Next Article in Journal
Rho GTPases: Big Players in Breast Cancer Initiation, Metastasis and Therapeutic Responses
Next Article in Special Issue
Potential Drug Candidates to Treat TRPC6 Channel Deficiencies in the Pathophysiology of Alzheimer’s Disease and Brain Ischemia
Previous Article in Journal
Correction: Nemade, H.; et al. Cyclooxygenases Inhibitors Efficiently Induce Cardiomyogenesis in Human Pluripotent Stem Cells. Cells 2020, 9, 554
Previous Article in Special Issue
Mitochondrial Calcium Deregulation in the Mechanism of Beta-Amyloid and Tau Pathology
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Presenilin-2 and Calcium Handling: Molecules, Organelles, Cells and Brain Networks

1
Department of Biomedical Sciences, University of Padua, Via U. Bassi 58/B, 35131 Padua, Italy
2
Neuroscience Institute, Italian National Research Council (CNR), Via U. Bassi 58/B, 35131 Padua, Italy
3
Venetian Institute of Molecular Medicine (VIMM), Via G. Orus 2B, 35131 Padua, Italy
*
Authors to whom correspondence should be addressed.
Cells 2020, 9(10), 2166; https://doi.org/10.3390/cells9102166
Submission received: 31 August 2020 / Revised: 15 September 2020 / Accepted: 18 September 2020 / Published: 25 September 2020

Abstract

:
Presenilin-2 (PS2) is one of the three proteins that are dominantly mutated in familial Alzheimer’s disease (FAD). It forms the catalytic core of the γ-secretase complex—a function shared with its homolog presenilin-1 (PS1)—the enzyme ultimately responsible of amyloid-β (Aβ) formation. Besides its enzymatic activity, PS2 is a multifunctional protein, being specifically involved, independently of γ-secretase activity, in the modulation of several cellular processes, such as Ca2+ signalling, mitochondrial function, inter-organelle communication, and autophagy. As for the former, evidence has accumulated that supports the involvement of PS2 at different levels, ranging from organelle Ca2+ handling to Ca2+ entry through plasma membrane channels. Thus FAD-linked PS2 mutations impact on multiple aspects of cell and tissue physiology, including bioenergetics and brain network excitability. In this contribution, we summarize the main findings on PS2, primarily as a modulator of Ca2+ homeostasis, with particular emphasis on the role of its mutations in the pathogenesis of FAD. Identification of cell pathways and molecules that are specifically targeted by PS2 mutants, as well as of common targets shared with PS1 mutants, will be fundamental to disentangle the complexity of memory loss and brain degeneration that occurs in Alzheimer’s disease (AD).

1. Presenilin-2 in Physiology and Pathology

Presenilin-2 (PS2)—and its homolog presenilin-1 (PS1)—is a 50-kDa multi-pass membrane protein with nine helical transmembrane (TM) domains, and in humans it is encoded by a gene present on chromosome 1 (PSEN2) [1]. Both presenilins (PSs) mainly localize to the endoplasmic reticulum (ER) and Golgi apparatus (GA) membranes but also, although less abundantly, in plasma membrane (PM) and endosomes [2]. Their mRNAs are expressed in different human and mouse tissues, with the highest levels in the hippocampus and cerebellum [3].
Both PSs represent the catalytic core of the γ-secretase complex, the enzyme ultimately responsible for generation of Aβ peptides; they were both discovered in genetic analyses of families in which Alzheimer’s disease (AD) is transmitted as an autosomal dominant trait. In fact, as of now, about 300 mutations in PSEN1 and 58 mutations in PSEN2 have been described (https://www.alzforum.org/mutations), the majority of which are dominant, mostly missense, and have been associated with the inherited forms of the disease (familial Alzheimer’s disease (FAD)) [4,5]. Mutations in the gene for one of the substrate of the γ-secretase complex, the amyloid precursor protein (APP), are also responsible for FAD cases [6]. It has been proposed that FAD-PS mutations lead to a less precise γ-secretase cleavage of APP, in some cases decreasing the total production of Aβ but increasing the relative amount of the more amyloidogenic Aβ42 peptide, the seeding core of extracellular amyloid plaques, over the more soluble Aβ40 peptide [7,8].
The γ-secretase complex is part of the family of intramembrane-cleaving proteases (I-CliPs), which perform hydrolysis of protein domains embedded in the hydrophobic environment of the membrane. The family includes SP2 metalloproteases, serine proteases of the rhomboid family, and the aspartyl proteases to which γ-secretase belongs.
The γ-secretase has a central role in cellular biology, with about 150 different integral membrane proteins recognized as substrates [9]; the most studied are the Notch family of receptors, with a crucial role in signalling and cell differentiation, and APP [4,9]. The γ-secretase complex is composed of four subunits: PS1 or PS2; nicastrin, an integral membrane protein concerned with substrate recognition and selection [10]; PS enhancer-2 (PEN-2) that stabilizes the PS complex and has a role in its endoproteolytic cleavage [11,12,13]; and anterior pharynx defective 1 (APH1), which interacts with nicastrin, providing the initial scaffold to which PS1/2 and PEN-2 are added [14,15]. In humans, APH1 is encoded by two paralogous genes (APH1A and APH1B), and each protein can interact with either PS, resulting in the existence of four different γ-secretase complexes that might have slightly different specificities [16]. After its enclosure within the complex, PS undergoes an endoproteolytic cleavage that produces N- and C-terminal fragments; the two fragments remain associated and represent the biologically active form of the complex, each carrying one of the two key aspartic acid residues on TM6 and TM7, respectively [17,18].
PS1 and PS2 share about 66% of amino acidic sequence; one key difference is a motif in PS2 that interacts with activating protein-1 (AP-1) complexes in a phosphorylation-dependent manner and targets PS2 to the late endosome/lysosome compartment, leading to a different subcellular distribution of PS2 and perhaps to subtly different functions [19,20]. For example, it has been demonstrated that PS2-containing γ-secretase complexes are involved in the processing of premelanosome (PMEL) protein, which is involved in melanosome maturation and melanin deposition [19]. Indeed, PS2-null zebrafish showed defects in skin pigmentation [21]. Importantly, melanosome biogenesis seems to be Ca2+-dependent [22] (see also below).
Several γ-secretase-independent functions of PSs have emerged in the recent years, enriching the overall importance of these proteins in cell biology. For example, PSs bind to glycogen synthase kinase 3β (GSK3β), a key protein of the Wnt signalling pathway, and to its substrate β-catenin, a transcription regulator [23,24]. The interaction of PSs with GSK3β and β-catenin is independent of γ-secretase activity [25] and influences β-catenin phosphorylation and turnover [26], as well as the activity of kinesin-1 and dynein and thus axonal transport of type 1 transmembrane receptors [27]. PSs have been implicated also in autophagy (see below) and protein trafficking [28].
Last, but not least, the regulation of cellular Ca2+ homeostasis has emerged as a key PS function, independent of γ-secretase activity, with relevant implications in multiple Ca2+-regulated cell processes. In the present review, we summarize the central role played by PS2 in cellular Ca2+ homeostasis, highlighting divergent and convergent aspects of PS2 vs. PS1 pathophysiology.

2. PS2 and Ca2+ Homeostasis

2.1. Alterations of Ca2+ Homeostasis in FAD-PS2 Cell Models

According to the so-called “Ca2+ overload” hypothesis for AD, FAD-PS mutations increase the ER Ca2+ content and cause excessive cytosolic Ca2+ release upon cell stimulations that, in turn, alters APP processing and sensitizes neurons to Ca2+-dependent cell death mechanisms [29]. Indeed, an excessive release of Ca2+ from the ER has been reported in different cell models expressing various FAD-PS mutations, as well as in neurons from transgenic (Tg) mice carrying FAD-PS1 mutations [30,31,32,33,34,35]. FAD-linked mutations in PS2 have also been reported to potentiate ER Ca2+ release from both ryanodine receptors (RyRs) [36] and inositol trisphosphate (IP3) receptors (IP3Rs) in Xenopus oocytes [30] and neurons from Tg mice expressing the PS2-N141I mutation [37]. Moreover, it has been proposed that wild type (WT) PSs form constitutively active ER Ca2+ leak channels whereas FAD-PS mutations disrupt the channel functionality; as a result of the reduced leak, the ER Ca2+ level increases and more Ca2+ is released upon stimulation [38].
In contrast, we showed that FAD patient-derived fibroblasts carrying the PS2-M239I mutation, as well as HeLa and HEK293 cells stably or transiently expressing the same PS2 mutant, show a decreased ER Ca2+ release when stimulated by IP3-generating agonists [39] (Figure 1); this result was confirmed in FAD patient-derived fibroblasts carrying another PS2 mutation (T122R) [40]. Of note, in this study, we analysed two monozygotic twins, one with overt signs of disease at the time of biopsy, whereas the other one was still asymptomatic; nevertheless, both cell samples shared a similar Ca2+ handling defect, strongly suggesting that Ca2+ dysregulation represents an early event in the pathogenesis of AD [40].
To clarify the divergent results, we directly monitored Ca2+ dynamics within intracellular stores. We employed aequorin-based Ca2+ probes targeted to ER and GA in cells expressing different PS1 and PS2 mutants. In several cell lines [SH-SY5Y, HeLa, HEK293 and Mouse Embryonic Fibroblast (MEF) cells], expressing the ER (or GA)-targeted aequorin together with a number of PS1 (P117L, M146L, L286V, and A246E) or PS2 (M239I, T122R, and N141I) mutants, we analysed Ca2+ concentrations and dynamics in the two organelles. By this more specific approach, we confirmed lower ER and GA Ca2+ levels in the presence of all the analysed FAD-PS2 mutants, and unchanged or slightly decreased ER/GA Ca2+ concentrations when PS1 mutants were expressed [41]. Similar results were obtained in FAD patient-derived fibroblasts and rat primary neurons, expressing either PS1 or PS2 mutants and loaded with the Ca2+ sensor fura-2, confirming the capability of PS to modify Ca2+ homeostasis, but questioning the “Ca2+ overload” hypothesis for AD [41,42]. Indeed, it was also shown that FAD-PS are associated with IP3R hyperactivity [43,44], providing an alternative explanation to the “Ca2+ overload” hypothesis based on increased ER Ca2+ release findings previously reported in FAD-PS-expressing cells. In particular, Foskett and co-workers showed that FAD-PS1/2 mutants, by physically interacting with the IP3R, modulate the channel gating, causing an exaggerated ER Ca2+ release regardless of its Ca2+ content [43,44]. Furthermore, by employing ER- and GA-targeted Ca2+ indicators, the same group subsequently confirmed that cells expressing FAD-PS1 mutants do not present ER Ca2+ overload, arguing against the previously proposed role of PS as ER Ca2+ leak channels [45].
Similarly, by using newly developed genetically encoded Ca2+ indicators, the Förster resonance energy transfer (FRET)-based probe targeted to ER (D4ER [46]), medial-GA [47], and trans-GA [48], we showed that (i) in SH-SY5Y and Baby Hamster Kidney (BHK) cells, expressing the FAD-PS2-T122R mutant, and in PS2-N141I patient-derived fibroblasts, there is a clear reduction in ER Ca2+ concentration; (ii) in cells expressing the FAD-PS1-A246E mutant, instead, no change was observed [46,49]; (iii) the expression of FAD-PS2 mutants induced a selective decrease in the medial-GA Ca2+ content, but not in that of the trans-GA; (iv) in contrast, the expression of the FAD-PS1 mutant was ineffective on both GA sub-compartments [49] (Figure 1).
Concerning the molecular mechanism through which FAD-PS2 alters intracellular Ca2+ store dynamics, researchers have shown that the Ca2+ phenotype is caused by the holoprotein and that it is independent of γ-secretase activity [38,39,42,44,49,50,51]. Moreover, it has been shown that the protein directly interacts with the IP3R, sensitizing it to lower IP3 concentrations [43], the RyR [36], the RyR-regulating protein sorcin [52], and the SERCA2b [34], inhibiting its activity [51]. This latter result is consistent with the differential effect of FAD-PS2 mutants on GA sub-compartments (see above), given that the trans-GA, where FAD-PS2 mutants are ineffective, relies only on the secretory pathway Ca2+ ATPase 1 (SPCA1) for Ca2+ uptake [48]. Finally, in the presence of FAD-PS1/2 mutations, increased expression levels and activity of RyRs have been reported [53,54,55], suggesting a RyR-dependent Ca2+ hyperexcitability in AD that is antagonized by the channel inhibitor dantrolene (see [56] for an extensive discussion of this issue; see also [57] for the involvement of IP3Rs).
Intracellular Ca2+ stores, mainly the ER, are functionally and physically coupled to mitochondria with which they jointly operate modulating several cell functionalities, such as lipid synthesis and Ca2+ homeostasis. Specific ER membrane domains tightly juxtaposed to mitochondria, called mitochondria-associated membranes (MAM; [58]), represent signalling platforms and play a key role in these processes [59]. Interestingly, MAM appear to be altered in AD samples [42,54,59,60,61,62,63,64]; in addition PS1/2, as well as the other components of the γ-secretase complex and APP, are enriched in these domains [63,65,66]. Only FAD-PS2 mutants, however, are able to increase the interaction between the two organelles, facilitating ER–mitochondria Ca2+ transfer [42,54,63] by binding to mitofusin-2 [63] and thus removing its negative modulation on organelle tethering [67] (Figure 1).
The other Ca2+ signalling pathway affected by FAD-PS2 is the store-operated Ca2+ entry (SOCE) [68,69]. In particular, it has been shown that several FAD-PS2 mutants reduce SOCE activity in different cell types [40,41,49,70]. Interestingly, this effect is shared with FAD-PS1 mutants, which similarly reduce this Ca2+ influx [41,49,70,71] (Figure 1). Accordingly, SOCE is potentiated in cells where PS levels are reduced [70,72]. In PS double knock out (KO) MEFs and in B-lymphocytes derived from patients expressing FAD-PS mutants [73], researchers have found that the levels of the key SOCE components Stromal interaction molecule (STIM) STIM1 and STIM2 [68,69] are reduced. Of note, alterations in SOCE and STIM1 protein level have also been reported in sporadic AD (SAD) patients [74]. It has been proposed that SOCE is regulated by a γ-secretase-dependent mechanism, with STIM1 being a substrate of PS1-containing γ-secretase complexes [71]. Nevertheless, we found lower SOCE and STIM1 protein levels in both FAD-PS1- and FAD-PS2-expressing cells treated with the γ-secretase inhibitor DAPT [49] (Figure 1).
Of note, the overexpression of WT-PS2 often mimics the effect of its FAD mutants on Ca2+ homeostasis, although higher levels of WT-PS2 are required to obtain the alterations in Ca2+ homeostasis elicited by FAD-PS2 mutants [40]. This latter finding could be relevant for SAD forms of the disease, where an upregulation of the endogenous PS2 has been reported in brain AD samples due to the loss of repressor element 1-silencing transcription factor (REST) [75]. It can be speculated that an abnormal accumulation of PS2 holoprotein could cause Ca2+ signalling dysregulation, also typically observed in SAD cases.

2.2. Calcium Handling in AD Mouse Models Expressing PS2-N141I

The findings reported above led us to investigate Ca2+ handling and brain network functionality in Tg mouse lines based on FAD-PS2 mutants by means of in vitro and in vivo approaches. We took advantage of two homozygous mouse lines expressing the PS2-N141I mutant, as described in detail in Box 1: the double Tg line B6.152H, also known as B6.PS2APP, and the single Tg line PS2.30H [76]. Here, we simply refer to these two lines as 2TG and TG, respectively.
We were firstly interested to verify whether the same Ca2+ changes found in FAD-PS2-expressing cell lines were also detectable in primary neuronal cultures and in acute hippocampal slices from 2-week-old animals. At this age, total brain Aβ levels in 2TG mice are still very low, but are already detectable and higher when compared to WT and TG mice [54]. Both TG and 2TG neurons, in culture or in situ, show a reduction in the ER Ca2+ content, when estimated indirectly, through Ca2+ release induced by IP3-generating agonists, or directly, with the Cameleon probe D4ER [46,54].
In acute hippocampal slices, upon stimulation with IP3-generating agonists, Ca2+ release was dramatically reduced not only in neurons but also in astrocytes of TG and 2TG mice, suggesting defective store Ca2+ content, as well as Ca2+ entry, in these latter cell types [54]. Importantly, these changes occur precociously and independently of APP overexpression and brain Aβ load, being found equally in TG and 2TG mice; thus, they reflect the intrinsic capability of modulating Ca2+ handling of FAD-PS2 mutants.
Studying neurons in vitro and in situ allowed us to also highlight relevant network properties brought about by the PS2 mutant. In both conditions, neuronal cells, when exposed to picrotoxin, a γ-aminobutyric acid (GABA)-A receptor antagonist, showed synchronous Ca2+ spiking activity that was higher in TG and 2TG mice with respect to WT [54]. This type of Ca2+ spiking is independent of Ca2+ stores and likely due to an imbalance between excitatory and inhibitory inputs that represent an early sign of network dysfunction [77,78].

3. Functional Effects of Ca2+ Dysregulation by FAD-PS2

3.1. Autophagy

Macroautophagy (hereafter autophagy) is a process in which double-membrane vesicles (called autophagosomes) engulf different cellular components (including misfolded proteins, portions of cytosol, and damaged organelles) and target them to lysosomes, where they are degraded into simpler molecular constituents.
In 2004, two seminal papers firstly suggested that PSs might be involved in autophagy modulation [79,80]. Specifically, Esselens and co-workers reported telencephalin accumulation within autophagosomes in PS1-KO hippocampal neurons as a result of a defective fusion of these vesicles with lysosomes. Similarly, Wilson and colleagues observed that PS1 deficiency, in fibroblasts and primary cortical neurons, resulted in the formation of enlarged lysosomes, with accumulation of α- and β-synuclein. Importantly, this phenomenon was likely associated with SOCE augmentation, suggesting that altered Ca2+ signalling may underpin the effect of PSs on the autophagy pathway [79]. Additional investigations, mostly focused on PS1, consistently reported that PSs modulate autophagosome-lysosome fusion. Nevertheless, consensus has not been reached on the underlying mechanism. Indeed, either a defective lysosomal acidification [81], a reduced lysosomal Ca2+ content [82], or altered expression of key genes belonging to the coordinated lysosomal expression and regulation (CLEAR) network [83,84] have been suggested as possible mechanisms (reviewed in [85]).
As far as FAD-PS mutants are concerned, some of the discrepancies might be linked to mutation-specific effects. Nevertheless, most studies converge on the lack of involvement of the γ-secretase activity, whereas Ca2+ signalling dysregulation has been frequently reported as a common feature among different FAD-PS mutants [85]. Recently, we observed that the reduced ER Ca2+ content, consistently observed in different FAD-PS2 cell models, affects the fusion of autophagosomes with lysosomes, thus inducing autophagosome accumulation [86] (Figure 2). Specifically, the phenomenon appears linked to the generation of lower cytosolic Ca2+ rises upon IP3-induced release of ER Ca2+, given that it can be mimicked by increasing the cytosolic Ca2+-buffering capacity (loading cells with the permeable forms of Ca2+ chelating agents). Mechanistically, we found that alterations of cytosolic Ca2+ dynamics affect the recruitment to autophagosomes of Ras-associated binding protein RAB7, a small GTPase whose association with both autophagosomes and lysosomes tunes their fusion in the final steps of the autophagy pathway [87]. Importantly, at variance with previous studies focused on FAD-PS1 [81,82,88], neither the pH of lysosomes nor their Ca2+ content were found to be affected by FAD-PS2 mutants [86]. Taken together these observations suggest that slightly different mechanisms might underlie the effects of FAD-PS1 and FAD-PS2 on the autophagy flux, with an altered Ca2+ signalling (though by distinct pathways) being a common feature.

3.2. Cell Metabolism and Bioenergetics

The first piece of evidence that WT-PS2 modulates mitochondrial metabolism was found in 2006, when lower mitochondrial respiration and decreased mitochondrial membrane potential (Δψm) were observed in PSEN2−/−, but not in PSEN1−/− MEFs [89]. Later, similar results were obtained by Contino and co-investigators [90], who found reduced basal and maximal mitochondrial oxygen consumption in PSEN2−/− and PS double KO MEFs (but not in PSEN1−/− MEFs), associated with an altered morphology of the mitochondrial cristae and a dampened expression of different subunits of the mitochondrial respiratory chain. Interestingly, in both studies, the ATP/ADP ratio was not significantly altered by PSEN2 ablation, likely because of a compensatory upregulation of the glycolytic flux [90]. Recently, we obtained data suggesting that, in primary cortical neurons from PS2−/− mice (see Box 1), reduced mitochondrial respiration is associated with a defective mitochondrial Ca2+ signal (Rossi et al., in preparation). This finding suggests that the Ca2+-mediated modulation of mitochondrial metabolism has a key role in the effects reported above [91].
It is well established that mitochondrial activity is critical for brain health—not only is the majority of neuronal ATP synthesized by mitochondria, but also the rate of ATP synthesis matches synaptic activity [92]. Therefore, the effects of FAD-PS on mitochondria metabolism might be relevant to FAD pathogenesis.
Alterations of mitochondrial activity have been reported in different AD models, mostly in Tg mouse models harboring FAD-PS1 and FAD-APP mutations. However, consensus has not been reached on the underlying molecular mechanisms. Indeed, either defective assembly/expression/activity of different subunits of the mitochondrial respiratory chain [93], altered mitochondrial Ca2+ signals [94,95], or organelle positioning/transport [96] have been suggested to contribute to the observed alterations. In contrast, only a few studies have focused on FAD-PS2 mutants. In primary cortical neurons from 2TG mice (see Box 1), we observed a reduced mitochondrial respiratory capacity [97]. This defect is not due to any intrinsic alteration of the respiratory chain, but rather depends on an impaired glycolytic flux, in turn affecting nutrient supply to mitochondria and thus organelle metabolism. However, considering that 2TG mice also express the FAD-APP mutant, it is not clear to what extent FAD-PS2 contributes to this phenotype. Recently, however, in different FAD-PS2-expressing cells, we observed a lower mitochondrial activity associated with a reduced ATP synthesis [98]. Mechanistically, these alterations depend in part on reduced mitochondrial Ca2+ signalling (due to partial depletion of ER Ca2+ content; see above), and in part on defective mitochondrial pyruvate uptake, caused by alterations in a signalling pathway driven by hyperactive GSK3β [98] (Figure 2), a feature commonly reported in AD [99]. Importantly, when compared to WT, in primary cortical neurons from TG mice (see Box 1), basal ATP levels are not significantly affected, whereas a faster ATP decrease is observed in cells exposed to ATP-consuming stimuli. In addition, we found that these metabolic alterations are associated with an increased susceptibility of FAD-PS2-N141I neurons to excitotoxicity induced by glutamate at physiological concentrations [98]. Overall, these results suggest that subtle mitochondrial alterations may be tolerated for a long time until specific stress conditions, imposing a high energy-demand, unveil their pathological potential. This might be relevant in neurological disorders characterized by a late onset, such as AD.

3.3. Brain Network Activity

Ca2+ dysregulation and altered APP processing, the two major hits linked to PS2-N141I expression, could affect neural circuit dynamics during the progression of amyloidosis. By studying brain oscillatory activity of adult 2TG mice under anesthesia, we observed that these mice develop a condition of hippocampal hyperactivity, with increased power in the gamma frequency range (45–90 Hz), as measured by spontaneous local field potential (LFP) signals. Curiously, age-matched TG mice also show a similar increase in the gamma power [100]. This hyperactivity is thus independent of Aβ production given that TG mice, unlike the 2TG animals, show neither plaque deposition nor gliosis, and Aβ42 levels are not significantly different from those found in WT mice [100]. This also suggests that, in 2TG mice, network hyperactivity is not due to compensatory, protective mechanisms and likely exerts a pathogenic role in the disease [78]. Of note, in humans, mild cognitive impairment (MCI) is marked by hyperactivity in the hippocampus, as well as in other cortical regions, that disappears with overt AD [101]. 2TG mice also present hyper-synchronicity, which is detectable as early as 3 months of age [100]. This aspect is likely attributable to the early phase of Aβ accumulation and represents a common feature in AD, often in the form of silent seizures, especially in FAD cases that show a higher incidence of epilepsy [102,103,104].
Studies on the brain electrical activity of both AD patients and mouse models have recently been focused on slow oscillations, which are directly involved in memory consolidation during sleep and unconsciousness [105]. By detecting mesoscale Ca2+ signals at the mouse brain level, Busche and coworkers elegantly demonstrated that functional connectivity in the slow-wave range (0.1–3 Hz) is severely reduced in the neocortex, thalamus, and hippocampus of different AD mouse models, also on the basis of PS1 [106]. Interestingly, slow-wave manipulation restores the functionality of brain circuits, rescues neuronal Ca2+ [107], and enhances memory consolidation in both types of mice [106,107].
Given that PS2-N141I alters neuronal and astrocytic Ca2+ homeostasis, it might also disturb hippocampal and cortical oscillatory activity in the slow-wave range. In mice under anesthesia, the oscillatory activity of different brain depths can be measured by simultaneously recording LFP signals with a multi-site linear probe. We used this approach to study brain rhythmicity at the cortical and hippocampal levels. In both TG and 2TG mice, the total power, which mostly reflects spontaneous activity in the low frequency range (0.1–5 Hz), is reduced, particularly at the hippocampal level, suggesting that the PS2 mutant by itself alters the brain electrical activity [108].
Another interesting feature shared by both TG and 2TG mice is the disruption of cortico-hippocampal oscillation coupling (Figure 2, [108]). The phenomenon, also known as phase-amplitude coupling (PAC), occurs when the phase of slower rhythms influences the amplitude of faster ones, and it has been found to be involved in memory consolidation and information transfer [105,109].
Unique features of 2TG mice help to mark the progression of Aβ accumulation and deposition—loss of functional connectivity in the slow-wave range marks the onset of Aβ accumulation, similarly to what reported in PS1-based AD mice [106], whereas low/high power imbalances characterize Aβ deposition in plaque-seeding mice [108]. Since Aβ42 oligomers are associated with Ca2+ homeostasis dysregulation [110,111,112,113], it is tempting to speculate that, in 2TG mice, Ca2+ handling alterations, due to PS2-N141I, sum up or synergize with defects linked to Aβ accumulation.

4. Concluding Remarks and Possible Therapeutic Targets

At the brain circuit level, the FAD-linked PS2-N141I mutant increases excitability [54,100] and disrupts the coupling of cortical slow-waves to hippocampal fast gamma frequencies [108]. Altogether, these findings are consistent with the high frequency of seizures and behavioral changes found in both FAD-PS2-N141I patients [1] and other mouse models expressing PS2-N141I [114,115]. From an pathogenic point of view, major alterations are expected in subpopulations of fast spiking interneurons that control the excitability of neuronal microcircuits, as reported in AD mouse models [103,116,117,118]. These highly active cells are likely more susceptible to the metabolic failure brought about by the aforementioned defective mitochondrial function [97,98]. One should also consider that, in these mouse models, only the PS2 mutant is expressed in both neurons and glial cells. In particular, astrocytes are good candidates to explain circuit dysfunctions given that, through spontaneous Ca2+ oscillations and intercellular Ca2+ waves, they can control the excitability of large neuronal networks [119,120], as well as modulate neighboring neurons by glio-transmission [121,122]. Furthermore, Ca2+ dysregulation and metabolic impairment in a cell type can also affect the closest cells, thus necessitating their investigation at the in situ and in vivo level.
It can be speculated that defects in metabolic and autophagic pathways, directly dependent on Ca2+ dysregulation (see above), are responsible for the described network hyperexcitability and excitation/inhibition imbalances, which has also been reported in other AD models [77,103,123].
As for Ca2+ dysregulation, a common denominator between FAD-PS1 and -PS2 mutations is SOCE. Nevertheless, up until now, only a few studies have addressed the role of this Ca2+ pathway in neurons, mainly because of technical problems, i.e., the difficulty of distinguishing between activation of SOCE and voltage-operated Ca2+ channels (VOCCs) [124,125]. Recently, STIM2 and ORAI2, two key players in SOCE machinery, have emerged as key components of neuronal SOCE, being implicated in SOCE impairment in mushroom spines of hippocampal neurons from FAD-PS1-M146V knock-in mice [126,127,128]. Although the role of SOCE in neurons is still unclear, it is important to stress that, in excitable cells, STIM and ORAI components might also play non-canonical roles—STIM1 binds to L-type VOCCs, inhibits their gating, and induces channel internalization [129,130] while ORAI1 increases neuronal excitability [131,132]. At variance with neuronal cells, it is now largely accepted that SOCE is crucial for the Ca2+-based excitability that characterizes glial cells both in vitro [133,134] and in vivo [135]. Nonetheless, studies that specifically address the role of FAD-PS2 in glial SOCE modulation are still lacking. Considering also the complexity of microglia involvement in the onset and progression of AD [113,136], it is conceivable that these cells might also be primarily affected in the SOCE pathway, given that PS2 is the major core component of γ-secretase complexes expressed in this cell type [137].
We have recently shown that there is an inverse relationship between SOCE level and Aβ42 accumulation [138], consistent with data obtained in neurons [70] and other model cells [139]. These observations suggest the possibility of rescuing the SOCE defect in neural cells while antagonizing Aβ42 production. It has been demonstrated that, in mouse lymphocytes, SOCE is increased by knockout of ORAI2, a channel subunit and a negative modulator of SOCE that is responsible for the Ca2+ release-activated Ca2+ current [140]. In Aβ42-secreting neuroglioma cells, ORAI2 downregulation also increases SOCE and reduces the Aβ42/Aβ40 ratio [138] (Figure 2). Of note, astrocytes actively participate in Aβ production and clearance [141]. We do not know yet whether ORAI2 can play a similar role in neurons; up until now, it looks unlikely, given that recent data by Betzprozvanny’s group favor the hypothesis that ORAI2 is a component of a specific type of neuronal SOCE that is based on transient receptor potential canonical 6 (TRPC6) channel and regulated by diacylglycerol [127]. What is clear is that investigating Ca2+ dysregulation in AD allows the design of alternative therapeutic approaches to this devastating disease.
Additional therapeutic approaches could be suggested on the basis of altered bioenergetic and autophagy pathways. Impaired mitochondria, unable to supply cellular ATP demand, cause alterations in neuronal excitability, eventually leading to Ca2+ overload and cell death [142]. Moreover, the accumulation of damaged mitochondria (and misfolded proteins), due to defective autophagy, further contributes to dysfunctional neurons, causing, over the long term, neurodegeneration. Indeed, mitochondrial alterations, and in particular defects in bioenergetic pathways, have been widely reported to be key factors not only in AD but also in other neurodegenerative diseases [142,143]. Importantly, bioenergetic alterations are reported in different SAD and FAD samples, appearing at the early stage of the disease, before Aβ plaque formation [144].
The bioenergetic state of neurons is a crucial determinant of their response to glutamate, with cells containing defective mitochondria undergoing bioenergetic crises, Ca2+ mishandling, and excitotoxicity. The Food and Drug Administration (FDA)-approved molecule memantine targets glutamate receptors and is among the few pharmacological treatments that provide modest benefits in AD patients, in addition to cholinesterase inhibitors [145]. Targeting Ca2+ defects, at multiple levels, was suggested as a possible therapeutic strategy, especially in the form of drug repurposing. Among the best candidates, there are dantrolene, a RyR modulator [146], and isradipine, a VOCC inhibitor, as reviewed by Chakraborty and Stutzmann [147]. Attention has to be payed to the fact that dihydropyridines, especially nimodipine, also increase Aβ42 secretion [148]. None of these drugs are in the pipeline yet, and thus additional interventions aiming at supporting other pathways, such as mitochondrial performance, are desirable. In line with this, we showed that GSK3β inhibition rescues the FAD-PS2-linked bioenergetic defect [98]. Interestingly, both PS and Aβ oligomers have been reported to interact with the kinase, favoring its activity [99]. Considering the fact that GSK3β activity has been observed at MAM [149], where PSs are also enriched and Aβ peptides are generated [63,65,66], a MAM-targeted intervention might represent a useful therapeutic strategy [98].
Finally, the impairment in mitochondrial bioenergetics described in AD models is likely linked to a metabolic rewiring, possibly resulting in systemic alterations in the concentration of specific metabolites. Thus, the detailed metabolic profiling of AD patient-derived peripheral samples (blood and cerebrospinal fluid) might offer the possibility to discover new biomarkers that are helpful for early AD diagnosis, as has been previously suggested [150,151].

5. Box 1: AD Mouse Models Based on PS2

Several mouse models have been developed to understand the pathogenesis of AD, however, none of them are capable of reproducing the full spectrum of the human disease. The large majority of the most used AD models are double-Tg mice based on human FAD-APP and -PS1 mutations, both required to obtain fast amyloid accumulation, plaque deposition, and gliosis between 2 and 8 months of age. These Tg mice are widely considered to be adequate models of Aβ amyloidosis and its inflammatory process; they allow us to study the initial stages of the disease, according to the vision that places Aβ toxicity among the first hits in the AD cascade [6,152,153]. Nonetheless, the latter appears necessary but not sufficient in terms of causing neurodegeneration, with other concomitant and downstream factors playing a key role [7]. Neurodegeneration, linked to tau aggregation, is in fact mainly present in 3xTg-AD mice, which host three human mutant genes encoding PS1, APP, and tau [154].
Curiously, only the PS2-N141I mutation has been used to generate AD mouse models based on PSEN2. In terms of the latter, we used two homozygous lines: the double Tg (2TG) B6.152H, also known as B6.PS2APP, and the single Tg (TG) PS2.30H [76]. The latter line expresses the human PS2-N141I under the prion protein promoter, with background C57Bl/6 > 90% [155]. The B6.152H line was instead obtained by co-injection of human PSEN2, carrying the N141I mutation—under the mouse prion protein promoter—and the human APP isoform 751, carrying the APP-KM670/671NL Swedish mutation—under the Thy1.2 promoter—into zygotes of the C57Bl/6 strain (background C57Bl/6 100%) [76].
The PS2.30H line was originally used to obtain hemizygous PS2APP mice by crossing PS2.30H females with APP-Swedish males of the BD.AD147.71H line, with background C57Bl/6 > 90% [155]. Up to 12 months of age, TG mice show neither plaques nor Aβ accumulation in the brain [100]. The histopathological traits of PS2APP and B6.PS2APP are very similar, showing an exponential growth of Aβ accumulation and plaques at 3 and 6 months of age, respectively [76,155]. Plaque deposition starts in the frontal cortex, subiculum, and hippocampus; increases for up to 12–16 months of age; and correlates with the level of human APP transcript [76,155]. Behavioral deficits have only been characterized thoroughly in PS2APP mice, with spatial learning (Morris water maze) and memory defects appearing at 8 months [155]. Biochemical and functional differences between the two closely related models are also present [156,157]. In our studies, TG and 2TG mice are maintained and used in homozygosity, a condition that allows for the reduction of the variability of APP expression [76]. The two lines express PS2 at a similar level, about twice that found in C57Bl/6 WT mice, used as controls [54].
B6.152H mice have also been used in hemizygosity (B6.152) to study different aspects of the AD phenotype [158], or to generate TauPS2APP triple Tg mice, upon crossing the B6.152H line with the Tau-overexpressing pR5 line [159,160] that expresses the human tau-40 isoform under the Thy1.2 promoter [161]. Of note, a PS2-/- mouse line has been obtained by neomycin insertion in the C57Bl/6 × 129Sv genetic background [162,163]. This line does not show alterations of the endogenous APP processing and it is useful in terms of studying the physiological role of PS2 and possible loss-of-function defects associated with PS2-N141I expression [100,108]. It was also used to produce PSEN conditional double KO mice and study neurodegeneration and memory impairments due to PS deficiency [164].
Other AD mouse models, based on PS2-N141I, have been generated under different promoters and genetic backgrounds. Comparison of their histopathological, functional, and behavioral properties is beyond the scope of this review. The latest generation of AD mouse models avoids the overexpression of FAD mutations and is focused on risk genes, such as Triggering receptor expressed on myeloid cells 2 (TREM2) [165] and Apolipoprotein E4 (APOE4) [166]. More than 200 AD mouse models are now available; detailed information about these AD animal models is available at the Alzforum website (https://www.alzforum.org/). It is also necessary to mention that doubts have recently been raised on the use of Tg mice to study human AD, given that, at variance with the trascriptomic profiles of physiological human and rodent brain aging, which appear very similar, those of AD brains are largely different between humans and rodents, and even between different Tg AD mouse lines [167].

Author Contributions

Writing—original draft preparation, P.P., E.B., R.F., E.G., A.L., D.P., N.R., M.R., N.V., T.P., C.F. Review and editing P.P., T.P., C.F. All authors have read and agreed to the published version of the manuscript.

Funding

Authors’ work was funded by the following grants: Consiglio Nazionale delle Ricerche (CNR) “Premio Scientifico 2018 DSB-CNR” to R.F.; the University of Padova, Italy (SID 2019), the Italian Ministry of University and Scientific Research (PRIN2017XA5J5N), and EU Joint Programme in Neurodegenerative Disease, 2015–2018 (grant CeBioND) to P.P.; Fondazione Cassa di Risparmio di Padua e Rovigo (CARIPARO Foundation) Excellence project 2017 (2018/113), Veneto Region ([RISIB Project]), Consiglio Nazionale delle Ricerche (CNR) Special Project Aging, Euro Bioimaging Project Roadmap/ESFRI from European Commission, and Telethon Italy Grant GGP16029A to TP; PRIN-20175C22WM to C.F.; UNIPD Funds for Research Equipment-2015. As post-docs, N.R. and A.L. were supported by fellowships from PRIN20175C22WM to C.F. and BIRD 2017 to C.F., respectively.

Acknowledgments

PS2.30H and B6.152H mouse lines used for the authors’ research were kindly donated by L. Ozmen and F. Hoffmann-La Roche Ltd. (Basel, Switzerland).

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Jayadev, S.; Leverenz, J.B.; Steinbart, E.; Stahl, J.; Klunk, W.; Yu, C.E.; Bird, T.D. Alzheimer’s disease phenotypes and genotypes associated with mutations in presenilin 2. Brain 2010, 133, 1143–1154. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Brunkan, A.L.; Goate, A.M. Presenilin function and gamma-secretase activity. J. Neurochem. 2005, 93, 769–792. [Google Scholar] [CrossRef] [PubMed]
  3. Lee, M.K.; Slunt, H.H.; Martin, L.J.; Thinakaran, G.; Kim, G.; Gandy, S.E.; Seeger, M.; Koo, E.; Price, D.L.; Sisodia, S.S. Expression of presenilin 1 and 2 (PS1 and PS2) in human and murine tissues. J. Neurosci. 1996, 16, 7513–7525. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Chévez-Gutiérrez, L.; Bammens, L.; Benilova, I.; Vandersteen, A.; Benurwar, M.; Borgers, M.; Lismont, S.; Zhou, L.; Van Cleynenbreugel, S.; Esselmann, H.; et al. The Mechanism of γ-Secretase Dysfunction in Familial Alzheimer Disease. EMBO J. 2012, 31, 2261–2274. [Google Scholar] [CrossRef]
  5. Chávez-Gutiérrez, L.; Szaruga, M. Mechanisms of neurodegeneration-insights from familial alzheimer’s disease. In Seminars in Cell and Developmental Biology; Elsevier Ltd.: Amsterdam, The Netherlands, 2020; pp. 75–85. [Google Scholar]
  6. Selkoe, D.J.; Hardy, J. The amyloid hypothesis of alzheimer’s disease at 25years. EMBO Mol. Med. 2016, 8, 595–608. [Google Scholar] [CrossRef]
  7. Long, J.M.; Holtzman, D.M. Alzheimer disease: An update on pathobiology and treatment strategies. Cell 2019, 179, 312–339. [Google Scholar] [CrossRef]
  8. Walker, E.S.; Martinez, M.; Brunkan, A.L.; Goate, A. Presenilin 2 familial alzheimer’s disease mutations result in partial loss of function and dramatic changes in abeta 42/40 ratios. J. Neurochem. 2005, 92, 294–301. [Google Scholar] [CrossRef]
  9. Güner, G.; Lichtenthaler, S.F. The substrate repertoire of γ-secretase/presenilin. In Seminars in Cell and Developmental Biology; Elsevier Ltd.: Amsterdam, The Netherlands, 2020; pp. 27–42. [Google Scholar]
  10. Bolduc, D.M.; Montagna, D.R.; Gu, Y.; Selkoe, D.J.; Wolfe, M.S. Nicastrin functions to sterically hinder γ-secretase-substrate interactions driven by substrate transmembrane domain. Proc. Natl. Acad. Sci. USA 2016, 113, E509–E518. [Google Scholar] [CrossRef] [Green Version]
  11. Holmes, O.; Paturi, S.; Selkoe, D.J.; Wolfe, M.S. Pen-2 is essential for γ-secretase complex stability and trafficking but partially dispensable for endoproteolysis. Biochemistry 2014, 53, 4393–4406. [Google Scholar] [CrossRef]
  12. Kim, S.H.; Sisodia, S.S. Evidence that the “NF” motif in transmembrane domain 4 of presenilin 1 is critical for binding with PEN-2. J. Biol. Chem. 2005, 280, 41953–41996. [Google Scholar] [CrossRef] [Green Version]
  13. Prokop, S.; Shirotani, K.; Edbauer, D.; Haass, C.; Steiner, H. Requirement of PEN-2 for stabilization of the presenilin n-/c-terminal fragment heterodimer within the γ-secretase complex. J. Biol. Chem. 2004, 279, 23255–23261. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Pardossi-Piquard, R.; Yang, S.P.; Kanemoto, S.; Gu, Y.; Chen, F.; Böhm, C.; Sevalle, J.; Li, T.; Wong, P.C.; Checler, F.; et al. APH1 polar transmembrane residues regulate the assembly and activity of presenilin complexes. J. Biol. Chem. 2009, 284, 16298–16307. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Gu, Y.; Chen, F.; Sanjo, N.; Kawarai, T.; Hasegawa, H.; Duthie, M.; Li, W.; Ruan, X.; Luthra, A.; Mount, H.T.; et al. APH-1 interacts with mature and immature forms of presenilins and nicastrin and may play a role in maturation of presenilin.nicastrin complexes. J. Biol. Chem. 2003, 278, 7374–7380. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Serneels, L.; Van Biervliet, J.; Craessaerts, K.; Dejaegere, T.; Horré, K.; Van Houtvin, T.; Esselmann, H.; Paul, S.; Schäfer, M.K.; Berezovska, O.; et al. γ-Secretase heterogeneity in the aph1 subunit: Relevance for alzheimer’s disease. Science 2009, 324, 629–632. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Wolfe, M.S.; Xia, W.; Ostaszewski, B.L.; Diehl, T.S.; Kimberly, W.T.; Selkoe, D.J. Two transmembrane aspartates in presenilin-1 required for presenilin endoproteolysis and g-secretase activity. Nature 1999, 398, 513–517. [Google Scholar] [CrossRef]
  18. Wolfe, M.S. Substrate recognition and processing by γ-secretase. Biochim. Biophys. Acta Biomembr. 2020, 1862, 183016. [Google Scholar] [CrossRef]
  19. Sannerud, R.; Esselens, C.; Ejsmont, P.; Mattera, R.; Rochin, L.; Tharkeshwar, A.K.; De Baets, G.; De Wever, V.; Habets, R.; Baert, V.; et al. Restricted location of PSEN2/gamma-secretase determines substrate specificity and generates an intracellular abeta pool. Cell 2016, 166, 193–208. [Google Scholar] [CrossRef] [Green Version]
  20. Meckler, X.; Checler, F. Presenilin 1 and presenilin 2 target gamma-secretase complexes to distinct cellular compartments. J. Biol. Chem. 2016, 291, 12821–12837. [Google Scholar] [CrossRef] [Green Version]
  21. Jiang, H.; Newman, M.; Lardelli, M. The zebrafish orthologue of familial alzheimer’s disease gene PRESENILIN 2 is required for normal adult melanotic skin pigmentation. PLoS ONE 2018, 13, e0206155. [Google Scholar] [CrossRef] [Green Version]
  22. Zhang, Z.; Gong, J.; Sviderskaya, E.V.; Wei, A.; Li, W. Mitochondrial NCKX5 regulates melanosomal biogenesis and pigment production. J. Cell Sci. 2019, 132, jcs232009. [Google Scholar] [CrossRef] [Green Version]
  23. Murayama, M.; Tanaka, S.; Palacino, J.; Murayama, O.; Honda, T.; Sun, X.; Yasutake, K.; Nihonmatsu, N.; Wolozin, B.; Takashima, A. Direct association of presenilin-1 with β-catenin. FEBS Lett. 1998, 433, 73–77. [Google Scholar] [CrossRef] [Green Version]
  24. Zhang, Z.; Hartmann, H.; Do, V.M.; Abramowski, D.; Sturchler-Pierrat, C.; Staufenbiel, M.; Sommer, B.; Van De Wetering, M.; Clevers, H.; Saftig, P.; et al. Destabilization of β-Catenin by mutations in Presenilin-1 potentiates neuronal apoptosis. Nature 1998, 395, 698–702. [Google Scholar] [CrossRef]
  25. Soriano, S.; Kang, D.E.; Fu, M.; Pestell, R.; Chevallier, N.; Zheng, H.; Koo, E.H. Presenilin 1 negatively regulates β-Catenin/T cell factor/Lymphoid enhancer factor-1 signaling independently of β-Amyloid precursor protein and notch processing. J. Cell Biol. 2001, 52, 785–794. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Kang, D.E.; Soriano, S.; Frosch, M.P.; Collins, T.; Naruse, S.; Sisodia, S.S.; Leibowitz, G.; Levine, F.; Koo, E.H. Presenilin 1 facilitates the constitutive turnover of β-Catenin: Differential activity of alzheimer’s disease-linked PS1 mutants in the β-catenin-signaling pathway. J. Neurosci. 1999, 19, 4229–4237. [Google Scholar] [CrossRef] [PubMed]
  27. Dolma, K.; Iacobucci, G.J.; Hong Zheng, K.; Shandilya, J.; Toska, E.; White, J.A.; Spina, E.; Gunawardena, S. Presenilin influences glycogen synthase kinase-3 beta (GSK-3beta) for Kinesin-1 and dynein function during axonal transport. Hum. Mol. Genet. 2014, 23, 1121–1133. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Barthet, G.; Dunys, J.; Shao, Z.; Xuan, Z.; Ren, Y.; Xu, J.; Arbez, N.; Mauger, G.; Bruban, J.; Georgakopoulos, A.; et al. Presenilin mediates neuroprotective functions of ephrinb and brain-derived neurotrophic factor and regulates ligand-induced internalization and metabolism of EphB2 and TrkB receptors. Neurobiol. Aging 2013, 34, 499–510. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. LaFerla, F.M. Calcium dyshomeostasis and intracellular signalling in alzheimer’s disease. Nat. Rev. Neurosci. 2002, 3, 862–872. [Google Scholar] [CrossRef]
  30. Leissring, M.A.; Paul, B.A.; Parker, I.; Cotman, C.W.; LaFerla, F.M. Alzheimer’s presenilin-1 mutation potentiates inositol 1,4,5-trisphosphate-mediated calcium signaling in xenopus oocytes. J. Neurochem. 1999, 72, 1061–1068. [Google Scholar] [CrossRef]
  31. Chan, S.L.; Mayne, M.; Holden, C.P.; Geiger, J.D.; Mattson, M.P. Presenilin-1 Mutations increase levels of ryanodine receptors and calcium release in PC12 cells and cortical neurons. J. Biol. Chem. 2000, 275, 18195–18200. [Google Scholar] [CrossRef] [Green Version]
  32. Leissring, M.A.; Akbari, Y.; Fanger, C.M.; Cahalan, M.D.; Mattson, M.P.; LaFerla, F.M. Capacitative calcium entry deficits and elevated luminal calcium content in mutant presenilin-1 knockin mice. J. Cell Biol. 2000, 149, 793–798. [Google Scholar] [CrossRef] [Green Version]
  33. Smith, I.F.; Boyle, J.P.; Vaughan, P.F.; Pearson, H.A.; Cowburn, R.F.; Peers, C.S. Ca2+ stores and capacitative Ca2+ entry in human neuroblastoma (SH- SY5Y) cells expressing a familial alzheimer’s disease presenilin-1 mutation. Brain Res. 2002, 949, 105–111. [Google Scholar] [CrossRef]
  34. 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] [PubMed] [Green Version]
  35. Cai, D.; Netzer, W.J.; Zhong, M.; Lin, Y.; Du, G.; Frohman, M.; Foster, D.A.; Sisodia, S.S.; Xu, H.; Gorelick, F.S.; et al. Presenilin-1 uses phospholipase d1 as a negative regulator of β-amyloid formation. Proc. Natl. Acad. Sci. USA 2006, 103, 1941–1946. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Hayrapetyan, V.; Rybalchenko, V.; Rybalchenko, N.; Koulen, P. The N-terminus of presenilin-2 increases single channel activity of brain ryanodine receptors through direct protein-protein interaction. Cell Calcium 2008, 44, 507–518. [Google Scholar] [CrossRef] [PubMed]
  37. Schneider, I.; Reverse, D.; Dewachter, I.; Ris, L.; Caluwaerts, N.; Kuiperi, C.; Gilis, M.; Geerts, H.; Kretzschmar, H.; Godaux, E.; et al. Mutant presenilins disturb neuronal calcium homeostasis in the brain of transgenic mice, decreasing the threshold for excitotoxicity and facilitating long-term potentiation. J. Biol. Chem. 2001, 276, 11539–11544. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. 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]
  39. Zatti, G.; Ghidoni, R.; Barbiero, L.; Binetti, G.; Pozzan, T.; Fasolato, C.; Pizzo, P. The presenilin 2 M239I mutation associated with familial alzheimer’s disease reduces Ca2+ release from intracellular stores. Neurobiol. Dis. 2004, 15, 269–278. [Google Scholar] [CrossRef]
  40. Giacomello, M.; Barbiero, L.; Zatti, G.; Squitti, R.; Binetti, G.; Pozzan, T.; Fasolato, C.; Ghidoni, R.; Pizzo, P. Reduction of Ca2+ stores and capacitative Ca2+ entry is associated with the familial alzheimer’s disease presenilin-2 T122R mutation and anticipates the onset of dementia. Neurobiol. Dis. 2005, 18, 638–648. [Google Scholar] [CrossRef]
  41. Zatti, G.; Burgo, A.; Giacomello, M.; Barbiero, L.; Ghidoni, R.; Sinigaglia, G.; Florean, C.; Bagnoli, S.; Binetti, G.; Sorbi, S.; et al. Presenilin mutations linked to familial alzheimer’s disease reduce endoplasmic reticulum and golgi apparatus calcium levels. Cell Calcium 2006, 39, 539–550. [Google Scholar] [CrossRef]
  42. 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] [Green Version]
  43. Cheung, K.H.; Shineman, D.; Muller, M.; Cardenas, C.; Mei, L.; Yang, J.; Tomita, T.; Iwatsubo, T.; Lee, V.M.; 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] [PubMed] [Green Version]
  44. Cheung, K.H.; Mei, L.; Mak, D.O.; Hayashi, I.; Iwatsubo, T.; Kang, D.E.; Foskett, J.K. Gain-of-function enhancement of IP3 receptor modal gating by familial alzheimer’s disease-linked presenilin mutants in human cells and mouse neurons. Sci. Signal. 2010, 3, ra22. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Shilling, D.; Mak, D.O.; Kang, D.E.; Foskett, J.K. Lack of evidence for presenilins as endoplasmic reticulum Ca2+ leak channels. J. Biol. Chem. 2012, 14, 10933–10944. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Greotti, E.; Wong, A.; Pozzan, T.; Pendin, D.; Pizzo, P. Characterization of the ER-targeted low affinity Ca2+ probe D4ER. Sensors 2016, 16, 1419. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Lissandron, V.; Podini, P.; Pizzo, P.; Pozzan, T. Unique characteristics of Ca2+ homeostasis of the trans-golgi compartment. Proc. Natl. Acad. Sci. USA 2010, 107, 9198–9203. [Google Scholar] [CrossRef] [Green Version]
  48. Wong, A.K.; Capitanio, P.; Lissandron, V.; Bortolozzi, M.; Pozzan, T.; Pizzo, P. Heterogeneity of Ca2+ handling among and within golgi compartments. J. Mol. Cell Biol. 2013, 5, 266–276. [Google Scholar] [CrossRef]
  49. Greotti, E.; Capitanio, P.; Wong, A.; Pozzan, T.; Pizzo, P.; Pendin, D. Familial alzheimer’s disease-linked presenilin mutants and intracellular Ca2+ handling: A single-organelle, FRET-based analysis. Cell Calcium 2019, 79, 44–56. [Google Scholar] [CrossRef]
  50. Nelson, O.; Tu, H.; Lei, T.; Bentahir, M.; de Strooper, B.; Bezprozvanny, I. Familial alzheimer disease-linked mutations specifically disrupt Ca2+ leak function of presenilin 1. J. Clin. Investig. 2007, 117, 1230–1239. [Google Scholar] [CrossRef] [Green Version]
  51. Brunello, L.; Zampese, E.; Florean, C.; Pozzan, T.; Pizzo, P.; Fasolato, C. Presenilin-2 dampens intracellular Ca2+ stores by increasing Ca2+ leakage and reducing Ca2+ uptake. J. Cell. Mol. Med. 2009, 13, 3358–3369. [Google Scholar] [CrossRef] [Green Version]
  52. Pack-Chung, E.; Meyers, M.B.; Pettingell, W.P.; Moir, R.D.; Brownawell, A.M.; Cheng, I.; Tanzi, R.E.; Kim, T.W. Presenilin 2 interacts with sorcin, a modulator of the ryanodine receptor. J. Biol. Chem. 2000, 275, 14440–14445. [Google Scholar] [CrossRef] [Green Version]
  53. Chakroborty, S.; Goussakov, I.; Miller, M.B.; Stutzmann, G.E. Deviant ryanodine receptor-mediated calcium release resets synaptic homeostasis in presymptomatic 3×Tg-AD mice. J. Neurosci. 2009, 29, 9458–9470. [Google Scholar] [CrossRef] [PubMed]
  54. Kipanyula, M.J.; Contreras, L.; Zampese, E.; Lazzari, C.; Wong, A.K.C.; Pizzo, P.; Fasolato, C.; Pozzan, T. Ca2+ dysregulation in neurons from transgenic mice expressing mutant presenilin 2. Aging Cell 2012, 11, 885–893. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Smith, I.F.; Hitt, B.; Green, K.N.; Oddo, S.; LaFerla, F.M. Enhanced caffeine-induced Ca2+ release in the 3×Tg-AD mouse model of alzheimer’s disease. J. Neurochem. 2005, 94, 1711–1718. [Google Scholar] [CrossRef]
  56. Del Prete, D.; Checler, F.; Chami, M. Ryanodine receptors: Physiological function and deregulation in alzheimer disease. Mol. Neurodegener. 2014, 9, 21. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Shilling, D.; Muller, M.; Takano, H.; Mak, D.O.; 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]
  58. Vance, J.E. Phospholipid synthesis in a membrane fraction associated with mitochondria. J. Biol. Chem. 1990, 265, 7248–7256. [Google Scholar]
  59. Filadi, R.; Theurey, P.; Pizzo, P. The endoplasmic reticulum-mitochondria coupling in health and disease: Molecules, functions and significance. Cell Calcium 2017, 62, 1–15. [Google Scholar] [CrossRef]
  60. Hedskog, L.; Pinho, C.M.; Filadi, R.; Ronnback, 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] [Green Version]
  61. Area-Gomez, E.; Del Carmen Lara Castillo, M.; Tambini, M.D.; Guardia-Laguarta, C.; de Groof, A.J.; 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]
  62. 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] [Green Version]
  63. 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] [PubMed] [Green Version]
  64. Liu, Y.; Zhu, X. Endoplasmic reticulum-mitochondria tethering in neurodegenerative diseases. Transl. Neurodegener. 2017, 6, 1–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Area-Gomez, E.; de Groof, A.J.; 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] [PubMed] [Green Version]
  66. Schreiner, B.; Hedskog, L.; Wiehager, B.; Ankarcrona, M. Amyloid-beta peptides are generated in mitochondria-associated endoplasmic reticulum membranes. J. Alzheimer’s Dis. 2015, 43, 369–374. [Google Scholar] [CrossRef]
  67. Filadi, R.; Greotti, E.; Turacchio, G.; Luini, A.; Pozzan, T.; Pizzo, P. Mitofusin 2 ablation increases endoplasmic reticulum-mitochondria coupling. Proc. Natl. Acad. Sci. USA 2015, 112, E2174–E2181. [Google Scholar] [CrossRef] [Green Version]
  68. Hogan, P.G.; Rao, A. Store-operated calcium entry: Mechanisms and modulation. Biochem. Biophys. Res. Commun. 2015, 460, 40–49. [Google Scholar] [CrossRef] [Green Version]
  69. Putney, J.W. Forms and functions of store-operated calcium entry mediators, stim and orai. Adv. Biol. Regul. 2018, 68, 88–96. [Google Scholar] [CrossRef]
  70. Yoo, A.S.; Cheng, I.; Chung, S.; Grenfell, T.Z.; Lee, H.; Pack-Chung, E.; Handler, M.; Shen, J.; Xia, W.; Tesco, G.; et al. Presenilin-mediated modulation of capacitative calcium entry. Neuron 2000, 27, 561–572. [Google Scholar] [CrossRef] [Green Version]
  71. Tong, B.C.; Lee, C.S.; Cheng, W.H.; Lai, K.O.; Foskett, J.K.; Cheung, K.H. Familial alzheimer’s disease-associated presenilin 1 mutants promote gamma-secretase cleavage of STIM1 to impair store-operated Ca2+ entry. Sci. Signal. 2016, 9, ra89. [Google Scholar] [CrossRef] [Green Version]
  72. Herms, J.; Schneider, I.; Dewachter, I.; Caluwaerts, N.; Kretzschmar, H.; Van Leuven, F. Capacitive calcium entry is directly attenuated by mutant presenilin-1, independent of the expression of the amyloid precursor protein. J. Biol. Chem. 2003, 278, 2484–2489. [Google Scholar] [CrossRef] [Green Version]
  73. Bojarski, L.; Pomorski, P.; Szybinska, A.; Drab, M.; Skibinska-Kijek, A.; Gruszczynska-Biegala, J.; Kuznicki, J. Presenilin-dependent expression of STIM proteins and dysregulation of capacitative Ca2+ entry in familial Alzheimer’s disease. Biochim. Biophys. Acta 2009, 1793, 1050–1057. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Pascual-Caro, C.; Berrocal, M.; Lopez-Guerrero, A.M.; Alvarez-Barrientos, A.; Pozo-Guisado, E.; Gutierrez-Merino, C.; Mata, A.M.; Martin-Romero, F.J. STIM1 deficiency is linked to Alzheimer’s disease and triggers cell death in SH-SY5Y cells by upregulation of L-Type voltage-operated Ca2+ entry. J. Mol. Med. 2018, 96, 1061–1079. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Lu, T.; Aron, L.; Zullo, J.; Pan, Y.; Kim, H.; Chen, Y.; Yang, T.H.; Kim, H.M.; Drake, D.; Liu, X.S.; et al. Rest and stress resistance in ageing and Alzheimer’s disease. Nature 2014, 507, 448–454. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Ozmen, L.; Albientz, A.; Czech, C.; Jacobsen, H. Expression of transgenic APP MRNA is the key determinant for beta-amyloid deposition in PS2APP transgenic mice. Neurodegener. Dis. 2009, 6, 29–36. [Google Scholar] [CrossRef]
  77. Zott, B.; Busche, M.A.; Sperling, R.A.; Konnerth, A. What happens with the circuit in Alzheimer’s disease in mice and humans? Annu. Rev. Neurosci. 2018, 41, 277–297. [Google Scholar] [CrossRef]
  78. Stargardt, A.; Swaab, D.F.; Bossers, K. Storm before the quiet: Neuronal hyperactivity and ab in the presymptomatic stages of alzheimer’s Disease. Neurobiol. Aging 2015, 36, 1–11. [Google Scholar] [CrossRef]
  79. Wilson, C.A.; Murphy, D.D.; Giasson, B.I.; Zhang, B.; Trojanowski, J.Q.; Lee, V.M. Degradative organelles containing mislocalized alpha-and beta-synuclein proliferate in presenilin-1 null neurons. J. Cell Biol. 2004, 165, 335–346. [Google Scholar] [CrossRef]
  80. Esselens, C.; Oorschot, V.; Baert, V.; Raemaekers, T.; Spittaels, K.; Serneels, L.; Zheng, H.; Saftig, P.; De Strooper, B.; Klumperman, J.; et al. Presenilin 1 mediates the turnover of telencephalin in hippocampal neurons via an autophagic degradative pathway. J. Cell Biol. 2004, 166, 1041–1054. [Google Scholar] [CrossRef]
  81. 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] [Green Version]
  82. Coen, K.; Flannagan, R.S.; Baron, S.; Carraro-Lacroix, L.R.; Wang, D.; Vermeire, W.; Michiels, C.; Munck, S.; Baert, V.; Sugita, S.; et al. Lysosomal calcium homeostasis defects, not proton pump defects, cause endo-lysosomal dysfunction in PSEN-deficient cells. J. Cell Biol. 2012, 198, 23–35. [Google Scholar] [CrossRef]
  83. Zhang, X.; Garbett, K.; Veeraraghavalu, K.; Wilburn, B.; Gilmore, R.; Mirnics, K.; Sisodia, S.S. A role for presenilins in autophagy revisited: Normal acidification of lysosomes in cells lacking PSEN1 and PSEN2. J. Neurosci. 2012, 32, 8633–8648. [Google Scholar] [CrossRef] [PubMed]
  84. Reddy, K.; Cusack, C.L.; Nnah, I.C.; Khayati, K.; Saqcena, C.; Huynh, T.B.; Noggle, S.A.; Ballabio, A.; Dobrowolski, R. Dysregulation of nutrient sensing and CLEARance in presenilin deficiency. Cell Rep. 2016, 14, 2166–2179. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Filadi, R.; Pizzo, P. Defective autophagy and alzheimer’s disease: Is calcium the key? In Neural Regeneration Research; Wolters Kluwer Medknow Publications: Mumbai, India, 2019; pp. 2081–2082. [Google Scholar] [CrossRef]
  86. 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, 1–19. [Google Scholar] [CrossRef] [PubMed]
  87. Gutierrez, M.G.; Munafo, D.B.; Beron, W.; Colombo, M.I. Rab7 is required for the normal progression of the autophagic pathway in mammalian cells. J. Cell Sci. 2004, 117, 2687–2697. [Google Scholar] [CrossRef] [Green Version]
  88. Lee, J.H.; McBrayer, M.K.; Wolfe, D.M.; Haslett, L.J.; Kumar, A.; Sato, Y.; Lie, P.P.; Mohan, P.; Coffey, E.E.; Kompella, U.; et al. Presenilin 1 maintains lysosomal Ca2+ homeostasis via TRPML1 by regulating VATPase-mediated lysosome acidification. Cell Rep. 2015, 12, 1430–1444. [Google Scholar] [CrossRef] [Green Version]
  89. Behbahani, H.; Shabalina, I.G.; Wiehager, B.; Concha, H.; Hultenby, K.; Petrovic, N.; Nedergaard, J.; Winblad, B.; Cowburn, R.F.; Ankarcrona, M. Differential role of presenilin-1 and -2 on mitochondrial membrane potential and oxygen consumption in mouse embryonic fibroblasts. J. Neurosci. Res. 2006, 84, 891–902. [Google Scholar] [CrossRef]
  90. Contino, S.; Porporato, P.E.; Bird, M.; Marinangeli, C.; Opsomer, R.; Sonveaux, P.; Bontemps, F.; Dewachter, I.; Octave, J.N.; Bertrand, L.; et al. Presenilin 2-dependent maintenance of mitochondrial oxidative capacity and morphology. Front. Physiol. 2017, 8, 796. [Google Scholar] [CrossRef] [Green Version]
  91. Rossi, A.; Pizzo, P.; Filadi, R. Calcium, mitochondria and cell metabolism: A functional triangle in bioenergetics. Biochim. Biophys. Acta Mol. Cell Res. 2019, 1866, 1068–1078. [Google Scholar] [CrossRef]
  92. Rangaraju, V.; Calloway, N.; Ryan, T.A. Activity-driven local ATP synthesis is required for synaptic function. Cell 2014, 156, 825–835. [Google Scholar] [CrossRef] [Green Version]
  93. Weidling, I.W.; Swerdlow, R.H. Mitochondria in alzheimer’s disease and their potential role in alzheimer’s proteostasis. In Experimental Neurology; Academic Press Inc.: Cambridge, MA, USA, 2020. [Google Scholar]
  94. Jadiya, P.; Kolmetzky, D.W.; Tomar, D.; Di Meco, A.; Lombardi, A.A.; Lambert, J.P.; Luongo, T.S.; Ludtmann, M.H.; Praticò, D.; Elrod, J.W. Impaired mitochondrial calcium efflux contributes to disease progression in models of alzheimer’s disease. Nat. Commun. 2019, 10, 3885. [Google Scholar] [CrossRef]
  95. Toglia, P.; Cheung, K.H.; Mak, D.O.D.; Ullah, G. Impaired mitochondrial function due to familial alzheimer’s disease-causing presenilins mutants via Ca2+ disruptions. Cell Calcium 2016, 59, 240–250. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Flannery, P.J.; Trushina, E. Mitochondrial dynamics and transport in alzheimer’s disease. In Molecular and Cellular Neuroscience; Academic Press Inc.: Cambridge, MA, USA, 2019; pp. 109–120. [Google Scholar]
  97. Theurey, P.; Connolly, N.M.C.; Fortunati, I.; Basso, E.; Lauwen, S.; Ferrante, C.; Moreira Pinho, C.; Joselin, A.; Gioran, A.; Bano, D.; et al. Systems biology identifies preserved integrity but impaired metabolism of mitochondria due to a glycolytic defect in alzheimer’s disease neurons. Aging Cell 2019, 18, e12924. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Rossi, A.; Rigotto, G.; Valente, G.; Giorgio, V.; Basso, E.; Filadi, R.; Pizzo, P. Defective mitochondrial pyruvate flux affects cell bioenergetics in alzheimer’s disease-related models. Cell Rep. 2020, 30, 2332–2348. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Llorens-Martin, M.; Jurado, J.; Hernandez, F.; Avila, J. GSK-3beta, a pivotal kinase in alzheimer disease. Front. Mol. Neurosci. 2014, 7, 46. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Fontana, R.; Agostini, M.; Murana, E.; Mahmud, M.; Scremin, E.; Rubega, M.; Sparacino, G.; Vassanelli, S.; Fasolato, C. Early hippocampal hyperexcitability in PS2APP mice: Role of mutant PS2 and APP. Neurobiol. Aging 2017, 50, 64–76. [Google Scholar] [CrossRef]
  101. Dickerson, B.C.; Salat, D.H.; Greve, D.N.; Chua, E.F.; Rand-Giovannetti, E.; Rentz, D.M.; Bertram, L.; Mullin, K.; Tanzi, R.E.; Blacker, D.; et al. Increased hippocampal activation in mild cognitive impairment compared to normal aging and AD. Neurology 2005, 65, 404–411. [Google Scholar] [CrossRef] [Green Version]
  102. Bakker, A.; Albert, M.S.; Krauss, G.; Speck, C.L.; Gallagher, M. Response of the medial temporal lobe network in amnestic mild cognitive impairment to therapeutic intervention assessed by FMRI and memory task performance. Neuroimage Clin. 2015, 7, 688–698. [Google Scholar] [CrossRef] [Green Version]
  103. Palop, J.J.; Mucke, L. Network abnormalities and interneuron dysfunction in alzheimer disease. Nat. Rev. Neurosci. 2016, 17, 777–792. [Google Scholar] [CrossRef]
  104. Lam, A.D.; Deck, G.; Goldman, A.; Eskandar, E.N.; Noebels, J.; Cole, A.J. Silent hippocampal seizures and spikes identified by foramen ovale electrodes in alzheimer’s disease. Nat. Med. 2017, 23, 678–680. [Google Scholar] [CrossRef]
  105. Mitra, A.; Snyder, A.Z.; Hacker, C.D.; Pahwa, M.; Tagliazucchi, E.; Laufs, H.; Leuthardt, E.C.; Raichle, M.E. Human cortical–hippocampal dialogue in wake and slow-wave sleep. Proc. Natl. Acad. Sci. USA 2016, 113, E6868–E6876. [Google Scholar] [CrossRef] [Green Version]
  106. Busche, M.A.; Grienberger, C.; Keskin, A.D.; Song, B.; Neumann, U.; Staufenbiel, M.; Forstl, H.; Konnerth, A. Decreased amyloid-beta and increased neuronal hyperactivity by immunotherapy in alzheimer’s models. Nat. Neurosci. 2015, 18, 1725–1727. [Google Scholar] [CrossRef] [PubMed]
  107. Kastanenka, K.V.; Hou, S.S.; Shakerdge, N.; Logan, R.; Feng, D.; Wegmann, S.; Chopra, V.; Hawkes, J.M.; Chen, X.; Bacskai, B.J. Optogenetic restoration of disrupted slow oscillations halts amyloid deposition and restores calcium homeostasis in an animal model of alzheimer’s disease. PLoS ONE 2017, 12, e0170275. [Google Scholar] [CrossRef] [PubMed]
  108. Leparulo, A.; Mahmud, M.; Scremin, E.; Pozzan, T.; Vassanelli, S.; Fasolato, C. Dampened slow oscillation connectivity anticipates amyloid deposition in the PS2APP mouse model of alzheimer’s disease. Cells 2019, 9, 54. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Born, J. Slow-wave sleep and the consolidation of long-term memory. World J. Biol. Psychiatry 2010, 11, 16–21. [Google Scholar] [CrossRef] [PubMed]
  110. Agostini, M.; Fasolato, C. When, Where and How? Focus on neuronal calcium dysfunctions in alzheimer’s disease. Cell Calcium 2016, 60, 289–298. [Google Scholar] [CrossRef] [Green Version]
  111. Lazzari, C.; Kipanyula, M.J.; Agostini, M.; Pozzan, T.; Fasolato, C. Ab42 oligomers selectively disrupt neuronal calcium release. Neurobiol. Aging 2015, 36, 877–885. [Google Scholar] [CrossRef]
  112. Tong, B.C.; Wu, A.J.; Li, M.; Cheung, K.H. Calcium signaling in alzheimer’s disease & therapies. Biochim. Biophys. Acta Mol. Cell Res. 2018, 1865, 1745–1760. [Google Scholar] [CrossRef] [PubMed]
  113. Brawek, B.; Garaschuk, O. Network-wide dysregulation of calcium homeostasis in alzheimer’s disease. In Cell and Tissue Research; Springer: Berlin/Heidelberg, Germany, 2014; pp. 427–438. [Google Scholar]
  114. Lee, Y.J.; Choi, I.S.; Park, M.H.; Lee, Y.M.; Song, J.K.; Kim, Y.H.; Kim, K.H.; Hwang, D.Y.; Jeong, J.H.; Yun, Y.P.; et al. 4-O-methylhonokiol attenuates memory impairment in presenilin 2 mutant mice through reduction of oxidative damage and inactivation of astrocytes and the ERK pathway. Free Radic. Biol. Med. 2011, 50, 66–77. [Google Scholar] [CrossRef]
  115. Yuk, D.Y.; Lee, Y.K.; Nam, S.Y.; Yun, Y.W.; Hwang, D.Y.; Choi, D.Y.; Oh, K.W.; Hong, J.T. Reduced anxiety in the mice expressing mutant (N141I) presenilin 2. J. Neurosci. Res. 2009, 87, 522–531. [Google Scholar] [CrossRef] [PubMed]
  116. Verret, L.; Mann, E.O.; Hang, G.B.; Barth, A.M.; Cobos, I.; Ho, K.; Devidze, N.; Masliah, E.; Kreitzer, A.C.; Mody, I.; et al. Inhibitory interneuron deficit links altered network activity and cognitive dysfunction in alzheimer model. Cell 2012, 149, 708–721. [Google Scholar] [CrossRef] [Green Version]
  117. Hahn, T.T.G.; Sakmann, B.; Mehta, M.R. Phase-locking of hippocampal interneurons’ membrane potential to neocortical up-down states. Nat. Neurosci. 2006, 9, 1359–1361. [Google Scholar] [CrossRef] [PubMed]
  118. Hijazi, S.; Heistek, T.S.; van der Loo, R.; Mansvelder, H.D.; Smit, A.B.; van Kesteren, R.E. Hyperexcitable parvalbumin interneurons render hippocampal circuitry vulnerable to amyloid beta. Iscience 2020, 23, 101271. [Google Scholar] [CrossRef] [PubMed]
  119. Fellin, T.; Halassa, M.M.; Terunuma, M.; Succol, F.; Takano, H.; Frank, M.; Moss, S.J.; Haydon, P.G. Endogenous nonneuronal modulators of synaptic transmission control cortical slow oscillations in vivo. Proc. Natl. Acad. Sci. USA 2009, 106, 15037–15042. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Poskanzer, K.E.; Yuste, R. Astrocytic regulation of cortical Up states. Proc. Natl. Acad. Sci. USA 2011, 108, 18453–18458. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Araque, A.; Carmignoto, G.; Haydon, P.G.; Oliet, S.H.; Robitaille, R.; Volterra, A. Gliotransmitters travel in time and space. Neuron 2014, 81, 728–739. [Google Scholar] [CrossRef] [Green Version]
  122. Durkee, C.A.; Araque, A. Diversity and specificity of astrocyte-neuron communication. Neuroscience 2019, 396, 73–78. [Google Scholar] [CrossRef]
  123. Busche, M.A.; Chen, X.; Henning, H.A.; Reichwald, J.; Staufenbiel, M.; Sakmann, B.; Konnerth, A. Critical role of soluble amyloid-beta for early hippocampal hyperactivity in a mouse model of alzheimer’s disease. Proc. Natl. Acad. Sci. USA 2012, 109, 8740–8745. [Google Scholar] [CrossRef] [Green Version]
  124. Lu, B.; Fivaz, M. Neuronal SOCE: Myth or reality? In Trends in Cell Biology; Elsevier Ltd.: Amsterdam, The Netherlands, 2016; pp. 890–893. [Google Scholar]
  125. Majewski, L.; Kuznicki, J. SOCE in Neurons: Signaling or just refilling? In Biochimica et Biophysica Acta-Molecular Cell Research; Elsevier: Amsterdam, The Netherlands, 2014; pp. 1940–1952. [Google Scholar]
  126. Sun, S.; Zhang, H.; Liu, J.; Popugaeva, E.; Xu, N.J.; Feske, S.; White, C.L.; Bezprozvanny, I. Reduced synaptic STIM2 expression and impaired store-operated calcium entry cause destabilization of mature spines in mutant presenilin mice. Neuron 2014, 82, 79–93. [Google Scholar] [CrossRef] [Green Version]
  127. Zhang, H.; Sun, S.; Wu, L.; Pchitskaya, E.; Zakharova, O.; Tacer, K.F.; Bezprozvanny, I. Store-operated calcium channel complex in postsynaptic spines: A new therapeutic target for alzheimer’s disease treatment. J. Neurosci. 2016, 36, 11837–11850. [Google Scholar] [CrossRef] [Green Version]
  128. Zhang, H.; Wu, L.; Pchitskaya, E.; Zakharova, O.; Saito, T.; Saido, T.; Bezprozvanny, I. Neuronal store-operated calcium entry and mushroom spine loss in amyloid precursor protein knock-in mouse model of alzheimer’s disease. J. Neurosci. 2015, 35, 13275–13286. [Google Scholar] [CrossRef] [Green Version]
  129. Park, C.Y.; Shcheglovitov, A.; Dolmetsch, R. The CRAC channel activator STIM1 binds and inhibits L-Type voltage-gated calcium channels. Science 2010, 330, 101–105. [Google Scholar] [CrossRef] [PubMed]
  130. Wang, Y.J.; Deng, X.X.; Mancarella, S.; Hendron, E.; Eguchi, S.; Soboloff, J.; Tang, X.A.D.; Gill, D.L. The calcium store sensor, STIM1, reciprocally controls orai and Ca(V)1.2 channels. Science 2010, 330, 105–109. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  131. Moccia, F.; Zuccolo, E.; Soda, T.; Tanzi, F.; Guerra, G.; Mapelli, L.; Lodola, F.; D’Angelo, E. Stim and orai proteins in neuronal Ca2+ signaling and excitability. Front. Cell. Neurosci. 2015, 9, 153. [Google Scholar] [CrossRef] [PubMed]
  132. Dou, Y.; Ia, J.; Gao, R.; Gao, I.; Munoz, F.M.; Wei, D.; Tian, Y.; Barrett, J.E.; Ajit, S.; Meucci, O.; et al. Orai1 plays a crucial role in central sensitization by modulating neuronal excitability. J. Neurosci. 2018, 38, 887–900. [Google Scholar] [CrossRef] [Green Version]
  133. Pizzo, P.; Burgo, A.; Pozzan, T.; Fasolato, C. Role of capacitative calcium entry on glutamate-induced calcium influx in Type-I rat cortical astrocytes. J. Neurochem. 2001, 79, 98–109. [Google Scholar] [CrossRef]
  134. Okubo, Y.; Iino, M.; Hirose, K. Store-operated Ca2+ entry-dependent Ca2+ refilling in the endoplasmic reticulum in astrocytes. Biochem. Biophys. Res. Commun. 2020, 522, 1003–1008. [Google Scholar] [CrossRef]
  135. Toth, A.B.; Hori, K.; Novakovic, M.M.; Bernstein, N.G.; Lambot, L.; Prakriya, M. CRAC channels regulate astrocyte Ca2+ signaling and gliotransmitter release to modulate hippocampal GABAergic transmission. Sci. Signal. 2019, 12, 5450. [Google Scholar] [CrossRef]
  136. Verkhratsky, A.; Rodríguez-Arellano, J.J.; Parpura, V.; Zorec, R. Astroglial calcium signalling in alzheimer’s disease. In Biochemical and Biophysical Research Communications; Elsevier, B.V.: Amsterdam, The Netherlands, 2017; pp. 1005–1012. [Google Scholar]
  137. Jayadev, S.; Case, A.; Eastman, A.J.; Nguyen, H.; Pollak, J.; Wiley, J.C.; Möller, T.; Morrison, R.S.; Garden, G.A. Presenilin 2 is the predominant γ-secretase in microglia and modulates cytokine release. PLoS ONE 2010, 5, e15743. [Google Scholar] [CrossRef] [Green Version]
  138. Scremin, E.; Agostini, M.; Leparulo, A.; Pozzan, T.; Greotti, E.; Fasolato, C. ORAI2 down-regulation potentiates SOCE and decreases Aβ42 accumulation in human neuroglioma cells. Int. J. Mol. Sci. 2020, 21, 5288. [Google Scholar] [CrossRef]
  139. Zeiger, W.; Vetrivel, K.S.; Buggia-Prevot, V.; Nguyen, P.D.; Wagner, S.L.; Villereal, M.L.; Thinakaran, G. Ca2+ influx through store-operated Ca2+ channels reduces alzheimer disease b-amyloid peptide secretion. J. Biol. Chem. 2013, 288, 26955–26966. [Google Scholar] [CrossRef] [Green Version]
  140. Vaeth, M.; Yang, J.; Yamashita, M.; Zee, I.; Eckstein, M.; Knosp, C.; Kaufmann, U.; Karoly Jani, P.; Lacruz, R.S.; Flockerzi, V.; et al. ORAI2 modulates store-operated calcium entry and T cell-mediated immunity. Nat. Commun. 2017, 8, 14714. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  141. Frost, G.R.; Li, Y.M. The role of astrocytes in amyloid production and alzheimer’s disease. In Open Biology; Royal Society Publishing: London, UK, 2017. [Google Scholar]
  142. Filadi, R.; Pizzo, P. Mitochondrial calcium handling and neurodegeneration: When a good signal goes wrong. Curr. Opin. Physiol. 2020. [Google Scholar] [CrossRef]
  143. Golpich, M.; Amini, E.; Mohamed, Z.; Azman Ali, R.; Mohamed Ibrahim, N.; Ahmadiani, A. Mitochondrial dysfunction and biogenesis in neurodegenerative diseases: Pathogenesis and treatment. CNS Neurosci. Ther. 2017, 23, 5–22. [Google Scholar] [CrossRef] [PubMed]
  144. Swerdlow, R.H.; Burns, J.M.; Khan, S.M. The alzheimer’s disease mitochondrial cascade hypothesis: Progress and perspectives. Biochim. Biophys. Acta 2014, 1842, 1219–1231. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Bazzari, F.H.; Abdallah, D.M.; El-Abhar, H.S. Pharmacological interventions to attenuate alzheimer’s disease progression: The story so far. Curr. Alzheimer Res. 2019, 16, 261–277. [Google Scholar] [CrossRef]
  146. 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-b load and memory impairments in Tg2576 mouse model of alzheimer disease. J. Neurosci. 2012, 32, 11820–11834. [Google Scholar] [CrossRef]
  147. Chakroborty, S.; Stutzmann, G.E. Calcium channelopathies and alzheimer’s disease: Insight into therapeutic success and failures. Eur. J. Pharmacol. 2014, 739, 83–95. [Google Scholar] [CrossRef]
  148. Facchinetti, F.; Fasolato, C.; Del Giudice, E.; Burgo, A.; Furegato, S.; Fusco, M.; Basso, E.; Seraglia, R.; D’Arrigo, A.; Leon, A. Nimodipine selectively stimulates b-amyloid 1-42 secretion by a mechanism independent of calcium influx blockage. Neurobiol. Aging 2006, 27, 218–227. [Google Scholar] [CrossRef]
  149. Bantug, G.R.; Fischer, M.; Grahlert, J.; Balmer, M.L.; Unterstab, G.; Develioglu, L.; Steiner, R.; Zhang, L.; Costa, A.S.H.; Gubser, P.M.; et al. Mitochondria-endoplasmic reticulum contact sites function as immunometabolic hubs that orchestrate the rapid recall response of memory CD8(+) T cells. Immunity 2018, 48, 542–555. [Google Scholar] [CrossRef] [Green Version]
  150. Trushina, E.; Dutta, T.; Persson, X.M.; Mielke, M.M.; Petersen, R.C. Identification of altered metabolic pathways in plasma and CSF in mild cognitive impairment and alzheimer’s disease using metabolomics. PLoS ONE 2013, 8, e63644. [Google Scholar] [CrossRef] [Green Version]
  151. Parnetti, L.; Gaiti, A.; Polidori, M.C.; Brunetti, M.; Palumbo, B.; Chionne, F.; Cadini, D.; Cecchetti, R.; Senin, U. Increased cerebrospinal fluid pyruvate levels in alzheimer’s disease. Neurosci. Lett. 1995, 199, 231–233. [Google Scholar] [CrossRef]
  152. Sasaguri, H.; Nilsson, P.; Hashimoto, S.; Nagata, K.; Saito, T.; De Strooper, B.; Hardy, J.; Vassar, R.; Winblad, B.; Saido, T.C. APP mouse models for alzheimer’s disease preclinical studies. EMBO J. 2017, 36, 2473–2487. [Google Scholar] [CrossRef] [PubMed]
  153. Jankowsky, J.L.; Zheng, H. Practical considerations for choosing a mouse model of alzheimer’s disease. Mol. Neurodegener. 2017, 12, 1–22. [Google Scholar] [CrossRef] [PubMed]
  154. Oddo, S.; Caccamo, A.; Shepherd, J.D.; Murphy, M.P.; Golde, T.E.; Kayed, R.; Metherate, R.; Mattson, M.P.; Akbari, Y.; LaFerla, F.M. Triple-transgenic model of alzheimer’s disease with plaques and tangles: Intracellular abeta and synaptic dysfunction. Neuron 2003, 39, 409–421. [Google Scholar] [CrossRef] [Green Version]
  155. Richards, J.G.; Higgins, G.A.; Ouagazzal, A.M.; Ozmen, L.; Kew, J.N.; Bohrmann, B.; Malherbe, P.; Brockhaus, M.; Loetscher, H.; Czech, C.; et al. PS2APP transgenic mice, coexpressing HPS2mut and HAPPswe, show age-related cognitive deficits associated with discrete brain amyloid deposition and inflammation. J. Neurosci. 2003, 23, 8989–9003. [Google Scholar] [CrossRef] [Green Version]
  156. Poirier, R.; Veltman, I.; Pflimlin, M.C.; Knoflach, F.; Metzger, F. Enhanced dentate gyrus synaptic plasticity but reduced neurogenesis in a mouse model of amyloidosis. Neurobiol. Dis. 2010, 40, 386–393. [Google Scholar] [CrossRef]
  157. Weidensteiner, C.; Metzger, F.; Bruns, A.; Bohrmann, B.; Kuennecke, B.; Von Kienlin, M. Cortical hypoperfusion in the B6.PS2APP mouse model for alzheimer’s disease: Comprehensive phenotyping of vascular and tissular parameters by MRI. Magn. Reson. Med. 2009, 62, 35–45. [Google Scholar] [CrossRef]
  158. Brendel, M.; Jaworska, A.; Grießinger, E.; Rötzer, C.; Burgold, S.; Gildehaus, F.J.; Carlsen, J.; Cumming, P.; Baumann, K.; Haass, C.; et al. Cross-sectional comparison of small animal [18f]-florbetaben amyloid-PET between transgenic AD mouse models. PLoS ONE 2015, 10, e0116678. [Google Scholar] [CrossRef] [Green Version]
  159. Grueninger, F.; Bohrmann, B.; Czech, C.; Ballard, T.M.; Frey, J.R.; Weidensteiner, C.; von Kienlin, M.; Ozmen, L. Phosphorylation of Tau at S422 is enhanced by Aβ in TauPS2APP triple transgenic mice. Neurobiol. Dis. 2010, 37, 294–306. [Google Scholar] [CrossRef]
  160. Loreth, D.; Ozmen, L.; Revel, F.G.; Knoflach, F.; Wetzel, P.; Frotscher, M.; Metzger, F.; Kretz, O. Selective degeneration of septal and hippocampal GABAergic neurons in a mouse model of amyloidosis and tauopathy. Neurobiol. Dis. 2012, 47, 1–12. [Google Scholar] [CrossRef]
  161. Kins, S.; Crameri, A.; Evans, D.R.H.; Hemmings, B.A.; Nitsch, R.M.; Götz, J. Reduced protein phosphatase 2A activity induces hyperphosphorylation and altered compartmentalization of Tau in transgenic mice. J. Biol. Chem. 2001, 276, 38193–38200. [Google Scholar] [CrossRef] [PubMed]
  162. Nyabi, O.; Pype, S.; Mercken, M.; Herreman, A.; Saftig, P.; Craessaerts, K.; Serneels, L.; Annaert, W.; De Strooper, B. No endogenous ab production in presenilin-deficient fibroblasts. Nat. Cell Biol. 2002, 4, E164. [Google Scholar] [CrossRef] [PubMed]
  163. Herreman, A.; Hartmann, D.; Annaert, W.; Saftig, P.; Craessaerts, K.; Serneels, L.; Umans, L.; Schrijvers, V.; Checler, F.; Vanderstichele, H.; et al. Presenilin 2 deficiency causes a mild pulmonary phenotype and no changes in amyloid precursor protein processing but enhances the embryonic lethal phenotype of presenilin 1 deficiency. Proc. Natl. Acad. Sci. USA 1999, 96, 11872–11877. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Lee, S.H.; Lutz, D.; Mossalam, M.; Bolshakov, V.Y.; Frotscher, M.; Shen, J. Presenilins regulate synaptic plasticity and mitochondrial calcium homeostasis in the hippocampal mossy fiber pathway. Mol. Neurodegener. 2017, 12, 48. [Google Scholar] [CrossRef] [PubMed]
  165. Gratuze, M.; Leyns, C.E.G.; Sauerbeck, A.D.; St-Pierre, M.-K.; Xiong, M.; Kim, N.; Remolina Serrano, J.; Tremblay, M.-È.; Kummer, T.T.; Colonna, M.; et al. Impact of TREM2R47H variant on tau pathology-induced gliosis and neurodegeneration. J. Clin. Investig. 2020. [Google Scholar] [CrossRef]
  166. Balu, D.; Karstens, A.J.; Loukenas, E.; Maldonado Weng, J.; York, J.M.; Valencia-Olvera, A.C.; LaDu, M.J. The role of APOE in transgenic mouse models of AD. Neurosci. Lett. 2019, 707, 134285. [Google Scholar] [CrossRef]
  167. Hargis, K.E.; Blalock, E.M. Transcriptional signatures of brain aging and alzheimer’s disease: What are our rodent models telling us? Behav. Brain Res. 2017, 322, 311–328. [Google Scholar] [CrossRef]
Figure 1. Familial Alzheimer’s disease (FAD)-presenilin-2 (PS2) alters multiple Ca2+ signalling pathways. The cartoon represents different intracellular membrane localizations of FAD-PS2, its interactions with several components of the molecular Ca2+ toolkit, and multiple Ca2+ signalling pathways that are altered by its action. See text for details. ER, endoplasmic reticulum; mGA, medial-Golgi apparatus; tGA, trans-Golgi apparatus; MIT, mitochondrion.
Figure 1. Familial Alzheimer’s disease (FAD)-presenilin-2 (PS2) alters multiple Ca2+ signalling pathways. The cartoon represents different intracellular membrane localizations of FAD-PS2, its interactions with several components of the molecular Ca2+ toolkit, and multiple Ca2+ signalling pathways that are altered by its action. See text for details. ER, endoplasmic reticulum; mGA, medial-Golgi apparatus; tGA, trans-Golgi apparatus; MIT, mitochondrion.
Cells 09 02166 g001
Figure 2. Functional consequences of dysregulated Ca2+ signalling induced by FAD-PS2. The cartoon represents the major dysfunctions linked to the expression of FAD-PS2 mutants at both the cellular and brain network levels. (A) Decreased store-operated Ca2+ entry (SOCE) potentiates amyloid precursor protein (APP) processing and Aβ42 production. (B) FAD-PS2-N141I-based mice show altered neuronal circuits (decreased phase-amplitude coupling between cortical slow oscillations and hippocampal fast gamma frequencies). (C) Decreased mitochondrial Ca2+ signalling and pyruvate uptake impair mitochondrial metabolism and cell bioenergetics. (D) Reduced endoplasmic reticulum (ER) Ca2+ release blocks the recruitment to autophagosomes of the Ras-associated binding protein RAB7 and their subsequent fusion with lysosomes. See text for details.
Figure 2. Functional consequences of dysregulated Ca2+ signalling induced by FAD-PS2. The cartoon represents the major dysfunctions linked to the expression of FAD-PS2 mutants at both the cellular and brain network levels. (A) Decreased store-operated Ca2+ entry (SOCE) potentiates amyloid precursor protein (APP) processing and Aβ42 production. (B) FAD-PS2-N141I-based mice show altered neuronal circuits (decreased phase-amplitude coupling between cortical slow oscillations and hippocampal fast gamma frequencies). (C) Decreased mitochondrial Ca2+ signalling and pyruvate uptake impair mitochondrial metabolism and cell bioenergetics. (D) Reduced endoplasmic reticulum (ER) Ca2+ release blocks the recruitment to autophagosomes of the Ras-associated binding protein RAB7 and their subsequent fusion with lysosomes. See text for details.
Cells 09 02166 g002

Share and Cite

MDPI and ACS Style

Pizzo, P.; Basso, E.; Filadi, R.; Greotti, E.; Leparulo, A.; Pendin, D.; Redolfi, N.; Rossini, M.; Vajente, N.; Pozzan, T.; et al. Presenilin-2 and Calcium Handling: Molecules, Organelles, Cells and Brain Networks. Cells 2020, 9, 2166. https://doi.org/10.3390/cells9102166

AMA Style

Pizzo P, Basso E, Filadi R, Greotti E, Leparulo A, Pendin D, Redolfi N, Rossini M, Vajente N, Pozzan T, et al. Presenilin-2 and Calcium Handling: Molecules, Organelles, Cells and Brain Networks. Cells. 2020; 9(10):2166. https://doi.org/10.3390/cells9102166

Chicago/Turabian Style

Pizzo, Paola, Emy Basso, Riccardo Filadi, Elisa Greotti, Alessandro Leparulo, Diana Pendin, Nelly Redolfi, Michela Rossini, Nicola Vajente, Tullio Pozzan, and et al. 2020. "Presenilin-2 and Calcium Handling: Molecules, Organelles, Cells and Brain Networks" Cells 9, no. 10: 2166. https://doi.org/10.3390/cells9102166

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