NCBI Bookshelf. A service of the National Library of Medicine, National Institutes of Health.
Riddle DR, editor. Brain Aging: Models, Methods, and Mechanisms. Boca Raton (FL): CRC Press/Taylor & Francis; 2007.
I. Ca2+ HYPOTHESIS OF AGING
It is now more than two decades since the first proposal of a “Ca2+ hypothesis of aging” [1]. In its mature formulation, the hypothesis aimed to provide a working hypothesis for explaining not only the aging process but also Alzheimer’s disease (AD). Drawing on many neurophysiological processes and mechanisms, the hypothesis contained six interrelated key elements [2], providing an extremely wide explanatory blanket. The exposition of these elements is worth repeating not only for historical reasons and to illustrate the breath of the proposal, but also because it still represents the source of some misunderstandings.
The first element of Khachaturian’s construct was simply that the cellular mechanisms that participate in Ca2+ homeostasis also play an important role in the process of aging and neurodegeneration. Consequently, altered Ca2+ homeostasis can account for a number of age-related changes associated with either aging or AD. The second element was that normal aging and the process of neurodegeneration, as in AD, must be seen as part of a continuum of processes associated with development and restructuring of the nervous system throughout life. Thus, the morphological changes in AD, such as dendritic pruning and loss of synapses and neurons, involve the same mechanisms that are responsible for neuroplasticity in the developing or mature brain. As an extension of the previous point, the third element of the hypothesis concentrated on the plasticity of the nervous system architecture seen, at any time point, as a balance between growth/regeneration and decaying/degeneration processes. In many of these processes the intracellular Ca2+ ([Ca2+]i) plays a crucial role and, from this perspective, the neuronal dysfunction of the aged or the AD brain results from an imbalance between the two types of opposing processes, an imbalance that can result from a breakdown of the homeostatic mechanisms regulating [Ca2+]i. The next, rather contentious element dealt with the temporal dimension of the aging/neurodegenerative processes and proposed that the functional product of the perturbation in [Ca2+]i homeostasis and time is a constant. Thus, small changes in [Ca2+]i (as those resulting, for example, from a less efficient Ca2+ extrusion pump, or an enhanced Ca2+ channel activity) exerted over a long period of time would have the same functional or morphological effects as a massive, but acute insult. It is clear now that the actions of Ca2+ in different temporal domains (i.e., over milliseconds, seconds, minutes, or days) generate very different types of responses and functional outcomes. A fifth element took into account the cellular effects of Ca2+ signaling and proposed that Ca2+-mediated processes are an important part of the final common pathway that leads to neuronal dysfunction and cell death. The model presented at this point an extremely broad and flexible perspective, and almost any element along the path leading from the triggers of Ca2+ dysregulation to various Ca2+ homeostatic mechanisms to, eventually, any of the Ca2+-dependent processes could fit the bill of explaining the involvement of Ca2+ in the process of aging and neurodegeneration. This point was even further expanded by the sixth element, which proposed that both the age-related and the AD-associated changes are not single factorial events but could be initiated by a wide variety of instances, acting single or in combination, simultaneous or sequentially, over a long period of time.
Underlying this hypothesis were two powerful ideas that reflected the predominant views of the time. One was that the difference between aging and neurodegeneration is just a matter of degree, both existing along a continuous path. The other idea was that intracellular Ca2+ signaling is one of the most important transducing processes involved in a wide range of cellular processes [3, 4]. Linking the two was the concept of Ca2+ regulation of neuronal death through the process of excitotoxicity. The term “excitotoxicity” was first coined by Olney in 1970 to describe the excessive activation of glutamate receptors that follows neuronal deprivation of oxygen and/or nutrients and resulting in neuronal death [5]. The subsequent demonstration in 1987 of the involvement of Ca2+ in mediating the excitotoxic outcome [6] strengthened the “Ca2+ hypothesis of aging.”
Today, of the two central ideas discussed above, only the latter remains unchallenged; the issue of understanding the aging as a neurodegenerative process is very much disputed. From animal studies it is clear that the age-related changes in functional and biochemical characteristics of the cortical circuitry that might underlie the moderate cognitive decline do not involve frank neuronal death but a significant amount of synaptic restructuring [7]. Some of the reasons why, from the late 1960s to the mid-to-late 1980s, the dogma of aging centered on the idea of associated significant neuronal loss were technological (the modern unbiased, dissector method for counting neurons was introduced only in 1984 [8] and helped demonstrate significant differences between neuronal numbers in normal aging and in neurodegenerative diseases [9]; see also Chapter 4) and methodological, involving the possible inclusion in some studies of material coming from subjects already affected by neurodegenerative diseases, (cf. [10]). Thus, in its initial formulation that links dysregulation of Ca2+ homeostasis to neuronal loss, the “Ca2+ hypothesis” is not directly applicable anymore to the process of normal aging, but retains much of its explanatory power in the context of Alzheimer’s disease (see [11] for a recent review). What is undisputed, however, is the fact that intracellular Ca2+ signaling and its regulation are central to neuronal physiology, linking with a variety of other processes (Figure 14.1).
II. Ca2+ HOMEOSTASIS
A. CA2+ Entry and CA2+ Extrusion Pathways
At any time point, the values of intracellular free Ca2+, as measured routinely by Ca2+-sensitive fluorescent dyes with different affinities, represent the functional balance between processes that increase and those that decrease [Ca2+]i (Figure 14.2). Neuronal activation usually is associated with increases in [Ca2+]i. For neurons, the source for these activation-induced Ca2+ responses is the extracellular medium, with Ca2+ ions flowing along a very steep (10,000-fold) concentration gradient (at rest [Ca]i is around 50 to 100 nM, whereas the extracellular Ca2+ is around 1.5 to 2 mM). This process involves a number of different types of ion channels; some are activated by the depolarization of the plasma membrane and have exclusive selectivity for Ca2+ ions (the voltage-operated Ca2+ channels, VOCCs). These channels exist as several types and are functionally divided into two categories: (1) the low voltage-activated (LVA) channels (prototype, the former T-type VOCC and current Cav3 family — for new VOCC nomenclature, see [12]), activated by smaller depolarizations of the plasma membrane; and (2) the high voltage-activated (HVA) channels. Of these HVAs, the best-known functionally is the L-type Ca2+ channel (current Cav1 family), but the family also comprises the N-, P-, Q-, and R-types (all part of the Cav2 family) [13]. Additional types of channels involved in generating Ca2+ signals in neuronal cytosol are the ones activated by specific ligands (ligand operated Ca2+ channels, LOCs), such as NMDA, AMPA (or, in vivo, glutamate) or ATP (purinergic P2X receptors). These are nonspecific cationic channels, with a Ca2+ permeability that is usually lower than that for Na+ (the latter ion having, nevertheless, a much lower electrochemical gradient). Another type of neuronal plasma membrane channel with Ca2+ permeability is represented by the second messenger-operated Ca2+ channels (SMOCs), with cyclic nucleotides or arachidonate as activators. Finally, a ubiquitous and ever-expanding family of nonselective plasma membrane channels is that of the transient receptor potential (TRP) channels, initially described during the study of the photoreceptor transduction process.
Recovery of resting [Ca]i values following such activation-evoked Ca2+ signals, which can generate increases of [Ca]i well in excess of 10 μM, taking into account the confined small volumes where these signals appear (synaptic boutons, dendritic spines), is an energy-dependent process involving either pumps (ATPases) or specific transporters. The pumps remove cytosolic Ca2+ either to the extracellular medium (the plasma membrane Ca2+ ATPase, PMCA) or into the endoplasmic reticulum, which is a very important Ca2+ storing organelle (through the sarco(endo)plasmic Ca2+ ATPase, SERCA). A specific transporter is the Na+/Ca2+ exchanger (NCX), which harnesses the electrochemical gradient of Na+. These transport systems have different thresholds for activity. While the pumps have lower transport rates, they have a higher affinity for Ca2+ and thus are able to respond to smaller changes in [Ca]i and ultimately to set the basal Ca2+ levels.
B. Intracellular Organelles
Many intracellular organelles participate in Ca2+ homeostasis and, as discussed in more detail below, another important mechanism for removal of cytosolic Ca2+ is the Ca2+ uniporter of the mitochondria. This system has a huge transport capacity (depending on conditions, mitochondria can accumulate millimolar Ca2+) with an optimal affinity in the micromolar range, levels easily reached at the peak of Ca2+ responses during activation. The system is also able, in a sub-mode of action or through a separate uptake system (the rapid uptake mode, RaM), to take up cytosolic Ca2+ following smaller, submicromolar [Ca]i increases [14].
Endoplasmic reticulum (ER) is a multifunctional organelle, intimately involved in Ca2+ homeostasis, acting both as a direct source of Ca2+ signaling (e.g., metabotropic glutamate stimulation, mediating release of Ca2+ from the ER through the generation of InsP3 and activation of the specific InsP3 receptors) or as an important amplifier of the Ca2+ signal generated at the plasma membrane, through the process commonly known as the Ca2+-induced Ca2+ release (CICR) [15]. As mentioned above, through the action of the SERCA pumps, the ER also can participate in the recovery of the resting [Ca2+]i values following neuronal activation. While this role of the ER is very well established in nonexcitable cells and in some neurons of the peripheral nervous system, its participation in the intense Ca2+ signaling of central nervous system neurons is more restricted [16]. Irrespective of the cell type under consideration, the ER is always the site of protein synthesis and maturation, and many of these processes are controlled by proteins sensitive to Ca2+ [17].
C. CA2+ Buffers
All the mechanisms briefly described above involve transporting proteins, embedded in membranes, which are able to bind Ca2+ with various affinities and mediate the transmembrane movement of ions. Another set of proteins with an important role in cellular Ca2+ homeostasis is known under the general name of Ca2+-binding proteins (CaBP). Many of these proteins are soluble, and this higher degree of mobility allows them to interact with Ca2+ with different affinities at various locations, and consequently shape the Ca2+ signal (either by transporting Ca2+ away from or by significant Ca2+ buffering at the site of generation). One of the largest families of CaBP is characterized by the EF-hand Ca2+-binding domain, consisting of two perpendicular 10- to 12-residue, alpha-helix regions separated by a 12-residue loop region; EF-hand domains are often found in single or multiple pairs, giving rise to a huge variety of structural/functional variations in the proteins that form a family comprising more than 600 members [18]. The CaBP are divided into two functional, well-established categories. One, having calmodulin as the reference member, acts as a Ca2+ sensing mechanism that translates graded changes in [Ca2+]i that follow cell stimulation into a graded response initiated by the binding of these CaBP to various Ca2+-sensitive target enzymes. Within this category, a group of proteins, called the neuronal calcium-sensor (NCS) proteins, has recently acquired prominence due to the relatively specific localization and/or set of functions they perform [19]. One of these proteins, calsenilin, could be very relevant to the processes of neurodegeneration because, among other things, it can modify the processing of presenilins and regulate the processing of the amyloid precursor protein (APP), both intimately involved in the pathophysiological processes of Alzheimer’s disease [20].
The other functional category, represented by proteins such as parvalbumin, calbindin, and calretinin (PB, CB, and CR, respectively), is that of the Ca2+ buffers. These proteins are widely expressed in the central nervous system, less in principal neurons, and more in the inhibitory GABAergic interneurons, each type being characterized by a predominant form of CaBP, and associated with a specific constellation of associated neurotransmitters, cell surface markers, and receptors [21]. Furthermore, experimental data and theoretical modeling, coupled with information on the biochemical properties of these CaBP buffers, show that they could exert profound effects on neuronal activity, from shaping the Ca2+ signal at local levels in the presynaptic active zone or postsynaptic densities, thus initiating the transfer of information in the neuronal networks [22] to modulate neuronal vulnerability. Many studies have shown the neuroprotective effects of CaBP. For example, in hippocampal primary cultures, neurons expressing CB are more effective at recovering [Ca2+]i levels after stimulation [23]; whereas for cortical neurons, expression of CR appears to play this protective role [24]. Calbindin can protect neurons from both oxidative stress and reduce apoptosis [25]. Significantly more motor neurons survived a standard severe peripheral injury in parvalbumin-overexpressing transgenic mice than in control, wild-type littermates [26].
At the other end of the functional spectrum, the distribution and types of intra-cellular Ca2+ buffers can affect significantly the electrical activity of neurons. For example, a strong correlation has been found between the expression of parvalbumin and the firing pattern of neurons: the PV-positive GABAergic interneurons are capable of maintaining a higher rate of repetitive activity, an observation explained by the fact that PV is a “slow” Ca2+ buffer and thus able to decrease more effectively the [Ca2+]i during higher frequency discharges [27]. Overall, it is important to mention in this discussion that a more general parameter, the Ca2+-binding ratio (Ks), is used to measure and compare the overall capacity of various cells to buffer Ca2+, independent of the type of Ca2+ buffer proteins involved. Experimental measurements of this important parameter show significant heterogeneity between different neuronal types — in hippocampus, the inhibitory interneurons have a two to three times larger buffering power (i.e., higher Ks) than the excitatory interneurons, and thus are able to withstand larger Ca2+ challenges [28]. In the spinal cord, the Ks of motor neurons is only about 50 (indicating that 1 in 50 Ca2+ ions entering the cells or 2% of the Ca2+ challenge remains free, unbound by the cytosolic buffers) [29], whereas the Ks of the sympathetic superior cervical ganglia (SCG) varies between 250 and 500, depending on the value of [Ca2+]i at a given time [30]. Interestingly, the value of Ks appears to be developmentally regulated [31].
III. EFFECTS OF AGING ON Ca2+ HOMEOSTATIC SYSTEMS
There are relatively few direct experimental data analyzing and describing the putative age-related changes of the various Ca2+ homeostatic mechanisms. Overall, they tend to show dysfunction but, as discussed at length elsewhere, the interpretation is rendered difficult by the use of a variety of cellular preparations and by the heterogeneity of neuronal composition and ratio of glia to neurons in various regions of the brain [32].
A. Age-Dependent Effects on CA2+ Entry Pathways
The age-dependent changes in the activity of various Ca2+ entry systems illustrate the problems in interpretation. Some of the complications are intrinsic to the model; there are several pathways allowing Ca2+ entry, each of them composed of various subunits with different Ca2+ permeabilities and regulating different sets of neuronal activities and functions. The other source of complication is the regional heterogeneity.
1. Ca2+ Channels
With respect to the effect of aging on the voltage-operated Ca2+ channels, two regions have been studied more intensely: (1) the hippocampus, in particular the pyramidal neurons of CA1 region; and (2) the cortical neurons of the basal forebrain.
In the hippocampus, one of the most consistent findings is that of an age-dependent increase in Ca2+-dependent afterhyperpolarization (AHP) [33–35], due to a demonstrated increase in the number of L-type VOCCs [34–36]. A somewhat similar conclusion was reached for the neurons of the basal forebrain, although the mechanisms mediating an increased Ca2+ load within aging appear different. In these latter neurons, the two large families of VOCCs — that is, the LVA and HVA — appear to respond differently to aging. While the LVA show an increased current density (i.e., larger current values per unit of surface area) with similar activation/inactivation kinetics [37], the HVA show the opposite — the same current density but with decreased inactivation [38]. While most of this electrophysiological evidence would indicate that in the CNS, aging might be associated with an increase in Ca2+ uptake through VOCCs, once again the peripheral neurons seem different; and for the dorsal root ganglia neurons, both LVA and HVA currents are decreased by aging [39]. Other lines of evidence, based on older biochemical studies, suggest an opposite functional outcome. An assessment of the binding sites for a typical, very specific VOCC antagonist (i.e., dihyropiridine) showed an age-dependent decrease in the binding affinity and sensitivity [40]. Work on another type of nervous system preparation, synaptosomes, also showed conflicting results in response to depolarization (the assumption being that depolarization by KCl will activate the VOCC present in the nerve terminals); some reported an increased Ca2+ uptake [41], whereas others showed an inhibition [42] or no effect [43]. Expanding the panoply of methods used for assessing the effects of aging on VOCC, one study [44] looked at the age-dependent changes in the α subunit composition of the VOCCs, using the cerebellum as a study model (for which, however, not much electrophysiology of the aging brain exists) and showed a consistent decrease in the αD subunit (encoding the L-type VOCC) in the Purkinje neurons. While the details of such investigations appear rather confusing, in general it appears that there is no unique effect of aging on VOCCs; and that for each type of VOCC, conclusions need to be drawn by direct experimentation on multiple neuronal types.
2. Glutamate-Operated Channels
The other path of Ca2+ entry during neuronal activation is through one or another of the various ionotropic glutamatergic receptor/channels; both the NMDA and AMPA receptors have a sizable Ca2+ permeability [45]. The main receptor subunit of the NMDA receptor is the NR1 subunit, which has three alternative splicing regions [46] and its presence is required in all expressed NMDA receptors, as being essential in establishing the functional homo- or heteromeric complexes [47]. Two other subunits have been described: (1) the NR2 (A-D forms) and (2) the NR3 (A and B) [46]. The NMDA receptors also have a special set of properties that give the receptor important functional characteristics at rest; the NMDA channel pore is blocked in a voltage-dependent manner by Mg+ ions [48]. This block is relieved by depolarization, but this induces a slight delay in the activation of the channel, generating a slower kinetic response [48]. There is an overwhelming body of evidence implicating the NMDA receptors in two functions that are affected by the aging process: (1) synaptic plasticity and memory, particularly in the hippocampus, and (2) excitotoxicity phenomena across the CNS. The overall view of the literature describing the changes in NMDA activity with aging is that of a decrease in effectiveness of NMDA signaling. As reviewed by Magnusson [49], the majority of studies have found an age-related decline in the number of NMDA receptor sites (as measured by a reduced Bmax value), associated, in some of these studies, with no change in the affinity of the receptor for glutamate. In addition, the subunit composition of the expressed NMDA receptors appears to have an important developmental component. At embryonic stages, the NR2B subunit is present across the brain, with NR2D present only in the diencephalon and brain stem. After birth, the mRNA for NR2A becomes evident and prominent and, in the specific case of the cerebellum, prenatal expression of NR2B on the granule cell is changed to the NR2C subunit type [50]. At the other extreme of the lifespan, a decrease in the expression of the NR2B mRNA and protein has been reported as an age-related phenomenon, while the expression of NR2A was not affected [51]. These data are corroborated by electrophysiological evidence for an age-related decrease in NMDA responsiveness in neostriatal slices of both aged rats (24 months old) and cats (16 years old) [52]. Others, however, have reported that the age-related decrease in the overall NMDA-mediated response is due to a reduction in the input to the NMDA receptor, rather than to the change in the electrophysiological NMDA response (rat CA1 region of the hippocampus, [53]). Experiments using short-term cell cultures derived from aged rats (25 to 26 months old) showed increased NMDA peak current when compared to middle-aged (9 to 10 months old) animals, associated with an increased Ca2+ signal [54]. Probably quite relevant in the attempt to systematize the results in the literature, in this last study, the NMDA response in the aged animals was larger than the one obtained in the middle-aged, but dramatically smaller (both NMDA and Ca2+ signal) than the signal recorded from cells derived from embryonic tissues [54].
AMPA receptor channels exist as homo- or, more common in their native form, as heteromers of the four possible subunits GluR1-GluR4 (or GluR-A to GluR-D). The AMPA channels are cation selective and are mainly permeable to the monovalent ions Na+ and K+. The permeability to Ca2+ depends on the inclusion and properties of the GluR2 subunit in the channel structure, such that AMPA-Rs assembled from combinations of GluR1, GluR2, and/or GluR4 are permeable to Ca2+ (and Zn2+) and show a voltage-dependent block by intracellular polyamines [55]. The permeability properties of the GluR2 subunit have been traced to a single amino acid residue located in the pore-forming domain [56]. A positively charged arginine residue at this position prevents Ca2+ passage, whereas the presence of the neutral amino acid residue glutamine results in significant Ca2+ permeability [57]. Arginine is not encoded by the GluR2 genomic DNA and is introduced at this site by subsequent RNA editing, which in adult brains is very efficient [58]. As a result, the majority of AMPA receptors in the CNS have a low Ca2+ permeability, although some neurons displaying AMPA receptors with high Ca2+ permeability were reported in both the hippocampus [59] and neocortex [60]. AMPA binding does not appear to show significant changes with aging [61], although in many regions there are possible subtler age-associated changes, as shown for C57Bl/6 mice [62]. Despite this seemingly minor effect of age on AMPA binding, there are several brain regions for which a significant correlation between cognitive tests and AMPA binding densities was reported [63]. The lack of significant effect of aging on AMPA binding is consistent with the electrophysiological studies of the AMPA response showing no age-related changes in either the mean quantal sizes or the unitary synaptic response [64]. Another important issue, with relevance for [Ca2+]i homeostasis, is that of the effect of age on GluR2 expression. The efficiency and fidelity of the RNA editing processing, ultimately conferring the Ca2+ resistance, is well conserved with aging in all the major regions of the brain [65]. In one of the few studies that analyzed specifically the distribution of GluR2 in the aged brain, analyzing primarily the short and long corticocortical projections, a notable downregulation in the expression of GluR2, as well as NMDA-R1, was demonstrated [66].
B. Age-Dependent Effects on CA2+ Extrusion Pathways
A number of studies have indicated that both the activity and the expression levels of the PMCA in synaptosomal preparations are significantly reduced as a function of aging both in the Fisher-344 and in the longer-lived Fisher344/BNF1 rats [67, 68]. The functional decrease in activity could be explained by the inhibitory effects of free radicals (reactive oxygen species, ROS) on the PMCA, either as a direct effect on the pump activity [69] or on calmodulin, the essential CaBP that activates, in its Ca2+-bound form, the PMCA [70]. Nevertheless, other studies using peripheral neurons (rat adrenergic neurons from the superior cervical ganglia) showed PMCA activity maintained with age and even a capacity to upregulate function (in conjunction with the other extrusion system — the Na+/Ca2+ exchanger) in order to meet higher extrusion demands [71]. This probably is an important feature, because the PMCA has been reported as the major Ca2+ clearing system in this neuronal cell type [30]. In these types of neurons, it was the SERCA pumps, which mediate the re-uptake of Ca2+ into the endoplasmic reticulum, that appear to be affected by the process of aging, as shown in experiments using either specific SERCA inhibitors (e.g., thapsigargin) or inhibitors of other extrusion pathways, thus isolating the SERCA activity for inspection [72]. It is interesting to note that it is in these peripheral neurons that the intracellular Ca2+ stores, such as the ER, appear to play a more important role in Ca2+ homeostasis [16].
C. Age-Dependent Effects on CA2+ Buffers
Immunocytochemical determinations of calbindin (CB) and parvalbumin (PV) expression in animal models showed that age decreases the number of CB-positive neurons in the rabbit hippocampus without affecting the PV-positive neurons, although for the latter CaBP there were some subtler differences in subcellular distribution, with a proposed decrease in PV expression in the neurites [73]. A similar picture came from studies of human cortex, which showed a consistent trend for age-related decreases in the calbindin- and calretinin (CR)-positive neurons that attained a level of statistical significance only in few regions, while no change was recorded for parvalbumin (PV)-positive neurons [74]. These studies expanded on the previous one from the same group that showed a dramatic 60% loss of CB from the basal forebrain cholinergic neurons (BFCN) [75]. It is entirely possible, however, that the changes in CaBP expression are region specific and influenced by levels of activity. Thus, when assessing the number of PV- and CB-expressing neurons in the dorsal cochlear nucleus of mice, an age-related increase has been reported [76] This was consistent with the observation that excessive stimulation of neurons in the cochlear nucleus by noise exposure resulted in increased expression of the CaBP [77]. Similarly in human specimens, it was found that the number of CB-positive neurons in the temporal cortex increased with age, in contrast to the situation for the AD-afflicted specimens [78].
Functionally, intracellular Ca2+ buffering was reported to decline with aging in adrenergic nerves [79], based on the observation that norepinephrine release increased with age and that this effect was reversed by an increase of intracellular Ca2+ buffering (with BAPTA-AM). Yet again, as for many others Ca2+ homeostasis parameters, the effect of aging on Ca2+ buffering is not simple, and other groups have described an increased Ca2+ buffering with aging, as in the case of the basal forebrain neurons [80]. When discussing Ca2+ buffering and the effects of this parameter on the values of [Ca2+]i,, one should take into account not only the specific proteins that bind Ca2+, the CaBPs discussed above, but also the participation of the internal organelles (e.g., the endoplasmic reticulum and mitochondria) with a Ca2+ uptake potential in shaping the Ca2+ signal. Thus, in assessing the dynamics of the [Ca2+]i increase following stimulation (i.e., the shape of the Ca2+ signal), two phases normally are described — a rapid phase that limits the amplitude of the [Ca2+]i response, followed by a second, slower phase that involves the Ca2+ extrusion systems, either Ca2+ removal or Ca2+ uptake into the intracellular compartments [81]. If the stimulus persists, then during this second phase, a new steady state (the Ca2+ plateau) is achieved, reflecting the balance between Ca2+ entry and Ca2+ removal from the cytosol. If the stimulation ceases, then this second, slower phase reflects the process of recovery of the resting [Ca2+]i values. The surprising observation in the case of the basal forebrain neurons is that the reported increased Ca2+ buffering is manifest only during the rapid phase, and is not determined by changes in the function of either ER or the mitochondrial fraction (the latter being, in fact, reduced in the aged neurons [82]) and takes place in neurons with a decreased expression of CaBP [83].
D. Effects of Aging on Intracellular CA2+ Compartments
1. Endoplasmic Reticulum
The endoplasmic reticulum (ER) is an important intracellular Ca2+ store and, depending on its refill status, can act either as a “Ca2+ sink,” buffering through SERCA pump-mediated Ca2+ reuptake into the ER, or as a “Ca2+ source,” through a process that involves Ca2+-induced Ca2+ release (CICR; [15]). When discussing the role of ER Ca2+ stores in neuronal physiology, a number of issues must be addressed, including:
- The morphology of the ER and its distribution is not uniform throughout the neuron. It is particularly patchy in the dendritic territory and is influenced by the morphology of the spine and the age of the preparation [84]. Thus, in adult animals, less than 25% of the thin spines contain SER, whereas almost 90% of the larger mushroom spines have SER. When all types of spines are analyzed, SER is present only in 50% of the spines (slightly more in the young preparations: 58% in immature brains vs. 48% in the adult spines). The important implication of these observations is that in about half of the spines, where synaptic activity presumably is still taking place, the entry of Ca2+ is buffered through mechanisms that do not involve ER Ca2+ reuptake. Where present, however, the ER is ideally adapted for Ca2+ uptake as its surface-to-ratio area is very large, with the SER occupying only about 3% of the volume but up to 40% of the surface of the spine territory.
- In addressing the issue of the participation of the CICR in the control and modulation of the Ca2+ synaptic response, there appear to be differences between the various types of CNS neurons, particularly the cerebellar Purkinje neurons and the hippocampal pyramidal neurons, in respect to the mechanisms of release of Ca2+ from intracellular ER stores. In Purkinje neurons, the main ER Ca2+ release pathway is the InsP3 receptor. As demonstrated in 1998 by two independent, but almost simultaneous, studies [85, 86], parallel fiber stimulation activates postsynaptically, even in the absence of membrane depolarization, a Ca2+ signal mediated by the activation of the mGluR1 receptors and which is able to trigger LTD. For the hippocampal neurons, the picture is more complex. For a start, the distribution of the Ca2+ release channels on the ER varies with the localization for the CA1 neurons; there are no InsP3 receptors, only RynR, in the spines, whereas in the dendritic shaft both receptors are present [87]. Furthermore, whereas some studies asserted a significant role for the Ca2+-induced Ca2+ release (CICR) process in postsynaptic Ca2+ signaling [88], others failed to detect a significant contribution [89].
- Intracellular Ca2+ stores can serve as “sinks” or “sources,” and the role played by the ER Ca2+ stores depends to a significant degree on their filling status. It was shown in earlier studies that Ca2+ responses that involve mobilization from intracellular stores were significantly enhanced by depolarization protocols that allowed their filling [90, 91], indicating that the stores are functionally deplete and require constant uptake, and that the ER can play a dominant role as a Ca2+ buffering organelle following a Ca2+ challenge, acting on a much longer time-scale than the mitochondria [92]. This action explains why inhibition of the ER Ca2+ reuptake system can result in amplification of the Ca2+ signals [93].
Information about age-related changes in the function and activity of the ER Ca2+ stores is rather limited, with an overall view that the size of the caffeine-releasable ER Ca2+ store is reduced by aging, as shown both in granule neurons of the cerebellar slices [94] and in aged acutely dissociated basal forebrain neurons [80], but not in cultured hippocampal neurons [95]. There are several mechanisms that can explain a decrease in the size of the caffeine-sensitive Ca2+ stores, including a decreased efficiency of the reuptake mechanisms through the SERCA pumps or an increased Ca2+ leakage. When investigated directly, in acutely dissociated basal forebrain neurons, aging did not affect the rate of spontaneous depletion (i.e., leakage) but decreased the efficiency of the ER loading [96]. In contrast, in long-term (30 days in vitro) cultured hippocampal neurons, glutamate stimulation activated a sustained Ca2+ response that was inhibited by rynanodine, a blocker of CICR, thus indicating an increased ER Ca2+ leak [95]. It is not clear yet if this increased Ca2+ leak, in effect a sustained CICR, was due to an alteration in the function of the rynanodine receptors or was related to a possible coupling between the L-type VOCCs and the rynanodine Ca2+-release channels [97], similar to the coupling between these types of Ca2+ channels in the muscle and which is known to be affected by the aging process [98]. An interesting observation on the role of intracellular Ca2+ stores in alterations in the normal Ca2+ homeostasis in the aged neurons has been recently reported with respect to the process of hippocampal LTP induction. Release of Ca2+ from the intracellular stores is an important participant in LTP induction for ranges of stimulation near the threshold of induction [99, 100]; but in the aged slices, inhibition of the Ca2+ release from the stores by either SERCA inhibitors or by rynanodine activated, rather than inhibited, LTP [101]. The explanation proposed for this paradoxical effect takes into account another well-established effect of Ca2+ in the aged neurons: activation of a slow afterhyperpolarization (AHP) current [102, 103], which in turn will affect the level of synaptic depolarization required for NMDA receptor and consequent LTP induction. The fact that aging is associated with a decline in LTP and synaptic plasticity [104, 105] might indicate an increase with age of the functional release of Ca2+ from the intracellular stores in these hippocampal neurons.
This area of discussion, the regulation of ER Ca2+ stores, is probably the most important one for appreciating the functional differences between normal physiological aging and the process of neurodegeneration, particularly AD, for which significant alterations in the ER Ca2+ handling were described [11]. Both bulwarks of pathology in AD (i.e., the Aβ protein fragments and the presenilins) are associated with or induce alterations in Ca2+ homeostasis [106, 107], although is not yet clear whether the changes in Ca2+ homeostasis precede the protein alterations in AD or vice versa [108]. One of the best-established mechanisms through which Aβ affects [Ca2+]i regulation is by the formation by the Aβ fragment of a cation-selective ion channel [109] with Ca2+ permeability. One effect of aggregated Aβ is the increase of the oxidative stress, a mechanisms that has been linked with increases in the resting [Ca2+]i [110]. The other culprits in AD pathogenesis are the presenilins (PS), which act as essential constituents of the β-secretase complex that cleaves a number of membrane proteins, including the amyloid precursor protein (APP), the latter generating the Aβ fragments [111]. Experimental manipulations of expression of either of the two genes controlling PS (PS1 and PS2) in various animal models resulted in phenotypes with disrupted Ca2+ homeostasis characterized, in the main, by an ER Ca2+ store overload [11]. As a result, neurons expressing such mutations show increased vulnerability to glutamatergic stimulation as shown, for example, in the case of hippocampal neurons, both in slices [112] and in primary cell cultures [107]. The overloading of the internal stores appear to affect the release of Ca2+ mediated either by the rynanodine receptor [113] or by the InsP3 receptor [114]. However, such drastic effects of presenilin mutations on ER Ca2+ homeostasis might indicate a possible physiological role, and it has been shown recently that PS-1 deficiency (−/− homozygotes) can impair the glutamate-evoked Ca2+ signals, although the resting [Ca2+]i levels were not affected [115]. Overall, these data illustrate the differences between normal aging and AD. In the former case, ER Ca2+ homeostasis is largely normal, with some tendency toward a reduction in the size of the Ca2+ store. In contrast, in the AD models, the ER stores are overfilled and the major dysfunctional proteins in AD are interfering with cellular Ca2+ homeostasis.
2. Mitochondria
On the back of the “free radicals” theory of aging [116], and involving the effects of free radicals on mitochondrial DNA, Miquel [117] proposed the “mitochondrial theory” of aging more than two decades ago. In the intervening years, the role and participation of mitochondria in cellular Ca2+ homeostasis became clearer, and this relationship came to be included among the important processes that underlie the functional changes of normal brain aging [35, 118]. The role of mitochondria as regulators of the cell death mode, dictating the progression toward necrosis or apoptosis, has been well established for some time now [119] but, as discussed above, normal neuronal aging is not about cellular death and the role of apoptosis in brain aging is relatively small in contrast to the situation in neurodegenerative diseases [119–121]. Instead, the role of mitochondria as the primary site of neuronal energy production and major source of metabolically linked release of reactive oxygen and nitrogen species (ROS and RNS) is much more important for the process of cellular aging [122] (Figure 14.3). Also, it should be noted that mitochondria, through their capacity of taking up or releasing Ca2+, can have a dramatic sculpting effect on the cytosolic Ca2+ signal.
The main drive for Ca2+ uptake is the steep mitochondrial membrane potential (−180 mV) generated by the activity of the electron transporters in the respiratory chain that couple electron transfer to proton extrusion (Figure 14.3). The transport of Ca2+ is implemented by a Ca2+ uniporter, which has only very recently been identified as a channel protein [123], that has a high affinity for Ca2+ (nanomolar range); is impermeable to K+ and Mg2+, ions that are abundant in the cytosol, and is inhibited by ruthenium red. The issue of Ca2+ affinity is problematic because functional studies have shown this mitochondrial Ca2+ uptake pathway as a low-affinity, high-capacity Ca2+ removal system [124]. It might then be more useful to talk about a Ca2+ uptake set point, at which the uptake and efflux of Ca2+ from the mitochondrial matrix are balanced, and above which net uptake of Ca2+ takes place [125]. One of the very important features of mitochondrial Ca2+ uptake is the potentially huge capacity of this Ca2+ store, a property that is based on the co-accumulation of phosphate together with Ca2+ to form a salt complex that is both rapidly dissociable and also osmotically inactive. In isolated mitochondria incubated in physiological buffers that mimic intracellular solutions, Ca2+ can reach concentrations of around 100 mM (i.e., 50 times larger than the extracellular Ca2+ concentrations), while maintaining a normal metabolic state [126]. More recent measurements of Ca2+ changes in intact cellular systems (lizard motor nerve terminals) confirmed the earlier data and showed that the accumulation of mitochondrial Ca2+ takes place while the concentration of free Ca2+ in the mitochondrial matrix ([Ca2+]mito) is maintained at stable levels [127].
This maintenance of a tight control on [Ca2+]mito is essential as this parameter is a sword with two edges. Moderate increases in [Ca2+]mito activate three important mitochondrial enzymes (pyruvate-, isocitrate-, and 2-oxoglutarate-dehydorgenases) that are involved in the Krebs (citric acid) cycle. The consequent increase in the reduction state of the NADH/NAD system will induce an increased electron transport and thus couple a cytosolic event that generated a Ca2+ response to a temporary increase in ATP supply, a process called “metabolic coupling” [128]. At the other extreme, significant increases in [Ca2+]mito will activate, together with other factors, the permeability transition pore (mPTP) and result in catastrophic consequences for the mitochondria [129], among which mitochondrial swelling is an early indicator. It is important to note that there is a likely tissue-specific heterogeneity in the activation of the mPTP, with liver mitochondria appearing to be more susceptible to Ca2+-dependent activation than brain mitochondria, at least in isolated mitochondria preparations [130].
While the opening of the mPTP would lead to a rapid, unregulated redistribution of the Ca2+ accumulated in the matrix back to the cytosol, mitochondria have other transport systems that effect mitochondrial Ca2+ release. Two main pathways have been described, with a certain tissue specificity: an Na+/Ca2+ exchange system that is predominant in the brain, heart, and skeletal muscle, whereas in the liver, kidney, and smooth muscles, Ca2+ efflux is predominantly Na+ independent [131].
In discussing the effects of mitochondria on cellular Ca2+ homeostasis, the spatial dimension also must be take into account because, as demonstrated mainly in non-excitable cells, mitochondria, acting on limited spatial domains (functional “micro-domains”), can have significant effects on the Ca2+ release from the ER. This effect also takes into account the fact that cytosolic Ca2+ has a biphasic effect on ER Ca2+ release, both at the InsP3 and the rynanodine receptors [16, 132]. Removal of Ca2+ from the immediate vicinity of the ER Ca2+ release channels should enhance the release and amplify the signal, and indeed, in Xenopus oocytes, enhancement of mitochondrial respiration increases the amplitude and velocity of the InsP3-generated cytosolic Ca2+ waves [133]. In a similar fashion, the mitochondrial Ca2+ uptake could influence the entry of Ca2+ from the extracellular medium. One path responsible for Ca2+ entry and sensitive to the energetic status of the mitochondria, strategically placed in the submembranar space, is the capacitative Ca2+ entry pathway [134], activated by the depletion of the intracellular Ca2+ stores and inhibited either by ER store refilling or by permeating Ca2+ ions. To date, the evidence for such a pathway effective in the central nervous system’s neurons is rather thin [135]. Although not yet investigated in detail, the same subplasmalemmal mitochondria also could modulate the activity of other Ca2+ channels that are inhibited by cytosolic Ca2+ increases, such as the voltage-operated Ca2+ channels [136]. Not only can Ca2+ uptake modulate neuronal function, but the mitochondrial release of Ca2+ can also affect neuronal physiology. Thus, tetanic stimulation of Xenopus motorneurons induces a potentiation of neurotransmitter release that is sensitive to the Ca2+ release from the mitochondria, but not from the ER [137].
To date, there is almost no information about the effects of aging on the biochemical properties of the mitochondrial Ca2+ transporters, either for uptake or for release, and this should be an important area of future research, both in the field of normal aging and for various neurodegenerative instances. However, other well-documented changes in mitochondrial status with age are affecting mitochondrial Ca2+ homeostasis. Of these, a chronic change in the mitochondrial membrane potential that progresses with age is one of the better-established features. Using a fluorescent dye (rhodamine 123) that equilibrates across cellular compartments as a function of the electrical potential across these membranes (Nerstian distribution), and thus accumulates preferentially in the mitochondria [138], it has been shown that in the aged preparations of isolated mitochondria, there is a population of mitochondria that is significantly depolarized [139]. Because the use of rhodamine has a number of important limitations — in particular, overloading of the mitochondrial compartment with a consequent quenching [138, 140] — other methods are available, including the use of a recording electrode sensitive to an ion with Nerstian distribution, tetraphenylphosphonium [141]. Use of such more precise techniques confirmed the chronic depolarization of aged mitochondria [142]. Because these experiments used preparations of isolated mitochondria that are susceptible to selective enrichment in viable mitochondria, and thus a potential underestimating bias, particularly for the more fragile aged cells [143], it was important to assess the situation in intact cellular preparations. Using the properties of protonophores to collapse the mitochondrial membrane potential and thus release the rhodamine accumulated in the mitochondria, it has been shown that in the aged neurons there is a significant age-dependent decrease in the amount of dye accumulated [143], in line with the idea of mitochondria in the aged cells being more depolarized. Metabolic control analysis showed, consistent with these ideas, that aged mitochondria have an increased mitochondrial proton leak, but it is difficult to assess if this observation is a cause or effect of reported mitochondrial depolarization, because in these experiments the aged mitochondria did not show a decreased membrane potential [144].
Clearly, a decrease in the mitochondrial potential would result in a decrease in the electrochemical force driving the mitochondrial Ca2+ uptake. Using simultaneous measurements of both cytosolic Ca2+ and mitochondrial depolarization response, it was estimated that in the aged cerebellar neurons there was an increase in the [Ca2+]i threshold required for triggering mitochondrial uptake [14, 145]. Associated with this question of the threshold for mitochondrial uptake is the issue of the size and dynamics of the mitochondrial Ca2+ pool. It has been shown previously that Ca2+ mitochondrial loading (as assessed by a protonophore-induced Ca2+ release) depends on the level and duration of stimulation and that this loading is dynamic, with a spontaneous discharge over a period of minutes [146]. Also, the extent of mitochondrial depolarization resulting from mitochondrial loading depends on the type of stimulus, with significant differences between KCl and glutamate [147], or even among the various types of glutamatergic agonists [148]. The existence of possible changes with age in these parameters of mitochondrial Ca2+ storage currently is under investigation. Some early results show, for the cerebellar granule neurons, that following glutamatergic stimulation the major determinant of the size of the mitochondrial Ca2+ accumulation is the cytoplasmic Ca2+ levels irrespective of the age of the preparations, and thus, by extension, of the level of chronic mitochondrial depolarization [149].
Neuronal stimulation means mitochondrial depolarization but our understanding of the effects of this mitochondrial depolarization on the state of the cell is evolving. In some studies, mitochondrial depolarization that follows glutamate stimulation appears to be the early event that initiates excitotoxic death [150–152]. However, other studies have reported an opposite effect, and mitochondrial depolarization preceding the glutamatergic stimulation proved a very effective neuroprotective protocol [153, 154]. Rather than being exclusive, these data, taken together, indicate that the mitochondrial Ca2+ — either as amount of Ca2+ loading or as level of free matrix Ca2+ — is a dynamic parameter that controls the physiological state of the cell. One target of mitochondrial Ca2+ is the regulation of the mPTP opening, mentioned above, and not surprisingly inhibition of the mPTP proved an effective neuroprotective strategy, effective in a variety of pathological instances but also improving the status of aged mitochondria [155].
Another potential target for increased mitochondrial Ca2+ is the production of ROS, which has been demonstrated to follow NMDA stimulation [156], in a manner sensitive to removal of Ca2+ [157]. However, the interpretation of studies reporting measurements of ROS production is problematic, because the acute, real-time detection of free radicals (mainly through the use of fluorescent dyes) is notoriously difficult because of (1) the selectivity of various dyes for different ROS species, (2) the different sites at which the ROS are produced and their lifetime, and (3) the fact that the fluorescence of the dyes could be influenced by other enzymatic and non-enzymatic processes [158]. In addition, the bioenergetic process that leads from increased matrix Ca2+ to increased free radicals is not well established. Electron leak, generating free radicals, occurs predominantly within complex III in the respiratory chain, during the so-called “Q cycle,” although complex I can also participate in the ROS generation process [159]. The crucial feature of this process is that it requires high mitochondrial membrane potential [160], and thus the Ca2+ uptake should inhibit rather than evoke an increase of ROS. One proposed explanation invokes (1) the stimulatory effect of Ca2+ on the citric acid cycle, mentioned above and that should accelerate the electron transfer rate; and (2) an effect of Ca2+ on nitric oxide (NO) generation, which in turn would inhibit the activity at the complex IV in the respiratory chain creating further favorable conditions for electron leak at complex III [161]. However, with respect to the effect of the respiration rate on the generation of ROS, a point of subtlety was made in a recent review by Nicholls [162]. The cytosolic environment is well hypoxic with respect to the values of the partial pressure of oxygen in the circulation or in the tissues, due to the increased diffusion pat, but even in these conditions oxygen, as a substrate for the final four-electron reduction to water, is in excess, and in sufficient supply. Under these conditions, the single electron reduction that generates free radicals is not simply a proportional slippage of the main electron carrier path, but rather the contrary. The generation of free radicals is inversely proportional to the rate of respiration, such that the higher the respiration rate and the lower the membrane potential, the less time will the electrons spend at the sites of leakage [162].
IV. AGING AND NEURODEGENERATION
Many times, “aging” and “neurodegeneration” are mentioned in the same breath, as if they are two state-points along a continuum, leading from the normality of old age to the state of confusion of senile dementia. Indeed, Terry and Katzman made this explicit prediction: they started from the observation that dementia occurs when there is a loss of about 40% of neocortical synapses and combined this with their estimations of the rate of age-dependent synaptic loss, as assessed by their analyses in human postmortem samples, extrapolating the data to the current estimations of human lifespan [163]. Apart from the anecdotal evidence of many centenarians that age successfully [164], including the Guinness Book of Records’ oldest certified human being, Madame Calment, who died in 1997 at the age of over 122 years without showing any signs of clinical dementia, the set of data that underlies the provocative hypothesis mentioned above might have been corrupted, unknowingly, by the inclusion in the analysis of subclinical instances of neurodegenerative pathology. Indeed, a similar error crept in the early sets of data that led to the conclusion that normal aging is associated with significant neuronal loss in all regions of the brain, a conclusion that only recently has been disproved [10]. Even the term “neurodegeneration” is problematic, despite the obvious etymology, because the term has been used in reference to a large group of neurological diseases, with heterogeneous clinical and pathological manifestations, affecting specific, but different regions of the nervous system [165]. Another experimental observation that complicates the issues is that the pathological markers associated with various neurodegenerative diseases, such as Lewy bodies, neurofibrillary tangles (NFTs), senile plaques, or other protein depositions, can be detected in the brain of aged asymptomatic individuals [166]. These observations obviously raise the question of whether such individuals were just aging normally or were sampled at a presymptomatic stage of a neurodegenerative disease — there are, as yet, no definite answers to these issues. The study of the relationship between normal aging and neurode-generation can be significantly helped by studies on animal models, because none of the human neurodegenerative diseases appear naturally in the animals. Through genetic manipulations, various animal models have been constructed that mimic one or another of the neurodegenerative phenotypes [167]. At the same time, behavioral studies allow a more and more detailed assessment of the cognitive status of the animals, leading to the generation of sophisticated models of learning and memory and the study of the effects of aging on such processes [105].
The animal studies also allowed a very detailed analysis of the metabolic changes associated with the process of normal aging, as reviewed in this chapter. What these studies indicate is that normal neuronal aging is a physiological process, characterized primarily by a decrease in the neuronal homeostatic reserve, where this homeostatic reserve is defined as the capacity of the cells to oppose the destabilizing effects of various metabolic stressors (Figure 14.4). A detailed discussion of the evidence was presented in earlier articles [122, 149]. Briefly, this hypothesis is based on the view that central to the changes in the cellular physiology of the aged neurons is a dysfunction of the metabolic triad: Ca2+ – mitochondria – ROS. It also states that the functional deficits of the aged neurons become evident only in a use-dependent manner, when the metabolic demands become excessive. This view accounts also for the lack of significant levels of neuronal loss with aging. On this weakened homeostatic reserve, any significant pathological instance that requires a strong metabolic response, be it trauma, stroke, or any idiopathic neurodegenerative process, would result in a significant level of neuronal death. In this model, it is the decreased homeostatic reserve that explains the increased neuronal vulnerability with age, an increased vulnerability that manifests itself in a variety of pathological scenarios.
The fact that aging is “just” a functional state should be very encouraging. We might not be in a position to bring on immortality through engineering negligible senescence, but a better understanding of the metabolism of the aged neuron could open important avenues for therapeutic intervention that will impede or delay the possible cognitive decline of the aged.
ACKNOWLEDGMENTS
The author wishes to acknowledge the BBSRC, which provided, through the SAGE Initiative, a significant part of the financial support for the work performed in this laboratory. The author is also grateful to Professor A. Verkhratsky for discussions on various aspects of the work reported here. Siddhartha helped by providing a new perspective.
REFERENCES
- 1.
- Khachaturian ZS. Towards theories of brain aging. In: Kay D, Burrows G, editors. Handbook of Studies on Psychiatry and Old Age. Elsevier; Amsterdam: 1984. p. 7.
- 2.
- Khachaturian ZS. Calcium hypothesis of Alzheimer’s disease and brain aging. Ann NY Acad Sci. 1994;747:1. [PubMed: 7847664]
- 3.
- Campbell AK. Intracellular Calcium — Its Universal Role as Regulator. John Wiley & Sons; New York: 1983.
- 4.
- Berridge MJ, Galione A. Cytosolic calcium oscillators. FASEB J. 1988;2:3074. [PubMed: 2847949]
- 5.
- Olney JW, Ho OL. Brain damage in infant mice following oral intake of glutamate, aspartate or cysteine. Nature. 1970;227:609. [PubMed: 5464249]
- 6.
- Choi DW, Maulucci-Gedde M, Kriegstein AR. Glutamate neurotoxicity in cortical cell culture. J Neurosci. 1987;7:357. [PMC free article: PMC6568898] [PubMed: 2880937]
- 7.
- Hof PR, Morrison JH. The aging brain: morphomolecular senescence of cortical circuits. Trends Neurosci. 2004;27:607. [PubMed: 15374672]
- 8.
- Sterio DC. The unbiased estimation of number and sizes of arbitrary particles using the disector. J Microsci. 1984;134:127. [PubMed: 6737468]
- 9.
- Gómez-Isla T, et al. Profound loss of layer II entorhinal cortex neurons occurs in very mild Alzheimer’s disease. J Neurosci. 1996;16:4491. [PMC free article: PMC6578866] [PubMed: 8699259]
- 10.
- Morrison J, Hof P. Life and death of neurons in the aging brain. Science. 1997;278:412. [PubMed: 9334292]
- 11.
- LaFerla FM. Calcium dyshomeostasis and intracellular signaling in Alzheimer disease. Nat Rev Neurosci. 2002;3:862. [PubMed: 12415294]
- 12.
- Catterall WA, et al. International Union of Pharmacology. XL. Compendium of voltage-gated ion channels: calcium channels. Pharmacol Rev. 2003;55:579. [PubMed: 14657414]
- 13.
- Trimmer JS, Rhodes KJ. Localization of voltage-gated ion channels in mammalian brain. Annu Rev Physiol. 2004;66:477. [PubMed: 14977411]
- 14.
- Xiong J, et al. Mitochondrial polarisation status and [Ca 2+]i signaling in rat cerebellar granule neurons aged in vitro. Neurobiol. Aging. 2004;25:349. [PubMed: 15123341]
- 15.
- Solovyova N, et al. Ca2+ dynamics in the lumen of the endoplasmic reticulum in sensory neurons: direct visualization of Ca2+-induced Ca2+ release triggered by physiological Ca(2+) entry. Embo J. 2002;21:622. [PMC free article: PMC125857] [PubMed: 11847110]
- 16.
- Verkhratsky A. Physiology and pathophysiology of the calcium store in the endoplasmic reticulum of neurons. Physiol Rev. 2005;85:201. [PubMed: 15618481]
- 17.
- Paschen W. Endoplasmic reticulum: a primary target in various acute disorders and degenerative diseases of the brain. Cell Calcium. 2003;34:365. [PubMed: 12909082]
- 18.
- Carafoli E. Calcium signaling: a tale for all seasons. Proc Natl Acad Sci USA. 2002;99:1115. [PMC free article: PMC122154] [PubMed: 11830654]
- 19.
- Burgoyne RD, et al. Neuronal Ca2+-sensor proteins: multitalented regulators of neuronal function. Trends Neurosci. 2004;27:203. [PubMed: 15046879]
- 20.
- Buxbaum JD. A role for calsenilin and related proteins in multiple aspects of neuronal function. Biochem Biophys Res Commun. 2004;322:1140. [PubMed: 15336961]
- 21.
- Baimbridge KG, Celio MR, Rogers JH. Calcium-binding proteins in the nervous system. Trends Neurosci. 1992;15:303. [PubMed: 1384200]
- 22.
- Augustine GJ, Santamaria F, Tanaka K. Local calcium signaling in neurons. Neuron. 2003;40:331. [PubMed: 14556712]
- 23.
- Mattson MP, et al. Evidence for calcium-reducing and excito-protective roles for the calcium-binding protein calbindin-D28k in cultured hippocampal neurons. Neuron. 1991;6:41. [PubMed: 1670921]
- 24.
- Lukas W, Jones KA. Cortical neurons containing calretinin are selectively resistant to calcium overload and excitotoxicity in vitro. Neuroscience. 1994;61:307. [PubMed: 7969911]
- 25.
- Dowd DR. Calcium regulation of apoptosis. Adv Second Messenger Phosphoprotein Res. 1995;30:255. [PubMed: 7695993]
- 26.
- Dekkers J, et al. Over-expression of parvalbumin in transgenic mice rescues motoneurons from injury-induced cell death. Neuroscience. 2004;123:459. [PubMed: 14698753]
- 27.
- Lee SH, Schwaller B, Neher E. Kinetics of Ca2+ binding to parvalbumin in bovine chromaffin cells: implications for [Ca2+] transients of neuronal dendrites. J. Physiol. 2000;525(Pt. 2):419. [PMC free article: PMC2269947] [PubMed: 10835044]
- 28.
- Lee SH, et al. Differences in Ca2+ buffering properties between excitatory and inhibitory hippocampal neurons from the rat. J. Physiol. 2000;525(Pt. 2):405. [PMC free article: PMC2269951] [PubMed: 10835043]
- 29.
- Palecek J, Lips MB, Keller BU. Calcium dynamics and buffering in motoneurons of the mouse spinal cord. J. Physiol. 1999;520(Pt. 2):485. [PMC free article: PMC2269591] [PubMed: 10523417]
- 30.
- Wanaverbecq N, et al. The plasma membrane calcium-ATPase as a major mechanism for intracellular calcium regulation in neurons from the rat superior cervical ganglion. J Physiol. 2003;550:83. [PMC free article: PMC2343008] [PubMed: 12879862]
- 31.
- Fierro L, Llano I. High endogenous calcium buffering in Purkinje cells from rat cerebellar slices. J. Physiol. 1996;496(Pt. 3):617. [PMC free article: PMC1160850] [PubMed: 8930830]
- 32.
- Verkhratsky A, Toescu E. Calcium and neuronal aging. Trends Neurosci. 1998;21:2. [PubMed: 9464677]
- 33.
- Pitler TA, Landfield PW. Aging-related prolongation of calcium spike duration in rat hippocampal slice neurons. Brain Res. 1990;508:1. [PubMed: 2337778]
- 34.
- Thibault O, et al. Calcium dysregulation in neuronal aging and Alzheimer’s disease: history and new directions. Cell Calcium. 1998;24:417. [PubMed: 10091010]
- 35.
- Toescu EC, Verkhratsky A. Ca2+ and mitochondria as substrates for deficits in synaptic plasticity in normal brain aging. J Cell Mol Med. 2004;8:181. [PMC free article: PMC6740225] [PubMed: 15256066]
- 36.
- Thibault O, Landfield PW. Increase in single L-type calcium channels in hippocampal neurons during aging. Science. 1996;272:1017. [PubMed: 8638124]
- 37.
- Murchison D, Griffith WH. Low-voltage activated calcium currents increase in basal forebrain neurons from aged rats. J Neurophysiol. 1995;74:876. [PubMed: 7472390]
- 38.
- Murchison D, Griffith WH. High-voltage-activated calcium currents in basal forebrain neurons during aging. J Neurophysiol. 1996;76:158. [PubMed: 8836216]
- 39.
- Kostyuk P, et al. Calcium currents in aged rat dorsal root ganglion neurons. J Physiol Lond. 1993;461:467. [PMC free article: PMC1175267] [PubMed: 8394426]
- 40.
- Govoni S, et al. Age-related reduced affinity in 3H nitrendipine labeling of brain voltage-dependent calcium channels. Brain Res. 1985;333:374. [PubMed: 2581663]
- 41.
- Alshuaib WB, Fahim MA. Aging increases calcium influx at motor nerve terminal. Int J Dev Neurosci. 1990;8:655. [PubMed: 2126908]
- 42.
- Martinez-Serrano A, et al. Reduction of K+-stimulated 45Ca2+ influx in synaptosomes with age involves inactivating and noninactivating calcium channels and is correlated with temporal modifications in protein dephosphorylation. J Neurochem. 1989;52:576. [PubMed: 2463338]
- 43.
- Farrar RP, et al. Neurosci Lett. Vol. 100. 1989. Aging does not alter cytosolic calcium levels of cortical synaptosomes in Fischer 344 rats; p. 319. [PubMed: 2761782]
- 44.
- Chung YH, et al. Differential alterations in the distribution of voltage-gated calcium channels in aged rat cerebellum. Brain Res. 2001;903:247. [PubMed: 11382411]
- 45.
- Burnashev N. Calcium permeability of ligand-gated channels. Cell Calcium. 1998;24:325. [PubMed: 10091002]
- 46.
- Dingledine R, et al. The glutamate receptor ion channels. Pharmacol Rev. 1999;51:7. [PubMed: 10049997]
- 47.
- Perez-Otano I, et al. Assembly with the NR1 subunit is required for surface expression of NR3A-containing NMDA receptors. J Neurosci. 2001;21:1228. [PMC free article: PMC6762235] [PubMed: 11160393]
- 48.
- Nowak L, et al. Magnesium gates glutamate-activated channels in mouse central neurons. Nature. 1984;307:462. [PubMed: 6320006]
- 49.
- Magnusson KR. The aging of the NMDA receptor complex. Front. Biosci. 1998;3:e70. [PubMed: 9576682]
- 50.
- Monyer H, et al. Developmental and regional expression in the rat brain and functional properties of four NMDA receptors. Neuron. 1994;12:529. [PubMed: 7512349]
- 51.
- Clayton DA, Browning MD. Deficits in the expression of the NR2B subunit in he hippocampus of aged Fisher 344 rats. Neurobiol Aging. 2001;22:165. [PubMed: 11164294]
- 52.
- Cepeda C, Li Z, Levine MS. Aging reduces neostriatal responsiveness to N-methyl-D-aspartate and dopamine: an in vitro electrophysiological study. Neuroscience. 1996;73:733. [PubMed: 8809794]
- 53.
- Barnes CA, Rao G, Shen J. Age-related decrease in the N-methyl-D-aspartate-mediated excitatory postsynaptic potential in hippocampal region CA1. Neurobiol Aging. 1997;18:445. [PubMed: 9330977]
- 54.
- Cady C, Evans MS, Brewer GJ. Age-related differences in NMDA responses in cultured rat hippocampal neurons. Brain Res. 2001;921:1. [PubMed: 11720706]
- 55.
- Verdoorn TA, et al. Structural determinants of ion flow through recombinant glutamate receptor channels. Science. 1991;252:1715. [PubMed: 1710829]
- 56.
- Mishina M, et al. A single amino acid residue determines the Ca2+ permeability of AMPA-selective glutamate receptor channels. Biochem Biophys Res Commun. 1991;180:813. [PubMed: 1659403]
- 57.
- Seeburg PH. The TINS/TiPS Lecture. The molecular biology of mammalian glutamate receptor channels. Trends Neurosci. 1993;16:359. [PubMed: 7694406]
- 58.
- Lomeli H, et al. Control of kinetic properties of AMPA receptor channels by nuclear RNA editing. Science. 1994;266:1709. [PubMed: 7992055]
- 59.
- Isa T, et al. Distribution of neurons expressing inwardly rectifying and Ca2+-permeable AMPA receptors in rat hippocampal slices. J Physiol. 1996;491:719. [PMC free article: PMC1158813] [PubMed: 8815206]
- 60.
- Itazawa SI, Isa T, Ozawa S. Inwardly rectifying and Ca2+-permeable AMPA-type glutamate receptor channels in rat neocortical neurons. J Neurophysiol. 1997;78:2592. [PubMed: 9356409]
- 61.
- Magnusson KR. Influence of dietary restriction on ionotropic glutamate receptors during aging in C57B1 mice. Mech Aging Dev. 1997;95:187. [PubMed: 9179830]
- 62.
- Magnusson KR, Cotman CW. Age-related changes in excitatory amino acid receptors in two mouse strains. Neurobiol Aging. 1993;14:197. [PubMed: 8391661]
- 63.
- Magnusson KR. Aging of glutamate receptors: correlations between binding and spatial memory performance in mice. Mech Aging Dev. 1998;104:227. [PubMed: 9818728]
- 64.
- Barnes CA, et al. Region-specific age effects on AMPA sensitivity: electrophysiological evidence for loss of synaptic contacts in hippocampal field CA1. Hippocampus. 1992;2:457. [PubMed: 1284976]
- 65.
- Carlson NG, et al. RNA editing (Q/R site) and flop/flip splicing of AMPA receptor transcripts in young and old brains. Neurobiol Aging. 2000;21:599. [PubMed: 10924778]
- 66.
- Tanaka H, et al. The AMPAR subunit GluR2: still front and center-stage. Brain Res. 2000;886:190. [PubMed: 11119696]
- 67.
- Michaelis M, et al. Decreased plasma membrane calcium transport activity in aging brain. Life Sci. 1996;59:405. [PubMed: 8761328]
- 68.
- Zaidi A, et al. Age-related decrease in brain synaptic membrane Ca2+-ATPase in F344/BNF1 rats. Neurobiol Aging. 1998;19:487. [PubMed: 9880051]
- 69.
- Zaidi A, Michaelis ML. Effects of reactive oxygen species on brain synaptic plasma membrane Ca2+-ATPase. Free Radic Biol Med. 1999;27:810. [PubMed: 10515585]
- 70.
- Michaelis ML, et al. Decreased plasma membrane calcium transport activity in aging brain. Life Sci. 1996;59:405. [PubMed: 8761328]
- 71.
- Tsai H, et al. Adrenergic nerve smooth endoplasmic reticulum calcium buffering declines with age. Neurobiol Aging. 1998;19:89. [PubMed: 9562509]
- 72.
- Pottorf WJ, Duckles SP, Buchholz JN. Aging and calcium buffering in adrenergic neurons. Auton Neurosci. 2002;96:2. [PubMed: 11911497]
- 73.
- De Jong GI, et al. Age-related loss of calcium binding proteins in rabbit hippocampus. Neurobiol Aging. 1996;17:459. [PubMed: 8725908]
- 74.
- Bu J, et al. Age-related changes in calbindin-D28k, calretinin, and parvalbumin-immunoreactive neurons in the human cerebral cortex. Exp Neurol. 2003;182:220. [PubMed: 12821392]
- 75.
- Geula C, et al. Loss of calbindin-D28k from aging human cholinergic basal forebrain: relation to neuronal loss. J Comp Neurol. 2003;455:249. [PubMed: 12454989]
- 76.
- Idrizbegovic E, et al. The total number of neurons and calcium binding protein positive neurons during aging in the cochlear nucleus of CBA/CaJ mice: a quantitative study. Hear Res. 2001;158:102. [PubMed: 11506942]
- 77.
- Idrizbegovic E, Bogdanovic N, Canlon B. Modulating calbindin and parvalbumin immunoreactivity in the cochlear nucleus by moderate noise exposure in mice. A quantitative study on the dorsal and posteroventral cochlear nucleus. Brain Res. 1998;800:86. [PubMed: 9685593]
- 78.
- Greene JR, et al. Accumulation of calbindin in cortical pyramidal cells with aging; a putative protective mechanism which fails in Alzheimer’s disease. Neuropathol Appl Neurobiol. 2001;27:339. [PubMed: 11679085]
- 79.
- Tsai H, et al. Intracellular calcium buffering declines in aging adrenergic nerves. Neurobiol Aging. 1997;18:229. [PubMed: 9258901]
- 80.
- Murchison D, Griffith W. Increased calcium buffering in basal forebrain neurons during aging. J Neurophysiol. 1998;80:350. [PubMed: 9658056]
- 81.
- Griffith WH, et al. Modification of ion channels and calcium homeostasis of basal forebrain neurons during aging. Behav Brain Res. 2000;115:219. [PubMed: 11000422]
- 82.
- Murchison D, Zawieja DC, Griffith WH. Reduced mitochondrial buffering of voltage-gated calcium influx in aged rat basal forebrain neurons. Cell Calcium. 2004;36:61. [PubMed: 15126057]
- 83.
- Wu CK, Geula C. Selective loss of calbindin D-28k from cholinergic neurons of the human basal forebrain in aging and Alzheimer’s disease. In: Igbal K, Winblad B, Nishimura T, Takeda M, Wisniewski HM, editors. Alzheimer’s Disease Biology Diagnosis and Therapeutics. John Wiley & Sons; New York: 1997. p. 297.
- 84.
- Spacek J, Harris KM. Three-dimensional organization of smooth endoplasmic reticulum in hippocampal CA1 dendrites and dendritic spines of the immature and mature rat. J Neurosci. 1997;17:190. [PMC free article: PMC6793680] [PubMed: 8987748]
- 85.
- Finch EA, Augustine GJ. Local calcium signaling by inositol-1,4,5-trisphos-phate in Purkinje cell dendrites. Nature. 1998;396:753. [PubMed: 9874372]
- 86.
- Takechi H, Eilers J, Konnerth A. A new class of synaptic response involving calcium release in dendritic spines. Nature. 1998;396:757. [PubMed: 9874373]
- 87.
- Svoboda K, Mainen ZF. Synaptic [Ca2+]: intracellular stores spill their guts. Neuron. 1999;22:427. [PubMed: 10197523]
- 88.
- Emptage N, Bliss TVP, Fine A. Single synaptic events evoke NMDA receptor-mediated release of calcium from internal stores in hippocampal dendritic spines. Neuron. 1999;22:115. [PubMed: 10027294]
- 89.
- Kovalchuk Y, et al. NMDA receptor-mediated subthreshold Ca2+ signals in spines of hippocampal neurons. J Neurosci. 2000;20:1791. [PMC free article: PMC6772937] [PubMed: 10684880]
- 90.
- Garaschuk O, Yaari Y, Konnerth A. Release and sequestration of calcium by ryanodine-sensitive stores in rat hippocampal neurons. J Physiol. 1997;502:13. [PMC free article: PMC1159569] [PubMed: 9234194]
- 91.
- Irving AJ, Collingridge GL. A characterization of muscarinic receptor-mediated intracellular Ca2+ mobilization in cultured rat hippocampal neurons. J. Physiol. 1998;511(Pt. 3):747. [PMC free article: PMC2231161] [PubMed: 9714857]
- 92.
- Pozzo-Miller LD, Connor JA, Andrews SB. Microheterogeneity of calcium signaling in dendrites. J. Physiol. 2000;525(Pt. 1):53. [PMC free article: PMC2269918] [PubMed: 10811724]
- 93.
- Toescu E. Intraneuronal Ca2+ stores act mainly as a ’Ca2+ sink’ in cerebellar granule neurons. Neuroreport. 1998;9:1227. [PubMed: 9601699]
- 94.
- Kirischuk S, Verkhratsky A. Calcium homeostasis in aged neurons. Life Sci. 1996;59:451. [PubMed: 8761333]
- 95.
- Clodfelter GV, et al. Sustained Ca2+-induced Ca2+-release underlies the post-glutamate lethal Ca2+ plateau in older cultured hippocampal neurons. Eur J Pharmacol. 2002;447:189. [PubMed: 12151011]
- 96.
- Murchison D, Griffith WH. Age-related alterations in caffeine-sensitive calcium stores and mitochondrial buffering in rat basal forebrain. Cell Calcium. 1999;25:439. [PubMed: 10579055]
- 97.
- Chavis P, et al. Functional coupling between ryanodine receptors and L-type calcium channels in neurons. Nature. 1996;382:719. [PubMed: 8751443]
- 98.
- Renganathan M, Messi ML, Delbono O. Dihydropyridine receptor-ryanodine receptor uncoupling in aged skeletal muscle. J Membr Biol. 1997;157:247. [PubMed: 9178612]
- 99.
- Harvey J, Collingridge GL. Thapsigargin blocks the induction of long-term potentiation in rat hippocampal slices. Neurosci Lett. 1992;139:197. [PubMed: 1319014]
- 100.
- Raymond CR, Redman SJ. Different calcium sources are narrowly tuned to the induction of different forms of LTP. J Neurophysiol. 2002;88:249. [PubMed: 12091550]
- 101.
- Kumar A, Foster TC. Enhanced long-term potentiation during aging is masked by processes involving intracellular calcium stores. J Neurophysiol. 2004;91:2437. [PubMed: 14762159]
- 102.
- Landfield PW, Pitler TA. Prolonged Ca2+-dependent afterhyperpolarizations in hippocampal neurons of aged rats. Science. 1984;226:1089. [PubMed: 6494926]
- 103.
- Shah M, Haylett DG. Ca2+ channels involved in the generation of the slow afterhyperpolarization in cultured rat hippocampal pyramidal neurons. J Neurophysiol. 2000;83:2554. [PubMed: 10805657]
- 104.
- Foster TC. Involvement of hippocampal synaptic plasticity in age-related memory decline. Brain Res Brain Res Rev. 1999;30:236. [PubMed: 10567726]
- 105.
- Rosenzweig ES, Barnes CA. Impact of aging on hippocampal function: plasticity, network dynamics, and cognition. Prog Neurobiol. 2003;69:143. [PubMed: 12758108]
- 106.
- Larson J, et al. Alterations in synaptic transmission and long-term potentiation in hippocampal slices from young and aged PDAPP mice. Brain Res. 1999;840:23. [PubMed: 10517949]
- 107.
- Guo Q, et al. Increased vulnerability of hippocampal neurons to excitotoxic necrosis in presenilin-1 mutant knock-in mice. Nat. Med. 1999 Jan 5:101. [PubMed: 9883847]
- 108.
- Leissring MA, et al. A physiologic signaling role for the gamma-secretase-derived intracellular fragment of APP. Proc Natl Acad Sci USA. 2002;99:4697. [PMC free article: PMC123710] [PubMed: 11917117]
- 109.
- Kagan BL, et al. The channel hypothesis of Alzheimer’s disease: current status. Peptides. 2002;23:1311. [PubMed: 12128087]
- 110.
- Mark RJ, et al. Amyloid beta-peptide impairs ion-motive ATPase activities: evidence for a role in loss of neuronal Ca2+ homeostasis and cell death. J Neurosci. 1995;15:6239. [PMC free article: PMC6577674] [PubMed: 7666206]
- 111.
- Thinakaran G, Parent AT. Identification of the role of presenilins beyond Alzheimer’s disease. Pharmacol Res. 2004;50:411. [PubMed: 15304238]
- 112.
- Schneider I, 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. [PubMed: 11278803]
- 113.
- Chan SL, et al. Presenilin-1 mutations increase levels of ryanodine receptors and calcium release in PC12 cells and cortical neurons. J Biol Chem. 2000;275:18195. [PubMed: 10764737]
- 114.
- Stutzmann GE, et al. Dysregulated IP3 signaling in cortical neurons of knock-in mice expressing an Alzheimer’s-linked mutation in presenilin1 results in exaggerated Ca2+ signals and altered membrane excitability. J Neurosci. 2004;24:508. [PMC free article: PMC6729995] [PubMed: 14724250]
- 115.
- Yang Y, Cook DG. Presenilin-1 deficiency impairs glutamate-evoked intracellular calcium responses in neurons. Neuroscience. 2004;124:501. [PubMed: 14980721]
- 116.
- Harman D. Free radical theory of aging: dietary implications. Am J Clin Nutr. 1972;25:839. [PubMed: 5046729]
- 117.
- Miquel J, et al. Mitochondrial role in cell aging. Exp Gerontol. 1980;15:575. [PubMed: 7009178]
- 118.
- Toescu EC, Verkhratsky A, Landfield PW. Ca2+ regulation and gene expression in normal brain aging. Trends Neurosci. 2004;27:614. [PubMed: 15374673]
- 119.
- Mattson MP, Kroemer G. Mitochondria in cell death: novel targets for neuroprotection and cardioprotection. Trends Mol Med. 2003;9:196. [PubMed: 12763524]
- 120.
- Drew B, Leeuwenburgh C. Aging and subcellular distribution of mitochondria: role of mitochondrial DNA deletions and energy production. Acta Physiol Scand. 2004;182:333. [PubMed: 15569094]
- 121.
- Melov S. Modeling mitochondrial function in aging neurons. Trends Neurosci. 2004;27:601. [PubMed: 15374671]
- 122.
- Toescu EC, Verkhratsky A. Neuronal aging from an intraneuronal perspective: roles of endoplasmic reticulum and mitochondria. Cell Calcium. 2003;34:311. [PubMed: 12909078]
- 123.
- Kirichok Y, Krapivinsky G, Clapham DE. The mitochondrial calcium uniporter is a highly selective ion channel. Nature. 2004;427:360. [PubMed: 14737170]
- 124.
- Gunter K, Gunter T. Transport of calcium by mitochondria. J Bioenerg Biomembr. 1994;26:471. [PubMed: 7896763]
- 125.
- Nicholls D, Budd S. Mitochondria and neuronal survival. Physiol Rev. 2000;80:315. [PubMed: 10617771]
- 126.
- Nicholls DG, Scott ID. The regulation of brain mitochondrial calcium-ion transport. The role of ATP in the discrimination between kinetic and membrane-potential-dependent calcium-ion efflux mechanisms. Biochem J. 1980;186:833. [PMC free article: PMC1161720] [PubMed: 7396840]
- 127.
- David G. Mitochondrial clearance of cytosolic Ca2+ in stimulated lizard motor nerve terminals proceeds without progressive elevation of mitochondrial matrix [Ca(2+)] J Neurosci. 1999;19:7495. [PMC free article: PMC6782502] [PubMed: 10460256]
- 128.
- McCormack J, Denton R. Signal transduction by intramitochondrial Ca2+ in mammalian energy metabolism. News Physiol Sci. 1994;9:71.
- 129.
- Zoratti M, Szabo I. The mitochondrial permeability transition. Biochim Biophys Acta. 1995;1241:139. [PubMed: 7640294]
- 130.
- Andreyev A, Fiskum G. Calcium induced release of mitochondrial cytochrome c by different mechanisms selective for brain versus liver. Cell Death Differ. 1999;6:825. [PubMed: 10510464]
- 131.
- Gunter T, Pfeiffer D. Mechanisms by which mitochondria transport calcium. Am. J. Physiol. 1990;258:C755. [PubMed: 2185657]
- 132.
- Bezprozvanny I, Watras J, Ehrlich BE. Bell-shaped calcium-response curves of Ins(1,4,5)P3- and calcium-gated channels from endoplasmic reticulum of cerebellum. Nature. 1991;351:751. [PubMed: 1648178]
- 133.
- Jouaville L, et al. Synchronization of calcium waves by mitochondrial substrates in Xenopus laevis oocytes. Nature. 1995;377:438. [PubMed: 7566122]
- 134.
- Parekh AB. Store-operated Ca2+ entry: dynamic interplay between endoplasmic reticulum, mitochondria and plasma membrane. J Physiol. 2003;547:333. [PMC free article: PMC2342659] [PubMed: 12576497]
- 135.
- Putney JW Jr. Capacitative calcium entry in the nervous system. Cell Calcium. 2003;34:339. [PubMed: 12909080]
- 136.
- Peterson BZ, et al. Calmodulin is the Ca2+ sensor for Ca2+-dependent inactivation of L-type calcium channels. Neuron. 1999;22:549. [PubMed: 10197534]
- 137.
- Yang F, et al. Ca2+ influx-independent synaptic potentiation mediated by mitochondrial Na(+)-Ca2+ exchanger and protein kinase C. J Cell Biol. 2003;163:511. [PMC free article: PMC2173636] [PubMed: 14610054]
- 138.
- Toescu E, Verkhratsky A. Assessment of mitochondrial polarization status in living cells based on analysis of the spatial heterogeneity of rhodamine 123 fluorescence staining. Pfluegers Archiv. 2000;440:941. [PubMed: 11041562]
- 139.
- Hagen T, et al. Mitochondrial decay in hepatocytes from old rats: membrane potential declines, heterogeneity and oxidants increase. Proc Natl Acad Sci USA. 1997;94:3064. [PMC free article: PMC20322] [PubMed: 9096346]
- 140.
- Ward M, et al. Mitochondrial membrane potential and glutamate excitotoxicity in cultured cerebellar granule cells. J Neurosci. 2000;20:7208. [PMC free article: PMC6772767] [PubMed: 11007877]
- 141.
- Kamo N, et al. Membrane potential of mitochondria measured with an electrode sensitive to tetraphenyl phosphonium and relationship between proton electrochemical potential and phosphorylation potential in steady state. J Membr Biol. 1979;49:105. [PubMed: 490631]
- 142.
- Kokoszka JE, et al. Increased mitochondrial oxidative stress in the Sod2 (+/−) mouse results in the age-related decline of mitochondrial function culminating in increased apoptosis. Proc Natl Acad Sci USA. 2001;98:2278. [PMC free article: PMC30129] [PubMed: 11226230]
- 143.
- Wilson PD, Franks LM. The effect of age on mitochondrial ultrastructure and enzymes. Adv Exp Med Biol. 1975;53:171. [PubMed: 164102]
- 144.
- Harper ME, et al. Age-related increase in mitochondrial proton leak and decrease in ATP turnover reactions in mouse hepatocytes. Am. J. Physiol. 1998;275:E197. [PubMed: 9688619]
- 145.
- Xiong J, Verkhratsky A, Toescu EC. Changes in mitochondrial status associated with altered Ca2+ homeostasis in aged cerebellar granule neurons in brain slices. J Neurosci. 2002;22:10761. [PMC free article: PMC6758428] [PubMed: 12486169]
- 146.
- Brocard JB, Tassetto M, Reynolds IJ. Quantitative evaluation of mitochondrial calcium content in rat cortical neurons following a glutamate stimulus. J Physiol. 2001;531:793. [PMC free article: PMC2278496] [PubMed: 11251059]
- 147.
- Duchen M. Contributions of mitochondria to animal physiology: from homeostatic sensor to calcium signaling and cell death. J Physiol. 1999;516:1. [PMC free article: PMC2269224] [PubMed: 10066918]
- 148.
- Rego A, Ward M, Nicholls D. Mitochondria control AMPA/Kainate receptor-induced cytoplasmic calcium deregulation in rat cerebellar granule cells. J Neurosci. 2001;21:1893. [PMC free article: PMC6762594] [PubMed: 11245674]
- 149.
- Toescu EC. Normal brain aging: models and mechanisms. Philos Trans R Soc Lond. 2005;360:2347. [PMC free article: PMC1569594] [PubMed: 16321805]
- 150.
- Ankarcrona M, et al. Glutamate-induced neuronal death: a succession of necrosis or apoptosis depending on mitochondrial function. Neuron. 1995;15:961. [PubMed: 7576644]
- 151.
- Khodorov B. Glutamate-induced deregulation of calcium homeostasis and mitochondrial dysfunction in mammalian central neurons. Prog Biophys Mol Biol. 2004;86:279. [PubMed: 15288761]
- 152.
- White R, Reynolds I. Mitochondrial depolarization in glutamate-stimulated neurons — an early signal specific to excitotoxin exposure. J Neurosci. 1996;16:5688. [PMC free article: PMC6578963] [PubMed: 8795624]
- 153.
- Budd SL, Nicholls DG. Mitochondria, calcium regulation, and acute glutamate excitotoxicity in cultured cerebellar granule cells. J Neurochem. 1996;67:2282. [PubMed: 8931459]
- 154.
- Stout AK, et al. Glutamate-induced neuron death requires mitochondrial calcium uptake. Nat Neurosci. 1998;1:366. [PubMed: 10196525]
- 155.
- Crompton M. Mitochondria and aging: a role for the permeability transition? Aging Cell. 2004;3:3. [PubMed: 14965348]
- 156.
- Lafon-Cazal M, et al. Nitric oxide, superoxide and peroxynitrite: putative mediators of NMDA-induced cell death in cerebellar granule cells. Neuropharmacology. 1993;32:1259. [PubMed: 7509050]
- 157.
- Atlante A, et al. Glutamate neurotoxicity in rat cerebellar granule cells: a major role for xanthine oxidase in oxygen radical formation. J Neurochem. 1997;68:2038. [PubMed: 9109530]
- 158.
- Nicholls DG, et al. Interactions between mitochondrial bioenergetics and cytoplasmic calcium in cultured cerebellar granule cells. Cell Calcium. 2003;34:407. [PubMed: 12909085]
- 159.
- St-Pierre J, et al. Topology of superoxide production from different sites in the mitochondrial electron transport chain. J Biol Chem. 2002;277:44784. [PubMed: 12237311]
- 160.
- Skulachev VP. Role of uncoupled and non-coupled oxidations in maintenance of safely low levels of oxygen and its one-electron reductants. Q Rev Biophys. 1996;29:169. [PubMed: 8870073]
- 161.
- Brookes PS, et al. Calcium, ATP, and ROS: a mitochondrial love-hate triangle. Am. J. Physiol. Cell. Physiol. 2004;287:C817. [PubMed: 15355853]
- 162.
- Nicholls DG. Mitochondrial membrane potential and aging. Aging Cell. 2004;3:35. [PubMed: 14965354]
- 163.
- Terry RD, Katzman R. Life span and synapses: will there be a primary senile dementia? Neurobiol Aging. 2001;22:347. [PubMed: 11378236]
- 164.
- Vaillant GE, Mukamal K. Successful aging. Am J Psychiatry. 2001;158:839. [PubMed: 11384887]
- 165.
- Przedborski S, Vila M, Jackson-Lewis V. Neurodegeneration: what is it and where are we? J Clin Invest. 2003;111:3. [PMC free article: PMC151843] [PubMed: 12511579]
- 166.
- Anderton BH. Aging of the brain. Mech Aging Dev. 2002;123:811. [PubMed: 11869738]
- 167.
- Teter B, Finch CE. Caliban’s heritance and the genetics of neuronal aging. Trends Neurosci. 2004;27:627. [PubMed: 15374675]
- 168.
- Verkhratsky A, Toescu EC. Integrative Aspects of Calcium Signaling. Plenum Press; New York: 1998. p. 408.
- 169.
- Xia Z, Storm DR. The role of calmodulin as a signal integrator for synaptic plasticity. Nat Rev Neurosci. 2005;6:267. [PubMed: 15803158]
- 170.
- Huang PL. Nitric oxide and cerebral ischemic preconditioning. Cell Calcium. 2004;36:323. [PubMed: 15261488]
- 171.
- Chern Y. Regulation of adenylyl cyclase in the central nervous system. Cell Signal. 2000;12:195. [PubMed: 10781926]
- 172.
- Hudmon A, Schulman H. Neuronal Ca2+/calmodulin-dependent protein kinase II: the role of structure and autoregulation in cellular function. Annu Rev Biochem. 2002;71:473. [PubMed: 12045104]
- 173.
- Groth RD, Dunbar RL, Mermelstein PG. Calcineurin regulation of neuronal plasticity. Biochem Biophys Res Commun. 2003;311:1159. [PubMed: 14623302]
- 174.
- Sudhof TC. The synaptic vesicle cycle. Annu Rev Neurosci. 2004;27:509. [PubMed: 15217342]
- 175.
- Catterall WA. Structure and regulation of voltage-gated Ca2+ channels. Annu Rev Cell Dev Biol. 2000;16:521. [PubMed: 11031246]
- 176.
- Faber ES, Sah P. Calcium-activated potassium channels: multiple contributions to neuronal function. Neuroscientist. 2003;9:181. [PubMed: 15065814]
- 177.
- Berridge MJ, Lipp P, Bootman MD. The versatility and universality of calcium signaling. Nat Rev Mol Cell Biol. 2000;1:11. [PubMed: 11413485]
- 178.
- Farooqui AA, Horrocks LA. Brain phospholipases A2: a perspective on the history. Prostaglandins Leukot Essent Fatty Acids. 2004;71:161. [PubMed: 15253885]
- 179.
- Chan SL, Mattson MP. Caspase and calpain substrates: roles in synaptic plasticity and cell death. J Neurosci Res. 1999;58:167. [PubMed: 10491581]
- 180.
- Catterall W. Structure and function of neuronal Ca2+ channels and their role in neurotransmitter release. Cell Calcium. 1998;2:5. [PubMed: 10091001]
- 181.
- Berridge MJ, Bootman MD, Roderick HL. Calcium signaling: dynamics, homeostasis and remodelling. Nat Rev Mol Cell Biol. 2003;4:517. [PubMed: 12838335]
- 182.
- Davare MA, et al. A beta2 adrenergic receptor signaling complex assembled with the Ca2+ channel Cav1.2. Science. 2001;293:98. [PubMed: 11441182]
- 183.
- Gubellini P, et al. Metabotropic glutamate receptors and striatal synaptic plasticity: implications for neurological diseases. Prog Neurobiol. 2004;74:271. [PubMed: 15582223]
- 184.
- Verkhratsky A, Orkand RK, Kettenmann H. Glial calcium: homeostasis and signaling function. Physiol Rev. 1998;78:99. [PubMed: 9457170]
- 185.
- Gunter T, et al. Mitochondrial calcium transport: mechanisms and functions. Cell Calcium. 2000;28:285. [PubMed: 11115368]
- Review Calcium and neuronal injury in Alzheimer's disease. Contributions of beta-amyloid precursor protein mismetabolism, free radicals, and metabolic compromise.[Ann N Y Acad Sci. 1994]Review Calcium and neuronal injury in Alzheimer's disease. Contributions of beta-amyloid precursor protein mismetabolism, free radicals, and metabolic compromise.Mattson MP. Ann N Y Acad Sci. 1994 Dec 15; 747:50-76.
- Alterations in calcium homeostasis as biological marker for mild Alzheimer's disease?[Physiol Res. 2004]Alterations in calcium homeostasis as biological marker for mild Alzheimer's disease?Rípová D, Platilová V, Strunecká A, Jirák R, Höschl C. Physiol Res. 2004; 53(4):449-52.
- Review Expansion of the calcium hypothesis of brain aging and Alzheimer's disease: minding the store.[Aging Cell. 2007]Review Expansion of the calcium hypothesis of brain aging and Alzheimer's disease: minding the store.Thibault O, Gant JC, Landfield PW. Aging Cell. 2007 Jun; 6(3):307-17. Epub 2007 Apr 26.
- Review Network-wide dysregulation of calcium homeostasis in Alzheimer's disease.[Cell Tissue Res. 2014]Review Network-wide dysregulation of calcium homeostasis in Alzheimer's disease.Brawek B, Garaschuk O. Cell Tissue Res. 2014 Aug; 357(2):427-38. Epub 2014 Feb 20.
- Review (Re-)activation of neurons in aging and dementia: lessons from the hypothalamus.[Exp Gerontol. 2011]Review (Re-)activation of neurons in aging and dementia: lessons from the hypothalamus.Swaab DF, Bao AM. Exp Gerontol. 2011 Feb-Mar; 46(2-3):178-84. Epub 2010 Sep 15.
- Altered Calcium Homeostasis in Old Neurons - Brain AgingAltered Calcium Homeostasis in Old Neurons - Brain Aging
Your browsing activity is empty.
Activity recording is turned off.
See more...