and SH-SY5Y cells was shown to impair Ca²+ homeostasis associated with reduced levels of CACNA1A, IP3R1, and IP3R3, leading to decreased levels of Ca²+-dependent mitochondrial transport proteins (i.e., MIRO1 and KIF5C), and enhanced tethering of ER and mitochondria [176]. Finally, PS1-FAD fibroblasts (A246E, M233T, H163Y, M146L and L392V) were also described to impact autophagy by impairing autolysosome acidification and cathepsin D activation that are caused by the failure of PS1 to target the v-ATPase V0a1 subunit to lysosomes [177]. Together, these results suggest that PS1 is likely directly or indirectly involved in MAMs structure and functions.
To conclude, these data support the hypothesis that different pathways regulated by PSs converge towards MAMs-linked Ca²+ homeostasis, autophagy, and mitochondria function, the alterations of which represent common functional defects in AD.

3.3. APP-Derived C-Terminal Fragments (APP-CTFs) and Aβ

APP participates neuronal migration, neurites outgrowth, synaptogenesis, synapses stabilization, and synaptic plasticity [178]. As stated in the introduction, APP is synthesized in the ER and transits through the Golgi apparatus, where it undergoes several post-translational modifications such as N- and O-glycosylations, sialylations, phosphorylations, and sulfations before arriving at the plasma membrane in its mature form [178]. Several studies suggest that APP is also targeted to mitochondria [179,180] (Figure 3 and Table 2). Importantly, it was demonstrated that full-length APP interacts with the mitochondrial import machinery formed by TOMM40 and TIMM23, inhibiting Cyt c oxidase activity and increasing mitochondrial ROS and dysfunction [181]. This finding was further supported by the presence of APP catabolites (i.e., Aβ, C99, and AICD) in the mitochondria [182,183,184,185,186,187]. Supporting APP localization and its processing in MAMs, we and others demonstrated the presence of APP-CTFs and Aβ in MAMs and also of different components of the γ-secretase complex in ER and mitochondria [183,187,188,189,190]. Consistently, we also showed in vitro activities of both β- and γ- secretases in MAMs of AD mice brains [187]. It has been proposed that MAMs behave as a detergent-resistant lipid raft-like domain [190], sustaining β- and γ-secretase the activities in MAMs [191,192,193].
Most importly, we and others also evidenced an increase of ER and mitochondria contact sites in SH-SY5Y cells overexpressing APPswe [187], in fibroblasts from SAD and FAD individuals, as well as in MEFs isolated from mice simple KO for either PS1 or PS2 or both [190]. Increased ER and mitochondria tethering is consistent with the hypothesis that Aβ is produced at the interface between ER and mitochondria [194] and its subsequent transport into mitochondria via the translocase of the OMM (TOMM) machinery [194], trigerring mitochondria dysfunctions (i.e., enhanced ROS, reduced mitochondrial respiratory chain complex I activity, and interaction with several mitochondrial proteins) [195,196,197,198] (Figure 3). In addition, Aβ was shown to regulate fatty-acid metabolism, another function of MAMs, as revealed by neutral lipids accumulation found in neuroblastoma and CHO cells expressing APPswe or APPLDN mutations [187] and in human-derived SAD and FAD fibroblasts [190]. While the toxicity of Aβ towards mitochondria is undoubtful, Pera et al. also demonstrated the toxic charge of C99 fragment in MAMs. Indeed, the authors showed in PSs double KO MEFs treated with DAPT (a γ-secretase inhibitor), an accumulation of C99 peptide triggering MAMs deregulation, as manifested by an upregulation of sphingolipid metabolism translated by a higher ceramide and sphingomyelin synthesis in mitochondrial membranes [186,199]. Interestingly, ceramide has been shown to alter mitochondrial lipid composition, disrupting ΔΨm and mitochondrial membrane permeability [200]. It was also suggested that mitochondrial ceramide elevation could disrupt the assembly and the activity of mitochondrial respiratory complexes and provoke mitochondrial dysfunctions and apoptosis [186]. Another study confirmed the contribution of C99 to lipid metabolism [201]. Hence, pathogenic accumulation of C99 in the ER induces excessive uptake of extracellular cholesterol, trafficking from the plasma membrane to the ER, resulting in the continuous formation, activation, and turnover of MAMs in AD [201] (Table 2). In a recent study, we demonstrated that APP-CTFs accumulates in mitochondria in cellular AD models (SH-SY5Y cells expressing either APPswe, or C99 fragment) as well as in brains of 3xTg-AD mice treated with γ-secretase inhibitor, and in C99-expressing mice [202]. We also demonstrated that APP-CTFs, independently from Aβ, triggers excessive mitochondrial cristae disorganization, enhances mitochondrial ROS production, and elicits basal mitophagy failure associated with deficient fusion of mitochondria with lysosomes. Importantly, we reported a correlation between APP-CTFs accumulation in mitochondria and the mitophagy failure molecular signature in human post-mortem SAD brains [202] (Table 2).
Taken together, these results led us to assume that APP-CTFs (particularly C99) is directly generated in MAMs, or targeted through an unknown mechanism to MAMs after being produced in endosomes. The overaccumulation of APP-CTFs in MAMs consequently triggers mitochondrial and lipid dysfunctions, and a mitophagy failure phenotype contributing to neuronal demise in AD.

3.4. Tau

Tau is an associated-microtubule protein contributing to axonal transport and synaptic plasticity [203]. Tau is involved in tauopathies such as FTD, fronto-temporal lobar degeneration with Pick bodies, corticobasal degeneration, and AD [204]. During the pathophysiological process of AD, Tau is abnormally hyperphosphorylated and loses its affinity for microtubules. pTau is then redistributed in the somato-dendritic compartment, where it aggregates in insoluble filaments forming NFTs and leading to neuronal dysfunctions through axonal transport alteration [203,205].
Even if there is no description of FAD linked to Tau mutations, “artificial” Tau mutations are used in fundamental research to yield Tau hyperphosphorylation and to mimic, at least in part, some aspects of AD-related Tau pathology. Interestingly, P301L mutation of Tau (Tau P301L), which is known to cause Tau hyperphosphorylation, was reported to decrease mitochondrial complex V levels, leading to impaired mitochondrial respiration and ATP synthesis as well as increased ROS production [206]. The same mutation in the N2a cell line was also shown to interact with Parkin and sequester it in the cytosol, triggering impaired mitophagy [207]. Besides, Tau overexpression in HEK293 cells and rat primary hippocampal neurons enhances the fusion of mitochondria by increasing the expression of fusion proteins OPA1, MFN1, and MFN2 triggering an alteration of mitochondrial dynamics and functions illustrated by decreased ATP level, ATP/ADP ratio, and complex I activity [208]. Perrault et al. revealed that Tau P301L increases the number of ER and mitochondria contacts in the spinal cord motor neurons. They also notice enhanced colocalization of Tau P301L with ER membranes in this model and also in AD brains [209], likely increasing ER-mitochondria tethering. In another study, Tau overexpression in CHO or N2a cells was reported to alter the distribution of mitochondria and ER to the cell periphery [210]. Afreen et al. then reported that calpain-mediated cleavage of Tau generating Tau45-230 fragment (occurring downstream of increased Ca2+ influx after exposure of hippocampal neurons to Aβo) is associated with the cytoskeleton plus membrane-bound organelles and reduces the number of mitochondria transported along hippocampal axons [211].
Together, these data suggest that the hyperphosphorylation and elevated intraneuronal concentration of Tau would affect the transport of mitochondria or ER dynamics, thus impairing the formation of ER and mitochondria contact sites in AD.
Figure 3. APP and its derived fragments APP-CTFs and Aβ; presenilins (PSs), pTau and APOE4 are involved in the impairment of the expression and activity of several mitochondrial proteins (highlighted in red) and interfere with UPR in the ER. APP localizes and is processed by β- and γ-secretases in MAMs. AD Molecular partners also disturb MAMs structure and function.
Figure 3. APP and its derived fragments APP-CTFs and Aβ; presenilins (PSs), pTau and APOE4 are involved in the impairment of the expression and activity of several mitochondrial proteins (highlighted in red) and interfere with UPR in the ER. APP localizes and is processed by β- and γ-secretases in MAMs. AD Molecular partners also disturb MAMs structure and function.
Ijms 21 09521 g003
Table 2. Alterations of mitochondria- and MAMs-structure and functions associated with AD-related proteins. These alterations include: ↓ reduced or ↑ increased expression/activity, ⟷ interaction observed in AD study models (cells and mice brains) and human derived samples (SAD brains and SAD/FAD fibroblasts). Mito: mitochondria, KO: knock-out; PSDKO: PS1 and PS2 double KO; KI: knock-in; “SI”: silencing; “OE”: overexpressing; PS: phosphatidylserine; CO: cholesterol oleate; CE: cholesterol ester; CM: conditioned media, WT: wild type; CGC: cerebellar granule cells; ROS: reactive oxygen species.
Table 2. Alterations of mitochondria- and MAMs-structure and functions associated with AD-related proteins. These alterations include: ↓ reduced or ↑ increased expression/activity, ⟷ interaction observed in AD study models (cells and mice brains) and human derived samples (SAD brains and SAD/FAD fibroblasts). Mito: mitochondria, KO: knock-out; PSDKO: PS1 and PS2 double KO; KI: knock-in; “SI”: silencing; “OE”: overexpressing; PS: phosphatidylserine; CO: cholesterol oleate; CE: cholesterol ester; CM: conditioned media, WT: wild type; CGC: cerebellar granule cells; ROS: reactive oxygen species.
ProteinsAlterationsStudy ModelsRef
APOEε4↓DRP1, ↑MFN1
Mito fusion (↑COX1, ↑TOMM40)
Mitophagy failure (↓PINK1, ↑Parkin, ↑p62)
AD brains
APOEε4 KI mice
[160]
↑VDAC1, ↑IP3R, ↑BAP31, ↑Fis1, ↑MFN2, ↓Grp75N2acells “OE” APOE4 [161]
↑PS, ↑CO, ↑CE, ↑Lipid dropletsAPOE4 KI mice astrocytes CM treatment[163]
PS1 & PS2↑LC3-II and ↑autophagosomesMEFs (PS1KO, PS2KO, & PSDKO)
HepG2 cells (PS1“SI”)
[165,166]
PS2⟷Mfn2
↑MAMs
MEFs (MFN2 KO), Human FAD fibroblasts, SHSY-5Y & primary neurons “OE” MFN2[46,172]
↑ER-mitochondria coupling
Autophagy blockade (↑LC3-II, ↑p62)
FAD PS2 (T122R, N141I) &.
SHSY-5Y “OE” FAD PS2
[174]
↓IP3R1/3, ↓MIRO1, ↓KIF5C
↑MAMs
Purkinje & SHSY-5Y “OE” FAD PS1(E280A)[176]
Autophagy impairmentFAD PS1 (A246E, M233T, H163Y, M146L, L392V) fibroblasts[177]
↑MAMs
↑ MAMs function (↑CE, ↑PS, ↑PE, ↑lipid droplets)
PS1 KO, PS2 KO, PSDKO
FAD and SAD patient fibroblasts
[190]
APP, Aβ
& APP-CTFs
APP in mitochondriaHCN-1A, COS cells “OE” APPWT or APP chimeric mutants & APPswe mice[169,180,181]
APP, APP-CTFs, Aβ in MAMs
↑MAMs
↑ceramide, ↓sphingomyelin,↑lipid droplets (Aβ dependant)
SHSY-5Y “OE” APPswe, APPLDN
Mice brains (WT,APP23 & APP23/PS45)
PSDKO +/− DAPT
[185,186,187,190,194]
↑mito ROS, ↓complex I &IV activities (Aβ-dependant) AD brains; Aβ-treated CGCs [195,196]
↑mito ROS, ↓complex I activity
Mitophagy failure (APP-CTFs dependant)
3xTgAD, 2xTgAD, AAV-C99 mice, SHSY-5Y cells “OE”APPswe or C99 +/− γ-secretase inhibitor
AD brains
[202]
Tau↓complex V, ↓ATP, ↑mitROSTauP301L mice[206]
Mitophagy failure (↓Parkin mitochondrial translocation)N2a “OE” TauP301L[207]
↑OPA1, ↑MFN1/2, Mitochondrial fragmentation, ↓complex I activity, ↓ATPHEK293 & rat primary hippocampal neurons “OE” Tau[208]
Mitochondria & ER distributionsCHO and N2a cells “OE” Tau[210]
↓Mito axonal movementHippocampal neurons “OE” Tau45-230 [211]
↑MAMsMotor neurons “OE” TauP301L[209]
↓MAMsTauP301L mice[212]

4. MAMs as a Potential Therapeutic Strategy

The studies highlited above tend to demonstrate that MAMs structure and function alterations could play a role during the pathophysiological process of AD. Taking this into account, we may assume that MAMs should constitute a new therapeutic target in AD. Up to now, different assays were driven to target either ER or mitochondrial dysfunctions in neurodegenerative diseases, as an indirect way to alleviate MAMs-associated dysfunctions. Among them, Salubrinal, a selective inhibitor eIF2α phosphatase enzymes, prevents ceramide-induced apoptotic cell death through the inhibition of IRE1α /apoptosis signal regulating kinase 1 (ASK1)/c-Jun N-terminal kinase (JNK) phosphorylation [213] and decreases apoptosis in neurons and microglia exposed to Aβ [214,215]. Salubrinal also maintains eIF2α phosphorylated, thus inhibiting the synthesis of ER secretory proteins and decreasing proteoglycans and other profibrotic proteins, then facilitating axonal regeneration [216]. Celastrol and Tanespimycin, which target the heat shock protein response, are other potential drugs showing promising results in transgenic mouse models of AD by preventing NFκB activation, inhibiting BACE1 expression and reducing Aβ40, Aβ42 levels as well as C99 and sAPPβ production and NFTs [217,218].
Among mitochondrial targets, mitochondrial division inhibitor-1 (mdivi-1), a selective inhibitor of mitochondrial fission through the targeting of DRP1, has been reported to display neuroprotective properties by rescuing propofol-induced toxicity in stem cell-derived neurons [219]. In microglial cells, inhibition of mitochondrial fission by mdivi-1 significantly reduced Aβ-induced expression of CD11b (a pan-macrophage marker) and reversed cell death. Importantly, mdivi-1 treatment also markedly reversed mitochondrial dysfunction (loss of ΔΨm and Cyt c release) and caspase-3 activation) [220]. These results were further confirmed in neuroblastoma N2a cells subjected to Aβ42 treatment, showing that mdivi-1 enhanced mitochondrial fusion activity, lowered fission machinery, and increased biogenesis and synaptic proteins. These features were associated with elevated cell viability in cells exposed to Aβ and treated with mdivi-1 [221]. The same group reported synergistic protective effects of a mitochondria-targeted antioxidant SS31 and mdivi-1 in N2a cells expressing APPswe/LDN mutations by increasing mitochondrial DNA copy number and cell survival, reducing mitochondrial dysfunctions. Altogether, these data suggest that mdivi-1 exerts neuroprotective effects against Aβ-induced mitochondrial structure and function alterations and apoptosis [222]. Another mitochondrial fission inhibitor, P110, abrogating the interaction between DRP1 and FIS1, was shown to inhibit excessive fission as well as to restore mitochondrial function [223]; however, its therapeutic potential is still not demonstrated in the context of AD.
Few studies pointed to MAMs as a potential therapeutic target. Arismendi-Morillo et al. showed that restriction of glucose or ketogenesis influences the structure and function of MAMs and could be used as a therapeutic approach in astrocytic tumors [224]. Lynes et al. also found that palmitoylation of cysteine residues could assemble ER membrane proteins in MAMs [225]. The use of artificial linkers to control the connections between ER and mitochondria was also proposed. The expression of an ER-mitochondria linker in a Drosophila model of AD increases physical coupling between both organelles, raises Ca²+ levels in mitochondria known to be detrimental to their function. Indeed, authors also noticed an increase of ROS production in flies expressing the linker. However, while the expression of ER-mitochondria linker in flies reverses electron transport and extends Drosophila lifespan, authors report negative outcomes in mice [226]. Rapamycin-inducible artificial linkers with different lengths [227] allow to demonstrate that distinct processes may require an optimal spatial ER-mitochondria distance. Indeed, while a loose connection is needed for Ca²+ delivery, tighter connections are required for lipid and protein transfer between organelles [227]. A recent study proposed urolithin A (produced in the gut microbiome) as a pharmacological tool to control ER-mitochondria contacts and Ca2+ homeostasis in AD [228]. Urolithin A lowered mitochondrial Ca2+ influx and significantly alleviated high glucose-induced mitochondrial ROS accumulation. Most importantly urolithin A reduces the expression of APP, BACE1, as well as Aβ production and Tau phosphorylation [228].

5. Conclusions

A large number of studies delineated the proteome composition of MAMs and deciphered the molecular and functional links between MAMs dysfunctions and AD, where disturbances of mitochondrial and ER Ca2+ signaling were consistently reported. Besides Ca2+ tunneling and apoptosis, MAMs are also the hub for autophagosome and mitophagosome membranes biosynthesis and lipid biosynthesis that are all cellular hallmarks that were reported to be perturbed in AD study models and most importantly in human AD-derived samples. About 30 million individuals are estimated to be affected by AD worldwide and to date, no effective treatment exists to arrest disease progression. This review highlights the relevance of MAMs dysfunctions in AD development and pinpoints the possibility to target MAMs molecular components as therapeutical targets to treat or alleviate or slow down AD pathogenesis.

Author Contributions

Writing—original draft preparation F.E., P.F.K., A.M., L.V.-B. Writing—review and editing, F.E., P.F.K., and M.C. Editing, M.C., and F.C. Supervision, M.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work has been developed and supported through the LABEX (excellence laboratory, program investment for the future) DISTALZ (Development of Innovative Strategies for a Transdisciplinary approach to Alzheimer’s disease) to FC, and by Fondation Vaincre Alzheimer to MC and by São Paulo Research Foundation (FAPESP, Grant #2019/26084-2) to P.F.K.

Acknowledgments

We thank Franck Aguila for figures drawing.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Checler, F. Processing of the beta-amyloid precursor protein and its regulation in Alzheimer’s disease. J. Neurochem. 1995, 65, 1431–1444. [Google Scholar] [CrossRef] [PubMed]
  2. Nhan, H.S.; Chiang, K.; Koo, E.H. The multifaceted nature of amyloid precursor protein and its proteolytic fragments: Friends and foes. Acta Neuropathol. 2015, 129, 1–19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Corder, E.H.; Saunders, A.M.; Strittmatter, W.J.; Schmechel, D.E.; Gaskell, P.C.; Small, G.W.; Roses, A.D.; Haines, J.L.; Pericak-Vance, M.A. Gene dose of apolipoprotein E type 4 allele and the risk of Alzheimer’s disease in late onset families. Science 1993, 261, 921–923. [Google Scholar] [CrossRef] [PubMed]
  4. 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] [PubMed] [Green Version]
  5. Kunkle, B.W.; Grenier-Boley, B.; Sims, R.; Bis, J.C.; Damotte, V.; Naj, A.C.; Boland, A.; Vronskaya, M.; van der Lee, S.J.; Amlie-Wolf, A.; et al. Genetic meta-analysis of diagnosed Alzheimer’s disease identifies new risk loci and implicates Aβ, tau, immunity and lipid processing. Nat. Genet. 2019, 51, 414–430. [Google Scholar] [CrossRef] [Green Version]
  6. Bellenguez, C.; Grenier-Boley, B.; Lambert, J.C. Genetics of Alzheimer’s disease: Where we are, and where we are going. Curr. Opin. Neurobiol. 2020, 61, 40–48. [Google Scholar] [CrossRef]
  7. Tanzi, R.E.; Bertram, L. Twenty years of the Alzheimer’s disease amyloid hypothesis: A genetic perspective. Cell 2005, 120, 545–555. [Google Scholar] [CrossRef]
  8. Van Cauwenberghe, C.; Van Broeckhoven, C.; Sleegers, K. The genetic landscape of Alzheimer disease: Clinical implications and perspectives. Genet. Med. 2016, 18, 421–430. [Google Scholar] [CrossRef] [Green Version]
  9. Hardy, J.A.; Higgins, G.A. Alzheimer’s disease: The amyloid cascade hypothesis. Science 1992, 256, 184–185. [Google Scholar] [CrossRef]
  10. Panza, F.; Lozupone, M.; Logroscino, G.; Imbimbo, B.P. A critical appraisal of amyloid-β-targeting therapies for Alzheimer disease. Nat. Rev. Neurol. 2019, 15, 73–88. [Google Scholar] [CrossRef]
  11. Flammang, B.; Pardossi-Piquard, R.; Sevalle, J.; Debayle, D.; Dabert-Gay, A.S.; Thévenet, A.; Lauritzen, I.; Checler, F. Evidence that the amyloid-β protein precursor intracellular domain, AICD, derives from β-secretase-generated C-terminal fragment. J. Alzheimer’s Dis. 2012, 30, 145–153. [Google Scholar] [CrossRef] [PubMed]
  12. Willem, M.; Tahirovic, S.; Busche, M.A.; Ovsepian, S.V.; Chafai, M.; Kootar, S.; Hornburg, D.; Evans, L.D.B.; Moore, S.; Daria, A.; et al. n-Secretase processing of APP inhibits neuronal activity in the hippocampus. Nature 2015, 526, 443–447. [Google Scholar] [CrossRef] [PubMed]
  13. Zhang, Z.; Song, M.; Liu, X.; Su Kang, S.; Duong, D.M.; Seyfried, N.T.; Cao, X.; Cheng, L.; Sun, Y.E.; Ping Yu, S.; et al. Delta-secretase cleaves amyloid precursor protein and regulates the pathogenesis in Alzheimer’s disease. Nat. Commun. 2015, 6, 1–16. [Google Scholar] [CrossRef] [PubMed]
  14. Baranger, K.; Marchalant, Y.; Bonnet, A.E.; Crouzin, N.; Carrete, A.; Paumier, J.M.; Py, N.A.; Bernard, A.; Bauer, C.; Charrat, E.; et al. MT5-MMP is a new pro-amyloidogenic proteinase that promotes amyloid pathology and cognitive decline in a transgenic mouse model of Alzheimer’s disease. Cell. Mol. Life Sci. 2016, 73, 217–236. [Google Scholar] [CrossRef] [Green Version]
  15. Bourgeois, A.; Lauritzen, I.; Lorivel, T.; Bauer, C.; Checler, F.; Pardossi-Piquard, R. Intraneuronal accumulation of C99 contributes to synaptic alterations, apathy-like behavior, and spatial learning deficits in 3 × TgAD and 2 × TgAD mice. Neurobiol. Aging 2018, 71, 21–31. [Google Scholar] [CrossRef]
  16. Lauritzen, I.; Pardossi-Piquard, R.; Bauer, C.; Brigham, E.; Abraham, J.D.; Ranaldi, S.; Fraser, P.; St-George-Hyslop, P.; Le Thuc, O.; Espin, V.; et al. The β-secretase-derived C-terminal fragment of βAPP, C99, but not Aβ, is a key contributor to early intraneuronal lesions in triple-transgenic mouse hippocampus. J. Neurosci. 2012, 32, 16243–16255. [Google Scholar] [CrossRef]
  17. Lauritzen, I.; Pardossi-Piquard, R.; Bourgeois, A.; Pagnotta, S.; Biferi, M.G.; Barkats, M.; Lacor, P.; Klein, W.; Bauer, C.; Checler, F. Intraneuronal aggregation of the β-CTF fragment of APP (C99) induces Aβ-independent lysosomal-autophagic pathology. Acta Neuropathol. 2016, 132, 257–276. [Google Scholar] [CrossRef] [Green Version]
  18. Kwart, D.; Gregg, A.; Scheckel, C.; Murphy, E.A.; Paquet, D.; Duffield, M.; Fak, J.; Olsen, O.; Darnell, R.B.; Tessier-Lavigne, M. Erratum: A large panel of isogenic APP and PSEN1 mutant human iPSC neurons reveals shared endosomal abnormalities mediated by APP β-CTFs, not Aβ. Neuron 2019, 104, 256–270. [Google Scholar] [CrossRef]
  19. Berridge, M.J. The endoplasmic reticulum: A multifunctional signaling organelle. Cell Calcium 2002, 32, 235–249. [Google Scholar] [CrossRef]
  20. Osellame, L.D.; Blacker, T.S.; Duchen, M.R. Cellular and molecular mechanisms of mitochondrial function. Best Pract. Res. Clin. Endocrinol. Metab. 2012, 26, 711–723. [Google Scholar] [CrossRef] [Green Version]
  21. Giacomello, M.; Pellegrini, L. The coming of age of the mitochondria—ER contact: A matter of thickness. Cell Death Differ. 2016, 23, 1417–1427. [Google Scholar] [CrossRef] [PubMed]
  22. Rizzuto, R.; Pinton, P.; Carrington, W.; Fay, F.S.; Fogarty, K.E.; Lifshitz, L.M.; Tuft, R.A.; Pozzan, T. Close contacts with the endoplasmic reticulum as determinants of mitochondrial Ca2+ responses. Science 1998, 280, 1763–1766. [Google Scholar] [CrossRef] [PubMed]
  23. Csordás, G.; Renken, C.; Várnai, P.; Walter, L.; Weaver, D.; Buttle, K.F.; Balla, T.; Mannella, C.A.; Hajnóczky, G. Structural and functional features and significance of the physical linkage between ER and mitochondria. J. Cell Biol. 2006, 174, 915–921. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Walter, L.; Hajnóczky, G. Mitochondria and endoplasmic reticulum: The lethal interorganelle cross-talk. J. Bioenerg. Biomembr. 2005, 37, 191–206. [Google Scholar] [CrossRef]
  25. Veeresh, P.; Kaur, H.; Sarmah, D.; Mounica, L.; Verma, G.; Kotian, V.; Kesharwani, R.; Kalia, K.; Borah, A.; Wang, X.; et al. Endoplasmic reticulum–mitochondria crosstalk: From junction to function across neurological disorders. Ann. N. Y. Acad. Sci. 2019, 1457, 41–60. [Google Scholar] [CrossRef]
  26. Van Vliet, A.R.; Agostinis, P. Mitochondria-Associated membranes and ER stress. In Current Topics in Microbiology and Immunology; Springer Nature Switzerland AG: Cham, Switzerland, 2018; Volume 414, pp. 73–102. [Google Scholar]
  27. Gerakis, Y.; Hetz, C. Emerging roles of ER stress in the etiology and pathogenesis of Alzheimer’s disease. FEBS J. 2018, 285, 995–1011. [Google Scholar] [CrossRef] [Green Version]
  28. Hashimoto, S.; Saido, T.C. Critical review: Involvement of endoplasmic reticulum stress in the aetiology of Alzheimer’s disease. Open Biol. 2018, 8. [Google Scholar] [CrossRef]
  29. Esteras, N.; Abramov, A.Y. Mitochondrial calcium deregulation in the mechanism of beta-amyloid and tau pathology. Cells 2020, 9, 2135. [Google Scholar] [CrossRef]
  30. Wang, W.; Zhao, F.; Ma, X.; Perry, G.; Zhu, X. Mitochondria dysfunction in the pathogenesis of Alzheimer’s disease: Recent advances. Mol. Neurodegener. 2020, 15, 1–22. [Google Scholar] [CrossRef]
  31. Chen, H.; Chomyn, A.; Chan, D.C. Disruption of fusion results in mitochondrial heterogeneity and dysfunction. J. Biol. Chem. 2005, 280, 26185–26192. [Google Scholar] [CrossRef] [Green Version]
  32. Cipolat, S.; De Brito, O.M.; Dal Zilio, B.; Scorrano, L. OPA1 requires mitofusin 1 to promote mitochondrial fusion. Proc. Natl. Acad. Sci. USA 2004, 101, 15927–15932. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Pich, S.; Bach, D.; Briones, P.; Liesa, M.; Camps, M.; Testar, X.; Palacín, M.; Zorzano, A. The Charcot-Marie-Tooth type 2A gene product, Mfn2, up-regulates fuel oxidation through expression of OXPHOS system. Hum. Mol. Genet. 2005, 14, 1405–1415. [Google Scholar] [CrossRef] [PubMed]
  34. Youle, R.J.; Karbowski, M. Mitochondrial fission in apoptosis. Nat. Rev. Mol. Cell Biol. 2005, 6, 657–663. [Google Scholar] [CrossRef] [PubMed]
  35. Martorell-Riera, A.; Segarra-Mondejar, M.; Muñoz, J.P.; Ginet, V.; Olloquequi, J.; Pérez-Clausell, J.; Palacín, M.; Reina, M.; Puyal, J.; Zorzano, A.; et al. Mfn2 downregulation in excitotoxicity causes mitochondrial dysfunction and delayed neuronal death. EMBO J. 2014, 33, 2388–2407. [Google Scholar] [CrossRef] [PubMed]
  36. Tanaka, A.; Cleland, M.M.; Xu, S.; Narendra, D.P.; Suen, D.F.; Karbowski, M.; Youle, R.J. Proteasome and p97 mediate mitophagy and degradation of mitofusins induced by Parkin. J. Cell Biol. 2010, 191, 1367–1380. [Google Scholar] [CrossRef] [Green Version]
  37. Naon, D.; Zaninello, M.; Giacomello, M.; Varanita, T.; Grespi, F.; Lakshminaranayan, S.; Serafini, A.; Semenzato, M.; Herkenne, S.; Hernández-Alvarez, M.I.; et al. Critical reappraisal confirms that Mitofusin 2 is an endoplasmic reticulum-mitochondria tether. Proc. Natl. Acad. Sci. USA 2016, 113, 11249–11254. [Google Scholar] [CrossRef] [Green Version]
  38. Chen, H.; McCaffery, J.M.; Chan, D.C. Mitochondrial fusion protects against neurodegeneration in the cerebellum. Cell 2007, 130, 548–562. [Google Scholar] [CrossRef] [Green Version]
  39. Zhang, L.; Trushin, S.; Christensen, T.A.; Bachmeier, B.V.; Gateno, B.; Schroeder, A.; Yao, J.; Itoh, K.; Sesaki, H.; Poon, W.W.; et al. Altered brain energetics induces mitochondrial fission arrest in Alzheimer’s Disease. Sci. Rep. 2016, 6, 1–12. [Google Scholar] [CrossRef] [Green Version]
  40. Kim, Y.J.; Park, J.K.; Kang, W.S.; Kim, S.K.; Han, C.; Na, H.R.; Park, H.J.; Kim, J.W.; Kim, Y.Y.; Park, M.H.; et al. Association between mitofusin 2 gene polymorphisms and late-onset Alzheimer’s disease in the Korean population. Psychiatry Investig. 2017, 14, 81–85. [Google Scholar] [CrossRef] [Green Version]
  41. Wang, X.; Su, B.; Lee, H.G.; Li, X.; Perry, G.; Smith, M.A.; Zhu, X. Impaired balance of mitochondrial fission and fusion in Alzheimer’s disease. J. Neurosci. 2009, 29, 9090–9103. [Google Scholar] [CrossRef]
  42. Manczak, M.; Calkins, M.J.; Reddy, P.H. Impaired mitochondrial dynamics and abnormal interaction of amyloid beta with mitochondrial protein Drp1 in neurons from patients with Alzheimer’s disease: Implications for neuronal damage. Hum. Mol. Genet. 2011, 20, 2495–2509. [Google Scholar] [CrossRef] [PubMed]
  43. Park, J.; Choi, H.; Min, J.S.; Kim, B.; Lee, S.R.; Yun, J.W.; Choi, M.S.; Chang, K.T.; Lee, D.S. Loss of mitofusin 2 links beta-amyloid-mediated mitochondrial fragmentation and Cdk5-induced oxidative stress in neuron cells. J. Neurochem. 2015, 132, 687–702. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Gan, X.; Wu, L.; Huang, S.; Zhong, C.; Li, G.; Yu, H.; Swerdlow, R.H.; Chen, J.X.; Yan, S.S. Oxidative stress-mediated activation of extracellualr signal-regulated kinase contributes to mild cognitive impairment-related mitochondrial dysfuntion. Free Radic. Biol. Med. 2014, 75, 230–240. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Leal, N.S.; Schreiner, B.; Pinho, C.M.; Filadi, R.; Wiehager, B.; Karlström, H.; Pizzo, P.; Ankarcrona, M. Mitofusin-2 knockdown increases ER-mitochondria contact and decreases amyloid β-peptide production. J. Cell. Mol. Med. 2016, 20, 1686–1695. [Google Scholar] [CrossRef]
  46. Filadi, R.; Greotti, E.; Turacchio, G.; Luini, A.; Pozzan, T.; Pizzo, P. Presenilin 2 modulates endoplasmic reticulum-mitochondria coupling by tuning the antagonistic effect of mitofusin 2. Cell Rep. 2016, 15, 2226–2238. [Google Scholar] [CrossRef] [Green Version]
  47. Berridge, M.J. Inositol triphosphate and calcium signalling. Nature 1993, 361, 315–325. [Google Scholar] [CrossRef]
  48. Joseph, S.K.; Hajnóczky, G. IP3 receptors in cell survival and apoptosis: Ca2+ release and beyond. Apoptosis 2007, 12, 951–968. [Google Scholar] [CrossRef] [Green Version]
  49. Mikoshiba, K. IP3 receptor/Ca2+ channel: From discovery to new signaling concepts. J. Neurochem. 2007, 102, 1426–1446. [Google Scholar] [CrossRef]
  50. De Stefani, D.; Raffaello, A.; Teardo, E.; Szabò, I.; Rizzuto, R. A forty-kilodalton protein of the inner membrane is the mitochondrial calcium uniporter. Nature 2011, 476, 336–340. [Google Scholar] [CrossRef]
  51. Szabadkai, G.; Bianchi, K.; Várnai, P.; De Stefani, D.; Wieckowski, M.R.; Cavagna, D.; Nagy, A.I.; Balla, T.; Rizzuto, R. Chaperone-Mediated coupling of endoplasmic reticulum and mitochondrial Ca2+ channels. J. Cell Biol. 2006, 175, 901–911. [Google Scholar] [CrossRef] [Green Version]
  52. Yoo, B.C.; Kim, S.H.; Cairns, N.; Fountoulakis, M.; Lubec, G. Deranged expression of molecular chaperones in brains of patients with Alzheimer’s disease. Biochem. Biophys. Res. Commun. 2001, 280, 249–258. [Google Scholar] [CrossRef] [PubMed]
  53. Cuadrado-Tejedor, M.; Vilariño, M.; Cabodevilla, F.; Del Río, J.; Frechilla, D.; Pérez-Mediavilla, A. Enhanced expression of the voltage-dependent anion channel 1 (VDAC1) in Alzheimer’s disease transgenic mice: An insight into the pathogenic effects of amyloid-β. J. Alzheimers Dis. 2011, 23, 195–206. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Hedskog, L.; Pinho, C.M.; Filadi, R.; Rönnbäck, A.; Hertwig, L.; Wiehager, B.; Larssen, P.; Gellhaar, S.; Sandebring, A.; Westerlund, M.; et al. Modulation of the endoplasmic reticulum-mitochondria interface in Alzheimer’s disease and related models. Proc. Natl. Acad. Sci. USA 2013, 110, 7916–7921. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Manczak, M.; Reddy, P.H. Abnormal interaction of VDAC1 with amyloid beta and phosphorylated tau causes mitochondrial dysfunction in Alzheimer’s disease. Hum. Mol. Genet. 2012, 21, 5131–5146. [Google Scholar] [CrossRef] [PubMed]
  56. Fernandez-Echevarria, C.; Díaz, M.; Ferrer, I.; Canerina-Amaro, A.; Marin, R. Aβ promotes VDAC1 channel dephosphorylation in neuronal lipid rafts. Relevance to the mechanisms of neurotoxicity in Alzheimer’s disease. Neuroscience 2014, 278, 354–366. [Google Scholar] [CrossRef]
  57. Manczak, M.; Sheiko, T.; Craigen, W.J.; Reddy, P.H. Reduced VDAC1 protects against Alzheimer’s disease, mitochondria, and synaptic deficiencies. J. Alzheimer’s Dis. 2013, 37, 679–690. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Ito, E.; Oka, K.; Etcheberrigaray, R.; Nelson, T.J.; McPhie, D.L.; Tofel-Grehl, B.; Gibson, G.E.; Alkon, D.L. Internal Ca2+ mobilization is altered in fibroblasts from patients with Alzheimer disease. Proc. Natl. Acad. Sci. USA 1994, 91, 534–538. [Google Scholar] [CrossRef] [Green Version]
  59. Cheung, K.; Mei, L.; Mak, D.D.; Hayashi, I.; Iwatsubo, T.; Kang, D.E.; Foskett, J.K. Gain-of-function enhancement of InsP3 receptor modal gating by familial Alzheimer’s disease-linked presenilin mutants in human cells and mouse neurons. Sci. Signal. 2010, 3. [Google Scholar] [CrossRef] [Green Version]
  60. Demuro, A.; Parker, I. Cytotoxicity of intracellular Aβ42 amyloid oligomers involves Ca2+ release from the endoplasmic reticulum by stimulated production of inositol trisphosphate. J. Neurosci. 2013, 33, 3824–3833. [Google Scholar] [CrossRef] [Green Version]
  61. Shilling, D.; Müller, M.; Takano, H.; Mak, D.O.D.; Abel, T.; Coulter, D.A.; Foskett, J.K. Suppression of InsP3 receptor-mediated Ca2+ signaling alleviates mutant presenilin-linked familial Alzheimer’s disease pathogenesis. J. Neurosci. 2014, 34, 6910–6923. [Google Scholar] [CrossRef] [Green Version]
  62. Cheung, K.H.; Shineman, D.; Müller, M.; Cárdenas, C.; Mei, L.; Yang, J.; Tomita, T.; Iwatsubo, T.; Lee, V.M.Y.; Foskett, J.K. Mechanism of Ca2+ disruption in Alzheimer’s disease by presenilin regulation of InsP3 receptor channel gating. Neuron 2008, 58, 871–883. [Google Scholar] [CrossRef] [Green Version]
  63. Del Prete, D.; Checler, F.; Chami, M. Ryanodine receptors: Physiological function and deregulation in Alzheimer disease. Mol. Neurodegener. 2014, 9, 21. [Google Scholar] [CrossRef] [Green Version]
  64. Lacampagne, A.; Liu, X.; Reiken, S.; Bussiere, R.; Meli, A.C.; Lauritzen, I.; Teich, A.F.; Zalk, R.; Saint, N.; Arancio, O.; et al. Post-Translational remodeling of ryanodine receptor induces calcium leak leading to Alzheimer’s disease-like pathologies and cognitive deficits. Acta Neuropathol. 2017, 134, 749–767. [Google Scholar] [CrossRef] [PubMed]
  65. Bussiere, R.; Lacampagne, A.; Reiken, S.; Liu, X.; Scheuerman, V.; Zalk, R.; Martin, C.; Checler, F.; Marks, A.R.; Chami, M. Amyloid beta production is regulated by beta2-adrenergic signaling-mediated post-translational modifications of the ryanodine receptor. J. Biol. Chem. 2017, 292, 10153–10168. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Oules, B.; Del Prete, D.; Greco, B.; Zhang, X.; Lauritzen, I.; Sevalle, J.; Moreno, S.; Paterlini-Brechot, P.; Trebak, M.; Checler, F.; et al. Ryanodine receptor blockade reduces amyloid-beta load and memory impairments in Tg2576 mouse model of Alzheimer disease. J. Neurosci. 2012, 32, 11820–11834. [Google Scholar] [CrossRef] [PubMed]
  67. Stutzmann, G.E.; Smith, I.; Caccamo, A.; Oddo, S.; LaFerla, F.M.; Parker, I. Enhanced ryanodine receptor recruitment contributes to Ca2+ disruptions in young, adult, and aged Alzheimer’s disease mice. J. Neurosci. 2006, 26, 5180–5189. [Google Scholar] [CrossRef] [Green Version]
  68. Stutzmann, G.E.; Smith, I.; Caccamo, A.; Oddo, S.; Parker, I.; Laferla, F. Enhanced ryanodine-mediated calcium release in mutant PS1-expressing Alzheimer’s mouse models. Ann. N. Y. Acad. Sci. 2007, 1097, 265–277. [Google Scholar] [CrossRef]
  69. Li, C.; Li, L.; Yang, M.; Zeng, L.; Sun, L. PACS-2: A key regulator of mitochondria-associated membranes (MAMs). Pharmacol. Res. 2020, 160, 105080. [Google Scholar] [CrossRef]
  70. Werneburg, N.W.; Bronk, S.F.; Guicciardi, M.E.; Thomas, L.; Dikeakos, J.D.; Thomas, G.; Gores, G.J. Tumor necrosis factor-related apoptosis-inducing ligand (TRAIL) protein-induced lysosomal translocation of proapoptotic effectors is mediated by phosphofurin acidic cluster sorting protein-2 (PACS-2). J. Biol. Chem. 2012, 287, 24427–24437. [Google Scholar] [CrossRef] [Green Version]
  71. Simmen, T.; Aslan, J.E.; Blagoveshchenskaya, A.D.; Thomas, L.; Wan, L.; Xiang, Y.; Feliciangeli, S.F.; Hung, C.-H.; Crump, C.M.; Thomas, G. PACS-2 controls endoplasmic reticulum-mitochondria communication and Bid-mediated apoptosis. EMBO J. 2005, 24, 717–729. [Google Scholar] [CrossRef] [Green Version]
  72. Youker, R.T.; Shinde, U.; Day, R.; Thomas, G. At the crossroads of homoeostasis and disease: Roles of the PACS proteins in membrane traffic and apoptosis. Biochem. J. 2009, 421, 1–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Stone, S.J.; Vance, J.E. Phosphatidylserine synthase-1 and -2 are localized to mitochondria-associated membranes. J. Biol. Chem. 2000, 275, 34534–34540. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Iwasawa, R.; Mahul-Mellier, A.-L.; Datler, C.; Pazarentzos, E.; Grimm, S. Fis1 and Bap31 bridge the mitochondria-ER interface to establish a platform for apoptosis induction. EMBO J. 2011, 30, 556–568. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Pagliuso, A.; Cossart, P.; Stavru, F. The ever-growing complexity of the mitochondrial fission machinery. Cell. Mol. Life Sci. 2018, 75, 355–374. [Google Scholar] [CrossRef] [Green Version]
  76. Shen, Q.; Yamano, K.; Head, B.P.; Kawajiri, S.; Cheung, J.T.M.; Wang, C.; Cho, J.-H.; Hattori, N.; Youle, R.J.; van der Bliek, A.M. Mutations in Fis1 disrupt orderly disposal of defective mitochondria. Mol. Biol. Cell 2014, 25, 145–159. [Google Scholar] [CrossRef] [Green Version]
  77. Yu, R.; Jin, S.; Lendahl, U.; Nistér, M.; Zhao, J. Human Fis1 regulates mitochondrial dynamics through inhibition of the fusion machinery. EMBO J. 2019, 38. [Google Scholar] [CrossRef]
  78. Rojansky, R.; Cha, M.-Y.; Chan, D.C. Elimination of paternal mitochondria in mouse embryos occurs through autophagic degradation dependent on PARKIN and MUL1. Elife 2016, 5. [Google Scholar] [CrossRef]
  79. Namba, T. BAP31 regulates mitochondrial function via interaction with Tom40 within ER-mitochondria contact sites. Sci. Adv. 2019, 5. [Google Scholar] [CrossRef] [Green Version]
  80. Joshi, A.U.; Saw, N.L.; Shamloo, M.; Mochly-Rosen, D. Drp1/Fis1 interaction mediates mitochondrial dysfunction, bioenergetic failure and cognitive decline in Alzheimer’s disease. Oncotarget 2017, 9, 6128–6143. [Google Scholar] [CrossRef] [Green Version]
  81. Wang, T.; Chen, J.; Hou, Y.; Yu, Y.; Wang, B. BAP31 deficiency contributes to the formation of amyloid-β plaques in Alzheimer’s disease by reducing the stability of RTN3. FASEB J. 2019, 33, 4936–4946. [Google Scholar] [CrossRef]
  82. Manczak, M.; Reddy, P.H. Abnormal interaction between the mitochondrial fission protein Drp1 and hyperphosphorylated tau in Alzheimer’s disease neurons: Implications for mitochondrial dysfunction and neuronal damage. Hum. Mol. Genet. 2012, 21, 2538–2547. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Cho, D.-H.; Nakamura, T.; Fang, J.; Cieplak, P.; Godzik, A.; Gu, Z.; Lipton, S.A. S-Nitrosylation of Drp1 Mediates β-Amyloid-Related Mitochondrial Fission and Neuronal Injury. Science 2009, 324, 102–105. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. De Vos, K.J.; Mórotz, G.M.; Stoica, R.; Tudor, E.L.; Lau, K.-F.; Ackerley, S.; Warley, A.; Shaw, C.E.; Miller, C.C.J. VAPB interacts with the mitochondrial protein PTPIP51 to regulate calcium homeostasis. Hum. Mol. Genet. 2012, 21, 1299–1311. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Stoica, R.; Paillusson, S.; Gomez-Suaga, P.; Mitchell, J.C.; Lau, D.H.W.; Gray, E.H.; Sancho, R.M.; Vizcay-Barrena, G.; De Vos, K.J.; Shaw, C.E.; et al. ALS/FTD-associated FUS activates GSK-3β to disrupt the VAPB–PTPIP51 interaction and ER–mitochondria associations. EMBO Rep. 2016, 17, 1326–1342. [Google Scholar] [CrossRef]
  86. Gomez-Suaga, P.; Paillusson, S.; Stoica, R.; Noble, W.; Hanger, D.P.; Miller, C.C.J. The ER-Mitochondria tethering complex VAPB-PTPIP51 regulates autophagy. Curr. Biol. 2017, 27, 371–385. [Google Scholar] [CrossRef] [Green Version]
  87. Galmes, R.; Houcine, A.; Vliet, A.R.; Agostinis, P.; Jackson, C.L.; Giordano, F. ORP5/ORP8 localize to endoplasmic reticulum–mitochondria contacts and are involved in mitochondrial function. EMBO Rep. 2016, 17, 800–810. [Google Scholar] [CrossRef] [Green Version]
  88. Tokutake, Y.; Yamada, K.; Ohata, M.; Obayashi, Y.; Tsuchiya, M.; Yonekura, S. ALS-Linked P56S-VAPB mutation impairs the formation of multinuclear myotube in C2C12 Cells. Int. J. Mol. Sci. 2015, 16, 18628–18641. [Google Scholar] [CrossRef] [Green Version]
  89. Stoica, R.; De Vos, K.J.; Paillusson, S.; Mueller, S.; Sancho, R.M.; Lau, K.-F.; Vizcay-Barrena, G.; Lin, W.-L.; Xu, Y.-F.; Lewis, J.; et al. ER-Mitochondria associations are regulated by the VAPB–PTPIP51 interaction and are disrupted by ALS/FTD-associated TDP-43. Nat. Commun. 2014, 5. [Google Scholar] [CrossRef] [Green Version]
  90. Yamoah, A.; Tripathi, P.; Sechi, A.; Köhler, C.; Guo, H.; Chandrasekar, A.; Nolte, K.W.; Wruck, C.J.; Katona, I.; Anink, J.; et al. Aggregates of RNA binding proteins and ER chaperones linked to exosomes in granulovacuolar degeneration of the Alzheimer’s disease brain. J. Alzheimer’s Dis. JAD 2020, 139–156. [Google Scholar] [CrossRef] [Green Version]
  91. Llorens-Marítin, M.; Jurado, J.; Hernández, F.; Ávila, J. GSK-3β, a pivotal kinase in Alzheimer disease. Front. Mol. Neurosci. 2014, 7. [Google Scholar] [CrossRef] [Green Version]
  92. Lau, D.H.W.; Paillusson, S.; Hartopp, N.; Rupawala, H.; Mórotz, G.M.; Gomez-Suaga, P.; Greig, J.; Troakes, C.; Noble, W.; Miller, C.C.J. Disruption of endoplasmic reticulum-mitochondria tethering proteins in post-mortem Alzheimer’s disease brain. Neurobiol. Dis. 2020, 143, 105020. [Google Scholar] [CrossRef] [PubMed]
  93. Hayashi, T.; Su, T.P. σ-1 receptors (σ1 binding sites) form raft-like microdomains and target lipid droplets on the endoplasmic reticulum: Roles in endoplasmic reticulum lipid compartmentalization and export. J. Pharmacol. Exp. Ther. 2003, 306, 718–725. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Watanabe, S.; Ilieva, H.; Tamada, H.; Nomura, H.; Komine, O.; Endo, F.; Jin, S.; Mancias, P.; Kiyama, H.; Yamanaka, K. Mitochondria-associated membrane collapse is a common pathomechanism in SIGMAR 1- and SOD 1 -linked ALS. EMBO Mol. Med. 2016, 8, 1421–1437. [Google Scholar] [CrossRef] [PubMed]
  95. Hayashi, T.; Su, T.P. The potential role of sigma-1 receptors in lipid transport and lipid raft reconstitution in the brain: Implication for drug abuse. Life Sci. 2005, 77, 1612–1624. [Google Scholar] [CrossRef]
  96. Gorbatyuk, M.; Gorbatyuk, O. The Molecular Chaperone GRP78/BiP as a Therapeutic Target for Neurodegenerative Disorders: A Mini Review. J. Genet. Syndr. Gene Ther. 2013, 4. [Google Scholar] [CrossRef] [Green Version]
  97. Prasad, M.; Pawlak, K.J.; Burak, W.E.; Perry, E.E.; Marshall, B.; Whittal, R.M.; Bose, H.S. Mitochondrial metabolic regulation by GRP78. Sci. Adv. 2017, 3. [Google Scholar] [CrossRef] [Green Version]
  98. Hetz, C.; Papa, F.R. The unfolded protein response and cell fate control. Mol. Cell 2018, 69, 169–181. [Google Scholar] [CrossRef] [Green Version]
  99. Hoozemans, J.J.M.; van Haastert, E.S.; Nijholt, D.A.T.; Rozemuller, A.J.M.; Scheper, W. Activation of the unfolded protein response is an early event in Alzheimer’s and Parkinson’s disease. Neurodegener. Dis. 2012, 10, 212–215. [Google Scholar] [CrossRef]
  100. Uddin, M.S.; Tewari, D.; Sharma, G.; Kabir, M.T.; Barreto, G.E.; Bin-Jumah, M.N.; Perveen, A.; Abdel-Daim, M.M.; Ashraf, G.M. Molecular mechanisms of ER stress and UPR in the pathogenesis of Alzheimer’s disease. Mol. Neurobiol. 2020, 57, 2902–2919. [Google Scholar] [CrossRef]
  101. Katayama, T.; Imaizumi, K.; Sato, N.; Miyoshi, K.; Kudo, T.; Hitomi, J.; Morihara, T.; Yoneda, T.; Gomi, F.; Mori, Y.; et al. Presenilin-1 mutations downregulate the signalling pathway of the unfolded-protein response. Nat. Cell Biol. 1999, 1, 479–485. [Google Scholar] [CrossRef]
  102. Sadleir, K.R.; Popovic, J.; Vassar, R. ER stress is not elevated in the 5XFAD mouse model of Alzheimer’s disease. J. Biol. Chem. 2018, 293, 18434–18443. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Hayashi, T.; Su, T.P. Sigma-1 receptor chaperones at the ER-mitochondrion interface regulate Ca2+ signaling and cell survival. Cell 2007, 131, 596–610. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Hayashi, T.; Su, T.P. Regulating ankyrin dynamics: Roles of sigma-1 receptors. Proc. Natl. Acad. Sci. USA 2001, 98, 491–496. [Google Scholar] [CrossRef] [PubMed]
  105. Morin-Surun, M.P.; Collin, T.; Denavit-Saubie, M.; Baulieu, E.E.; Monnet, F.P. Intracellular σ1 receptor modulates phospholipase C and protein kinase C activities in the brainstem. Proc. Natl. Acad. Sci. USA 1999, 96, 8196–8199. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Tsai, S.Y.; Hayashi, T.; Harvey, B.K.; Wang, Y.; Wu, W.W.; Shen, R.F.; Zhang, Y.; Becker, K.G.; Hoffer, B.J.; Su, T.P. Sigma-1 receptors regulate hippocampal dendritic spine formation via a free radical-sensitive mechanism involving Rac1·GTP pathway. Proc. Natl. Acad. Sci. USA 2009, 106, 22468–22473. [Google Scholar] [CrossRef] [Green Version]
  107. Ryskamp, D.A.; Korban, S.; Zhemkov, V.; Kraskovskaya, N.; Bezprozvanny, I. Neuronal sigma-1 receptors: Signaling functions and protective roles in neurodegenerative diseases. Front. Neurosci. 2019, 13, 1–20. [Google Scholar] [CrossRef]
  108. Delprat, B.; Crouzier, L.; Su, T.-P.; Maurice, T. At the crossing of ER stress and MAMs: A key role of Sigma-1 receptor? Adv. Exp. Med. Biol. 2020, 1131, 699–718. [Google Scholar] [CrossRef]
  109. Jansen, K.L.R.; Faull, R.L.M.; Storey, P.; Leslie, R.A. Loss of sigma binding sites in the CA1 area of the anterior hippocampus in Alzheimer’s disease correlates with CA1 pyramidal cell loss. Brain Res. 1993, 623, 299–302. [Google Scholar] [CrossRef]
  110. Fehér, Á.; Juhász, A.; László, A.; Kálmán, J.; Pákáski, M.; Kálmán, J.; Janka, Z. Association between a variant of the sigma-1 receptor gene and Alzheimer’s disease. Neurosci. Lett. 2012, 517, 136–139. [Google Scholar] [CrossRef]
  111. Huang, Y.; Zheng, L.; Halliday, G.; Dobson-Stone, C.; Wang, Y.; Tang, H.-D.; Cao, L.; Deng, Y.-L.; Wang, G.; Zhang, Y.-M.; et al. Genetic polymorphisms in Sigma-1 receptor and apolipoprotein E interact to influence the severity of Alzheimers disease. Curr. Alzheimer Res. 2011, 8, 765–770. [Google Scholar] [CrossRef]
  112. Fisher, A.; Bezprozvanny, I.; Wu, L.; Ryskamp, D.A.; Bar-Ner, N.; Natan, N.; Brandeis, R.; Elkon, H.; Nahum, V.; Gershonov, E.; et al. AF710B, a Novel M1/sigma1 agonist with therapeutic efficacy in animal models of Alzheimer’s disease. Neurodegener. Dis. 2016, 16, 95–110. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Hall, H.; Iulita, M.F.; Gubert, P.; Flores Aguilar, L.; Ducatenzeiler, A.; Fisher, A.; Cuello, A.C. AF710B, an M1/sigma-1 receptor agonist with long-lasting disease-modifying properties in a transgenic rat model of Alzheimer’s disease. Alzheimers Dement. 2018, 14, 811–823. [Google Scholar] [CrossRef] [PubMed]
  114. Maurice, T.; Strehaiano, M.; Duhr, F.; Chevallier, N. Amyloid toxicity is enhanced after pharmacological or genetic invalidation of the σ1 receptor. Behav. Brain Res. 2018, 339, 1–10. [Google Scholar] [CrossRef] [PubMed]
  115. Lahmy, V.; Meunier, J.; Malmström, S.; Naert, G.; Givalois, L.; Kim, S.H.; Villard, V.; Vamvakides, A.; Maurice, T. Blockade of tau hyperphosphorylation and aβ 1-42 generation by the aminotetrahydrofuran derivative ANAVEX2-73, a mixed muscarinic and σ 1 receptor agonist, in a nontransgenic mouse model of Alzheimer’s disease. Neuropsychopharmacology 2013, 38, 1706–1723. [Google Scholar] [CrossRef]
  116. Lahmy, V.; Long, R.; Morin, D.; Villard, V.; Maurice, T. Mitochondrial protection by the mixed muscarinic/σ1 ligand ANAVEX2-73, a tetrahydrofuran derivative, in Aβ25–35 peptide-injected mice, a nontransgenic Alzheimer’s disease model. Front. Cell. Neurosci. 2015, 8, 1–11. [Google Scholar] [CrossRef] [Green Version]
  117. Tsai, S.Y.A.; Pokrass, M.J.; Klauer, N.R.; Nohara, H.; Su, T.P. Sigma-1 receptor regulates Tau phosphorylation and axon extension by shaping p35 turnover via myristic acid. Proc. Natl. Acad. Sci. USA 2015, 112, 6742–6747. [Google Scholar] [CrossRef] [Green Version]
  118. Verfaillie, T.; Rubio, N.; Garg, A.D.; Bultynck, G.; Rizzuto, R.; Decuypere, J.P.; Piette, J.; Linehan, C.; Gupta, S.; Samali, A.; et al. PERK is required at the ER-mitochondrial contact sites to convey apoptosis after ROS-based ER stress. Cell Death Differ. 2012, 19, 1880–1891. [Google Scholar] [CrossRef] [Green Version]
  119. Munoz, J.P.; Ivanova, S.; Sanchez-Wandelmer, J.; Martinez-Cristobal, P.; Noguera, E.; Sancho, A.; Diaz-Ramos, A.; Hernandez-Alvarez, M.I.; Sebastian, D.; Mauvezin, C.; et al. Mfn2 modulates the UPR and mitochondrial function via repression of PERK. EMBO J. 2013, 32, 2348–2361. [Google Scholar] [CrossRef] [Green Version]
  120. Toyofuku, T.; Okamoto, Y.; Ishikawa, T.; Sasawatari, S.; Kumanogoh, A. LRRK2 regulates endoplasmic reticulum-mitochondrial tethering through the PERK-mediated ubiquitination pathway. EMBO J. 2020, 39, e100875. [Google Scholar] [CrossRef]
  121. Hoozemans, J.J.; Veerhuis, R.; Van Haastert, E.S.; Rozemuller, J.M.; Baas, F.; Eikelenboom, P.; Scheper, W. The unfolded protein response is activated in Alzheimer’s disease. Acta Neuropathol. 2005, 110, 165–172. [Google Scholar] [CrossRef]
  122. Hoozemans, J.J.; van Haastert, E.S.; Nijholt, D.A.; Rozemuller, A.J.; Eikelenboom, P.; Scheper, W. The unfolded protein response is activated in pretangle neurons in Alzheimer’s disease hippocampus. Am. J. Pathol. 2009, 174, 1241–1251. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Devi, L.; Ohno, M. PERK mediates eIF2alpha phosphorylation responsible for BACE1 elevation, CREB dysfunction and neurodegeneration in a mouse model of Alzheimer’s disease. Neurobiol. Aging 2014, 35, 2272–2281. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Ma, T.; Trinh, M.A.; Wexler, A.J.; Bourbon, C.; Gatti, E.; Pierre, P.; Cavener, D.R.; Klann, E. Suppression of eIF2alpha kinases alleviates Alzheimer’s disease-related plasticity and memory deficits. Nat. Neurosci. 2013, 16, 1299–1305. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Page, G.; Rioux Bilan, A.; Ingrand, S.; Lafay-Chebassier, C.; Pain, S.; Perault Pochat, M.C.; Bouras, C.; Bayer, T.; Hugon, J. Activated double-stranded RNA-dependent protein kinase and neuronal death in models of Alzheimer’s disease. Neuroscience 2006, 139, 1343–1354. [Google Scholar] [CrossRef]
  126. Abisambra, J.F.; Jinwal, U.K.; Blair, L.J.; O’Leary, J.C., 3rd; Li, Q.; Brady, S.; Wang, L.; Guidi, C.E.; Zhang, B.; Nordhues, B.A.; et al. Tau accumulation activates the unfolded protein response by impairing endoplasmic reticulum-associated degradation. J. Neurosci. 2013, 33, 9498–9507. [Google Scholar] [CrossRef] [Green Version]
  127. Radford, H.; Moreno, J.A.; Verity, N.; Halliday, M.; Mallucci, G.R. PERK inhibition prevents tau-mediated neurodegeneration in a mouse model of frontotemporal dementia. Acta Neuropathol. 2015, 130, 633–642. [Google Scholar] [CrossRef] [Green Version]
  128. Segev, Y.; Barrera, I.; Ounallah-Saad, H.; Wibrand, K.; Sporild, I.; Livne, A.; Rosenberg, T.; David, O.; Mints, M.; Bramham, C.R.; et al. PKR inhibition rescues memory deficit and ATF4 overexpression in ApoE epsilon4 human replacement mice. J. Neurosci. 2015, 35, 12986–12993. [Google Scholar] [CrossRef] [Green Version]
  129. O’Connor, T.; Sadleir, K.R.; Maus, E.; Velliquette, R.A.; Zhao, J.; Cole, S.L.; Eimer, W.A.; Hitt, B.; Bembinster, L.A.; Lammich, S.; et al. Phosphorylation of the translation initiation factor eIF2alpha increases BACE1 levels and promotes amyloidogenesis. Neuron 2008, 60, 988–1009. [Google Scholar] [CrossRef] [Green Version]
  130. Yang, W.; Zhou, X.; Zimmermann, H.R.; Cavener, D.R.; Klann, E.; Ma, T. Repression of the eIF2alpha kinase PERK alleviates mGluR-LTD impairments in a mouse model of Alzheimer’s disease. Neurobiol. Aging 2016, 41, 19–24. [Google Scholar] [CrossRef] [Green Version]
  131. Costa-Mattioli, M.; Gobert, D.; Stern, E.; Gamache, K.; Colina, R.; Cuello, C.; Sossin, W.; Kaufman, R.; Pelletier, J.; Rosenblum, K.; et al. eIF2alpha phosphorylation bidirectionally regulates the switch from short- to long-term synaptic plasticity and memory. Cell 2007, 129, 195–206. [Google Scholar] [CrossRef] [Green Version]
  132. Baleriola, J.; Walker, C.A.; Jean, Y.Y.; Crary, J.F.; Troy, C.M.; Nagy, P.L.; Hengst, U. Axonally synthesized ATF4 transmits a neurodegenerative signal across brain regions. Cell 2014, 158, 1159–1172. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. DuBoff, B.; Götz, J.; Feany, M.B. Tau promotes neurodegeneration via DRP1 mislocalization In Vivo. Neuron 2012, 75, 618–632. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Biswas, J.; Gupta, S.; Verma, D.K.; Gupta, P.; Singh, A.; Tiwari, S.; Goswami, P.; Sharma, S.; Singh, S. Involvement of glucose related energy crisis and endoplasmic reticulum stress: Insinuation of streptozotocin induced Alzheimer’s like pathology. Cell. Signal. 2018, 42, 211–226. [Google Scholar] [CrossRef] [PubMed]
  135. Cui, W.; Wang, S.; Wang, Z.; Wang, Z.; Sun, C.; Zhang, Y. Inhibition of PTEN attenuates endoplasmic reticulum stress and apoptosis via activation of PI3K/AKT pathway in Alzheimer’s disease. Neurochem. Res. 2017, 42, 3052–3060. [Google Scholar] [CrossRef] [PubMed]
  136. Lin, L.; Liu, G.; Yang, L. Crocin improves cognitive behavior in rats with Alzheimer’s disease by regulating endoplasmic reticulum stress and apoptosis. Biomed. Res. Int. 2019, 2019. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Reinhardt, S.; Schuck, F.; Grosgen, S.; Riemenschneider, M.; Hartmann, T.; Postina, R.; Grimm, M.; Endres, K. Unfolded protein response signaling by transcription factor XBP-1 regulates ADAM10 and is affected in Alzheimer’s disease. FASEB J. 2014, 28, 978–997. [Google Scholar] [CrossRef]
  138. Gerakis, Y.; Dunys, J.; Bauer, C.; Checler, F. Aβ42 oligomers modulate β-secretase through an XBP-1s-dependent pathway involving HRD1. Sci. Rep. 2016, 6, 37436. [Google Scholar] [CrossRef] [Green Version]
  139. Bussiere, R.; Oules, B.; Mary, A.; Vaillant-Beuchot, L.; Martin, C.; El Manaa, W.; Vallee, D.; Duplan, E.; Paterlini-Brechot, P.; Alves Da Costa, C.; et al. Upregulation of the sarco-endoplasmic reticulum calcium ATPase 1 truncated isoform plays a pathogenic role in Alzheimer’s disease. Cells 2019, 8, 1539. [Google Scholar] [CrossRef] [Green Version]
  140. Mori, T.; Hayashi, T.; Hayashi, E.; Su, T.P. Sigma-1 receptor chaperone at the ER-Mitochondrion interface mediates the mitochondrion-ER-nucleus signaling for cellular survival. PLoS ONE 2013, 8. [Google Scholar] [CrossRef] [Green Version]
  141. Takeda, K.; Nagashima, S.; Shiiba, I.; Uda, A.; Tokuyama, T.; Ito, N.; Fukuda, T.; Matsushita, N.; Ishido, S.; Iwawaki, T.; et al. MITOL prevents ER stress-induced apoptosis by IRE1alpha ubiquitylation at ER-mitochondria contact sites. EMBO J. 2019, 38, e100999. [Google Scholar] [CrossRef]
  142. Son, S.M.; Byun, J.; Roh, S.E.; Kim, S.J.; Mook-Jung, I. Reduced IRE1alpha mediates apoptotic cell death by disrupting calcium homeostasis via the InsP3 receptor. Cell Death Dis. 2014, 5, e1188. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Urano, F.; Wang, X.; Bertolotti, A.; Zhang, Y.; Chung, P.; Harding, H.P.; Ron, D. Coupling of stress in the ER to activation of JNK protein kinases by transmembrane protein kinase IRE1. Science 2000, 287, 664–666. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Nishitoh, H.; Matsuzawa, A.; Tobiume, K.; Saegusa, K.; Takeda, K.; Inoue, K.; Hori, S.; Kakizuka, A.; Ichijo, H. ASK1 is essential for endoplasmic reticulum stress-induced neuronal cell death triggered by expanded polyglutamine repeats. Genes Dev. 2002, 16, 1345–1355. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Lee, J.H.; Won, S.M.; Suh, J.; Son, S.J.; Moon, G.J.; Park, U.J.; Gwag, B.J. Induction of the unfolded protein response and cell death pathway in Alzheimer’s disease, but not in aged Tg2576 mice. Exp. Mol. Med. 2010, 42, 386–394. [Google Scholar] [CrossRef] [Green Version]
  146. Cissé, M.; Duplan, E.; Lorivel, T.; Dunys, J.; Bauer, C.; Meckler, X.; Gerakis, Y.; Lauritzen, I.; Checler, F. The transcription factor XBP1s restores hippocampal synaptic plasticity and memory by control of the Kalirin-7 pathway in Alzheimer model. Mol. Psychiatry 2017, 22, 1562–1575. [Google Scholar] [CrossRef]
  147. Casas-Tinto, S.; Zhang, Y.; Sanchez-Garcia, J.; Gomez-Velazquez, M.; Rincon-Limas, D.E.; Fernandez-Funez, P. The ER stress factor XBP1s prevents amyloid-beta neurotoxicity. Hum. Mol. Genet. 2011, 20, 2144–2160. [Google Scholar] [CrossRef] [Green Version]
  148. Waldherr, S.M.; Strovas, T.J.; Vadset, T.A.; Liachko, N.F.; Kraemer, B.C. Constitutive XBP-1s-mediated activation of the endoplasmic reticulum unfolded protein response protects against pathological tau. Nat. Commun. 2019, 10, 4443. [Google Scholar] [CrossRef]
  149. Acosta-Alvear, D.; Zhou, Y.; Blais, A.; Tsikitis, M.; Lents, N.H.; Arias, C.; Lennon, C.J.; Kluger, Y.; Dynlacht, B.D. XBP1 controls diverse cell type- and condition-specific transcriptional regulatory networks. Mol. Cell 2007, 27, 53–66. [Google Scholar] [CrossRef]
  150. Chami, M.; Gozuacik, D.; Saigo, K.; Capiod, T.; Falson, P.; Lecoeur, H.; Urashima, T.; Beckmann, J.; Gougeon, M.L.; Claret, M.; et al. Hepatitis B virus-related insertional mutagenesis implicates SERCA1 gene in the control of apoptosis. Oncogene 2000, 19, 2877–2886. [Google Scholar] [CrossRef] [Green Version]
  151. Chami, M.; Gozuacik, D.; Lagorce, D.; Brini, M.; Falson, P.; Peaucellier, G.; Pinton, P.; Lecoeur, H.; Gougeon, M.L.; le Maire, M.; et al. SERCA1 truncated proteins unable to pump calcium reduce the endoplasmic reticulum calcium concentration and induce apoptosis. J. Cell Biol. 2001, 153, 1301–1314. [Google Scholar] [CrossRef] [Green Version]
  152. Chami, M.; Oules, B.; Szabadkai, G.; Tacine, R.; Rizzuto, R.; Paterlini-Brechot, P. Role of SERCA1 truncated isoform in the proapoptotic calcium transfer from ER to mitochondria during ER stress. Mol. Cell 2008, 32, 641–651. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Xu, Q.; Bernardo, A.; Walker, D.; Kanegawa, T.; Mahley, R.W.; Huang, Y. Profile and regulation of apolipoprotein E (ApoE) expression in the CNS in mice with targeting of green fluorescent protein gene to the ApoE locus. J. Neurosci. 2006, 26, 4985–4994. [Google Scholar] [CrossRef] [PubMed]
  154. Krasemann, S.; Madore, C.; Cialic, R.; Baufeld, C.; Calcagno, N.; El Fatimy, R.; Beckers, L.; O’Loughlin, E.; Xu, Y.; Fanek, Z.; et al. The TREM2-APOE pathway drives the transcriptional phenotype of dysfunctional microglia in neurodegenerative diseases. Immunity 2017, 47, 566–581.e9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Mahley, R.W. Apolipoprotein E: Cholesterol transport protein with expanding role in cell biology. Science 1988, 240, 622–630. [Google Scholar] [CrossRef] [PubMed]
  156. Lane-Donovan, C.; Wong, W.M.; Durakoglugil, M.S.; Wasser, C.R.; Jiang, S.; Xian, X.; Herz, J. Genetic restoration of plasma apoe improves cognition and partially restores synaptic defects in ApoE-deficient mice. J. Neurosci. 2016, 36, 10141–10150. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Neu, S.C.; Pa, J.; Kukull, W.; Beekly, D.; Kuzma, A.; Gangadharan, P.; Wang, L.-S.; Romero, K.; Arneric, S.P.; Redolfi, A.; et al. Apolipoprotein E genotype and sex risk factors for Alzheimer disease: A meta-analysis. JAMA Neurol. 2017, 74, 1178–1189. [Google Scholar] [CrossRef]
  158. Mahley, R.W.; Huang, Y. Apolipoprotein E sets the stage: Response to injury triggers neuropathology. Neuron 2012, 76, 871–885. [Google Scholar] [CrossRef] [Green Version]
  159. Castellano, J.M.; Kim, J.; Stewart, F.R.; Jiang, H.; DeMattos, R.B.; Patterson, B.W.; Fagan, A.M.; Morris, J.C.; Mawuenyega, K.G.; Cruchaga, C.; et al. Human apoE isoforms differentially regulate brain amyloid-β peptide clearance. Sci. Transl. Med. 2011, 3. [Google Scholar] [CrossRef] [Green Version]
  160. Simonovitch, S.; Schmukler, E.; Masliah, E.; Pinkas-Kramarski, R.; Michaelson, D.M. The Effects of APOE4 on mitochondrial dynamics and proteins in vivo. J. Alzheimer’s Dis. 2019, 70, 861–875. [Google Scholar] [CrossRef]
  161. Orr, A.L.; Kim, C.; Jimenez-Morales, D.; Newton, B.W.; Johnson, J.R.; Krogan, N.J.; Swaney, D.L.; Mahley, R.W. Neuronal apolipoprotein E4 expression results in proteome-wide alterations and compromises bioenergetic capacity by disrupting mitochondrial function. J. Alzheimer’s Dis. 2019, 68, 991–1011. [Google Scholar] [CrossRef] [Green Version]
  162. Hayashi, T.; Rizzuto, R.; Hajnoczky, G.; Su, T.-P. MAM: More than just a housekeeper. Trends Cell Biol. 2009, 19, 81–88. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Tambini, M.D.; Pera, M.; Kanter, E.; Yang, H.; Guardia-Laguarta, C.; Holtzman, D.; Sulzer, D.; Area-Gomez, E.; Schon, E.A. ApoE4 upregulates the activity of mitochondria-associated ER membranes. EMBO Rep. 2016, 17, 27–36. [Google Scholar] [CrossRef] [PubMed]
  164. Oikawa, N.; Walter, J. Presenilins and γ-secretase in membrane proteostasis. Cells 2019, 8, 209. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Neely, K.M.; Green, K.N.; LaFerla, F.M. Presenilin is necessary for efficient proteolysis through the autophagy-lysosome system in a γ-secretase-independent manner. J. Neurosci. 2011, 31, 2781–2791. [Google Scholar] [CrossRef] [Green Version]
  166. Száraz, P.; Bánhegyi, G.; Marcolongo, P.; Benedetti, A. Transient knockdown of presenilin-1 provokes endoplasmic reticulum stress related formation of autophagosomes in HepG2 cells. Arch. Biochem. Biophys. 2013, 538, 57–63. [Google Scholar] [CrossRef]
  167. Bezprozvanny, I.; Mattson, M.P. Neuronal calcium mishandling and the pathogenesis of Alzheimer’s disease. Trends Neurosci. 2008, 31, 454–463. [Google Scholar] [CrossRef] [Green Version]
  168. Tu, H.; Nelson, O.; Bezprozvanny, A.; Wang, Z.; Lee, S.F.; Hao, Y.H.; Serneels, L.; De Strooper, B.; Yu, G.; Bezprozvanny, I. Presenilins form ER Ca2+ leak channels, a function disrupted by familial Alzheimer’s disease-linked mutations. Cell 2006, 126, 981–993. [Google Scholar] [CrossRef] [Green Version]
  169. Green, K.N.; Demuro, A.; Akbari, Y.; Hitt, B.D.; Smith, I.F.; Parker, I.; LaFerla, F.M. SERCA pump activity is physiologically regulated by presenilin and regulates amyloid β production. J. Cell Biol. 2008, 181, 1107–1116. [Google Scholar] [CrossRef] [Green Version]
  170. Stutzmann, G.E.; Caccamo, A.; LaFerla, F.M.; Parker, I. Dysregulated IP3 signaling in cortical neurons of knock-in mice expressing an Alzheimer’s-linked mutation in Presenilin1 results in exaggerated Ca2+ signals and altered membrane excitability. J. Neurosci. 2004, 24, 508–513. [Google Scholar] [CrossRef] [Green Version]
  171. Area-Gomez, E.; De Groof, A.J.C.; Boldogh, I.; Bird, T.D.; Gibson, G.E.; Koehler, C.M.; Yu, W.H.; Duff, K.E.; Yaffe, M.P.; Pon, L.A.; et al. Presenilins are enriched in endoplasmic reticulum membranes associated with mitochondria. Am. J. Pathol. 2009, 175, 1810–1816. [Google Scholar] [CrossRef] [Green Version]
  172. Zampese, E.; Fasolato, C.; Kipanyula, M.J.; Bortolozzi, M.; Pozzan, T.; Pizzo, P. Presenilin 2 modulates endoplasmic reticulum (ER)-mitochondria interactions and Ca2+ cross-talk. Proc. Natl. Acad. Sci. USA 2011, 108, 2777–2782. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Kipanyula, M.J.; Contreras, L.; Zampese, E.; Lazzari, C.; Wong, A.K.C.; Pizzo, P.; Fasolato, C.; Pozzan, T. Ca 2+ dysregulation in neurons from transgenic mice expressing mutant presenilin 2. Aging Cell 2012, 11, 885–893. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Fedeli, C.; Filadi, R.; Rossi, A.; Mammucari, C.; Pizzo, P. PSEN2 (presenilin 2) mutants linked to familial Alzheimer disease impair autophagy by altering Ca2+ homeostasis. Autophagy 2019, 15, 2044–2062. [Google Scholar] [CrossRef] [PubMed]
  175. Nelson, O.; Tu, H.; Lei, T.; Bentahir, M.; De Strooper, B.; Bezprozvanny, I. Familial Alzheimer disease-linked mutations specifically disrupt Ca 2+ leak function of presenilin 1. J. Clin. Investig. 2007, 117, 1230–1239. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Sepulveda-Falla, D.; Barrera-Ocampo, A.; Hagel, C.; Korwitz, A.; Vinueza-Veloz, M.F.; Zhou, K.; Schonewille, M.; Zhou, H.; Velazquez-Perez, L.; Rodriguez-Labrada, R.; et al. Familial Alzheimer’s disease-associated presenilin-1 alters cerebellar activity and calcium homeostasis. J. Clin. Investig. 2014, 124, 1552–1567. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Lee, J.H.; Yu, W.H.; Kumar, A.; Lee, S.; Mohan, P.S.; Peterhoff, C.M.; Wolfe, D.M.; Martinez-Vicente, M.; Massey, A.C.; Sovak, G.; et al. Lysosomal proteolysis and autophagy require presenilin 1 and are disrupted by Alzheimer-related PS1 mutations. Cell 2010, 141, 1146–1158. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  178. Müller, U.C.; Zheng, H. Physiological functions of APP family proteins. Cold Spring Harb. Perspect. Med. 2012, 2, 1–17. [Google Scholar] [CrossRef] [Green Version]
  179. Anandatheerthavarada, H.K.; Devi, L. Mitochondrial translocation of amyloid precursor protein and its cleaved products: Relevance to mitochondrial dysfunction in Alzheimer’s disease. Rev. Neurosci. 2007, 18, 343–354. [Google Scholar] [CrossRef]
  180. Anandatheerthavarada, H.K.; Biswas, G.; Robin, M.A.; Avadhani, N.G. Mitochondrial targeting and a novel transmembrane arrest of Alzheimer’s amyloid precursor protein impairs mitochondrial function in neuronal cells. J. Cell Biol. 2003, 161, 41–54. [Google Scholar] [CrossRef]
  181. Devi, L.; Prabhu, B.M.; Galati, D.F.; Avadhani, N.G.; Anandatheerthavarada, H.K. Accumulation of amyloid precursor protein in the mitochondrial import channels of human Alzheimer’s disease brain is associated with mitochondrial dysfunction. J. Neurosci. 2006, 26, 9057–9068. [Google Scholar] [CrossRef] [Green Version]
  182. Pagani, L.; Eckert, A. Amyloid-beta interaction with mitochondria. Int. J. Alzheimer’s Dis. 2011, 2011. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Pavlov, P.F.; Wiehager, B.; Sakai, J.; Frykman, S.; Behbahani, H.; Winblad, B.; Ankarcrona, M. Mitochondrial γ-secretase participates in the metabolism of mitochondria-associated amyloid precursor protein. FASEB J. 2011, 25, 78–88. [Google Scholar] [CrossRef] [PubMed]
  184. Devi, L.; Ohno, M. Mitochondrial dysfunction and accumulation of the β-secretase-cleaved C-terminal fragment of APP in Alzheimer’s disease transgenic mice. Neurobiol. Dis. 2012, 45, 417–424. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Schreiner, B.; Hedskog, L.; Wiehager, B.; Ankarcrona, M. Amyloid-β peptides are generated in mitochondria-associated endoplasmic reticulum membranes. J. Alzheimer’s Dis. 2015, 43, 369–374. [Google Scholar] [CrossRef] [PubMed]
  186. Pera, M.; Larrea, D.; Guardia-Laguarta, C.; Montesinos, J.; Velasco, K.R.; Agrawal, R.R.; Xu, Y.; Chan, R.B.; Di Paolo, G.; Mehler, M.F.; et al. Increased localization of APP -C99 in mitochondria-associated ER membranes causes mitochondrial dysfunction in Alzheimer disease. EMBO J. 2017, 36, 3356–3371. [Google Scholar] [CrossRef] [PubMed]
  187. Del Prete, D.; Suski, J.M.; Oulès, B.; Debayle, D.; Gay, A.S.; Lacas-Gervais, S.; Bussiere, R.; Bauer, C.; Pinton, P.; Paterlini-Bréchot, P.; et al. Localization and processing of the amyloid-β protein precursor in mitochondria-associated membranes. J. Alzheimer’s Dis. 2017, 55, 1549–1570. [Google Scholar] [CrossRef] [Green Version]
  188. Hansson, C.A.; Frykman, S.; Farmery, M.R.; Tjernberg, L.O.; Nilsberth, C.; Pursglove, S.E.; Ito, A.; Winblad, B.; Cowburn, R.F.; Thyberg, J.; et al. Nicastrin, presenilin, APH-1, and PEN-2 form active γ-secretase complexes in mitochondria. J. Biol. Chem. 2004, 279, 51654–51660. [Google Scholar] [CrossRef] [Green Version]
  189. Ankarcrona, M.; Hultenby, K. Presenilin-1 is located in rat mitochondria. Biochem. Biophys. Res. Commun. 2002, 295, 766–770. [Google Scholar] [CrossRef]
  190. Area-Gomez, E.; Del Carmen Lara Castillo, M.; Tambini, M.D.; Guardia-Laguarta, C.; De Groof, A.J.C.; Madra, M.; Ikenouchi, J.; Umeda, M.; Bird, T.D.; Sturley, S.L.; et al. Upregulated function of mitochondria-associated ER membranes in Alzheimer disease. EMBO J. 2012, 31, 4106–4123. [Google Scholar] [CrossRef] [Green Version]
  191. Marquer, C.; Devauges, V.; Cossec, J.C.; Liot, G.; Lécart, S.; Saudou, F.; Duyckaerts, C.; Lévêque-Fort, S.; Potier, M.C. Local cholesterol increase triggers amyloid precursor protein-Bace1 clustering in lipid rafts and rapid endocytosis. FASEB J. 2011, 25, 1295–1305. [Google Scholar] [CrossRef]
  192. Cordy, J.M.; Hussain, I.; Dingwall, C.; Hooper, N.M.; Turner, A.J. Exclusively targeting β-secretase to lipid rafts by GPI-anchor addition up-regulates β-site processing of the amyloid precursor protein. Proc. Natl. Acad. Sci. USA 2003, 100, 11735–11740. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Vetrivel, K.S.; Cheng, H.; Lin, W.; Sakurai, T.; Li, T.; Nukina, N.; Wong, P.C.; Xu, H.; Thinakaran, G. Association of γ-secretase with lipid rafts in post-golgi and endosome membranes. J. Biol. Chem. 2004, 279, 44945–44954. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Hansson Petersen, C.A.; Alikhani, N.; Behbahani, H.; Wiehager, B.; Pavlov, P.F.; Alafuzoff, I.; Leinonen, V.; Ito, A.; Winblad, B.; Glaser, E.; et al. The amyloid-peptide is imported into mitochondria via the TOM import machinery and localized to mitochondrial cristae. Proc. Natl. Acad. Sci. USA 2008, 105, 13145–13150. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Vargas, T.; Ugalde, C.; Spuch, C.; Antequera, D.; Morán, M.J.; Martín, M.A.; Ferrer, I.; Bermejo-Pareja, F.; Carro, E. Aβ accumulation in choroid plexus is associated with mitochondrial-induced apoptosis. Neurobiol. Aging 2010, 31, 1569–1581. [Google Scholar] [CrossRef] [PubMed]
  196. Bobba, A.; Amadoro, G.; Valenti, D.; Corsetti, V.; Lassandro, R.; Atlante, A. Mitochondrial respiratory chain Complexes I and IV are impaired by β-amyloid via direct interaction and through Complex I-dependent ROS production, respectively. Mitochondrion 2013, 13, 298–311. [Google Scholar] [CrossRef] [PubMed]
  197. Lin, M.T.; Beal, M.F. Alzheimer’s APP mangles mitochondria. Nat. Med. 2006, 12, 1241–1243. [Google Scholar] [CrossRef]
  198. Pavlov, P.F.; Petersen, C.H.; Glaser, E.; Ankarcrona, M. Mitochondrial accumulation of APP and Aβ: Significance for Alzheimer disease pathogenesis. J. Cell. Mol. Med. 2009, 13, 4137–4145. [Google Scholar] [CrossRef] [Green Version]
  199. Agrawal, R.R.; Montesinos, J.; Larrea, D.; Area-Gomez, E.; Pera, M. The silence of the fats: A MAM’s story about Alzheimer. Neurobiol. Dis. 2020, 145, 105062. [Google Scholar] [CrossRef]
  200. Kogot-Levin, A.; Saada, A. Ceramide and the mitochondrial respiratory chain. Biochimie 2014, 100, 88–94. [Google Scholar] [CrossRef]
  201. Montesinos, J.; Pera, M.; Larrea, D.; Guardia-Laguarta, C.; Agrawal, R.R.; Velasco, K.R.; Yun, T.D.; Stavrovskaya, I.G.; Xu, Y.; Koo, S.Y.; et al. The Alzheimer’s disease-associated C99 fragment of APP regulates cellular cholesterol trafficking. EMBO J. 2020, 39. [Google Scholar] [CrossRef]
  202. Vaillant-Beuchot, L.; Mary, A.; Pardossi-Piquard, R.; Bourgeois, A.; Lauritzen, I.; Eysert, F.; Kinoshita, P.F.; Cazareth, J.; Badot, C.; Fragaki, K.; et al. Accumulation of amyloid precursor protein C-terminal fragments triggers mitochondrial structure, function, and mitophagy defects in Alzheimer’s disease models and human brains. Acta Neuropathol. 2020. [Google Scholar] [CrossRef]
  203. Ittner, L.M.; Ke, Y.D.; Delerue, F.; Bi, M.; Gladbach, A.; van Eersel, J.; Wölfing, H.; Chieng, B.C.; Christie, M.J.; Napier, I.A.; et al. Dendritic function of tau mediates amyloid-β toxicity in alzheimer’s disease mouse models. Cell 2010, 142, 387–397. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Arendt, T.; Stieler, J.T.; Holzer, M. Tau and tauopathies. Brain Res. Bull. 2016, 126, 238–292. [Google Scholar] [CrossRef] [PubMed]
  205. Roy, S.; Zhang, B.; Lee, V.M.Y.; Trojanowski, J.Q. Axonal transport defects: A common theme in neurodegenerative diseases. Acta Neuropathol. 2005, 109, 5–13. [Google Scholar] [CrossRef] [PubMed]
  206. David, D.C.; Hauptmann, S.; Scherping, I.; Schuessel, K.; Keil, U.; Rizzu, P.; Ravid, R.; Dröse, S.; Brandt, U.; Müller, W.E.; et al. Proteomic and functional analyses reveal a mitochondrial dysfunction in P301L tau transgenic mice. J. Biol. Chem. 2005, 280, 23802–23814. [Google Scholar] [CrossRef] [Green Version]
  207. Cummins, N.; Tweedie, A.; Zuryn, S.; Bertran-Gonzalez, J.; Götz, J. Disease-associated tau impairs mitophagy by inhibiting Parkin translocation to mitochondria. EMBO J. 2019, 38. [Google Scholar] [CrossRef]
  208. Li, X.C.; Hu, Y.; Wang, Z.H.; Luo, Y.; Zhang, Y.; Liu, X.P.; Feng, Q.; Wang, Q.; Ye, K.; Liu, G.P.; et al. Human wild-type full-length tau accumulation disrupts mitochondrial dynamics and the functions via increasing mitofusins. Sci. Rep. 2016, 6, 1–10. [Google Scholar] [CrossRef] [Green Version]
  209. Perreault, S.; Bousquet, O.; Lauzon, M.; Paiement, J.; Leclerc, N. Increased association between rough endoplasmic reticulum membranes and mitochondria in transgenic mice that express P301L Tau. J. Neuropathol. Exp. Neurol. 2009, 68, 503–514. [Google Scholar] [CrossRef] [Green Version]
  210. Ebneth, A.; Godemann, R.; Stamer, K.; Illenberger, S.; Trinczek, B.; Mandelkow, E.-M.; Mandelkow, E. Overexpression of Tau protein inhibits kinesin-dependent trafficking of vesicles, mitochondria, and endoplasmic reticulum: Implications for Alzheimer’s disease. J. Cell Biol. 1998, 143, 777–794. [Google Scholar] [CrossRef] [Green Version]
  211. Afreen, S.; Riherd Methner, D.N.; Ferreira, A. Tau45-230 association with the cytoskeleton and membrane-bound organelles: Functional implications in neurodegeneration. Neuroscience 2017, 362, 104–117. [Google Scholar] [CrossRef]
  212. Yoshiyama, Y.; Higuchi, M.; Zhang, B.; Huang, S.M.; Iwata, N.; Saido, T.C.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] [PubMed] [Green Version]
  213. Gong, T.; Wang, Q.; Lin, Z.; Chen, M.L.; Sun, G.Z. Endoplasmic reticulum (ER) stress inhibitor salubrinal protects against ceramide-induced SH-SY5Y cell death. Biochem. Biophys. Res. Commun. 2012, 427, 461–465. [Google Scholar] [CrossRef] [PubMed]
  214. Lee, D.Y.; Lee, K.S.; Lee, H.J.; Kim, D.H.; Noh, Y.H.; Yu, K.; Jung, H.Y.; Lee, S.H.; Lee, J.Y.; Youn, Y.C.; et al. Activation of PERK signaling attenuates Aβ-mediated ER stress. PLoS ONE 2010, 5, e10489. [Google Scholar] [CrossRef] [PubMed]
  215. Huang, X.; Chen, Y.; Zhang, H.; Ma, Q.; Zhang, Y.; Xu, H. Salubrinal attenuates β-amyloid-induced neuronal death and microglial activation by inhibition of the NF-κB pathway. Neurobiol. Aging 2012, 33, 1007.e9–1007.e17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  216. Barreda-Manso, M.A.; Yanguas-Casás, N.; Nieto-Sampedro, M.; Romero-Ramírez, L. Salubrinal inhibits the expression of proteoglycans and favors neurite outgrowth from cortical neurons in vitro. Exp. Cell Res. 2015, 335, 82–90. [Google Scholar] [CrossRef]
  217. Paris, D.; Ganey, N.J.; Laporte, V.; Patel, N.S.; Beaulieu-Abdelahad, D.; Bachmeier, C.; March, A.; Ait-Ghezala, G.; Mullan, M.J. Reduction of β-amyloid pathology by celastrol in a transgenic mouse model of Alzheimer’s disease. J. Neuroinflamm. 2010, 7, 1–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  218. Ho, S.W.; Tsui, Y.T.C.; Wong, T.T.; Cheung, S.K.K.; Goggins, W.B.; Yi, L.M.; Cheng, K.K.; Baum, L. Effects of 17-allylamino-17-demethoxygeldanamycin (17-AAG) in transgenic mouse models of frontotemporal lobar degeneration and Alzheimer’s disease. Transl. Neurodegener. 2013, 2, 1–9. [Google Scholar] [CrossRef] [Green Version]
  219. Twaroski, Y.; Zaja, C.; Bosnjak, B. Altered Mitochondrial dynamics contributes to propofol-induced cell death in human stem cell-derived neurons. Physiol. Behav. 2015, 123, 1067–1083. [Google Scholar] [CrossRef] [Green Version]
  220. Xie, N.; Wang, C.; Lian, Y.; Wu, C.; Zhang, H.; Zhang, Q. Inhibition of mitochondrial fission attenuates Aβ-induced microglia apoptosis. Neuroscience 2014, 256, 36–42. [Google Scholar] [CrossRef]
  221. Reddy, P.H.; Manczak, M.; Yin, X. Mitochondria-Division inhibitor 1 protects against amyloid-β induced mitochondrial fragmentation and synaptic damage in Alzheimer’s disease. J. Alzheimer’s Dis. 2017, 58, 147–162. [Google Scholar] [CrossRef] [Green Version]
  222. Yeon, J.Y.; Min, S.H.; Park, H.J.; Kim, J.W.; Lee, Y.H.; Park, S.Y.; Jeong, P.S.; Park, H.; Lee, D.S.; Kim, S.U.; et al. Mdivi-1, mitochondrial fission inhibitor, impairs developmental competence and mitochondrial function of embryos and cells in pigs. J. Reprod. Dev. 2015, 61, 81–89. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  223. Guo, X.; Disatnik, M.; Monbureau, M.; Shamloo, M.; Mochly-rosen, D.; Qi, X. Inhibition of mitochondrial fragmentation diminishes Huntington’s disease-associated neurodegeneration. J. Clin. Investig. 2013, 123, 5371–5388. [Google Scholar] [CrossRef] [PubMed]
  224. Arismendi-Morillo, G.; Castellano-Ramírez, A.; Seyfried, T.N. Ultrastructural characterization of the Mitochondria-associated membranes abnormalities in human astrocytomas: Functional and therapeutics implications. Ultrastruct. Pathol. 2017, 41, 234–244. [Google Scholar] [CrossRef] [PubMed]
  225. Lynes, E.M.; Bui, M.; Yap, M.C.; Benson, M.D.; Schneider, B.; Ellgaard, L.; Berthiaume, L.G.; Simmen, T. Palmitoylated TMX and calnexin target to the mitochondria-associated membrane. EMBO J. 2012, 31, 457–470. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  226. Garrido-Maraver, J.; Loh, S.H.Y.; Martins, L.M. Forcing contacts between mitochondria and the endoplasmic reticulum extends lifespan in a Drosophila model of Alzheimer’s disease. Biol. Open 2020, 9. [Google Scholar] [CrossRef] [Green Version]
  227. Csordás, G.; Várnai, P.; Golenár, T.; Roy, S.; Purkins, G.; Schneider, T.G.; Balla, T.; Hajnóczky, G. Imaging interorganelle contacts and local calcium dynamics at the ER-mitochondrial interface. Mol. Cell 2010, 39, 121–132. [Google Scholar] [CrossRef] [PubMed]
  228. Lee, H.J.; Jung, Y.H.; Choi, G.E.; Kim, J.S.; Chae, C.W.; Lim, J.R.; Kim, S.Y.; Yoon, J.H.; Cho, J.H.; Lee, S.J.; et al. Urolithin A suppresses high glucose-induced neuronal amyloidogenesis by modulating TGM2-dependent ER-mitochondria contacts and calcium homeostasis. Cell Death Differ. 2020. [Google Scholar] [CrossRef]
Figure 1. Bidirectional regulation of ER and mitochondria functions in MAMs. Mitochondria Ca2+ uptake (i.e., by the VDAC (voltage-dependent anion channel) located in the outer mitochondria membrane (OMM) and the mitochondrial Ca2+ uniporter complex (MCU) located in the inner mitochondria membrane (IMM)) increases the activity of mitochondrial enzymes (i.e., the pyruvate dehydrogenase (PDH), and the tricarboxylic acid cycle (TCA)) orchestrating the activity of the oxidative phosphorylation electron transport chain (ETC) producing ATP by the F1F0 ATP-synthase. ATP is then transported through the adenine nucleotide translocator (ANT). In return, ATP is used by the sarco-endoplasmic reticulum Ca2+ ATPase (SERCA) ensuring active Ca2+ storage in the ER, necessary for ER functions (i.e., protein synthesis and maturation). Mitochondrial Ca2+ homeostasis results from an equilibrium between Ca2+ uptake and extrusion by the Na2+/Ca2+ exchanger (NCX). Prolonged or excessive mitochondrial Ca²+ uptake triggers mitochondrial membrane permeability transition pore (PTP) opening, initiating cytochrome c (Cyt c) release, apoptosome formation, and the activation of apoptotic cell death. In particular, altered ER Ca2+ homeostasis is associated with the unfolded protein response (UPR), triggering excessive transfer of Ca2+ from the ER to mitochondria (i.e., through SERCA1 truncated isoform (S1T), or the inositol 1,4,5-trisphosphate receptors (IP3R)), which is tightly linked to ER stress-mediated mitochondria apoptotic cell death. In parallel, MAMs control phospholipids biosynthesis and transfer between organelles where phosphatidylcholine (PC); phosphatidylethanolamine (PE); and phosphatidylserine (PS) are the most abundant ones. MAMs are also the intersection site of transport and metabolism of cholesterol (Chol), which is catalyzed to generate pregnenolone (Preg) and released to the ER to produce progesterone (Prog). MAMs-associated lipid microdomains regulate the autophagic/mitophagic processes through the recruitment of autophagic proteins and autophagosome formation.
Figure 1. Bidirectional regulation of ER and mitochondria functions in MAMs. Mitochondria Ca2+ uptake (i.e., by the VDAC (voltage-dependent anion channel) located in the outer mitochondria membrane (OMM) and the mitochondrial Ca2+ uniporter complex (MCU) located in the inner mitochondria membrane (IMM)) increases the activity of mitochondrial enzymes (i.e., the pyruvate dehydrogenase (PDH), and the tricarboxylic acid cycle (TCA)) orchestrating the activity of the oxidative phosphorylation electron transport chain (ETC) producing ATP by the F1F0 ATP-synthase. ATP is then transported through the adenine nucleotide translocator (ANT). In return, ATP is used by the sarco-endoplasmic reticulum Ca2+ ATPase (SERCA) ensuring active Ca2+ storage in the ER, necessary for ER functions (i.e., protein synthesis and maturation). Mitochondrial Ca2+ homeostasis results from an equilibrium between Ca2+ uptake and extrusion by the Na2+/Ca2+ exchanger (NCX). Prolonged or excessive mitochondrial Ca²+ uptake triggers mitochondrial membrane permeability transition pore (PTP) opening, initiating cytochrome c (Cyt c) release, apoptosome formation, and the activation of apoptotic cell death. In particular, altered ER Ca2+ homeostasis is associated with the unfolded protein response (UPR), triggering excessive transfer of Ca2+ from the ER to mitochondria (i.e., through SERCA1 truncated isoform (S1T), or the inositol 1,4,5-trisphosphate receptors (IP3R)), which is tightly linked to ER stress-mediated mitochondria apoptotic cell death. In parallel, MAMs control phospholipids biosynthesis and transfer between organelles where phosphatidylcholine (PC); phosphatidylethanolamine (PE); and phosphatidylserine (PS) are the most abundant ones. MAMs are also the intersection site of transport and metabolism of cholesterol (Chol), which is catalyzed to generate pregnenolone (Preg) and released to the ER to produce progesterone (Prog). MAMs-associated lipid microdomains regulate the autophagic/mitophagic processes through the recruitment of autophagic proteins and autophagosome formation.
Ijms 21 09521 g001
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Eysert, F.; Kinoshita, P.F.; Mary, A.; Vaillant-Beuchot, L.; Checler, F.; Chami, M. Molecular Dysfunctions of Mitochondria-Associated Membranes (MAMs) in Alzheimer’s Disease. Int. J. Mol. Sci. 2020, 21, 9521. https://doi.org/10.3390/ijms21249521

AMA Style

Eysert F, Kinoshita PF, Mary A, Vaillant-Beuchot L, Checler F, Chami M. Molecular Dysfunctions of Mitochondria-Associated Membranes (MAMs) in Alzheimer’s Disease. International Journal of Molecular Sciences. 2020; 21(24):9521. https://doi.org/10.3390/ijms21249521

Chicago/Turabian Style

Eysert, Fanny, Paula Fernanda Kinoshita, Arnaud Mary, Loan Vaillant-Beuchot, Frédéric Checler, and Mounia Chami. 2020. "Molecular Dysfunctions of Mitochondria-Associated Membranes (MAMs) in Alzheimer’s Disease" International Journal of Molecular Sciences 21, no. 24: 9521. https://doi.org/10.3390/ijms21249521

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

Article Metrics

Back to TopTop