Skip to main content
The Journal of Physiology logoLink to The Journal of Physiology
. 2013 Jul 15;592(Pt 2):281–293. doi: 10.1113/jphysiol.2013.257527

Calcium regulation of neural rhythms, memory and Alzheimer's disease

Michael J Berridge 1
PMCID: PMC3922493  PMID: 23753528

Abstract

Alzheimer's disease (AD) begins with a decline in cognition followed by neuronal cell death and dementia. These changes have been linked to a deregulation of Ca2+ signalling caused by a progressive increase in the resting level of Ca2+, which may influence cognition by interfering with the rhythm rheostat that controls the sleep/wake cycle. The rise in resting levels of Ca2+ may not alter the processes of memory acquisition during consciousness (gamma and theta rhythms), but may duplicate some of the events that occur during the slow oscillations responsible for the twin processes of memory consolidation and memory erasure that occur during sleep. The persistent elevation in the resting level of Ca2+ induced by an accumulation of amyloid β (Aβ) oligomers duplicates a similar small global elevation normally restricted to the period of slow oscillations when memories are erased during sleep. In AD, such a rapid erasure of memories soon after they are acquired during the wake period means that they are not retained for consolidation during sleep. The Aβ deregulates Ca2+ signalling through direct effects on the neurons and indirectly by inducing inflammatory responses in the microglia and astrocytes. Some of these deleterious effects of Aβ may be alleviated by vitamin D.

Introduction

Neurons have highly developed Ca2+ signalling systems (Berridge, 2012) responsible for regulating neural functions such as brain rhythms, information processing, learning and memory. Remodelling of these Ca2+ signalling pathways that create inappropriate Ca2+ responses have been linked to many major neural diseases (Khachaturian 2013; LaFerla 1989; Stutzmann 2005; Thibault et al. 1992; Bezprozvanny & Mattson, 2012a; Berridge 1998, 2010,2011 b). In the case of Alzheimer's disease (AD), Ca2+ levels are set too high and this has an impact on many neural functions and particularly memory formation and consolidation. Why subtle deregulation of Ca2+ signalling can have such a marked effect on memory remains somewhat mysterious. Memory formation is a complex process that depends on memory acquisition during consciousness followed by memory consolidation and erasure during sleep. To obtain a better understanding of AD, therefore, the first part of this review explores the way Ca2+ signalling participates, in not only controlling brain rhythms, but also how it generates memories that occur during the markedly different brain rhythms that characterize the sleep–wake cycle. The remodelling of Ca2+ signalling in AD may explain the observation that this sleep–wake cycle is markedly disrupted in AD with an increase in wakefulness (Roh et al. 2008). The second part of the review explores the possibility that the deregulation of Ca2+ signalling that occurs in AD results in the continuous activation of a Ca2+-dependent memory erasure mechanism, which is normally restricted to the early phase of sleep. Such a mechanism could explain why memories are rapidly lost soon after they are acquired during periods of wakefulness.

Michael Berridge is an Emeritus Babraham Fellow at the Babraham Institute in Cambridge. He received his B.Sc. in 1960 from the University of Rhodesia and Nyasaland in Salisbury, Rhodesia and was awarded his Ph.D. from the University of Cambridge in 1964. His current research interest is focused on the role of inositol trisphosphate (InsP3) in controlling the spatiotemporal aspects of Ca2+ signalling with particular emphasis on neural signalling and neural diseases such as Alzheimer's disease, bipolar disorder and schizophrenia.

Neural rhythms and neural functions

The brain is highly rhythmical and the different neural rhythms that occur during the sleep–wake cycle regulate its multiple functions. These rhythms can be divided into the fast gamma (20–80 Hz), alpha (8–12 Hz) and theta (6–10 Hz) oscillations, which occur during the awake state, and the slower delta (1–4 Hz) and slow oscillations (<1 Hz) that occur during sleep (Fig. 1). These different oscillatory modes are regulated by the ascending arousal system that consists of a heterogeneous population of neurons that project their axons throughout the brain where transmitters such as orexin, acetylcholine, noradrenaline, 5-hydroxytryptamine, histamine and dopamine are released on to the excitatory and inhibitory neurons that constitute functional neural circuits (Pace-Schott & Hobson 2007; Datta 1996). These transmitters then act on receptors coupled to signalling pathways to adjust the level of the tonic excitatory drive that regulates these different oscillatory states (Fig. 1).

Figure 1.

Figure 1

Tonic excitatory drive and the control of neuronal rhythms. Neurons of the ascending arousal systems release transmitters such as orexin, acetylcholine (ACh), 5-hydroxytryptamine (5-HT), dopamine (DA) and noradrenaline (NA), which induce signalling systems to control neural rhythms that occur during the sleep/wake cycle. The tonic excitatory drive mechanism depends on membrane depolarization driven by different signalling mechanisms. Hydrolysis of the phospholipid PtdIns4,5P2 has two effects. First, it closes the KV 7.2/7.3 channels responsible for the M current. Switching off this M current depolarizes the membrane to increase neuronal activity. Secondly, the formation of inositol 1,4,5-trisphosphate (InsP3) releases Ca2+ that stimulates the Ca2+-activated non-selective cation (CAN) channel. The CAN channel can also be activated by Ca2+ entering through the voltage-operated Ca2+ (VOC) channel. The NA and DA act through the cyclic AMP signalling pathway to enhance the activity of the hyperpolarizing-activated cyclic nucleotide-gated (HCN) channel responsible for the depolarizing Ih current.

Variations in the activity of this tonic excitatory drive functions much like a rhythm rheostat in that it controls a hierarchy of rhythms with the lowest frequencies occurring during sleep that are then switched to the higher frequency rhythms of the wake state (Fig. 1). The Ca2+ signalling mechanisms control the rhythm rheostat and have to operate within the parameters of the ongoing rhythms to regulate memory formation, consolidation and erasure. Memory acquisition occurs during fast gamma rhythms during consciousness, while the slow oscillations mediate memory consolidation and erasure during sleep. The sleep–wake cycle is markedly disrupted in AD with an increase in wakefulness associated with a decrease in the slow oscillation responsible for non-rapid eye movement (NREM) sleep rhythms (Roh et al. 2008).

Gamma and theta rhythms and calcium signalling mechanisms

The fast theta and gamma oscillations, which are highly synchronous throughout the brain, are generated by a typical network oscillator consisting of fast spiking inhibitory interneurons and excitatory neurons that interact with each other through a positive/negative feedback loop (Fig. 2). Each inhibitory interneuron controls the activity of a large array of excitatory neurons all of which send axon collaterals back to the inhibitory interneuron. Most information is available for the gamma rhythms where the interneurons fire an action potential on each cycle and this induces a brief hyperpolarization in all the excitatory neurons that occurs synchronously over extensive areas of the brain (Hájos & Paulsen 2013). In contrast to this continuous firing of the inhibitory neurons, the excitatory neurons fire much less frequently with each action potential occurring within a narrow time window towards the end of each pacemaker depolarization (Fig. 2). This gamma oscillation synchronization provides a precise timing mechanism that enables excitatory neurons located in different parts of the brain to communicate with each other by firing together as part of a circuit as they process sensory information. This action potential coincidence is also crucial for triggering the input-specific Ca2+ transients responsible for memory formation.

Figure 2.

Figure 2

Neuronal network gamma oscillations. Most neural circuits consist of fast spiking inhibitory interneurons (red) and excitatory neurons (green) interacting with each other through a positive/negative feedback loop (see inset at the bottom). Each interneuron controls the activity of many excitatory neurons (red arrows) all of which send axon collaterals back to the inhibitory interneuron (green arrow). The interneuron fires an AP on each gamma cycle and this serves to induce synchronous hyperpolarizations in all the excitatory neurons. The excitatory neurons fire much less frequently towards the end of the pacemaker depolarization. The inhibitory interneuron registers each of these APs as a small EPSP that sum to activate the interneuron to fire an AP to initiate another gamma oscillatory cycle. ACh, acetylcholine; AP, action potential; EPSP, excitatory postsynaptic potential.

Important Ca2+ signalling events occur when excitatory neurons, which are connected together as part of a neuronal circuit, fire action potentials that are synchronized by the gamma oscillation rhythm. Such action potential coincidence (Fig. 2) is important for memory formation through a process of spike time-dependent plasticity, which favours the induction of long-term potentiation (LTP) (Caporale & Dan 2010). Depolarization of a neuron while it receives a pulse of glutamate from another neuron serves to open the NMDA receptors (NMDARs) on the activated spines to allow rapid entry of Ca2+ to trigger LTP (Fig. 3A). A form of NMDAR-independent synaptic plasticity is induced by Ca2+ entering through CaV1.2 channels (Moosmang et al. 2013). The brief high concentration spike of Ca2+, which is highly localized within the activated spine, induces at least three immediate biochemical events: phosphorylation of the AMPA receptors (AMPARs), exocytosis of vesicles containing new AMPARs and activation of actin polymerization resulting in a change in the shape of the spine. These biochemical events are the basis of new memories (Citri & Malenka 2008) that are then retained until either consolidated or erased during the slow oscillations that occur during NREM sleep. The mechanisms responsible for memory formation seem to operate normally in AD, but are not retained long enough to be consolidated during sleep. A clue as to why these memories are rapidly erased in AD has emerged from analysing the mechanisms responsible for memory erasure that are normally restricted to the period of slow oscillations that occur during NREM sleep.

Figure 3.

Figure 3

Ca2+-induced synaptic plasticity during memory formation (A) and memory erasure (B). A, the opening of NMDA receptors (NMDARs) or Cav1.2 L-type Ca2+ channels generate large local increases in Ca2+ within specific spines (see inset). This localized Ca2+ signal induces long-term potentiation (LTP) that depends on AMPA receptor (AMPAR) phosphorylation by CaMKII, AMPAR insertion that increases the number of receptors in the postsynaptic membrane, actin remodelling that changes spine morphology and protein synthesis at polyribosomes located at the base of each spine. B, lower levels of Ca2+ in the spines, which may depend on Ca2+ diffusing in from the global dendritic Ca2+ waves of the slow oscillations that occur during sleep, activates long-term depression (LTD) that depends on calcineurin (CaN) acting to reverse many of the processes that occur during LTP.

Slow oscillation and calcium signalling mechanisms

The slow (<1 Hz) oscillations, which occur during NREM sleep, have been linked to both memory consolidation and erasure (Diekelmann & Born 2011). The onset of these endogenous slow oscillations, which occurs when the tonic excitatory drive is reduced at the onset of sleep (Fig. 1), has typical UP states when the membrane potential (Vm) is depolarized to approximately –65 mV that alternates with DOWN states where the membrane is hyperpolarized by 10–15 mV. These slow oscillations occur in near synchrony throughout the brain and this enables neurons to communicate with each other as they fire rapid bursts of action potentials on the crest of each UP state (Massimini et al. 2013). This synchrony is achieved through a slow wave propagation mechanism whereby the action potentials that occur at the start of the UP state in one neuron entrains neighbouring neurons to initiate their slow oscillation thereby a wave of excitation propagates through the brain travelling at approximately 3 m s−1 in an anterior–posterior direction. The signalling mechanisms that occur during these slow oscillations are responsible for both memory consolidation and erasure.

A decline in the tonic excitatory drive that occurs during NREM sleep turns down the rhythm rheostat to set the stage for oscillations to occur (Fig. 1). A decrease in the activity of the metabotropic transmitters that stimulate the hydrolysis of PtdIns4,5P2 enables this lipid to accumulate such that it can open the KV7.2/KV7.3 channels resulting in an increase in the M current that provides the membrane hyperpolarization that drives the DOWN state. The endogenous oscillator that drives the alternating UP and DOWN states depends on an interaction between different channels (Fig. 4) (Crunelli et al. 2008, 2007). During the DOWN state, there is steady pacemaker depolarization that depends on both a hyperpolarizing-activated cyclic nucleotide-gated 1 (HCN1) channel and a Ca2+-dependent non-selective cation channel (CAN). The latter is activated directly by Ca2+ whereas the HCN1 is modulated indirectly by Ca2+ acting on adenylyl cyclase (AC) to provide cyclic AMP. This cyclic AMP does not activate HCN1 directly, but it alters the open probability by shifting the activation curve towards more depolarizing potentials.

Figure 4.

Figure 4

This hypothetical scheme describes the main ionic channels and signalling mechanisms that have been implicated in generating the slow oscillations characterized by periodic UP (red panels) and DOWN (green panels) states. The CaV3 T-type channels, which are located on the dendrites, are switched on at the beginning of an UP state to provide a low-threshold Ca2+ spike(s), but then inactivate to become a persistent inward current, also known as the T window current, to provide a tonic depolarization that helps to maintain the UP state. Opening of these T channels at the beginning of each UP state results in a global elevation of Ca2+ that spreads rapidly down the dendrites and into the spines where it may act to induce long-term depression (see Fig. 3B). ACh, acetylcholine; CaN, calcineurin; ER, endoplasmic reticulum; 5-HT, 5-hydroxytryptamine; InsP3, inositol 1,4,5-trisphosphate; HCN, hyperpolarizing-activated cyclic nucleotide-gated (channel).

The beginning of the UP state is characterized by the opening of CaV3 T-type channels that produce a low-threshold Ca2+ spike(s), but these channels then inactivate to become a persistent inward current, also known as the T window current, to provide a tonic depolarization that helps to maintain the UP state (Crunelli et al. 2008, 2007). The opening of these T type channels, which are located on the soma and dendrites, results in a global elevation of Ca2+ that not only spreads through the dendrites, but it also diffuses into the spines (Errington et al. 2011). During the course of a slow oscillation, therefore, there are two important Ca2+ signalling events taking place that may have a major impact on memory formation. First, the global elevation that spreads into all of the spines may function to activate long-term depression (LTD) to erase temporary memories (Fig. 3B). Secondly, it is likely that there are localized pulses of Ca2+ within specific spines during rapid action potentials that occur on the crest of each slow oscillation as memories that are reactivated in the hippocampus are redistributed to the cortex where they are consolidated into longer-term memories during sleep.

It has been proposed that memories acquired during the wake period are stored temporarily in the synaptic connections of relevant brain circuits located mainly in the hippocampus. Some of these temporary memories represent novel information that will be retained, but many other memory traces are erased during sleep to avoid cluttering up the brain with irrelevant information. This synaptic homeostasis hypothesis proposes that the redundant information stored in those potentiated synaptic connections are returned to baseline levels during sleep (Tononi & Cirelli, 2006; Vyazovskiy et al. 2009). The way in which some memories are consolidated and retained while others are erased is not known. The following speculation suggests that these two processes may run concurrently, even within individual neurons, due to a spatial separation of the Ca2+ signalling events during slow oscillations.

The fast action potentials that occur during the UP state of the slow oscillation (Fig. 4) probably reflects the neural activity responsible for memory consolidation. The spindles and ripples that appear in EEG recordings during sleep are thought to reflect the hippocampal–cortical dialogue as labile information stored in the hippocampus is transferred for more permanent storage in the cortex (Ji & Wilson 2013). The ripples represent high-frequency bursts of action potentials as memories are reactivated in the hippocampus, whereas the spindles result from the oscillatory firing of thalamocortical neuronal loops as this information is received and consolidated in the cortex (Diekelmann & Born 2011). Activation and transfer of these memories may depend on the same spike time-dependent plasticity mechanisms described earlier (Fig. 3A). The large and persistent pulsing of Ca2+, restricted to activated spines, will contribute to consolidation by stimulating protein synthesis by polysomes located at the base of each spine.

At the same time that novel memories are being consolidated, other memories are being deleted and it is conceivable that these processes may run concurrently in individual neurons. The erasure mechanism is not fully understood, but Errington et al. (2011) have suggested that the global elevation in Ca2+ that occurs during the slow oscillation might play a role in the mechanism of ‘homeostatic synaptic plasticity’. The most probable explanation for this is that the global elevation of Ca2+, which is known to invade the spines, will activate LTD. The specific proposal is that Ca2+ in the dendrites diffuses into the spine where it may reach levels in the 300–500 nm range that will activate LTD resulting in memory erasure (Fig. 3B). This Ca2+-dependent erasure depends on activation of calcineurin that reverses the three processes that occurred during LTP, i.e. AMPARs are dephosphorylated, AMPARs are retrieved from the spine surface through endocytosis and the actin filaments are depolymerized. In this way, the conundrum as to how memory consolidation and erasure can run concurrently can be resolved if high levels of Ca2+ necessary for LTP are localized to those spines undergoing consolidation, whereas all redundant memories can be erased en masse through the global elevation of lower levels of Ca2+ that pervade the spines to activate LTD. The loss of memory in AD may result from a similar erasure being activated continuously during the wake period when memories are acquired.

Dysregulation of Ca2+ signalling and Alzheimer's disease

The development of AD is driven by the accumulation of amyloid β (Aβ) oligomers, which are a neuron-derived pathogenic factor that brings about the loss of memory and neuronal cell death that characterizes the progression of AD. The Ca2+ hypothesis of AD suggests that these deleterious effects of Aβ depend on a dysregulation of Ca2+ signalling (Khachaturian 2013; LaFerla 1989; Stutzmann 2005; Thibault et al. 1992; Bezprozvanny & Mattson 2012a; Stutzmann & Mattson 2012; Berridge 2013, 1998, 2011b). The basic idea is that abnormal amyloid metabolism induces an upregulation of neuronal Ca2+ signalling that is responsible for the initial decline in memory and subsequent apoptosis. When Ca2+ is measured in the spines and dendrites of cortical pyramidal neurons of transgenic mice, there was a higher than normal resting level in those neurons located close to amyloid deposits (Kuchibhotla et al. 2007). Similarly, the resting level of Ca2+ in the cortical neurons of triple transgenic AD animals was 247 nmol l−1, which was twice that found in the non-transgenic controls (110 nmol l−1) (Lopez et al. 2011). In addition, there is increasing evidence that Aβ also acts on neighbouring microglial cells and astrocytes (Abramov et al. 2004; Saijo & Glass, 2008) to induce local inflammatory responses that contributes to Ca2+ signalling deregulation. The following sequence of events attempts to explain these multiple actions of Aβ on neurons, microglia and astrocytes to induce the upregulation of Ca2+ signalling that may be responsible for AD (the numbers on Fig. 5 correspond to those outlined below):

Figure 5.

Figure 5

Ca2+ hypothesis of AD. The development of AD is induced by the accumulation of Aβ oligomers that have a number of actions. The Aβ can act directly on the neurons to bring about elevations in Ca2+ that have been linked to the initial phase of memory loss and the subsequent increase in apoptosis that characterizes the development of AD. The Aβ can also induce inflammatory responses in neighbouring microglia and astrocytes that activate processes that enhance this dysregulation of Ca2+ signalling. Vitamin D3 may alleviate the development of AD by inhibiting the inflammatory responses and by increasing the expression of processes that reduce the elevation of Ca2+. Aβ, amyloid β; AD, Alzheimer's disease; GSH, glutathione; IL, interleukin; InsP3, inositol 1,4,5-trisphosphate; mTOR, mammalian target of rapamycin; PLC, phospholipase C; PMCA, plasma membrane Ca2+-ATPases; ROS, reactive oxygen species; RYR, ryanodine receptor; TNF, tumor necrosis factor; VDR, vitamin D receptor; VDRE, vitamin D response element.

  1. Neuronal Ca2+ signalling deregulation may depend on changes in both the entry of external Ca2+ and its release from internal stores. The Aβ oligomers that accumulate outside diseased neurons can bring about an elevation in Ca2+ through different mechanisms. They can be inserted into the membrane to form channels (Demuro et al. 2010) or they can activate the Ca2+-sensitive receptor (CaSR) to increase the level of InsP3 (Ye et al. 2006; Chiarini et al. 2008; Armato et al. 2004). The CaSR is coupled to phospholipase C through the G protein Gq, which is inhibited by the regulator of G protein signalling 4 (RGS4). The level of RGS4 is reduced in the human AD brain and this may further enhance the generation of InsP3 (Emilsson et al. 2010). Injection of Aβ into Xenopus oocytes stimulated the production of InsP3 through a G protein-dependent activation of phospholipase C (Demuro & Parker 2007) providing further evidence that some of the actions of Aβ might be mediated by an increase in the InsP3/Ca2+ signalling pathway.

  2. An increase in the formation of InsP3 will enhance the amount of Ca2+ being released from the endoplasmic reticulum by the InsP3 receptors (InsP3Rs). Indeed, a feature of AD is an increase in the activity of InsP3Rs (Cheung et al. 2013; Müller et al. 2006). This elevation in InsP3-dependent Ca2+ signalling, which needs to be turned down for slow oscillations to occur during sleep, may explain the observation that the sleep–wake cycle is markedly disrupted in AD with an increase in wakefulness (Roh et al. 2008). In transgenic mice, an increase in amyloids may induce global elevations of Ca2+ through a burst firing-mediated mechanism that depends on glutamate activating an InsP3/Ca2+ signalling pathway (Czarnecki et al. 2006). Expression of the Cav1.2 L-type Ca2+ channel, which has been implicated in LTP induction (Moosmang et al. 2013) (Fig. 3A), is induced by Aβ (Webster et al. 2012; Dursun et al. 2009) and this will enhance the release of Ca2+ from ryanodine receptors (RYRs). Such an action would be enhanced further by the increased expression of the RYR, particularly the RYR3 isoform (Supnet et al. 2007). Neuronal levels of the Ca2+ buffer calbindin-28 k are known to be reduced in AD (Sutherland et al. 2011). In addition, Aβ may also reduce Ca2+ extrusion from the cell by inhibiting both the plasma membrane Ca2+-ATPase (PMCA) and the Na+/K+-ATPase that maintains the Na+/Ca2+ exchanger (NCX) (Mark et al. 2011). Thus, there are a number of mechanisms that could contribute to the upregulation of Ca2+ signalling to account for persistent elevation in the resting level of Ca2+ (Kuchibhotla et al. 2007; Lopez et al. 2011).

  3. This dysregulation of neuronal Ca2+ signalling seems to be exacerbated by Aβ-induced neuroinflammation that occurs during AD (Saijo & Glass 2008). Aβ-induced Ca2+ signals can enhance microglial inflammatory responses by increasing the release of cytokines and reactive oxygen species (ROS) (Farber & Kettenmann 2006). The Aβ also acts through CaSRs to produce InsP3 that then releases Ca2+ from internal stores (Lee et al. 2008). Depletion of these stores then triggers store-operated Ca2+ entry through the Orai1 channel (Ohana et al. 2011) that is maintained by the hyperpolarization induced by the calcium-activated potassium channel KCa3.1. Microglial-dependent neurotoxicity could be reduced in vivo by inhibiting these KCa3.1 channels with triarylmethane-34 (Kaushal et al. 2007; Maezawa et al. 2008) thus emphasizing the significance of Ca2+ in regulating neuroinflammation. Drugs thought to reduce inflammation such as Ro5-4864, which is a ligand for the translocator protein, reduce both the accumulation of Aβ and the decline in cognition in transgenic AD mice (Barron et al. 2000).

  4. One of the consequences of the Aβ-dependent elevation of microglial Ca2+ is activation of the inflammasome. The oligomers that are taken up in the phagosome vesicles enter the cytosol to increase NLRP3 activity resulting in stimulation of caspase-1 that cleaves pro-interleukin-1β (IL-1β) to form IL-1β (Fig. 5). The inflammasome inhibits phagocytic clearance of Aβ and has been strongly implicated in AD (Heneka et al. 2008). The triggering receptor expressed in myeloid cells 2 (TREM-2), which functions as a negative regulator of innate immunity, suppresses the ability of the microglia to release inflammatory mediators such as tumour necrosis factor (TNF)α. TREM2 is a transmembrane glycoprotein that associates with DNAX-activating protein 12. A variant of TREM2, which reduces the anti-inflammatory role of TREM2, is associated with a markedly increased risk of developing AD (Guerreiro et al. 2002; Jonsson et al. 2013).

  5. Aβ can also induce microglial inflammation by stimulating toll-like receptors (TLR-2 and TLR-4). Polymorphisms in these receptors have been associated with an increased susceptibility and progression of AD (Minoretti et al. 2011; Yu et al. 2009, 1997). Activation of the TLR-2/4 receptors can have both beneficial and deleterious actions. With regard to the former, TLR-2/4 receptors stimulate phagocytosis that removes and destroys Aβ. The deleterious effects depend upon activation of the NF-κB signalling pathway that increase release of proinflammatory mediators such as TNFα and ROS, all of which can enhance neuronal Ca2+ signalling. Increased susceptibility and progression of AD have been linked to polymorphisms in TNFα (Di Bona et al. 2013; Yang et al. 2004). The monophosphoryl lipid A has an interesting property of being able to promote the beneficial phagocytic mechanism while not inducing the deleterious inflammatory response and thus may be an effective treatment for AD (Michaud et al. 2004). Omega-3 fatty acids can also enhance phagocytosis of Aβ to reduce the formation of proinflammatory cytokines (Hjorth et al. 2009).

  6. The TNFα and IL-1β released from the microglia can have a number of actions. The TNFα binds to the TNF receptor to contribute to neuronal cell death by activating apoptosis. The TNFα can also enhance Ca2+ signalling by acting through the JNK signalling pathway to increase the expression of InsP3R1 by phosphorylating the transcription factor specificity protein 1 (Sp1) (Park et al. 2009). Both TNFα and IL-1β can also influence memory by altering the electrophysiological correlates of LTP and LTD (Cunningham et al. 2005; Albensi & Mattson 2004).

  7. The exogenous ROS, which diffuses into the neuron, will add to that being produced by the mitochondria and can enhance Ca2+ signalling in several ways. It can increase the sensitivity of both the InsP3Rs and RYRs to increase the release of Ca2+ while inhibiting the PMCA Ca2+ pump (Lock et al. 2012). These two effects of ROS, which will increase the resting level of Ca2+, will potentially set up a positive feedback system in that the excess Ca2+ will increase mitochondrial ROS formation (Müller et al. 2006). Inhibition of ROS formation by a mitochondrial-targeted antioxidant MitoQ prevents the cognitive decline in a transgenic mouse model of AD (McManus et al. 1995).

  8. Aβ acts on the astrocytes to induce an inflammatory response and the resulting increase in ROS decreases the level of the antioxidant glutathione (GSH), which has serious repercussions for the neuron as it receives its GSH from the astrocytes (Abramov et al. 2004). A decrease in neuronal GSH levels will enable ROS to have a greater impact on Ca2+ levels as outlined above.

  9. A number of mechanisms have been proposed to describe how the elevation in the resting level of Ca2+ accounts for symptoms of AD. The characteristic loss of memory that occurs in the early stages of AD may be driven by the abnormal resting levels of Ca2+, which is manifest either as a persistent elevation in the level of Ca2+ in the dendrites and spines (Kuchibhotla et al. 2007) or as an increase in spontaneous Ca2+ transients (Busche et al. 2008, 2006). This increase in Ca2+ signalling, may disrupt cognition by activating LTD through the mechanism described earlier (Fig. 3B) (Berridge, 2013, 1998). As the disease progresses, elevation of Ca2+ will begin to rise further to a point where it will activate apoptosis resulting in the neuronal cell loss responsible for the final stages of dementia.

  10. A marked feature of AD is a decline in autophagy (Son et al. 2011), which seems to be associated with an increase in the activity of the mammalian target of rapamycin (mTOR) (Caccamo et al. 2008, 2010). The InsP3R is known to play a role in autophagy by assembling a complex containing regulators such as Beclin-1, Bcl-2 and hVps34 (Criollo et al. 2009; Vicencio et al. 2006). The level of Beclin-1, which is a key component of the autophagy complex, is known to be reduced in AD (Pickford et al. 2009). The decline in autophagy in AD may be related to an increase in InsP3 that disrupts the autophagic complex by binding to the InsP3R. The drug Li+, which is known to reduce the risk of developing AD (Nunes et al. 2005), can reduce this inhibitory effect by lowering the level of InsP3 (Sarkar et al. 2012). Autophagy may also be reduced in AD by the elevated levels of Ca2+ that can disrupt the complex by activating hVps34 (Gulati et al. 2011). The activation of hVps34 may also account for the increase in mTOR (Gulati et al. 2011) that could explain the decline of autophagy in AD. The cognitive decline in mouse models of AD is reduced by rapamycin, which inhibits the activity of mTOR (Caccamo et al. 2008, 2010). Another role for mTOR is to phosphorylate Tau to increase its pathological role in AD. The elevation of Ca2+ can also stimulate CaMKK2 to increase the activity of AMPK that then enhances the phosphorylation of Tau thus contributing to the symptoms of AD (Mairet-Coello et al. 2013).

  11. There is considerable evidence for a link between vitamin D deficiency and the onset of AD and other neurodegenerative diseases such as multiple sclerosis and Parkinson's disease (Garcion et al. 2012; Tuohimaa et al. 2007; Berridge 2011b; Wang et al. 2009; Lu'o'ng & Nguyên 2011). Expression of the vitamin D receptor (VDR) is reduced in the hippocampus of patients with AD (Sutherland et al. 2011) and VDR polymorphisms have been identified as risk factors for AD (Lehmann et al. 2002; Wang et al. 2009).There is evidence that Aβ acts to both reduce the expression of VDR while increasing the expression of the Cav2.1 Ca2+ channel.

  1. Administration of vitamin D can reverse many of the changes induced by Aβ. In cultured primary neurons, vitamin D acts to increase the expression of the VDR and it reduces the expression of the Cav1.2 L-type Ca2+ channel (Brewer et al. 2012b; Taniura et al. 2006; Dursun et al. 2009; Gezen-Ak et al. 2006). In the intestine, vitamin D is known to increase the expression of proteins such as PMCA, NCX1 and Ca2+ buffers such as calbindin-28 k and parvalbumin (Wasserman 2008; Pérez et al. 2002). It seems reasonable to propose that vitamin D may reduce the risk of AD by promoting the expression of those proteins, such as PMCA and NCX1, that act to lower the level of intracellular Ca2+ (Berridge 2011b). Vitamin D can also have a beneficial effect by dampening down the microglial inflammatory responses to reduce the formation of TNFα.

In summary, the build-up of Aβ oligomers during the onset of AD has a profound effect on the activity of the local community of cells in the brain. The inflammatory response in both the microglia and astrocytes contribute to dysregulation of neural Ca2+ signalling that seems to be one of the major factors in the development of AD. It is argued that in the early stages of AD, this alteration in signalling is manifest as a persistent elevation of the resting level of Ca2+ that results in memories acquired during the wake period being rapidly erased before they can be consolidated during sleep. Vitamin D may play a critical role in memory retention by regulating the expression of the Ca2+ components necessary to maintain low resting levels of Ca2+.

Acknowledgments

I would like to thank Vincenzo Crunelli for helping me to understand the mechanism of slow oscillations and for suggesting improvements to some of the figures.

Glossary

AD

Alzheimer's disease

amyloid β

CAN

Ca2+-activated non-selective cation channel

CaSR

Ca2+-sensitive receptor

GSH

glutathione

HCN1

hyperpolarizing-activated cyclic nucleotide-gated 1

InsP3

inositol 1,4,5-trisphosphate;

LTD

long-term depression

LTP

long-term potentiation

mTOR

mammalian target of rapamycin

NCX

Na+/Ca2+ exchanger

NREM

non-rapid eye movement

NLRP

NACHT, LRR and PYD domains-containing protein

PMCA

plasma membrane Ca2+-ATPases

PtdIns4,5P2

phosphatidylinositol 4,5-bisphosphate

RGS4

regulator of G protein signalling 4

ROS

reactive oxygen species

RYR

ryanodine receptor

TREM-2

triggering receptor expressed in myeloid cells 2

TNFα

tumour necrosis factor α

VDR

vitamin D receptor

Competing interests

None declared.

References

  1. Abramov AY, Canevari L, Duchen MR. Calcium signals induced by amyloid β peptide and their consequences in neurons and astrocytes in culture. Biochim Biophys Acta. 1742:81–87. doi: 10.1016/j.bbamcr.2004.09.006. [DOI] [PubMed] [Google Scholar]
  2. Albensi BC, Mattson MP. Evidence for the involvement of TNF and NF-κB in hippocampal synaptic plasticity. Synapse. 35:151–159. doi: 10.1002/(SICI)1098-2396(200002)35:2<151::AID-SYN8>3.0.CO;2-P. [DOI] [PubMed] [Google Scholar]
  3. Armato U, Bonafini C, Chakravarthy B, Pacchiana R, Chiarini A, Whitfield JF, Dal Prà I. The calcium-sensing receptor: a novel Alzheimer's disease crucial target. J Neurol Sci. 2004;322:137–140. doi: 10.1016/j.jns.2012.07.031. [DOI] [PubMed] [Google Scholar]
  4. Barron AM, Garcia-Segura LM, Caruso D, Jayaraman A, Lee J-W, Melcangi RC, Pike CJ. Ligand for translocator protein reverses pathology in a mouse model of Alzheimer's disease. J Neurosci. 2000;33:8891–8897. doi: 10.1523/JNEUROSCI.1350-13.2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Berridge MJ. Neural calcium signalling. Neuron. 2012;21:13–26. doi: 10.1016/s0896-6273(00)80510-3. [DOI] [PubMed] [Google Scholar]
  6. Berridge MJ. Calcium hypothesis of Alzheimer's disease. Pflügers Arch. 2013;459:441–449. doi: 10.1007/s00424-009-0736-1. [DOI] [PubMed] [Google Scholar]
  7. Berridge MJ. Calcium signalling and Alzheimer's disease. Neurochem Res. 1998;36:1149–1156. doi: 10.1007/s11064-010-0371-4. [DOI] [PubMed] [Google Scholar]
  8. Berridge MJ. Calcium signalling remodelling and disease. Biochem Soc Trans. 2010;40:297–309. doi: 10.1042/BST20110766. [DOI] [PubMed] [Google Scholar]
  9. Berridge MJ. Dysregulation of neural calcium signalling in Alzheimer disease, bipolar disorder and schizophrenia. Prion. 2011;6:1–12. doi: 10.4161/pri.21767. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Bezprozvanny I, Mattson MP. Neuronal calcium mishandling and the pathogenesis of Alzheimer's disease. Trends Neurosci. 2012a;31:454–463. doi: 10.1016/j.tins.2008.06.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Brewer LD, Porter NM, Kerr DS, Landfield PW, Thibault O. Chronic 1α,25-(OH)2 vitamin D3 treatment reduces Ca2+-mediated hippocampal biomarkers of aging. Cell Calcium. 2012b;40:277–286. doi: 10.1016/j.ceca.2006.04.001. [DOI] [PubMed] [Google Scholar]
  12. Busche MA, Eichhoff G, Adelsberger H, Abramowski D, Wiederhold KH, Haass C, Staufenbiel M, Konnerth A, Garaschuk O. Clusters of hyperactive neurons near amyloid plaques in a mouse model of Alzheimer's disease. Science. 2008;321:1686–1689. doi: 10.1126/science.1162844. [DOI] [PubMed] [Google Scholar]
  13. Busche MA, Chen X, Henning HA, Reichwald J, Staufenbiel M, Sakmann B, Konnerth A. Critical role of soluble amyloid-β for early hippocampal hyperactivity in a mouse model of Alzheimer's disease. Proc Natl Acad Sci U S A. 2006;109:8740–8745. doi: 10.1073/pnas.1206171109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Caccamo A, Majumder S, Richardson A, Strong R, Oddo S. Molecular interplay between mammalian target of rapamycin (mTOR), amyloid-beta, and Tau: effects on cognitive impairments. J Biol Chem. 2008;285:13107–13120. doi: 10.1074/jbc.M110.100420. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Caccamo A, Magrì A, Medina DX, Wisely EV, López-Aranda MF, Silva AJ, Oddo S. mTOR regulates tau phosphorylation and degradation: implications for Alzheimer's disease and other tauopathies. Aging Cell. 2012;12:370–380. doi: 10.1111/acel.12057. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Caporale N, Dan Y. Spike timing-dependent plasticity: a Hebbian learning rule. Annu Rev Neurosci. 2010;31:25–46. doi: 10.1146/annurev.neuro.31.060407.125639. [DOI] [PubMed] [Google Scholar]
  17. Cheung K-H, Shineman D, Müller M, Cárdenas C, Mei L, Yang J, Tomita T, Iwatsubo T, Lee VM, Foskett JK. Mechanisms of Ca2+ disruption in Alzheimer's disease by presenilin regulation of InsP3 receptor channel gating. Neuron. 2013;58:871–883. doi: 10.1016/j.neuron.2008.04.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Chiarini A, Dal Pra I, Marconi M, Chakravarthy B, Whitfield JF, Armato U. Calcium-sensing receptor (CaSR) in human brain's pathophysiology: roles in late-onset Alzheimer's disease (LOAD) Curr Pharm Biotechnol. 2008;10:317–326. doi: 10.2174/138920109787847501. [DOI] [PubMed] [Google Scholar]
  19. Citri A, Malenka RC. Synaptic plasticity: Multiple forms, functions, and mechanisms. Neuropsychopharmacology. 2008;33:18–41. doi: 10.1038/sj.npp.1301559. [DOI] [PubMed] [Google Scholar]
  20. Criollo A, Maiuri MC, Tasdemir E, Vitale I, Fiebig AA, Andrews D, Molgó J, Díaz J, Lavandero S, Harper F, Pierron G, di Stefano D, Rizzuto R, Szabadkai G, Kroemer G. Regulation of autophagy by the inositol trisphosphate receptor. Cell Death Differ. 2009;14:1029–1039. doi: 10.1038/sj.cdd.4402099. [DOI] [PubMed] [Google Scholar]
  21. Crunelli V, Tóth TI, Cope DW, Blethyn K, Hughes S. The ‘window' T-type calcium current in brain dynamics of different behavioural states. J Physiol. 2008;562:121–129. doi: 10.1113/jphysiol.2004.076273. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Crunelli V, Cope DW, Hughes SW. Thalamic T-type Ca2+ channels and NREM sleep. Cell Calcium. 2007;40:175–190. doi: 10.1016/j.ceca.2006.04.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Cunningham AJ, Murray CA, O'Neill LAJ, Lynch MA, O'Connor Interleukin-1β (IL-1β) and tumour necrosis factor (TNF) inhibit long-term potentiation in the rat dentate Gyrus in vitro. Neurosci Lett. 2005;203:17–20. doi: 10.1016/0304-3940(95)12252-4. [DOI] [PubMed] [Google Scholar]
  24. Czarnecki A, Birtoli B, Ulrich D. Cellular mechanisms of burst firing-mediated long-term depression in rat neocortical pyramidal cells. J Physiol. 2006;578:471–479. doi: 10.1113/jphysiol.2006.123588. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Datta S. Cellular and chemical neuroscience of mammalian sleep. Sleep Med. 1996;11:431–440. doi: 10.1016/j.sleep.2010.02.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. 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. 2007;33:3824–3833. doi: 10.1523/JNEUROSCI.4367-12.2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Demuro A, Smith M, Parker I. Single-channel Ca2+ imaging implicates Abeta1–42 amyloid pores in Alzheimer's disease pathology. J Cell Biol. 2010;195:515–524. doi: 10.1083/jcb.201104133. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Di Bona D, Candore G, Franceschi C, Licastro F, Colonna-Romano G, Cammà C, Lio D, Caruso C. Systematic review by meta-analyses on the possible role of TNF-alpha polymorphisms in association with Alzheimer's disease. Brain Res Rev. 2013;61:60–68. doi: 10.1016/j.brainresrev.2009.05.001. [DOI] [PubMed] [Google Scholar]
  29. Diekelmann S, Born J. The memory function of sleep. Nat Rev Neurosci. 2011;11:114–126. doi: 10.1038/nrn2762. [DOI] [PubMed] [Google Scholar]
  30. Dursun E, Gezen-Ak D, Yilmazer S. A novel perspective for Alzheimer's disease: vitamin D receptor suppression by amyloid-β and preventing the amyloid-β induced alterations by vitamin D in cortical neurons. J Alzheimers Dis. 2009;23:207–219. doi: 10.3233/JAD-2010-101377. [DOI] [PubMed] [Google Scholar]
  31. Emilsson L, Saetre P, Jazin E. Alzheimer's disease: mRNA expression profiles of multiple patients show alterations of genes involved with calcium signalling. Neurobiol Dis. 2010;21:618–625. doi: 10.1016/j.nbd.2005.09.004. [DOI] [PubMed] [Google Scholar]
  32. Errington AC, Hughes SW, Crunelli V. Rhythmic dendritic Ca2+ oscillations in thalamocortical neurons during slow non-REM sleep-related activity in vitro. J Physiol. 2011;590:3691–3700. doi: 10.1113/jphysiol.2012.232132. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Farber K, Kettenmann H. Functional role of calcium signals for microglial function. Glia. 2006;54:656–665. doi: 10.1002/glia.20412. [DOI] [PubMed] [Google Scholar]
  34. Garcion E, Wion-Barbot N, Montero-Menei CN, Berger F, Wion D. New clues about vitamin D functions in the nervous system. Trends Endocrinol Metab. 2012;13:100–105. doi: 10.1016/s1043-2760(01)00547-1. [DOI] [PubMed] [Google Scholar]
  35. Gezen-Ak D, Dursun E, Yilmazer S. The effects of vitamin D receptor silencing on the expression of LVSCC-A1C and LVSCC-A1D and the release of NGF in cortical neurons. PLoS ONE. 2006;6:e17553. doi: 10.1371/journal.pone.0017553. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Guerreiro R, Wojtas A, Bras J, Carrasquillo M, Rogaeva E, Majounie E, Cruchaga C, Sassi C, et al. TREM2 variants in Alzheimer's disease. N Engl J Med. 2002;368:117–127. doi: 10.1056/NEJMoa1211851. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Gulati P, Gaspers LD, Dann SG, Joaquin M, Nobukuni T, Natt F, Kozma SC, Thomas AP, Thomas G. Amino acids activate mTOR complex 1 via Ca2+/CaM signaling to hVps34. Cell Metab. 2011;7:456–465. doi: 10.1016/j.cmet.2008.03.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Hájos N, Paulsen O. Network mechanisms of gamma oscillations in the CA3 region of the hippocampus. Neural Netw. 2013;22:1113–1119. doi: 10.1016/j.neunet.2009.07.024. [DOI] [PubMed] [Google Scholar]
  39. Heneka MT, Kummer MP, Stutz A, Delekate A, Schwartz S, Vieira-Saecker A, Griep A, Axt D, Remus A, Tzeng TC, Gelpi E, Halle A, Korte M, Latz E, Golenbock DT. NLRP3 is activated in Alzheimer's disease and contributes to pathology in APP/PS1 mice. Nature. 2008;493:674–678. doi: 10.1038/nature11729. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Hjorth E, Zhu M, Toro VC, Vedin I, Palmblad J. Omega-3 fatty acids enhance phagocytosis of Alzheimer's disease-related amyloid-β42 by human microglia and decrease inflammatory markers. J Alzheimers Dis. 2009;35:697–713. doi: 10.3233/JAD-130131. [DOI] [PubMed] [Google Scholar]
  41. Ji D, Wilson MA. Coordinated memory replay in the visual cortex and hippocampus during sleep. Nat Neurosci. 2013;10:100–107. doi: 10.1038/nn1825. [DOI] [PubMed] [Google Scholar]
  42. Jonsson T, Stefansson H, Steinberg S, Jonsdottir I, Jonsson PV, Snaedal J, Bjornsson S, Huttenlocher J, Levey AI, Lah JJ, Rujescu D, Hampel H, et al. Variant of TREM2 associated with the risk of Alzheimer's disease. N Engl J Med. 2013;368:107–116. doi: 10.1056/NEJMoa1211103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Kaushal V, Koeberle PD, Wang Y, Schlichter LC. The Ca2+-activated K+ channel KCNN4/KCa3.1 contributes to microglia activation and nitric oxide-dependent neurodegeneration. J Neurosci. 2007;27:234–244. doi: 10.1523/JNEUROSCI.3593-06.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Khachaturian ZS. Calcium, membranes, aging, and Alzheimer's disease. Introduction and overview. Ann N Y Acad Sci. 2013;568:1–4. doi: 10.1111/j.1749-6632.1989.tb12485.x. [DOI] [PubMed] [Google Scholar]
  45. Kuchibhotla KV, Goldman ST, Lattarulo CR, Wu HY, Hyman BT, Bacskai BJ. Abeta plaques lead to aberrant regulation of calcium homeostasis in vivo resulting in structural and functional disruption of neuronal networks. Neuron. 2007;59:214–225. doi: 10.1016/j.neuron.2008.06.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. LaFerla FM. Calcium dyshomeostasis and intracellular signaling in Alzheimer's disease. Nat Rev Neurosci. 1989;3:862–872. doi: 10.1038/nrn960. [DOI] [PubMed] [Google Scholar]
  47. Lee G-S, Subramanian N, Kim AI, Aksentijevich I, Goldbach-Mansky R, Sacks DB, Germain RN, Kastner DL, Chae JJ. The calcium-sensing receptor regulates the NLRP3 inflammasome through Ca2+ and cyclic AMP. Nature. 2008;492:123–127. doi: 10.1038/nature11588. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Lehmann DJ, Refsum H, Warden DR, Medway C, Wilcock GK, Smith AD. The vitamin D receptor gene is associated with Alzheimer's disease. Neur Lett. 2002;504:79–82. doi: 10.1016/j.neulet.2011.08.057. [DOI] [PubMed] [Google Scholar]
  49. Lock JT, Sinkins WG, Schilling WP. Effect of protein S-gluathionylation on Ca2+ homeostasis in cultured aortic endothelial cells. Am J Physiol Heart Circ Physiol. 2012;300:H493–H506. doi: 10.1152/ajpheart.01073.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. Lopez JR, Lyckman A, Oddo S, LaFerla FM, Querfurth HW, Shtifman A. Increased intraneuronal resting [Ca2+] in adult Alzheimer's disease mice. J Neurochem. 2011;105:262–271. doi: 10.1111/j.1471-4159.2007.05135.x. [DOI] [PubMed] [Google Scholar]
  51. Lu'o'ng KV, Nguyên LT. The role of vitamin D in Alzheimer's disease: possible genetic and cell signalling mechanisms. Am J Alzheimers Dis Other Demen. 2011;28:126–136. doi: 10.1177/1533317512473196. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Maezawa I, Zimin PI, Wulff H, Jin L-W. Amyloid-β protein oligomer at low nanomolar concentrations activates microglia and induces microglial neurotoxicity. J Biol Chem. 2008;286:3693–3706. doi: 10.1074/jbc.M110.135244. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Mairet-Coello G, Courchet J, Pieraut S, Virginie Courchet V, Maximov A, Polleux F. The CAMKK2-AMPK kinase pathway mediates the synaptotoxic effects of Aβ oligomers through tau phosphorylation. Neuron. 2013;78:94–108. doi: 10.1016/j.neuron.2013.02.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Mark RJ, Hensley K, Butterfield DA, Mattson MP. Amyloid beta-peptide impairs ion-motive ATPase activities: evidence for a role in loss of neuronal Ca2+ homeostasis and cell death. J Neurosci. 2011;15:6239–6249. doi: 10.1523/JNEUROSCI.15-09-06239.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Massimini M, Huber R, Ferrarelli F, Hill S, Tononi G. The sleep slow oscillation as a travelling wave. J Neurosci. 2013;24:6862–6870. doi: 10.1523/JNEUROSCI.1318-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. McManus J, Murphy MP, Franklin JL. The mitochondrial-targeted antioxidant MitoQ prevents loss of spatial memory retention and early neuropathology in a transgenic mouse mode4l of Alzheimer's disease. J Neurosci. 1995;31:15703–15715. doi: 10.1523/JNEUROSCI.0552-11.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Michaud J-P, Hallé M, Lampron A, Thériault P, Préfontaine P, Filali M, Tribout-Jover P, Lanteigne AM, Jodoin R, Cluff C, Brichard V, Palmantier R, Pilorget A, Larocque D, Rivest S. Toll-like receptor 4 stimulation with the detoxified ligand monophosphoryl lipid A improves Alzheimer's disease-related pathology. Proc Natl Acad Sci U S A. 2004;110:1941–1946. doi: 10.1073/pnas.1215165110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Minoretti P, Gazzaruso C, Vito CD, Emanuele E, Bianchi M, Coen E, Reino M, Geroldi D. Effect of the functional toll-like receptor 4 Asp299Gly polymorphism on susceptibility to late-onset Alzheimer's disease. Neurosci Lett. 2011;391:147–149. doi: 10.1016/j.neulet.2005.08.047. [DOI] [PubMed] [Google Scholar]
  59. Moosmang S, Haider N, Klugbauer N, Adelsberger H, Langwieser N, Müller J, Stiess M, Marais E, Schulla V, Lacinova L, Goebbels S, Nave KA, Storm DR, Hofmann F, Kleppisch T. Role of hippocampal Cav1.2 Ca2+ channels in NMDA receptor-independent synaptic plasticity and spatial memory. J Neurosci. 2013;25:9883–9892. doi: 10.1523/JNEUROSCI.1531-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Müller M, Cheung KH, Foskett JK. Enhanced ROS generation mediated by Alzheimer's disease presenilin regulation of InsP3 Ca2+ signaling. Antioxid Redox Signal. 2006;14:1225–1235. doi: 10.1089/ars.2010.3421. [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Nunes PV, Forlenza OV, Gattaz WF. Lithium and risk for Alzheimer's disease in elderly patients with bipolar disorder. Br J Psychiatry. 2005;190:359–360. doi: 10.1192/bjp.bp.106.029868. [DOI] [PubMed] [Google Scholar]
  62. Ohana L, Newell EW, Stanley EF, Schlichter LC. The Ca2+ release-activated Ca2+ current (ICRAC) mediates store-operated Ca2+ entry in rat microglia. Channels. 2011;3:129–139. doi: 10.4161/chan.3.2.8609. [DOI] [PubMed] [Google Scholar]
  63. Pace-Schott EF, Hobson JA. The neurobiology of sleep: genetics, cellular physiology and subcortical networks. Nat Rev Neurosci. 2007;3:591–605. doi: 10.1038/nrn895. [DOI] [PubMed] [Google Scholar]
  64. Park KM, Yule DI, Bowers WJ. Tumor necrosis factor-α-mediated regulation of the inositol 1,4,5-trisphosphate receptor promotor. J Biol Chem. 2009;284:27557–27566. doi: 10.1074/jbc.M109.034504. [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Pérez AV, Picotto G, Carpentieri AR, Rivoira MA, Peralta López ME, Tolosa de Talamoni NG. Minireview on regulation of intestinal calcium absorption. Emphasis on molecular mechanisms of transcellular pathway. Digestion. 2002;77:22–34. doi: 10.1159/000116623. [DOI] [PubMed] [Google Scholar]
  66. Pickford F, Masliah E, Britschgi M, Lucin K, Narasimhan R, Jaeger PA, Small S, Spencer B, Rockenstein E, Levine B, Wyss-Coray T. The autophagy-related protein beclin 1 shows reduced expression in early Alzheimer disease and regulates amyloid b accumulation in mice. J Clin Invest. 2009;118:2190–2199. doi: 10.1172/JCI33585. [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Roh JH, Huang Y, Bero AW, Kasten T, Stewart FR, Bateman RJ, Holtzman DM. Disruption of the sleep-wake cycle and diurnal fluctuation of amyloid-β in mice with Alzheimer's disease pathology. Sci Transl Med. 2008;4:150ra122. doi: 10.1126/scitranslmed.3004291. [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. Saijo K, Glass CK. Microglial cell origin and phenotypes in health and disease. Nat Rev Immunol. 2008;11:775–787. doi: 10.1038/nri3086. [DOI] [PubMed] [Google Scholar]
  69. Sarkar S, Floto RA, Berger Z, Imarisio S, Cordenier A, Pasco M, Cook LJ, Rubinsztein DC. Lithium induces autophagy by inhibiting inositol monophosphatase. J Cell Biol. 2012;170:1101–1111. doi: 10.1083/jcb.200504035. [DOI] [PMC free article] [PubMed] [Google Scholar]
  70. Son JH, Shim JH, Kim K-H, Ha J-Y, Han JY. Neuronal autophagy and neurodegenerative diseases. Exp Mol Med. 2011;44:89–98. doi: 10.3858/emm.2012.44.2.031. [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Stutzmann GE. The pathogenesis of Alzheimer's disease is it a lifelong ‘calciumopathy'. Neuroscientist. 2005;13:546–559. doi: 10.1177/1073858407299730. [DOI] [PubMed] [Google Scholar]
  72. Stutzmann GE, Mattson MP. Endoplasmic reticulum Ca2+ handling in cells in health and disease. Pharmacol Rev. 2012;63:700–727. doi: 10.1124/pr.110.003814. [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Supnet C, Grant J, Kong H, Westaway D, Mayne M. Amyloid-β-(1–42) increases ryanodine receptor-3 expression and function in neurons of TgCRND8 mice. J Biol Chem. 2007;281:38440–38447. doi: 10.1074/jbc.M606736200. [DOI] [PubMed] [Google Scholar]
  74. Sutherland MK, Somerville MJ, Yoong LKK, Bergeron C, Haussler MR, Mclachlan DRC. Reduction of vitamin D hormone receptor mRNA levels in Alzheimer as compared to Huntington hippocampus: correlation with calbindin-28 k mRNA levels. Mol Brain Res. 2011;13:239–250. doi: 10.1016/0169-328x(92)90032-7. [DOI] [PubMed] [Google Scholar]
  75. Taniura H, Ito M, Sanada N, Kuramoto N, Ohno Y, Nakamichi N, Yoneda Y. Chronic vitamin D3 treatment protects against neurotoxicity by glutamate in association with upregulation of vitamin D receptor mRNA expression in cultured rat cortical neurons. J Neurosci Res. 2006;83:1179–1189. doi: 10.1002/jnr.20824. [DOI] [PubMed] [Google Scholar]
  76. Thibault O, Gant JC, Landfield PW. Expansion of the calcium hypothesis of brain ageing and Alzheimer's disease: minding the store. Aging Cell. 1992;6:307–317. doi: 10.1111/j.1474-9726.2007.00295.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Tononi G, Cirelli C. Sleep function and synaptic homeostasis. Sleep Med Rev. 2006;10:49–62. doi: 10.1016/j.smrv.2005.05.002. [DOI] [PubMed] [Google Scholar]
  78. Tuohimaa P, Keisala T, Minasyan A, Cachat J, Kalueff A. Vitamin D, nervous system and aging. Psychoneuroendocrinology. 2007;34(Suppl 1):S278–S286. doi: 10.1016/j.psyneuen.2009.07.003. [DOI] [PubMed] [Google Scholar]
  79. Vicencio JM, Ortiz C, Criollo A, Jones AW, Kepp O, Galluzzi L, Joza N, Vitale I, Morselli E, Tailler M, Castedo M, Maiuri MC, Molgó J, Szabadkai G, Lavandero S, Kroemer G. The inositol 1,4,5-trisphosphate receptor regulates autophagy through its interaction with Beclin 1. Cell Death Differ. 2006;16:1006–1017. doi: 10.1038/cdd.2009.34. [DOI] [PubMed] [Google Scholar]
  80. Vyazovskiy VV, Cirelli C, Pfister-Genskow M, Faraguna U, Tononi G. Molecular and electrophysiological evidence for net synaptic potentiation in wake and depression in sleep. Nat Neurosci. 2009;11:200–208. doi: 10.1038/nn2035. [DOI] [PubMed] [Google Scholar]
  81. Wang L, Hara K, Van Baaren JM, Price JC, Beecham GW, Gallins PJ, Whitehead PL, Wang G, Lu C, Slifer MA, Züchner S, Martin ER, Mash D, Haines JL, Pericak-Vance MA, Gilbert JR. Vitamin D receptor and Alzheimer's disease: a genetic and functional study. Neurobiol Aging. 2009;33:1844.e1–1849e1. doi: 10.1016/j.neurobiolaging.2011.12.038. [DOI] [PubMed] [Google Scholar]
  82. Wasserman RH. Vitamin D and the dual processes of intestinal calcium absorption. J Nutr. 2008;134:3137–3139. doi: 10.1093/jn/134.11.3137. [DOI] [PubMed] [Google Scholar]
  83. Webster NJ, Ramsden M, Boyle JP, Pearson HA, Peers C. Amyloid peptides mediate hypoxic increase of L-type Ca2+ channels in central neurones. Neurobiol Aging. 2012;27:439–445. doi: 10.1016/j.neurobiolaging.2005.02.002. [DOI] [PubMed] [Google Scholar]
  84. Yang L, Lu R, Jiang L, Liu Z, Peng Y. Expression and genetic analysis of tumor necrosis factor-alpha (TNF-alpha) G-308A polymorphism in sporadic Alzheimer's disease in a Southern China population. Brain Res. 2004;1247:178–181. doi: 10.1016/j.brainres.2008.10.019. [DOI] [PubMed] [Google Scholar]
  85. Ye C, Ho-Pao CL, Kanazirska M, Quinn S, Rogers K, Seidman CE, Seidman JG, Brown EM, Vassilev PM. Amyloid-beta proteins activate Ca2+-permeable channels through calcium-sensing receptors. J Neurosci Res. 2006;47:547–554. doi: 10.1002/(sici)1097-4547(19970301)47:5<547::aid-jnr10>3.0.co;2-v. [DOI] [PubMed] [Google Scholar]
  86. Yu JT, Mou SM, Wang LZ, Mao CX, Tan L. Toll-like receptor 2 -196 to -174 del polymorphism influences the susceptibility of Han Chinese people to Alzheimer's disease. J Neuroinflammation. 2009;8:136–139. doi: 10.1186/1742-2094-8-136. [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Yu JT, Miao D, Cui WZ, Ou JR, Tian Y, Wu ZC, Zhang W, Tan L. Common variants in toll-like receptor 4 confer susceptibility to Alzheimer's disease in a Han Chinese population. Curr Alzheimer Res. 2012;9:458–466. doi: 10.2174/156720512800492495. [DOI] [PubMed] [Google Scholar]

Articles from The Journal of Physiology are provided here courtesy of The Physiological Society

RESOURCES