Keywords: buffer, calcium
Abstract
Calcium signaling underlies much of physiology. Almost all the Ca2+ in the cytoplasm is bound to buffers, with typically only ∼1% being freely ionized at resting levels in most cells. Physiological Ca2+ buffers include small molecules and proteins, and experimentally Ca2+ indicators will also buffer calcium. The chemistry of interactions between Ca2+ and buffers determines the extent and speed of Ca2+ binding. The physiological effects of Ca2+ buffers are determined by the kinetics with which they bind Ca2+ and their mobility within the cell. The degree of buffering depends on factors such as the affinity for Ca2+, the Ca2+ concentration, and whether Ca2+ ions bind cooperatively. Buffering affects both the amplitude and time course of cytoplasmic Ca2+ signals as well as changes of Ca2+ concentration in organelles. It can also facilitate Ca2+ diffusion inside the cell. Ca2+ buffering affects synaptic transmission, muscle contraction, Ca2+ transport across epithelia, and the killing of bacteria. Saturation of buffers leads to synaptic facilitation and tetanic contraction in skeletal muscle and may play a role in inotropy in the heart. This review focuses on the link between buffer chemistry and function and how Ca2+ buffering affects normal physiology and the consequences of changes in disease. As well as summarizing what is known, we point out the many areas where further work is required.
CLINICAL HIGHLIGHTS.
Intracellular calcium controls the function of all organs in the body.
Most of the intracellular calcium is not free; rather, it is bound to calcium buffers. This affects the amplitude and time course of intracellular calcium transients.
Disordered intracellular calcium regulation is linked to many pathological conditions. Abnormal calcium buffering, as opposed to changes in transmembrane fluxes, has been suggested as a primary pathological mechanism.
For example, in common conditions such as atrial fibrillation and heart failure, altered intracellular calcium buffering is thought to contribute to proarrhythmic behavior.
In rarer genetic conditions such as hypertrophic cardiomyopathy, altered calcium buffering caused by mutations in contractile proteins is thought to promote abnormal contraction time course and increased incidence of arrhythmias. Initial studies suggest that drugs targeting the abnormal calcium buffering may improve contraction kinetics and reduce arrhythmias.
In neurons, calcium-binding proteins contribute to the regulation of electrical excitability and short-term synaptic plasticity.
Altered neuronal expression of calcium-binding proteins has been reported in the context of neurological diseases including dementia, epilepsy, and ataxia, but whether this is the primary cause or a consequence of abnormal neural function is less established.
This review introduces the readership of Physiological Reviews to the basic chemistry of calcium buffers and discusses how this relates to their function in both health and disease. Although modulation of intracellular calcium buffering may represent a promising novel therapeutic route, challenges such as the complex interdependence of calcium signaling and the need for tissue-specific interventions need to be addressed in future research.
1. INTRODUCTION
The importance of Ca2+ ions in regulating cell and tissue function is well established. Life begins with an increase of intracellular calcium concentration ([Ca2+]i) at the moment of fertilization (1). Cell death (apoptosis) is also accompanied by a Ca2+ signal (2). In between, the function of virtually every cell and tissue is controlled or influenced by changes of [Ca2+]i, and many important diseases involve disorders of Ca2+ signaling (for reviews see, e.g., Refs. 3–10). This Ca2+-dependent regulation is mediated by Ca2+ ions binding to proteins and thereby changing their structure and function. However, a consequence of this binding is that changes in cytoplasmic free Ca2+ are strongly constrained or “buffered,” with typically only ∼1% being free and ionized at resting [Ca2+]i. The chemistry of ionic buffering was first described with respect to the hydrogen ion (H+) in the early twentieth century and quantified initially by Koppel and Spiro (11) (see also Ref. 12) and subsequently by van Slyke (13). One of the first physiological indications of Ca2+ buffering, noted >70 years ago, was that intracellular Ca2+ is relatively immobile (14). Subsequent work on the squid axon led to the concept that most cytoplasmic Ca2+ is bound to Ca2+-binding molecules rather than being freely ionized (15), and similar conclusions have been reached for other tissues.
This buffering function of Ca2+-binding molecules means that not only the amplitude but also the time course of changes of [Ca2+]i may depend as much on the properties of the Ca2+ buffers as on the magnitude of the underlying Ca2+ fluxes. As described in this review, Ca2+ buffers are essential regulators for such important and diverse functions as muscle contraction, neuronal excitability and synaptic facilitation, epithelial transport, and killing of bacteria. The effects of buffers are not restricted to the cytoplasm but also extend to endoplasmic reticulum (ER), mitochondrial, and nuclear Ca2+.
Much is now known about the molecules that buffer Ca2+ and how their chemical properties determine this buffering. For many tissues, however, the quantitative contribution of the various potential buffers still remains to be established. Buffers differ in the affinity and speed with which they bind and release Ca2+ ions, and this binding can also be affected by other ions such as Mg2+ and protons. It is therefore important to understand how functional properties of Ca2+ buffers relate to their structure and chemistry. Some Ca2+ buffers are mobile, and binding to these can accelerate the rate at which Ca2+ diffuses within the cytoplasm. Another important and underappreciated issue concerns the factors that determine the strength or power of Ca2+ buffering. For example, how does buffering depend on the Ca2+ concentration and the affinity of the buffer for Ca2+, and how is it different for buffers that bind more than one Ca2+ in a cooperative manner?
As well as being physiologically important, experimentally added Ca2+ buffers have been used to control Ca2+ concentrations and probe cellular mechanisms of Ca2+ handling. Ca2+-sensitive fluorescent indicators used to measure [Ca2+]i are Ca2+ buffers and accordingly alter Ca2+ signaling.
Despite its importance, buffering is much less well understood than are many other aspects of Ca2+ signaling. In this review, we first consider the chemistry of Ca2+ buffering and the factors that determine the extent and speed of buffering. We emphasize the importance of the quantitative properties of buffering, deriving these from physicochemical principles and simplifying assumptions to describe general relationships. We show how the physiology of cells as diverse as epithelial cells, neurons, and muscle and immune cells is shaped by the chemistry of Ca2+ buffering and how a given buffer can be used to control the function of diverse tissues. Finally, there is much that is still not understood about buffering, and we highlight new directions for research.
2. CHEMISTRY OF Ca2+ BUFFERING
In aqueous solution, Ca2+ ions electrostatically form a solvation sphere with 6–10 water molecules (16). Chelation is the process of replacing two or more of these water molecules with dative covalent bonds between the Ca2+ ion and a ligand molecule and is the mechanism through which Ca2+ ions in solution bind to both small synthetic molecules and various Ca2+-binding proteins. In biological environments, the energetically optimum coordination of water around Ca2+ is 7, with 5 points arranged almost in a planar pentagon and 2 on an orthogonal plane creating a pentagonal bipyramid (17) (FIGURE 1). Most synthetic and biological chelators replace 6 or 7 of the water molecules with a chelation structure. Molecules interact directly with Ca2+ via electronegative points on the ligand molecule, typically negatively charged oxygen on a carboxylate group, the electronegative carbonyl group, or the lone pair of electrons of amino or histidyl groups (18) (FIGURE 2). Synthetic or natural Ca2+-binding molecules present an array of these electronegative residues to form a clawlike structure around each Ca2+ ion. The dative covalent bonds can form on each of the X, Y, and Z planes around the ion, thereby excluding most of the water molecules in the solvation sphere (17, 18).
FIGURE 1.
Pentagonal bipyramid arrangement of water around the Ca2+ ion in aqueous solution. In the chelation process, each water molecule is replaced by an organic residue that has an electronegative pole to support a dative covalent bond with Ca2+. The common groups that participate in chelation are shown with the electronegative poles designated.
FIGURE 2.
Examples of chelate structures. A: the small molecule EDTA adapted from Ref. 19, with permission from Fibres and Textiles in Eastern Europe. B: the EF-hand domain in terms of the helix-loop-helix structure (left) and the detail of the chelation site for the canonical form (right). Adapted from Ref. 17, with permission from Biochemical Journal. C: the proposed reaction scheme for the polymerization of calsequestrin (CSQ). Ca2+ ions are denoted as yellow balls in A–C; adapted from Ref. 20, with permission from Journal of Biological Chemistry.
2.1. Kinetics and Affinity
reaction 1 |
The simplest Ca2+ binding scheme is described in reaction 1. More complicated reactions, involving cooperative binding to multiple sites or competition with other ions are considered below. Here, Ca2+ binds to the buffer (B) with a rate constant kon (in M−1·s−1) and dissociates with rate constant koff (in s−1). As described in sect. 3, for a total buffer concentration, [B]Tot, the equilibrium concentration of Ca2+ bound to buffer, [CaB], will be given by
(1) |
The equilibrium dissociation constant is given by Kd = koff/kon. In chemistry, it is more customary to describe the association reaction of reaction 1 with an association constant (Ka = kon/koff). However, Ka has units of M−1, and its inverse, the Kd, is easier to appreciate as it is equivalent to the concentration of free Ca2+ at which 50% of the buffer has Ca2+ bound. Interestingly, almost regardless of the structure, the rate of Ca2+ binding to a buffer is generally rapid and close to the diffusion limit of 103 to 104 µM−1·s−1 (109 to 1010 M−1·s−1) in aqueous media (21). Such rates are technically difficult to measure and involve measurements on isolated proteins/molecules under chemical conditions that are far from physiological. Dissociation rate constants and equilibrium constants are easier to measure accurately, and therefore very fast association rate constants are often calculated as the ratios of koff over Kd. Values of Kds vary greatly across a range of chelation structures, with low-affinity buffers resulting from structures with high koff values.
2.2. Competition with Mg2+ and Protons
A mechanism that is frequently involved in lowering the affinity of a buffer is a decreased effective kon of the chelation/association reaction due to sequestering of the free buffer by other competing cations. As discussed below and illustrated in reactions 2A and 2B, if Mg2+ or protons are bound to a buffer, these ions must dissociate before Ca2+ can bind. This additional step slows down the overall rate constant for the Ca2+-binding reaction and therefore decreases the “apparent” affinity of the buffer under these conditions. For some buffers (for example ATP), even in the presence of Mg2+ or H+, Ca2+ binding is still rapid, whereas for others the slowing is profound (TABLE 1).
reaction 2A |
reaction 2B |
Table 1.
Examples of Kd and rate constants (both measured and apparent), for buffers which have a significant affinity for Mg2+ and or H+ in addition to Ca2+
Buffer | Kd, μM | kon, µM−1·s−1 | koff, s−1 | Kd,app, µM (pH 7.2, 1 mM Mg2+) | kon,app, µM−1·s−1 (pH 7.2, 1 mM Mg2+) |
---|---|---|---|---|---|
EDTA | 2.5 × 10-5 a | 2,500a | 0.7a | 15.7a | 0.06a |
EGTA | 1 × 10-5 b | 2,500c | 0.5d | 0.18c | 2.7d |
BAPTA | 0.17d | 450d | 79d | 0.176d | 440d |
ATP | 170l | 150k | 30,000k | 2,496l | 12l |
PV | 0.012e,g | 100e,g | 1e,f | 0.16g | 6.2 |
CB-D28k | 0. 312i,h | 110i,h | 34h | 0.434i | 78i |
Isolated skeletal TnC | 3.3j | 157j | 342j | 5.9j | 45j |
The listed measured equilibrium constants (Kd) and rate constants kon and koff values apply to Ca2+ binding to the buffer in the presence of a range of conditions including very low H+ and Mg2+ concentrations. The apparent equilibrium constants (Kd,app) were taken from measured values at 1 mM Mg2+, pH 7.2. The rate constants kon,app were estimated under those conditions based on measured koff values. Note that parvalbumin (PV), calbindin-D28k (CB-D28k), and Troponin-C (TnC) have multiple Ca2+ binding sites, and the kon values represent the highest measured at a single site unless the dissociation constants are identical. The parameters for a single ligand can come from different studies: the measured Kd does not equal the expected ratio of the rate constants. aRef. 22, 25°C, no added Mg2+; bRef. 23; cRef. 24, 35°C, no added Mg2+; dRef. 25, 22°C, no added Mg2+; eRef. 26, whiting (fish) protein, kon is highest of two binding sites, ∼20°C, no added Mg2+; fRef. 27, bovine chromaffin cells, Kd and koff at 20°C, 0.14 mM Mg2+; gRef. 28, bullfrog skeletal muscle, Kd values for Ca2+ and Mg2+ (∼90 μM) [note that another study found a 2-fold lower value for Kd(Mg) ∼40 μM (29)]; hRef. 30, recombinant protein, highest kon of 2 binding sites, 20°C, no added Mg2+; iRef. 31, recombinant human protein, highest Kd value of 4 binding sites, 20°C; jRef. 32, isolated TnC in the absence and presence of 3 mM Mg2+; kRef. 33, measured in solution at 20°C in 0.1 M KCl, from a compilation of stability constants; lRef. 34, rate constants based on measurements in free solution but corrected for intracellular conditions.
The equilibria describing the reactions between protons, Mg2+, and Ca2+ with a buffer can be expressed by modifying Eq. 1 such that
(2) |
where
(3) |
Kd,Ca, Kd,Mg, and Kd,H are the equilibrium dissociation constants, for Ca2+, Mg2+, and H+, respectively, such that Kd,Ca = koff,Ca/kon,Ca, Kd,Mg = koff,Mg/kon,Mg, and similarly for H+.
In other words, as [Mg2+]i or [H+]i increases, the apparent Kd for Ca2+ increases. Equations 2 and 3 quantify the extent to which H+- or Mg2+-bound forms of the buffer will affect the overall Ca2+ equilibration position. The apparent equilibrium constant (Kd,app) can be used to calculate the apparent association rate constant kon,app = koff/Kd,app using the measured dissociation rate constant (koff), which only involves Ca2+ and the buffer and is therefore unaffected by Mg2+ or H+. Some examples of measurements that demonstrate the reduction of the rate constant are given in TABLE 1 and discussed in sects. 2.3 and 2.5. It should be noted that the numbers presented come from many different studies, not necessarily performed at the same temperature, pH, and ionic strength and using material from different sources. Therefore, one cannot expect quantitative agreement in all cases. As will become apparent in subsequent sections, uncertainty about Ca2+ binding under physiological conditions makes it difficult to predict the properties of Ca2+ buffering from the known concentrations of buffers, and it will be important to address this in future work.
2.3. Small Synthetic Molecule Ca2+ Buffers
Small synthetic molecule buffers have many uses in experimental biology, forming the basis of many Ca2+-indicator dyes (35, 36), and are used for determining or altering the free Ca2+ concentration of solutions (35, 37, 38). In addition, “caged” Ca2+ chelators can change their affinity for Ca2+ upon photolytic exposure to light, thereby either releasing (39–41) or chelating (42) Ca2+ ions. Furthermore, understanding how the different chelation structures of the small synthetic molecules correlate with changes in affinity, selectivity, and kinetics helps explain the range of properties of the various Ca2+-binding molecules seen in nature. The kinetics and equilibria of Ca2+ binding depend on the structure of the ligand relative to the ionic radius of the cation. In the case of ethylenediaminetetraacetic acid (EDTA; FIGURE 2A), the ligand structure chelates Ca2+ but also accommodates Mg2+, an abundant intracellular divalent cation with a smaller ionic radius and a larger solvated radius. Although the affinity of EDTA for Mg2+ is 100-fold lower than for Ca2+ cytosolic levels of Mg2+ are >1,000-fold higher, and thus Mg2+ binding dominates over Ca2+ in an intracellular environment. Under these conditions, buffering or binding Ca2+ inevitably involves displacement of bound Mg2+, and therefore the complete reaction scheme incorporates Mg2+ dissociation/association (reaction 2A). As discussed in sect. 2.2, this feature, together with the effect of H+ binding, reduces the affinity (by ∼400,000 fold) and slows the overall kinetics of Ca2+ buffering. Furthermore, Ca2+ binding to EDTA releases Mg2+ from the buffer molecule and potentially increases the free Mg2+ level. In turn, buffering by EDTA will depend on Mg2+, which is buffered by free ATP2− and therefore depends on the metabolic state of the cell.
The related ligand ethylene glycol-bis(β-aminoethyl ether)-N,N,N′,N′-tetraacetic acid (EGTA) has a higher selectivity for Ca2+ over Mg2+ (10,000 fold) because of a longer backbone between the two amino groups that is more suited to chelate the larger ionic radius of Ca2+. Therefore, Mg2+ binding is minimal and interferes less with the Ca2+ buffering action. This means that kon of Ca2+ binding to the unprotonated form of EGTA approaches the diffusion limit of 103 to 104 µM−1·s−1 (24). However, at neutral pH, >98.5% of EGTA is in the diprotonated form, with protons bound to the two amino groups that participate in chelation, and this form of the molecule has a much-reduced ability to bind Ca2+, with a Kd ∼105.6 times greater than in the absence of protons and therefore not significantly contributing to chelation. Significant amounts of the unprotonated form are only available to bind Ca2+ metal ions after the dissociation of protons (reaction 2B) (25); thus the effective kon for Ca2+ binding is ∼100-fold less than the diffusion limit because of the additional steps of proton dissociation as part of the chelation reaction. This property not only slows buffering but also makes it steeply pH dependent, and for every Ca2+ ion that binds to EGTA approximately two H+ ions are released (23). In consequence, EGTA-containing solutions will acidify upon addition of Ca2+. Therefore, solutions should be pH adjusted after adding Ca2+, and EGTA-loaded cells should be strongly pH buffered to avoid pH changes upon Ca2+ influx. It should be noted that around pH 7.2 a titration error of only 0.1 pH units will cause a 50–80% error in the apparent Kd of EGTA (23, 43, 44).
In contrast to EDTA and EGTA, for which the rate of Ca2+ binding is limited by the need for Mg2+ or protons to dissociate, 1,2-bis(2-aminophenoxy) ethane-N,N,N′,N′-tetraacetic acid (BAPTA) has fast Ca2+ binding kinetics at neutral pH (35). This molecule combines a high selectivity for Ca2+ over Mg2+ arising from the optimum backbone length and a low affinity for protons on the two amine groups, allowing it to be relatively unaffected by the Mg2+ concentration and pH of the solution. Ca2+ binding can therefore occur without any significant exchanges with Mg2+ or H+, and this explains the fast kon for Ca2+ association (>102 µM−1·s−1) (25). It is worth emphasizing that the intrinsic rate of Ca2+ binding by BAPTA is actually lower than for EDTA and EGTA (TABLE 1). It is the lack of competition with H+ or Mg2+ that ensures an effective fast binding rate under physiological conditions.
2.4. Small Biological Molecule Ca2+ Binding
A range of small-molecule ligands is endogenous to the cellular environment. One example is that of humic and fulvic acids, comprising a mixture of many different molecules resulting from the breakdown of organic matter, which buffer a range of environmental divalent cations including Ca2+, Cd2+, and Pb2+, with the Kd of humic acid for Ca2+ being of the order of 1 mM at neutral pH (45, 46). In biological systems, individual amino acids such as l-arginine and l-lysine have significant single-site Ca2+ affinities, whereas di- and tripeptides, e.g., Gly-Glu, bind Ca2+ with chelation structure and Kds (0.1–1 µM) similar to synthetic chelators such as EGTA and BAPTA (47). These Ca2+-peptide complexes may have a biological function, for example facilitating Ca2+ transport across the intestinal epithelium (48).
2.4.1. Histidyl dipeptides, including carnosine.
Another function of Ca2+ binding to dipeptides is suggested by studies of the histidyl dipeptides (HDPs, e.g., carnosine) in striated muscle (49). Although possessing a relatively low Ca2+ affinity (Kd ∼1 mM) (50, 51), the very high intracellular levels (10–20 mM) mean that these dipeptides may bind significant amounts of intracellular Ca2+. However, there is uncertainty over the apparent Ca2+ affinity of carnosine under physiological conditions, in particular over the role of an amino group with a pK of ∼9 (50), which will be protonated at normal intracellular pH. If this is involved in chelation, Ca2+ binding will be negligible at normal pH (34). However, a recent study suggests that carnosine does bind Ca2+ with millimolar affinity at pH 7.2 (52). That carnosine can bind Ca2+ at physiological pH is also suggested by the observation that acidification from 7.3 to 6.7 results in the release of Ca2+ ions from carnosine in a mock intracellular solution (49). This pH sensitivity is, however, greater than what would be expected from measured binding constants (52). Given the high intracellular concentration of carnosine and its potential importance as a buffer, it is essential to address these issues.
2.4.2. ATP.
Another important biological small-molecule buffer is ATP; this is almost entirely bound to Mg2+ under physiological conditions. The molecule also binds Ca2+ with a low affinity relative to intracellular Ca2+ concentrations [Kd ∼ 0.2 mM (33)], but the high intracellular concentration of ATP (∼5 mM) makes it a significant Ca2+ buffer in many cell types (34). Most importantly, the fact that ATP is freely diffusible enables it to facilitate diffusion of Ca2+ (see sects. 4.5, 6.3, 7.6, and 9.2).
2.4.3. Inorganic phosphate.
is in equilibrium with and , with the relative proportions being sensitive to pH. All forms have the ability to bind Ca2+ (33). At normal intracellular pH (∼7.0) the apparent Kd for Ca2+ binding is ∼16 mM, so for [Ca2+]i ≤ 100 nM, the total Ca2+ bound to 1 mM phosphate will be ≤10 nM. In other words, phosphate makes a negligible contribution to cytoplasmic Ca2+ binding but may be more relevant to Ca2+ buffering in mitochondria (sect. 2.6.2.1) or organelles with higher [Ca2+] such as sarcoplasmic reticulum (SR), endoplasmic reticulum, and the nuclear envelope.
2.4.4. Other intracellular anions including gluconate.
Intracellular anions include intracellular proteins (see above) and the phosphate and carbonyl groups of lipids in the inner leaflet of the plasma membrane (53, 54), as well as a series of small-molecular weight anions, the relative concentration of which varies depending on cell type. A surface array of negative charges represented by the phospholipids of the inner surface of the plasma membrane is thought to influence the physical chemistry of Ca2+ in cells in two ways: 1) by contributing to intracellular buffering, which can be approximated by a contribution to global cytoplasmic binding (54) [such a role of negative surface charges has featured in models of intracellular Ca2+ binding (55)], but its contribution to bulk Ca2+ buffering is estimated to be small, and 2) electrostatic interactions between Ca2+ and the surface of fixed negative charges may create a layer of 2- to 3-nm depth with a higher Ca2+ concentration than bulk concentration (56). This may influence the apparent affinity of Ca2+ binding to proteins within the membrane. Again, the influence on bulk buffering will be small apart from regions of the cell where the relative surface area of membrane to cytoplasmic volume is very high, e.g., the diadic/triadic clefts in striated muscle. Under these circumstances, in the presence of significant transmembrane Ca2+ fluxes, Ca2+ binding to nearby surface membranes may be a significant factor in determining the extent and time course of free Ca2+.
The inorganic anions include Cl−, , , /, and and the organic anions include amino acids and dipeptides, all of which associate via various forms of electrostatic interactions with the cations of the intracellular medium including Ca2+ and Mg2+ ions (57). Replicating this complex range of anions in mock intracellular solutions can be difficult, and unphysiological levels of some anions can have effects on intracellular processes, e.g., higher than normal intracellular chloride concentrations inhibit G protein-related reactions (58). Gluconate has become a popular choice for the major anion of pipette solutions in whole cell patch-clamp studies on neurons. Although earlier work had not assigned any Ca2+-binding capacity to this compound (59), a subsequent study (60) demonstrated low-affinity Ca2+ binding with a Kd of 57 mM. As shown in sect. 9.4, this cannot be neglected given the relatively high concentrations of this anion in pipette-filling solutions.
2.5. Structural Ca2+-Binding Motifs in Proteins
The reader is referred to several review articles on the general subject of Ca2+-binding proteins (CBPs) (61–64). Most proteins can bind Ca2+ at two types of sites: the electronegative sites of salt bridges and the hydrogen bonds formed between carbonyl/carboxylate side chains and an amino/imidazole residue of nearby amino acids. These interactions form loose chelate structures that bind ionized Ca2+ and Mg2+ with a range of low affinities such that the relationship between Ca2+ binding and concentration is approximately linear even at the millimolar levels of ionized Ca2+ in the extracellular space, as in the case of serum albumin (65). However, such low-affinity interactions produce little Ca2+ binding at cytoplasmic Ca2+ concentrations. At rest, intracellular [Ca2+] is ∼2,000- to 20,000-fold lower than extracellular, and in the presence of millimolar levels of Mg2+, specific Ca2+ binding requires precise chelate structures for selectivity and affinity. Within eukaryotes, the Ca2+-binding motifs of intracellular CBPs conform to a limited set of designs, discussed below in this section; variants of each form with different Ca2+ binding characteristics allow their use in a range of intracellular processes.
2.5.1. Ca2+ buffering by the EF-hand domain.
The EF hand is a helix-loop-helix motif that generates a chelation structure in the loop segment (FIGURE 2B); many CBPs contain multiple paired EF-hand domains allowing the structure to interact with intracellular Ca2+ over a specific concentration range. Ca2+ binding to an EF-hand domain changes the tertiary structure. A commonly applied distinction is between CBPs acting solely as buffers with little conformational change upon Ca2+ binding and those acting as signaling molecules (“Ca2+ sensors”). Although this distinction is not always justified (66), for some CBPs such as parvalbumin (PV; Ref. 565) and calbindin-D9k (CB-D9k), which function primarily as buffers or to facilitate Ca2+ diffusion, Ca2+ binding-induced conformational changes can be small. In contrast, many EF-hand proteins undergo large changes of tertiary structure on binding Ca2+, thus altering the properties of the protein containing the EF hand. This mechanism is used to control the activity of proteins in many aspects of cellular function including cell mitosis, movement, sensory function, and molecular memory. Examples include Ca2+ binding to the Troponin-C (TnC) subunit causing major structural changes in striated muscle thin filament proteins, thereby allowing cross-bridge activity and contraction, and Ca2+ binding to calmodulin (CaM), which exposes specific hydrophobic structures allowing binding to corresponding sites on the regulatory domain of CaM kinase, displacing the autoinhibitory domain and activating this enzyme. For instance, Ca2+/calmodulin kinase II (CaMKII) and its activation by Ca2+ are essential for induction of long-term synaptic plasticity and memory formation (67). The S100 protein group is a large family (25 to date) of related proteins with 2 EF-hand motifs/protein. As reviewed previously (68), their role is varied, with some entirely acting inside the cell, some with both intracellular and extracellular actions, and others with a purely extracellular role. Inside the cell, S100 proteins generally endow Ca2+ sensitivity to cellular processes; for example, the abovementioned CB-D9k is also known as S100G (68). Interestingly, some of the S100 proteins bind Zn2+, Cu2+, and Mn2+ at the interface of S100 dimers, i.e., sites distinct from the EF motifs. The sequestering or buffering of these transition metal ions modulates biological pathways; for example, Mn2+ binding by extracellular S100 proteins is thought to be responsible for aspects of “nutritional immunity” (see Ref. 69 for review and sect. 11).
The canonical EF motif structure has a 12-residue loop containing the amino acids aspartate and glutamate that commonly provides 4 carbonyl and 2 carboxylate groups for coordination to Ca2+; the seventh coordination site is provided by a water molecule hydrogen-bonded to the carboxylate of the aspartate at residue 9 (17). This structure was first discovered in PV almost 50 years ago (70), and >4,400 EF motif structures have since been identified in >1,600 proteins (71). When incorporated into a protein, the EF-hand motif exists in closely associated pairs that are not identical in structure; binding in one motif can influence the binding of the partner, resulting in cooperative Ca2+ binding characteristics (see sect. 3.3). CaM (72) and CB-D9k (73) are well-documented examples of this phenomenon. The two EF-hand motifs are positioned such that the Ca2+-binding domains face each other and in such proximity that the bound Ca2+ ions are only 10 Å apart (17). EF hands can have high affinities for Ca2+ even under intracellular conditions. For example (TABLE 1), CB-D28k has a measured Kd in the absence of Mg2+ of ∼0.12 µM (30, 31), which increases to ∼0.24 µM at intracellular Mg2+ levels.
2.5.1.1. Mg2+ binding by the ef domain.
For the reasons outlined above, Mg2+ would be expected to bind to the EF-hand motif (74). Since the canonical EF motif structure has a Kd for Mg2+ in the millimolar range, ∼50% of binding sites have Mg2+ bound at resting levels of [Ca2+]i. There are two forms of site-selective interactions between Ca2+, Mg2+, and the EF hand, which render the respective sites as either 1) Ca2+-selective or 2) “Ca/Mg” sites.
1) Ca2+-selective sites: The coordination of Mg2+ involves only four residues on the NH2-terminal loop region; none of the residues in the COOH-terminal section of the loop normally participates. With this type of interaction, Mg2+ binding acts to stabilize the apo state (without Ca2+ bound), the inactive form of the EF motif. In contrast, Ca2+ binding recruits both NH2- and COOH-terminal regions of the loop and activates the motif. Therefore, as with EDTA, the presence of intracellular Mg2+ will slow Ca2+ binding and decrease its affinity at those sites (75, 76). There are other potential consequences of Mg2+ binding: if it is to one site of a pair of adjacent EF sites, Mg2+ binding can act cooperatively to enhance Ca2+ binding to the other site (75, 77). Therefore, the effects of Mg2+ binding to the COOH-terminal section of these Ca2+-selective sites are complex, and the consequences for the Ca2+ sensitivity of the proteins with canonical EF-hand sites have yet to be fully explored (76, 77).
2) Ca/Mg sites: In some noncanonical forms of the EF-hand motif, the sequences in the loop region form a smaller chelate structure that engages both NH2-terminal and COOH-terminal regions when either Ca2+ or Mg2+ is bound, and therefore both ions activate the motif; these sites are classified as Ca/Mg sites. In many instances, there is little unbound buffer, and at low intracellular Ca2+ levels (100 nM), the majority of the sites have either Ca2+ or Mg2+ bound. This chronically activated EF-hand domain is important for protein shape (i.e., structural sites). This is the case for the EF-hand domains of sites III and IV in striated muscle TnC, the so-called “nonspecific sites.” Both are Ca/Mg sites and, in the resting/inactive muscle, are almost fully occupied with Ca2+ (∼49%) or Mg2+ (49%) (78, 79). Sustained increases of cytoplasmic Ca2+ will slowly displace the bound Mg2+ on sites III and IV, accounting for the slow kinetics of Ca2+ buffering (for discussion see Refs. 80–82). One of the best-characterized Ca/Mg sites is that on PV, where the slow Ca2+ binding and Mg2+ dissociation have important consequences for skeletal muscle physiology and synaptic function and plasticity (see sects. 7.2, 9.4, and 9.5).
Both the Ca/Mg and Ca2+-selective sites bind Mg2+, but only in the former does this binding mimic that of Ca2+ binding. Thus, in both Ca/Mg and Ca2+-selective EF-hand domains, Mg2+ will slow the kinetics of Ca2+ binding and reduce affinity. Interestingly, Mg2+-bound states are often not included in computational models of Ca-EF hand interactions (83–86), although it has been observed that Mg2+ is bound to the Ca2+-selective sites of troponin-C and may slow Ca2+ binding and thence the development of force (87).
Approximately 80 different subfamilies of EF-hand domains have been recognized (88). Insertions, deletions, and substitutions in the loop region chelation site are believed to be the cause of the range of Ca2+ affinities that these EF-hand structure variants display, ranging from a Kd of ∼10−9 M seen in some CBPs such as PV to Kd values close to 10−3 M in the CREC, a group of ER-based proteins (89) (see sect. 2.6.1.1). A systematic study of the various EF-hand chelation structures and associated Ca2+ affinities across a range of EF-hand domain sequences is lacking, preventing the prediction of affinity from structure alone (89).
Almost one-third of known EF-hand structures do not bind Ca2+ (88), supposedly because of changes in the amino acid composition of the loop region including the substitution of the amino acids with oxygen-containing side chains. A well-documented example of a nonbinding EF-hand motif is that of site I in the striated muscle protein troponin-C. In the fast-twitch skeletal form both sites I and II bind Ca2+, whereas in the cardiac and slow-twitch forms only site II can do so. It has been suggested that site I cannot bind Ca2+ because of the disruption of the chelation site by an insertion (V28) and two key Ca2+-binding amino acid substitutions (D29L and D31A) (90).
2.5.1.2. h+ influence on ef-hand buffering of ca2+.
As mentioned above, some small-molecule Ca2+ chelators have a steep dependence on pH due to the direct involvement of amino groups (with pKa close to neutral pH) in the coordination of Ca2+. In contrast, the chelation structure of EF-hand domains uses only the carboxylate groups on glutamate and aspartate residues to coordinate Ca2+. The amino groups on other residues do not participate directly in chelation. These carboxylate groups have a pKa of ∼2 and would therefore be expected to be fully dissociated at normal intracellular pH. Hydrogen ions, however, can form a loose association by bridging carboxylate and carbonyl groups of adjacent residues. The shared H+ has a pKa of ∼6.2, close to the physiological pH range, and is thought to be one of the main causes of pH dependence of Ca2+ binding (91) and for the stoichiometry of ∼1 H+ released per Ca2+ bound to an EF-hand domain at neutral pH (92). This interaction between Ca2+ and protons also means that a decrease of pH will release Ca2+ from the EF hand, and this has been suggested to account for the acid-induced increase of [Ca2+]i seen in cardiac muscle (49, 93).
2.5.2. Buffering due to cytosolic Ca2+ binding to P-type ATPases.
The P-type transport ATPases constitute a family of proteins involved in the pumping of cations including Ca2+ across plasma membranes using a binding and translocation process. Although they are included in some models of Ca2+ dynamics in myocytes (e.g., Ref. 80), it is often overlooked that, in addition to their role in active transport, the chelation sites of these proteins contribute to the static intracellular Ca2+ buffering component of the cell (94). The two main forms of Ca2+ pumps in mammalian cells are the sarco(endo)plasmic reticulum Ca2+-ATPase (SERCA) and the plasma membrane Ca2+ ATPase (PMCA) (for reviews see Refs. 6, 95). These pumps are further divided into isoforms (SERCA1–3 and PMCA1-4), with alternative splicing generating ∼15–20 variants. The two types of Ca2+ pumps have a common general structure composed of E1 and E2 conformational states. The E1 state binds Ca2+ on the cytosolic side, and the E2 state releases Ca2+ on the intraluminal (SERCA) or extracellular (PMCA) side (96, 97). As part of the Ca2+ pump cycle, the Ca2+ binding to SERCA follows a sequential cooperative reaction scheme involving the two binding sites. It can be described by a Hill slope of 1.9 and a Kapp ∼0.4 µM (see sect. 3.3). The two individual sites have quite different Kd values (20 nM and 8.3 µM, respectively) (98). The cooperative Ca2+ binding has consequences for the Ca2+ buffer characteristics of SERCA, reducing its buffer contribution at lower [Ca2+]i (see sect. 3.3). The buffer characteristics of SERCA interact with its function as a pump (94, 99, 100), leading to paradoxical consequences of altering SERCA expression. For example, increased expression of SERCA increases the Ca2+ content of the ER/SR and therefore the amount released on stimulation, but the accompanying increased cytoplasmic buffer power may limit the rise in free Ca2+ levels and curtail the peak of the Ca2+ transient (99, 101).
In contrast to SERCA, PMCA pumps bind only one Ca2+ per cycle/ATP with a Kd of ∼1 µM (102). Because of this simpler reaction scheme, the PMCA has a different [Ca2+]-buffer power relationship (see sect. 3.3). As with SERCA, the pump turnover complicates the analysis of the overall buffer function of the PMCA. Its peripheral location and a relatively limited PMCA expression mean that the direct contribution to buffering may be small compared to the role in control of local subplasmalemma Ca2+ levels within cells (103).
2.5.3. Ca2+ binding by the C2 domain.
The conserved domain 2 (C2) is another Ca2+ binding motif that endows Ca2+ sensitivity to a series of membrane-targeted signaling molecules that include membrane receptors, kinases, G proteins, and various synaptic proteins. The numbers of different C2 and EF-hand designs are comparable in eukaryotes (104). Although less studied in terms of the physical chemistry of Ca2+ binding than the EF-hand domain, the C2 domains are found in a range of signaling pathways including kinases that phosphorylate membrane proteins, vesicle-targeting proteins, enzymes that modify signaling phospholipids, transmembrane pore proteins, and Ca2+ sensors for neurotransmitter and hormone release. A C2 domain is made up of ∼140 amino acids and typically forms a three-loop structure that links two β-sheets (105), chelating normally two and sometimes three Ca2+ using carboxylates and carbonyl residues in adjacent regions of two loops with relatively low affinity. Binding of the second Ca2+ within a C2 site usually shows strong cooperativity and generates a steep Ca2+ dependence and a switchlike response to increased intracellular Ca2+ (105). The Ca2+ levels required to activate C2-dependent signaling tend to be higher than those for common EF-hand domain motifs (106–108). The overall lower affinity for Ca2+ and Mg2+ implies that the low-affinity C2 domains do not have significant Ca2+ or Mg2+ bound under resting conditions. The high-affinity C2 domain on the common PKC isoforms has a Kd of 1–5 µM compared to ∼20 µM on the second C2 domain (107). In summary, the C2 domains represent a chelation design used in various signaling proteins that require [Ca2+]i in the micromolar or tens of micromolar range to activate. These concentrations of free Ca2+ are observed in the cytoplasm of skeletal muscle during a tetanus but otherwise are typically seen only within local Ca2+ domains, for instance near open Ca2+ channels at active zones of synapses (109) (see sects. 4.7.2 and 9.4). Therefore, C2 domains do not contribute substantially to overall Ca2+ buffering power but respond to local Ca2+ signals that are shaped by the actions of Ca2+ flux and other dominant buffers.
2.5.4. Ca2+ binding to annexins.
Another common Ca2+ binding site that is distinct from EF-hand and C2-protein domains is the endonexin fold seen in the Annexin group of proteins (110, 111). This large group is expressed across many species and has the general property of mediating Ca2+-sensitive phospholipid binding that allows increased cytoplasmic Ca2+ to mediate events such as membrane fusion. Twelve types of annexins are expressed in mammalian cells, and although their intracellular roles have not been fully explored, their low expression levels and generally low Ca2+ affinity means they do not participate in intracellular Ca2+ buffering or diffusion (110, 111).
2.6. Ca2+ Buffering in Organelles
2.6.1. ER and SR Ca2+ buffers.
A major role of the endoplasmic and sarcoplasmic reticulum is to release Ca2+ ions into the cytoplasm through inositol trisphosphate (IP3) and ryanodine receptors (RyRs), respectively. The free Ca2+ concentration in the ER and SR is typically between 100 µM and 1 mM (112–115), i.e., 1,000 to 10,000 times greater than that in the cytoplasm, thereby requiring different Ca2+ buffers, which are discussed below in this section.
2.6.1.1. ef-hand buffers.
Reticulocalbin (44 kDa) was identified as a CBP of the ER lumen containing six EF hands (116, 117). Subsequent work showed the existence of a class of such proteins, the CREC family comprising also Calumenin, ER Ca2+-binding protein of 55 kDa, and Calumenin 1 (see Refs. 89, 118 for reviews). Calumenin (37 kDa) (119) binds Ca2+ with a low affinity (Kd ∼600 µM) at seven EF-hand sites (120). Binding of Ca2+ results in major changes of structure, from disordered at low Ca2+ to alpha-helical at higher (121). These proteins may also regulate the activity of other proteins including RyR1 in skeletal muscle (122) and SERCA2 in cardiac muscle (123). Another lumen-resident protein is Stromal Interacting Molecule (STIM 1 and 2). This uses an EF hand-based Ca2+ sensor to detect the decrease of luminal Ca2+ associated with ER-mediated Ca2+ release. Ca2+ free STIM interacts with the plasma membrane Ca2+ release-activated calcium channel protein ORAI1 to increase Ca2+ influx into the cell, a process known as store-operated calcium entry (SOCE) (see Ref. 124 for review). Because the concentration of these proteins is low in comparison to those discussed in sect, 2.6.1.2, they will not play a major role in buffering Ca2+ in the ER lumen (117).
2.6.1.2. calreticulin, calnexin, and calmegin.
The bulk of endoplasmic reticulum Ca2+ buffering results from proteins with structures very different from those discussed in the previous section. Calreticulin [CRT, molecular weight (MW) 46,000] is predominantly an ER luminal CBP and molecular chaperone that promotes protein folding but also has many other functions (125). There are three distinct regions in the protein: the N globular domain, the P-arm domain, and the C domain. None of these sites conforms to EF-domain or C2 protein configurations. The N domain binds Zn2+ and participates in chaperone interactions. The proline-rich P arm is key to the protein folding function and contains a single Ca2+ binding domain with a Kd of ∼1–10 µM (126). At the high Ca2+ concentration of the ER lumen, this site is normally fully Ca2+ bound and crucial for the tertiary structure of the protein. Finally, the COOH terminal of the protein ends in a region containing highly acidic residues (35 glutamate and aspartate residues out of 50), and this region can bind up to 25 Ca2+ ions per molecule with a Kd of ∼1 mM (for reviews see Refs. 125, 127). The degree to which the binding reactions involve cooperative interactions is not clear. The chelation mechanism uses two acidic groups on Asp and Glu residues to bind to Ca2+ with a kon close to the diffusion limit (∼103 µM−1·s−1) and a rapid koff (∼106 s−1). CRT accounts for ∼50% of the Ca2+ buffering within the ER. Ca2+ binding to the COOH-terminal sites not only buffers Ca2+ but also alters the structure of this region, which in turn alters the affinity of CRT for interacting proteins (chaperones) within the ER (125, 128). Through this mechanism, the intra-ER Ca2+ concentration modulates the types of proteins processed (129). Measurements of free Ca2+ concentration within the lumen of the ER with targeted sensors suggest a maximal concentration of 300–400 µM (113, 114). The minimal Ca2+ concentration when the ER is depleted has not been accurately measured, as few studies have used luminal indicators with appropriate sensitivities, but concentrations lower than 5–10 µM may destabilize the tertiary structure of CRT by dissociation of Ca2+ from the high-affinity structural site. The ER also contains two structurally related proteins, calnexin and calmegin. These differ from calreticulin in that they are membrane bound (see Ref. 130 for review). Interestingly, in calnexin the Ca2+ binding C domain is exposed to the cytoplasm (131) (for review see Ref. 125), raising the possibility that a single molecule can buffer Ca2+ on both sides of the ER membrane.
2.6.1.3. calsequestrin.
The related protein calsequestrin (CSQ; MW 44,000) is the main Ca2+ buffer in skeletal, cardiac, and smooth muscle sarcoplasmic reticulum and comprises two isoforms (CSQ1 and CSQ2). Aspects specific to skeletal muscle function are considered in sect. 7.5. Analogous to CRT, CSQ has a limited number (2 or 3) of high-affinity Ca2+ binding sites; occupancy of these is key to the tertiary structure of the CSQ monomer. However, most Ca2+ binding occurs at multiple low-affinity binding sites arising from the numerous aspartate and glutamate residues particularly on the COOH-terminal but also the NH2-terminal end of the monomer. These acidic groups can bind >40 Ca2+ ions per molecule, with Kd values ranging from 0.1 to 10 mM (132). The isoform CSQ2, present in slow-twitch skeletal, cardiac, and smooth muscle, has a modified COOH terminal and binds up to 20 Ca2+ ions per molecule (133). As with CRT, the coordination of Ca2+ by these low-affinity sites is via pairs of acidic resides binding a single Ca2+. An aspect unique to CSQ is the process through which Ca2+ binding to the monomer allows CSQ to polymerize and generate a structure with further Ca2+ binding sites (FIGURE 2C). This is a cooperative process, as described in sect. 3.3. In both skeletal (134) and cardiac (135, 136) muscle, in addition to buffering SR Ca2+, CSQ2 can interact with the RyR and thereby provide a mechanism for luminal Ca2+ controlling RyR opening. It can therefore be challenging to separate the consequences of CSQ2-RyR interactions from Ca2+ buffering effects.
Mg2+ can also bind to CSQ. However, evidence suggests that, unlike many chelation interactions, the Mg2+ binds at sites distinct from Ca2+ (137). The two sites interact such that Ca2+ binding to CSQ displaces a fraction of bound Mg2+ (137). In common with other CBPs that involve carbonyl-carboxylate interactions, a proton can bridge between these two residues in the absence of Ca2+ (pKa ∼6.2). Therefore Ca2+ binding to CSQ will displace significant amounts of protons (137, 138) and may change intraluminal pH, influencing other processes including SERCA and the Ca2+ release channel.
2.6.2. Buffering in other organelles.
The reader is referred to an earlier review for a discussion of this area (139). Here we provide brief comments on mitochondria and the nucleus.
2.6.2.1. mitochondria.
Regulation of mitochondrial matrix [Ca2+] ([Ca2+]mito) is important since it regulates various mitochondrial enzymes, and thus ATP production (140, 141). The introduction of Ca2+-sensitive indicators targeted to the mitochondrial matrix has suggested that [Ca2+]mito is normally similar to cytoplasmic [Ca2+]i, ∼100 nM (142, 143). As previously reviewed (144), there is controversy as to the extent to which mitochondrial matrix Ca2+ responds to brief changes of cytoplasmic Ca2+. For example, in cardiac muscle, some studies have reported that the cytoplasmic Ca2+ transients result in beat-to-beat changes of [Ca2+]mito (145–147), whereas other work has reported little or no change of [Ca2+]mito in response to fluctuations of cytoplasmic [Ca2+]i (148). Even when [Ca2+]mito transients are observed, they can decay very slowly (142), such that stimulation at normal rates results in a maintained increase of [Ca2+]mito with little or no beat-to-beat fluctuations (see Ref. 9 for review). There are two explanations for this slow decay of [Ca2+]mito: 1) a low activity of the mitochondrial sodium/calcium exchange (NCLX) that pumps Ca2+ out of the mitochondria (149) and 2) a high mitochondrial matrix buffer power. This emphasizes the need to quantify mitochondrial Ca2+ buffer power.
Most measurements have been made in isolated mitochondria, often at [Ca2+]mito greater than those thought to occur physiologically (see below); and they may be susceptible to changes resulting from mitochondrial isolation. One such study found that total Ca2+ concentration was ∼1,000 times greater than free in heart mitochondria (150). High Ca2+ buffer powers were also found in heart and liver mitochondria (151). Another study reported a value of up to ∼150,000, indicating a total mitochondrial Ca2+ concentration of the order of 100 mM (152). These high levels of total Ca2+ are thought to exist mainly as calcium phosphate crystals (153), in which case the relevance of such buffering will depend critically on the kinetics of crystal formation and dissolution. A more recent study on isolated mitochondria used changes in the free Ca2+ of the external fluid to calculate Ca2+ fluxes into and out of mitochondria and compared this with changes of [Ca2+]mito (154). Modeling of the data (155) suggested two classes of buffers. One was tentatively attributed to phospholipids and metabolites. The other was estimated to bind Ca2+ in a cooperative manner and was suggested to involve annexins acting as “nucleation factors” to promote binding of Ca2+ with phosphate. Another study on isolated guinea pig heart mitochondria found that phosphate is a significant buffer at elevated mitochondrial Ca2+ but that some, unidentified, mitochondrial buffer contributes at lower [Ca2+]mito levels (156).
Work on intact cells found that the change of [Ca2+]mito during a cardiac cytoplasmic Ca2+ transient was 2–10 nM (157). The authors pointed out that if the mitochondrial Ca2+ buffering power was 100 (similar to that of cytoplasm) the magnitude of the measured changes of [Ca2+]mito is consistent with the calculated contribution of mitochondria to Ca2+ removal from the cytoplasm (158). More precise measurements are required, but it is difficult to imagine how total mitochondrial Ca2+ can be measured in intact cells. Another obvious question relates to whether the mitochondrial Ca2+ buffering power is identical in cells from all tissues.
Given the challenges involved in direct measurements of mitochondrial Ca2+ buffering, estimating it from the known composition of the matrix is a useful alternative. One suggested contributor is inorganic phosphate (for review see Ref. 159). Considering the binding constants of the various forms of phosphate (33), and assuming a total matrix phosphate concentration of 1 mM, then at 100 nM [Ca2+]mito the contribution to buffer power will be only 1.5. There is also ∼5 mM ATP present, which will provide a buffer power of 7.2. This is greater than the expected buffer power of 3.6 for the cytoplasm of cardiac muscle (see TABLE 5) because of the more alkaline pH (8.0) in mitochondria. Thus, the anticipated Ca2+ buffer power of the mitochondria is considerably lower than that of cytosol, and it is not obvious how the known buffers can account for some of the above measurements. Further work is required to determine the identity of mitochondrial Ca2+ buffers.
Table 5.
Contributions of various compounds to calcium buffering in cardiac muscle
Buffer | Conc, µM | Kapp, µM | Hill Slope | Binding Sites | βi(Ca) 0.1 μM (ss) | βi(Ca) 0.2 μM (ss) | βi(Ca) 1.0 μM (ss) | βi(Ca) 10 μM (ss) | βi(ΔCa) 0.1–1.0 μM (in) | βi(ΔCa) 0.1–1.0 μM (ss) | βi(ΔCa) 0.1–10 μM (in) | βi(ΔCa) 0.1–10 μM (ss) | Ref. |
---|---|---|---|---|---|---|---|---|---|---|---|---|---|
Calmodulin | 6 | 2.7/12.7 | 1.8/1.9 | 4 | 0.6 | 1.0 | 2.8 | 0.7 | 2.0 | 2.0 | 1.6 | 1.6 | 187 |
SERCA | 48 | 0.60 | 1.9 | 2 | 56.8 | 89.5 | 36.3 | 0.1 | 73.9 | 73.9 | 9.3 | 9.3 | 98 |
TnC (reg) | 70 | 0.60 | 1 | 1 | 85.7 | 65.6 | 16.4 | 0.4 | 37.5 | 37.5 | 5.7 | 5.7 | 78 |
Carnosine | 10,000 | 1,350.00 | 1 | 1 | 7.4 | 7.4 | 7.4 | 7.3 | 7.4 | 7.4 | 7.3 | 7.4 | 49 |
ATP | 5,000 | 2,496 | 1 | 1 | 2.0 | 2.0 | 2.0 | 2.0 | 2.0 | 2.0 | 1.9 | 2.0 | 33 |
Myosin | 70 | 9.10 | 1 | 2 | 15.1 | 14.7 | 12.5 | 3.5 | 0.5 | 13.7 | 0.1 | 7.3 | 78 |
TnC (nonspec) | 70 | 0.10 | 1 | 2 | 349.9 | 157.2 | 11.9 | 0.1 | 2.3 | 64.5 | 0.2 | 7.0 | 78 |
Total | 517 | 338 | 89 | 14 | 126 | 201 | 26 | 40 |
From left to right: cytoplasmic concentration; Kapp; Hill slope (n); no. of binding sites; steady-state (ss) buffer power (βi) at 0.1, 0.2, 1.0, and 10 µM Ca2+, calculated with Eq. 8 or 16; and the buffer power for a change in calcium concentration ([Ca2+]) [β(ΔCa)] for 0.1–1 µM and 0.1–10 μM Ca2+, calculated from the change in Ca2+ binding (Eq. 5 or 15). To indicate the relative contributions of the fast and slow buffers, we show both instantaneous buffer power (in, assuming equilibration of Ca2+ binding but no change in the Mg2+ binding state) and steady-state buffer power (ss, assuming equilibrium at both Ca2+ and Mg2+ sites). The value of the Hill slope for the regulatory Troponin-C (TnC) site is taken from isolated TnC (78) and ignores the influence of possible cross-bridge interactions (sect. 3.3).SERCA, sarco(endo)plasmic reticulum Ca2+-ATPase.
2.6.2.2. nucleus.
One might expect the Ca2+ buffering of the nucleus to be different from that of cytoplasm. For example, troponin, which is a major cytoplasmic buffer in striated muscle, is absent from the nucleus and cannot contribute to buffering there. Some work has produced results consistent with lower Ca2+ buffering in the nucleus than in cytoplasm. For example, Ca2+ puffs, which are mediated by Ca2+ release from the ER through IP3 receptors and occur near the nuclear membrane, spread further in the nucleus than in the cytoplasm (160). Similar results have been seen for Ca2+ released from Golgi (161). The apparent diffusion coefficient for Ca2+ is similar in nucleus and cytoplasm (162, 163). It should, however, be noted that there is no simple relationship between Ca2+ diffusion and buffering: Mobile buffers can increase while fixed buffers decrease diffusion. Electron microprobe measurements have reported a higher concentration of total Ca2+ in the nucleus compared with the cytoplasm (164). Assuming that the free Ca2+ concentration is the same in both regions, this would imply greater Ca2+ buffering in the nucleus. The highest total Ca2+ concentrations were found in areas occupied by heterochromatin (condensed DNA). This is interesting, given that high (∼1 mM) Ca2+ has been reported to be bound to DNA and that this binding alters the structure of DNA (165). Separate studies on isolated chromosomes suggest that Ca2+ binding alters chromosome structure during mitosis (166). It should, however, be noted that the nuclear envelope, which contains high total Ca2+, invaginates the nucleus as the so-called “nucleoplasmic reticulum” (167, 168). It is possible that Ca2+ in this structure contributes to the high total Ca2+ measurements from the nucleus, whereas in the nuclear lumen total Ca2+ content and buffering power may in fact be lower.
The work described above does not provide quantitative measurements of nuclear Ca2+ buffering. Another approach is to look for differences in the concentrations of candidate Ca2+ buffers between nucleus and cytoplasm. For example, a study on smooth muscle cell lines derived from human aorta and jejunum characterized the distribution of various S100A proteins (see sect. 2.5.1). S100A2 and A6 were found mainly in the nucleus, whereas A1 and A4 were largely in the cytosol (169). The effects of increasing nuclear Ca2+ buffering have been studied in cultured hippocampal neurons by targeting parvalbumin to the nucleus. This decreased the rise of nuclear Ca2+ produced by electrical stimulation but had no effect on the cytoplasmic [Ca2+]i. A functional consequence of increased nuclear Ca2+ buffering was a change of cell morphology, in particular decreased dendrite length and complexity (170, 171). Work on human eggs and preimplantation embryos found that various CBPs including calreticulin and calsequestrin were distributed in the cytoplasm but not evident in the nucleus (172).
3. QUANTIFICATION OF BUFFER POWER
The degree of Ca2+ buffering can be defined as “buffer power.” This is illustrated for the case of a simple buffer in FIGURE 3. For a total buffer concentration [B]Tot, bound Ca2+ ([CaB]) and free Ca2+ ([Ca2+]) are described by Eq. 1 above.
FIGURE 3.
Dependence of buffer power on Ca2+. A, a: bound Ca2+ as a function of intracellular calcium concentration ([Ca2+]i). The black curve represents binding to a simple buffer according to Eq. 1, with an equilibrium dissociation constant (Kd) of 0.5 µM (vertical dotted line). Total buffer concentration ([B]Tot) is 100 µM. The red line is a fit to the initial part of the curve with a slope of [B]Tot/Kd. b: Buffer power as a function of [Ca2+]i. B: similar to A but for a buffer with a Kd of 2 µM (dotted line). The dashed curve in Bb shows the buffer power-[Ca2+]i relationship of the buffer in A superimposed for comparison. The gray shaded regions in all panels represent the [Ca2+]i range from 100 nM (resting [Ca2+]i) to 300 nM. For a [Ca2+]i transient covering this range, buffer power is predicted to decrease by 44% (arrowheads, from 139 to 78) for the Kd of 0.5 µM (A) and by 17% (arrowheads, from 45 to 38) for the Kd of 2.0 µM (B).
The total Ca2+ concentration ([Ca2+]Tot) is given by the sum of free and bound:
(4) |
(5) |
Buffer power has been defined by relating changes of either total or bound Ca2+ to those of free. In the former case it is quantified [by analogy with the original definition of pH buffering (11, 13)] as the ratio of the change in total to that of free concentration (d[Ca2+]total/d[Ca2+]free), which is often (including this article) represented by the symbol β and is commonly used in experimental muscle physiology.
By differentiating Eq. 5 with respect to [Ca2+] at constant Kd one obtains
(6) |
A slightly different definition of buffer power, often referred to in the neuroscience literature as differential Ca2+-binding ratio (173) and customarily designated as κ, describes the buffering by individual buffers or groups of buffers. It is the ratio of the change in bound to that of free (d[Ca2+]bound/d[Ca2+]free).
(7) |
That there are two definitions is largely a consequence of differences in experimental approaches. In many studies on muscle, changes of [Ca2+]i are compared with total changes of Ca2+ as measured from Ca2+-specific ionic currents. This corresponds to the buffer power, as defined by β (Eq. 6). In contrast, in the neuroscience literature buffer power is often measured by the “added buffer approach” (see below), which does not invoke total changes in Ca2+ but rather a comparison of effects of an endogenous buffer with those of a known buffer, usually that of a Ca2+-indicator dye. It should be noted (cf. Eqs. 6 and 7) that β is equal to 1 + κ. Therefore, for values of β or κ commonly measured in physiology (>20) there is very little difference between the two values.
As expressed in Eq. 6, β contains contributions from both free Ca2+ and Ca2+ bound to buffer (respectively, the first and second terms on the right-hand side). In the remainder of this review, it will be convenient to quantify the contribution to buffering of particular buffers. We therefore define βi as the contribution to buffering provided by the ith buffer.
(8) |
The contributions of the n buffers can then be summed and added to 1 (representing free Ca2+) to give the overall value of β:
(9) |
Likewise, κ can be annotated with a subscript i when individual buffers are considered, such as in Eq. 29 below. βi is identical to κi, and in this article, which aims to bring together literature across physiology, the terms are used interchangeably depending on the context.
3.1. Effects of Ca2+ Concentration on Buffer Power
Equation 8 shows that the buffer power contribution is greatest at very low [Ca2+]i, where it can be approximated by ([B]Tot/Kd), illustrated by the red lines that are tangents to the binding curves of FIGURE 3. As shown in FIGURE 3, bottom, as [Ca2+]i increases more buffer has Ca2+ bound and its contribution to buffer power declines eventually to a value very close to zero at high [Ca2+], when effectively all binding sites are occupied. Expressed as a multiple of [B]Tot/Kd, buffer power decreases from 1 at 0 [Ca2+] to 0.44 when [Ca2+] = 0.5 × Kd and 0.25 when [Ca2+] = Kd. When [Ca2+] rises to 2 × Kd, buffer power falls to 0.11.
If [Ca2+] changes over a finite but small range (≪Kd), βi can be taken as constant and calculated at the average level of [Ca2+]i. Therefore, relating changes in total Ca2+ bound, ΔCaB, to those of free Ca2+, Δ[Ca2+]:
(10) |
For larger values of Δ[Ca2+], βi can be replaced by (174)
(11) |
Here, [Ca2+]1 and [Ca2+]2 are [Ca2+] values before and after the addition of Ca2+ ions to a buffer system as in TABLE 5 and TABLE 6. For more complex reaction mechanisms, βi(ΔCa) can be calculated from the binding curves as shown in FIGURE 3 and FIGURE 5.
Table 6.
Contributions of various compounds to calcium buffering in skeletal muscle, fast-twitch fibers, and slow-twitch fibers in the mouse
Buffer | Conc, µM | Kapp, µM | Hill Slope | Binding Sites | βi(Ca) 50 nM (ss) | βi(Ca) 2.0 μM (ss) | βi(Ca) 20 μM (ss) | βi(ΔCa) 50 nM-2.0 μM (in) | βi(ΔCa) 50 nM-2.0 μM (ss) | βi(ΔCa) 50 nM-20 μM (in) | βi(ΔCa) 50 nM-20 μM (ss) | Ref. |
---|---|---|---|---|---|---|---|---|---|---|---|---|
Fast-twitch fibers | ||||||||||||
SERCA1a | 120 | 0.30 | 1.9 | 2 | 283.8 | 5.9 | 0.0 | 115.9 | 115.9 | 11.6 | 11.6 | 98 |
TnC (reg) | 120 | 1.30 | 1 | 2 | 171 | 28.7 | 0.7 | 70.0 | 70.0 | 10.8 | 10.9 | 84 |
Carnosine | 10,000 | 1,350 | 1 | 1 | 7.4 | 7.4 | 7.2 | 7.4 | 7.4 | 7.2 | 7.3 | 49 |
Parvalbumin | 75 | 0.14 | 1 | 2 | 574 | 4.7 | 0.1 | 4.6 | 51.9 | 0.5 | 5.5 | 83 |
ATP | 8,000 | 2,496 | 1 | 1 | 3.2 | 3.2 | 3.2 | 3.2 | 3.2 | 2.9 | 3.2 | 33 |
Myosin | 120 | 9.10 | 1 | 2 | 26.1 | 17.7 | 2.6 | 0.4 | 21.5 | 0.0 | 8.2 | 78 |
TnC (nonspec) | 120 | 0.10 | 1 | 2 | 1,060 | 5.5 | 0.1 | 1.6 | 76.6 | 0.2 | 8.0 | 78 |
Total | 2,125 | 73 | 14 | 203 | 347 | 33 | 55 | |||||
Slow-twitch fibers | ||||||||||||
SERCA2a | 48 | 0.60 | 1.9 | 2 | 31.9 | 7.6 | 0.0 | 44.3 | 44.3 | 4.8 | 4.8 | 98 |
TnC (reg) | 120 | 1.30 | 1 | 1 | 86 | 14.3 | 0.3 | 35.0 | 35.0 | 5.4 | 5.4 | 84 |
Carnosine | 10,000 | 1,350 | 1 | 1 | 7.4 | 7.4 | 7.2 | 7.4 | 7.4 | 7.2 | 7.3 | 49 |
Parvalbumin | 7.5 | 0.14 | 1 | 2 | 57 | 0.5 | 0.0 | 0.5 | 5.2 | 0.0 | 0.6 | 83 |
ATP | 5,000 | 2,496 | 1 | 1 | 2.0 | 2.0 | 2.0 | 2.0 | 2.0 | 1.8 | 2.0 | 33 |
Myosin | 120 | 9.10 | 1 | 2 | 26.1 | 17.7 | 2.6 | 0.4 | 21.5 | 0.0 | 8.2 | 78 |
TnC (nonspec) | 120 | 0.10 | 1 | 2 | 1,060 | 5.5 | 0.1 | 1.6 | 76.6 | 0.2 | 8.0 | 78 |
Total | 1,270 | 55 | 12 | 91 | 192 | 19 | 36 |
From left to right: cytoplasmic concentration; Kapp; Hill slope (n); no. of binding sites; steady-state (ss) buffer power [βi(Ca)] at 0.05, 2, and 20 µM Ca2+ (calculated with Eq. 8 or 16); and the buffer power for a change in [Ca2+] [βi(ΔCa)] for 0.05–2 µM and 0.05–20 μM Ca2+, calculated from the change in Ca2+ binding (Eq. 5 or 15). steady-state binding (ss) was calculated based on equilibrium at both Ca2+ and Mg2+ sites. Instantaneous buffer power (in) was calculated assuming equilibration of Ca2+ binding but no change in the Mg2+ binding state in order to indicate the relative extent of the fast and slow buffers. SERCA, sarco(endo)plasmic reticulum Ca2+-ATPase; TnC, Troponin-C.
FIGURE 5.
Dependence of buffer power on intracellular calcium concentration [Ca2+]i for buffers with cooperative Ca2+ binding. A and B: bound Ca2+ (top) and buffer power (bottom), both as a function of [Ca2+]. All buffers are present at a concentration giving a total of 100 µM binding sites and have a [Ca2+] at which half the binding sites are occupied (Kapp) of 0.5 µM. A: cooperative binding described by the Hill equation with values of Hill slope (n) of 1.0, 1.25, 1.5, and 2.0. B: comparison of higher values of n (4 and 8) with 1 and 2.
Does the dependence of buffer power on Ca2+ concentration have any physiological significance? One can speculate that a high buffer power at low [Ca2+] will stabilize the low resting levels of cytosolic Ca2+ whereas at elevated [Ca2+]i a smaller increase in total Ca2+ is required to produce a given rise of [Ca2+]i, and this may be energetically advantageous. Evidently, an increase of resting [Ca2+]i will decrease buffer power (but see sect. 3.3.1 for discussion of cooperative buffers). This has two consequences. 1) As discussed in the context of skeletal muscle and nerve physiology (sects. 7.4 and 9.5), an initial increase of [Ca2+]i will decrease buffer power such that a subsequent identical increase of total Ca2+ will increase [Ca2+]i more, thereby contributing to phenomena such as synaptic facilitation (175–178) and muscle contraction (82, 84, 179). 2) As discussed in sect. 7.1, some of the variation of buffer power reported for a given tissue may reflect the range of [Ca2+]i over which it was measured. In experimental studies, it is important to investigate whether changes of buffer power arise from alteration of the properties of the buffer or, alternatively, are a consequence of changes of [Ca2+]i. For example, a measured decrease of buffer power could be a consequence of either a change in the concentration or properties of buffers or, alternatively, an increase of resting [Ca2+]i.
3.2. Effects of Buffer Kd
The total buffering produced by a given concentration of any buffer is fixed, and changing Kd simply shifts the range of [Ca2+] over which buffering occurs. This is exemplified in FIGURE 3, bottom, for two buffer species with fourfold different Kd values. Whereas the maximum β is four times larger for the high-affinity buffer (FIGURE 3A), β decreases more rapidly with increasing [Ca2+] compared to the low-affinity buffer (FIGURE 3B). FIGURE 4A shows buffer power (β) as a function of Kd at various values of [Ca2+]. At each [Ca2+], increasing Kd first increases and then decreases buffer power for that buffer (82, 180). As Kd decreases below an optimal value, the buffer becomes increasingly saturated with Ca2+ and its ability to buffer is decreased. As Kd increases above the optimum, less Ca2+ binds for a given increase of [Ca2+] and, again, buffer power falls. Buffer power for an individual buffer is therefore zero at both limiting low and high values of Kd. This relationship can be seen in the contour plot of FIGURE 4B and can be examined by differentiating Eq. 8 with respect to Kd at constant [Ca2+]:
(12) |
FIGURE 4.
The dependence of buffer power on equilibrium dissociation constant (Kd) and calcium concentration ([Ca2+]). A: 3-dimensional (3-D) plot showing buffer power (z-axis) as a function of Kd at the various [Ca2+] values indicated. A buffer concentration of 100 µM was assumed. Note that at each [Ca2+] buffer power increases and then decreases with increasing Kd. B: contour plot showing buffer power as a function of Kd (x-axis) and [Ca2+] (y-axis). The horizontal dashed red line shows how buffer power changes with Kd at constant [Ca2+].
For a given [Ca2+], the maximum buffer power (βi,max) of an individual buffer will occur when and therefore Kd = [Ca2+]. At this value of [Ca2+], substitution into Eq. 8 shows that the maximum buffer power will be given by
(13) |
Therefore, the lower the Kd, the greater the maximum buffer power (cf. FIGURE 3, bottom). One important issue concerns the effect that a given change of Kd will have. Consider a case when a physiological or pathological change increases buffer Kd from Kd,1 to Kd,2. From Eq. 8 the buffer power will be identical when
This is satisfied when [Ca2+] equals the geometric mean of the two values of Kd, i.e.,
(14) |
For example, when comparing βi for various [Ca2+] for the two buffers exemplified in FIGURE 3, it is seen that their buffer power is identical at [Ca2+] = = 1 μM (FIGURE 3Bb). Below this level of [Ca2+] the increase of Kd from 0.5 to 2 µM will decrease buffer power, and above it buffer power will increase. As illustrated below (see FIGURE 10 and sect. 4.3), changes of buffer Kd are predicted to have complicated effects on the kinetics of decay of the [Ca2+] transient, depending on whether the level of [Ca2+] is below or above the geometric mean of the Kds.
FIGURE 10.
Predicted effects of changing the buffer affinity and amount of Ca2+ released on the decay of intracellular calcium concentration ([Ca2+]i). A: effects of Kd. Model similar to that of FIGURE 7 except that the buffer had different Kd values, as indicated. As for FIGURE 7, the buffer is assumed to bind Ca2+ instantaneously. a: [Ca2+]i. b: Normalized [Ca2+]i. c: Instantaneous rate constant of decay. d: Buffer power. A portion of the records at expanded timescale is displayed on right. Note the expanded vertical scale on the right graph of d. B: effects of changing the amount of Ca2+ released. The same saturating buffer (Kd = 500 nM) was used throughout. The increase of total cytoplasmic Ca2+ was changed from 25 (black) to 50 (red) µM.
3.3. Buffers with Cooperative Ca2+ Binding
Above, we have described the features of simple interactions between Ca2+ ions and small buffer molecules and the independent binding of Ca2+ ions to larger molecules. Frequently, however, Ca2+ binding to multiple sites on a protein is either positively or negatively cooperative: binding one Ca2+ ion enhances or diminishes the affinity of an adjacent site during a subsequent binding step. This type of interaction is commonly seen with Ca2+ binding to buffers such as EF-hand proteins and Ca2+ pump proteins. Both sequential and independent schemes can be summarized by reactions 3A and 3B:
reaction 3A |
reaction 3B |
Ba and Bb are two sites on the same protein with individual affinities described by Kd(a) and Kd(b). When two binding sites are in close proximity, the binding of Ca2+ to an array of negative charges may be expected to reduce the chances of a second Ca2+ binding purely on the basis of electrostatics, thus causing negative cooperativity (17). The fact that positive cooperativity is commonly seen in adjacent Ca2+ binding sites means the effect of favorable structural changes evoked by binding of the first Ca2+ ion must outweigh the electrostatic effects. Cooperative Ca2+ binding may also arise from interactions between linked protein molecules, for example in the case of Troponin-C (TnC), where binding of Ca2+ to one TnC will change the structure of the thin filament and increase binding to other TnC proteins bound to adjacent sites on the thin filament (32, 181, 182).
In many cases, the properties of the individual binding sites are unknown and the cooperative binding can be usefully approximated by reaction 4.
(reaction 4) |
As discussed previously (183), this can be described mathematically by the Hill equation (Eq. 15) (184). Note that in Eqs. 15–28 we abbreviate [Ca2+] to Ca for simplicity:
(15) |
Here, Kapp is the concentration of Ca2+ at which half the binding sites are occupied. n is the Hill slope, which represents the degree of cooperativity, with a value of 1 meaning no cooperativity. For a molecule with two sites that bind Ca2+ cooperatively, Kapp and n can be calculated from the individual dissociation constants [Kd(a) and Kd(b)] as follows (185): and . In the special case where Kd(b) ⪢ Kd(a), it can be seen that n approaches 2. We have used this method to calculate values of Kapp and n in TABLE 2.
Table 2.
Examples of individual Kd values from pairs of EF-hand domains in different Ca2+ binding proteins along with the associated Kapp value and the Hill slope
Buffer Name | Ka, µM | Kb, µM | Kapp, µM | Hill Slope | Ref. |
---|---|---|---|---|---|
CB-D28K | 0.41 | 0.24 | 0.31 | 1.1 | 30 |
CB-D9k | 0.31 | 0.15 | 0.21 | 1.2 | 186 |
CR | 2.8 | 0.068 | 1.38 | 1.5 | 185 |
CaM (COOH terminus) | 28 | 0.26 | 2.7 | 1.8 | 187 |
CaM (NH2 terminus) | 193 | 0.79 | 12.7 | 1.9 | 187 |
SERCA | 8.3 | 0.02 | 0.41 | 1.9 | 98 |
Ka represents the equilibrium constant for the initial binding and Kb the equilibrium constant created by the cooperative interaction. CalbindinD9k (CB-D9k) has 1 pair of Ca2+-binding EF-hand motifs and calretinin (CR) and calbindinD28k (CB-D28K) have 2 pairs; the values listed apply to both pairs in that molecule. Note that CR has an additional functional EF-hand site that is not described here. The pairs of EF-hand domains situated at the COOH- and NH2-terminal ends of calmodulin (CaM) are shown separately. SERCA, sarco(endo)plasmic reticulum Ca2+-ATPase.
Ca2+ binding domains can exist in pairs, and in many CBPs multiple pairs of domains exist. The structure and therefore function of these domains are not identical; some differ to the extent that Ca2+ cannot bind to one of the pairs of sites, in other domains Mg2+ or Ca2+ can bind and engage (Ca/Mg sites), and in others Mg2+ can bind but not activate the site (Ca2+-selective sites). In the case of some well-studied CBPs such as calmodulin (4 EF sites) (72), calretinin (5 EF sites) (185), and calbindin-D9k (2 EF sites) (186), the cooperativity is between the two adjacent EF-hand domains. TABLE 2 shows examples of several CBPs and the varying extent of cooperativity that has been measured. Detailed information on all individual Ca2+ binding sites in a single protein is only available for a few members of the large family of CBPs. More commonly, only general descriptions of Ca2+ binding in terms of the Hill Kapp and slope value have been reported, and we now consider cooperative buffering in these terms.
3.3.1. The effects of cooperativity on buffer power.
The effects of a cooperative Ca2+ binding scheme on the relationship between free [Ca2+], Ca2+ binding, and buffer power are shown in FIGURE 5. As the Hill slope (n) value increases, so does the steepness of the dependence of bound on free Ca2+ (FIGURE 5, top). An instructive comparison is between n values of 1.0 and 1.25 (FIGURE 5A). The difference in the binding curves (FIGURE 5A, top) would be difficult to distinguish experimentally, at least with the techniques applied to intact cells and tissues. However, it has marked effects on the dependence of buffer power on [Ca2+] (FIGURE 5A, bottom).
The effects of cooperative binding on Ca2+ buffering have been considered previously (188). Here, we calculate the buffer power as a function of [Ca2+] by differentiating Eq. 15:
(16) |
The dependence of the buffer power on [Ca2+] can be shown by differentiating βi with respect to [Ca2+] to obtain
(17) |
When dβi/dCa = 0, buffer power (β) will have a maximum value, which will be obtained at a value of [Ca2+] given by
(18) |
Consistent with Eq. 8, this predicts for n = 1 that the highest buffer power is obtained at zero [Ca2+]. In contrast, when the buffer is cooperative (n > 1), buffer power is low at low [Ca2+] and peaks at intermediate [Ca2+], before decreasing to zero at higher [Ca2+] (188) (FIGURE 5). Two other conclusions can be derived from FIGURE 5: 1) An increase of n reduces the range of [Ca2+] over which buffering occurs. Since the total amount of Ca2+ bound is unaffected, this means that the maximum buffer power is greatest at higher values of n. Indeed, for n = ∞ (not shown), the buffer power curve will have an infinitely high and narrow peak. 2) The value of [Ca2+] at which the maximum buffer power is obtained increases with n, approaching a value of Kapp at high n.
As regards the dependence of buffer power on Kapp, one can differentiate Eq. 16 with respect to Kapp and obtains
(19) |
For a given [Ca2+]i, the maximum buffer power is obtained when this derivative is equal to zero. Irrespective of the value of n, this occurs when Kapp = [Ca2+]. Substituting into Eq. 16 and expressing as buffer power (β) gives
(20) |
In other words, the maximum buffer power (at [Ca2+]i = Kapp) is proportional to n. For the simple case of n = 1 (no cooperativity), this reduces to Eq. 13.
It can also be shown that Eq. 14 holds for cooperative buffers. If Kapp increases from Kapp,1 to Kapp,2, there will be no change of buffer power at a [Ca2+] given by the geometric mean ().
A good example of the effects of cooperativity on buffer power is provided by calsequestrin (CSQ). At [Ca2+] < 50 µM, CSQ exists as monomers; an increase of free Ca2+ causes binding to acidic residues at the NH2 terminal and COOH terminals, changing their tertiary structure allowing NH2-to-NH2 terminal and COOH-to-COOH terminal region binding. This results initially in the formation of dimers, then as [Ca2+] increases tetramers, and at the highest Ca2+ concentration (above ∼5 mM) polymers. Recent work suggests that CSQ forms polymers of helically arranged CSQ monomers (138). This change in tertiary structure exposes further binding sites on the outer surface and the water-filled lumen of the helix. This highly cooperative reaction scheme means that, unlike the case of a noncooperative buffer, the buffer power of CSQ increases with Ca2+ binding. As shown in FIGURE 6, buffer power calculated from published data on Ca2+ binding to CSQ (133) shows a complex relationship between luminal Ca2+ concentration and Ca2+ buffer power as a consequence of the multiphasic and cooperative binding curve. The extent and kinetics of this reaction are not fully understood. There are no values for n in the literature, but a value of 10 has been used in modeling (189), and values of 5–9 fit the data of Park et al. (133) (see FIGURE 6). The properties of CSQ are discussed more fully in the context of skeletal muscle (sect. 7.5).
FIGURE 6.
Ca2+ dependence of buffering by calsequestrin (CSQ). The data points featured are measurements of Ca2+ binding to CSQ at different free calcium concentration ([Ca2+]) by Park et al (133). The solid black line represents a multiple summed sigmoidal relationship that best fit the data. The red line is the buffer power calculated from the black fit. Note that the multiphasic steep relationship results in limited epochs of Ca2+ buffer power over the range of luminal [Ca2+] expected.
For the different CBPs listed in TABLE 2, the cooperativity arises from the interaction between the two adjacent EF-hand domains on the same protein or, in the case of the binding sites on SERCA, from a sequential reaction scheme. For Ca2+ binding to isolated fast-twitch skeletal TnC, the adjacent EF-hand domains (sites I and II) have distinct Kd values (affinity of site II is ∼10 fold higher than that of site I) as a consequence of sequential cooperative binding (32, 190), i.e., Ca2+ binding to site II is necessary to generate the higher-affinity structure of site I. Note that slow-twitch skeletal and cardiac TnC binds Ca2+ only to site II, as site I is nonfunctional (181). Therefore, the cooperativity of Ca2+ binding to slow-twitch/cardiac TnC observed when TnC is bound to the troponin-thin filament-myosin complex (191) is a consequence of long-range effects of Ca2+ TnC binding transmitted through the thin filament to adjacent TnC sites. Other examples of sequential cooperative schemes can be seen in intraorganelle proteins such as calreticulin and calsequestrin, which can bind up to 20 Ca2+ ions per molecule (see sects. 2.6.1 and 7.5).
As mentioned above, detailed kinetic parameters for cooperative Ca2+ binding exist only for a few CBPs. A study of Ca2+ binding to calretinin revealed two pairs of sites with cooperative interactions and a single independent site (185). The cooperative sites consisted of a low-affinity rapid binding and a higher-affinity slower binding site. A protein with such kinetically heterogeneous sites will maintain rapid Ca2+ binding kinetics across a wide range of baseline Ca2+ concentrations, a feature not possible in simpler buffer systems (185). This unusual buffer feature is consistent with the experimental observation that calretinin modifies the time course of the IP3-evoked Ca2+ transient in ways that cannot be reproduced by either BAPTA or EGTA (192) and illustrates the importance of understanding other intracellular Ca2+ buffers in detail to fully appreciate the functional consequences of buffer action.
4. THE EFFECTS OF Ca2+ BUFFER PROPERTIES ON Ca2+ SIGNALING
4.1. Effects of Buffers on [Ca2+]i
It is important to consider the roles that Ca2+ buffers play in the regulation of cytoplasmic Ca2+ concentration ([Ca2+]i). Buffers decrease the change of [Ca2+]i produced by a given flux of Ca2+. A similar effect can be produced by transport of Ca2+ into organelles, and the combined effects have been referred to as “muffling” (193). In this review, however, we are concerned primarily with physicochemical buffering as opposed to transport into and out of intracellular organelles. It is important to note that, in contrast to the use of the term “buffer” in chemistry, Ca2+ buffers cannot alter the steady-state level of [Ca2+]i in an intact cell. At steady state, the level of [Ca2+]i reflects a simple balance between Ca2+ entry and efflux at the surface membrane (194–196), whereas buffers are equilibrated with Ca2+ according to their Kd. A caveat applies to whole cell patch-clamp recordings during which the intracellular medium is connected to a large volume of buffered pipette solution, which can influence the steady state [Ca2+]i. What buffering does under normal cellular conditions is to slow down and decrease the effects of changes of total Ca2+ on changes in free Ca2+. FIGURE 7 simulates the changes in free Ca2+ resulting from a bolus of Ca2+ entering the cytoplasm. [Ca2+]i increases in a steplike fashion and subsequently decays back to its basal value. It is assumed that Ca2+ is removed from the cytosol by a plasma membrane Ca2+ pump with activity proportional to [Ca2+]i. Four scenarios are shown, for βi equal to 25, 50, 100, or 200. In all cases, it is assumed that Ca2+ binds instantaneously to the buffer and that diffusion of Ca2+ and buffer across the volume of interest is rapid. Higher buffer power strongly attenuates the rise of free [Ca2+]i. Equally important is the prominent slowdown of the decay, since the pump will have to remove not only free Ca2+ but also the Ca2+ that is dissociating from buffers while free Ca2+ decays back to baseline values. Decay time constant τ and inverse of the peak amplitude (A−1) are plotted as a function of βi in FIGURE 7, B and C, respectively. These plots illustrate that the fraction by which the decay is slowed is identical to that by which the amplitude is reduced (see also sect. 4.7.1.1 and Eqs. 30–32). If Ca2+ removal is a linear function of [Ca2+]i, then the area under the curve (time integral A × τ) and correspondingly the time-averaged [Ca2+]i during the transient will be unaffected by the presence of buffers (FIGURE 7D). The effects of experimental alteration of buffer power on Ca2+ transients are illustrated in FIGURE 12, FIGURE 16B, and FIGURE 21A (see also sect. 9.7 for a discussion of the effects of changing the concentration of the buffer calretinin on the changes of [Ca2+]i). The impact of buffering on processes regulated by [Ca2+]i will depend on the activation mechanisms of these processes. For a process that is activated linearly with [Ca2+]i, the presence of a buffer will not alter the cumulative effect of Ca2+ transients. If the Ca2+-regulated process has a steep supralinear dependence on [Ca2+]i, then increased buffering will decrease the end effect, since high Ca2+ concentrations are ablated. On the other hand, a process with a saturating Ca2+ dependence will actually be enhanced by the presence of a buffer, since the latter attenuates the amplitudes of the [Ca2+]i transients, thereby preventing saturation of the Ca2+-dependent process, while prolonging the duration of Ca2+ action.
FIGURE 7.
Simulation of the effects of altering buffer power (βi) on amplitude and decay of a cytosolic intracellular calcium concentration ([Ca2+]i) transient. A: [Ca2+]i is controlled by a constant Ca2+ leak into the cell that is balanced by a Ca2+ pump with an activity equal to γ × [Ca2+]i where γ = 1,000 s−1. At 20 ms, 10 µmol of Ca2+ was added per liter of cytosol. The 4 traces denote simulations performed with the buffer powers indicated. A very low-affinity buffer was assumed. Note that increasing βi decreases the amplitude and slows the decay of the transient. The area under the [Ca2+]i transient above resting [Ca2+]i of 100 nM is unchanged (see also Eqs. 30–32). B–D: predicted dependence on βi of decay time constant τ (B); reciprocal of the [Ca2+]i transient amplitude A (C); and the product A × τ (D). Symbols represent the respective values for the sample traces in A with corresponding colors. The lines in B and C intersect with the x-axis at a value of βi of −1.
FIGURE 12.
Effects of exogenous buffers on cardiac myocytes. A: fast buffer. Top: measurements of intracellular calcium concentration [Ca2+]i; from left to right: control cell; after loading with the photolabile Ca2+ chelator nitr-5; after flash photolysis of nitr-5. Bottom: averaged, normalized traces. B: slow buffer. Top: recordings from a cell loaded with NP-EGTA before (left) and after (right) flash photolysis. Bottom: normalized traces. Figures redrawn from Ref. 199, with permission from Biophysical Journal.
FIGURE 16.
Estimating cytosolic Ca2+ buffering power by competition of endogenous with added exogenous buffers. A: somatic intracellular calcium concentration ([Ca2+]i) transients recorded in a cerebellar Purkinje neuron either shortly (top) or 18 min (middle) after establishing whole cell configuration with 500 µM fura-2 in the patch pipette. Bottom: the progressive loading of the cell with the dye. Amplitudes and decay kinetics of [Ca2+]i transients are resistant to fura-2 loading via a patch pipette. Modified from Ref. 443, with permission from Journal of Physiology. B: similar experiment as in A performed in a calyx of Held nerve terminal. a: Loading of the terminal with 100 µM fura-2 monitored with the fluorescence at the isosbestic excitation wavelength (top). Fluorescence signals at 385-nm excitation (F385; middle) and corresponding AP-evoked [Ca2+]i transients (bottom) are shown for 3 time points after establishing whole cell configuration (top; filled circles). With increasing dye concentration, [Ca2+]i transients decrease in amplitude and decay more slowly (cf. FIGURE 7). b: Decay time constant τ (top) and inverse of the peak amplitude A−1 (bottom) plotted as a function of the fura-2 Ca2+-binding ratio (exogenous buffer power) for various time points during dye loading. Regression lines through the τ and A−1 plots yield extrapolated y-axis intercepts for τ and A−1 at = 0 (no added buffer). Negative x-axis intercepts represent estimates for 1 + (endogenous buffer power). Modified from Ref. 440, with permission from Biophysical Journal. C: similar as in B except that Ca2+ influx was triggered by short trains of action potential (AP)-like depolarizations, a low-affinity dye was used, and plotted decay constants represent pooled data obtained from several terminals loaded with 3 different dye concentrations. Each data cluster corresponds to 1 of the 3 fura-6F concentrations. Modified from Ref. 248, with permission from Journal of Physiology.
FIGURE 21.
Endogenous Ca2+ buffers shape the time course of intracellular calcium concentration ([Ca2+]i) transients in dendrites and spines. A: [Ca2+]i transients evoked by single action potentials (APs), measured in the proximal dendrites of cortical pyramidal neurons that lack expression of parvalbumin (PV), CB-D28k, and calretinin (CR) and therefore exhibit a low endogenous Ca2+-binding ratio, somatically loaded with either 20, 80, 125, or 250 µM fura-2. Gray areas represent the integral A × τ, the product of peak amplitude A and time constant τ fitted to the decay, which is similar under all recording conditions. Modified from Ref. 441, with permission from Biophysical Journal. B: slow and biphasic decay kinetics of [Ca2+]i transients in dendritic shafts (top) and spines (bottom) of cerebellar Purkinje cells. The superimposed continuous lines represent double-exponential fits to the decay. The dashed lines represent an assumed resting [Ca2+]i of 45 nM. Insets show the fits with their corresponding fast (red) and slow (blue) components. In the spine, the fit yielded a similar slow component but a much larger fast component, likely reflecting fast efflux of Ca2+ from the spine into the dendrite. Modified from Ref. 510, with permission from Journal of Physiology.
For these conclusions to be quantitatively correct, it has to be assumed that the on and off rate constants are infinitely fast such that bound and free Ca2+ are always in equilibrium, that diffusion of buffers and Ca2+ across the cellular dimensions is rapid, and that buffering power is constant over the relevant range of changes in [Ca2+]i (see sects. 4.2 and 4.7.1).
4.1.1. Effects of buffers on the calculations of fluxes.
Fluxes of Ca2+ across cell membranes are often estimated by measuring the resulting changes of [Ca2+]i, and it is often overlooked that changes of Ca2+ buffering will alter the change of [Ca2+]i produced by a given flux. For example, an increase of buffer power will attenuate the increase of [Ca2+]i produced by a given influx of Ca2+ through a membrane Ca2+ channel. Furthermore, the activities of Ca2+ removal processes are often assessed from the time constant of decay of [Ca2+]i. However, as shown in FIGURE 7, A AND B, this time constant is also affected by Ca2+ buffering. This raises the possibility that an increase of the time constant of decay of [Ca2+]i could result from an increase of buffering as opposed to a decrease of pumping (197). Given these issues, it is best to calculate Ca2+ fluxes from changes of total Ca2+ estimated from the buffer properties and changes of [Ca2+]i (80, 197).
4.2. Effects of Buffer Kinetics
It is important to distinguish between instantaneous and steady-state buffering. As shown in reaction 1, a simple buffer can be characterized by binding (kon) and unbinding (koff) rate constants, which determine the time over which buffering occurs. FIGURE 8A shows the response of both free and bound Ca2+ to a step increase of total [Ca2+] for buffers with a similar Kd but different rate constants. For simplicity we assume that there is no pumping of Ca2+ and therefore changes of [Ca2+]i are determined solely by Ca2+ binding to and Ca2+ unbinding from the buffer. The resulting time courses of total, free, and bound Ca2+ are shown for a slow (EGTA) and a fast (BAPTA) buffer. In both cases, all added Ca2+ initially appears as free Ca2+. As Ca2+ progressively binds to the buffer, [Ca2+]i decreases, and this occurs more quickly with the faster buffer. Finally, in the case of BAPTA, [Ca2+]i approaches a steady state within a few milliseconds, at which it is in equilibrium with the bound state according to Eq. 1. EGTA is sufficiently slow that [Ca2+]i is still declining at the end of the period shown. Two further points are worth noting: 1) The transient overshoot of [Ca2+]i will only occur if the rise of total Ca2+ is very fast relative to kon; if slower, the fast phase of decay of free Ca2+ will be obscured. 2) Even if the fast phase of decay does occur, the ability to detect it will require a fast Ca2+ indicator and a fast acquisition rate.
FIGURE 8.
Effects of buffer kinetics. A: response to a bolus of total Ca2+. a: Total Ca2+ was increased by 10 µM. b: Free calcium concentration ([Ca2+]free). c: Ca2+ bound to buffer. Simulations are shown for both EGTA (black solid lines) and BAPTA (red dashed lines). Note the transient overshoot of intracellular [Ca2+] ([Ca2+]i) for both buffers, the decay of which is scarcely apparent at this timescale for the case of EGTA. B, a: a transient rise of [Ca2+]i was imposed. b and c: The calculated change of Ca2+ bound to BAPTA (b) and EGTA (c). The solid lines in d and e show hysteresis plots of bound Ca2+ plotted as a function of free Ca2+ calculated from the responses shown in b and c, respectively. The arrows indicate the direction of time. The dashed lines show the steady-state relationship between free and total Ca2+.
It is also instructive to see how Ca2+ binding to a buffer tracks a rise of free [Ca2+]. Comparison of FIGURE 8, Bb AND Bc, shows that binding to even the fast buffer BAPTA lags slightly behind free Ca2+ and the lag is much greater for EGTA. The hysteresis curves (FIGURE 8, Bd AND Be) plot bound as a function of free Ca2+, both for the rise and fall of [Ca2+]. For BAPTA, the hysteresis curve only deviates markedly from the steady-state relationship during the very fast rise of free [Ca2+]. In contrast, for the slower buffer EGTA, there is a marked difference between the instantaneous and steady-state relationships throughout. FIGURE 8Be also shows that less Ca2+ is bound at a given free [Ca2+] during the transient compared to the steady state (when bound and free Ca2+ are in equilibrium), indicating that EGTA buffers more weakly than BAPTA during a transient.
As discussed in sects. 2.2 and 2.5.1.1, most slow kinetics arise not from intrinsically slow kinetics of Ca2+ binding but rather from the time it takes for Mg2+ or H+ to dissociate, making binding sites available for Ca2+. If [Ca2+]i increases, free Ca2+ can quickly bind only to free buffer, which may be a small fraction of the total. By mass action, the resulting decrease in concentration of the free buffer will decrease the concentration of the Mg2+- or H+-bound forms, thereby regenerating free buffer for Ca2+ to bind to. Thus, the steady-state buffering is much greater than the instantaneous one. The speed at which this occurs is limited by the rate at which the competing ions dissociate. This can result in two phases of Ca2+ binding: rapid binding to the free buffer and delayed binding to that which had Mg2+ bound. This is illustrated in FIGURE 9, which also shows how a slow buffer (here PV as an example) and a Ca2+ pump combine to promote relaxation of [Ca2+]i. The results should be compared with the simulation of FIGURE 7 for fast buffering. The simulation illustrated by the black solid line in FIGURE 9 assumes a total concentration of 500 µM PV. Initially, only 25 µM PV is free, with 205 and 270 µM, respectively, bound to Ca2+ (FIGURE 9B) and Mg2+ (FIGURE 9C). The addition of 150 µM total Ca2+ results in a rise of [Ca2+]i (FIGURE 9A), leading to a rapid increase of Ca2+-bound PV (FIGURE 9B) (over the period indicated by i), as Ca2+ binds to available free PV. The kinetics of this increase are determined by the binding and dissociation rate constants of Ca2+ with PV. This is accompanied by an initial rapid fall of [Ca2+]i. A slower phase (ii) then ensues when Mg2+ dissociates from Mg2+-bound PV (FIGURE 9C) in exchange for Ca2+ binding (FIGURE 9B). There is also a decrease of free PV (not shown). On a longer timescale, beyond the period shown, the concentrations of all species will return to the initial levels. Such biphasic decays have been demonstrated when PV was added to adrenal chromaffin cells (27) and also recorded in cardiac muscle when slow buffers were added to the cytoplasm (198, 199).
FIGURE 9.
Comparison of a slow Ca2+ buffer with a Ca2+ pump. All simulations use a slow buffer, here taken to be parvalbumin (PV) at the concentrations indicated (500 or 1,000 µM). Steady-state intracellular calcium concentration ([Ca2+]i) was controlled by a leak flux opposed by a pump. Pump rate is given by pump·[Ca2+]i, where the rate constant of pump (units of s−1) is defined in the key. The leak was adjusted to give a resting [Ca2+]i of 100 nM in all cases. Total Ca2+ was increased by 150 µM at 20 ms. A: [Ca2+]i. B: Ca2+ bound to buffer. C: Mg2+ bound to buffer. D: Ca pump rate. The fast and slow phases of Ca2+ binding are denoted by i and ii. Note that the data for 1,000 µM PV case are off the top of the range shown in B and C.
FIGURE 9 also serves to compare the effects of PV with those of a Ca pump. The red dashed lines show the effects of reducing the pump to 1%. [Ca2+]i still decreases, and this is accompanied by larger changes of bound Mg2+ and Ca2+ indicating that PV is compensating for the loss caused by the pump. The two final simulations show that either increasing PV (dotted blue lines) or the pump (green solid lines) can accelerate the slow decay component of [Ca2+]i to similar levels, with the former resulting in more Ca2+ binding to PV and the latter in less. FIGURE 9D indicates the effects on pump activity, which are discussed in more detail in sect. 7.2.
4.3. Effects of Kd and Ca2+ Dependence of Buffering
The influence of buffer Kd on Ca2+ signaling can be appreciated from FIGURE 10A, which shows a simulation of the effects of adding the same amount of Ca2+ to the cytoplasm in the presence of fast buffers with different values of Kd. The effect of buffer Kd on the decay depends on the fact that (see sects. 3.1. and 3.2) buffer power is sensitive to both Kd and cytosolic [Ca2+]i. In contrast to the simulations of FIGURE 7, where buffer power was assumed to be constant (Δ[Ca2+]i ≪ Kd), the decays of [Ca2+]i are not single exponentials as shown (FIGURE 10Ab) by the curves intersecting. This is emphasized in FIGURE 10Ac, which plots the instantaneous rate constant of decay of [Ca2+]i. For a simple exponential decay this would be constant. However, the instantaneous rate constants of all [Ca2+]i transients shown are initially high and then decrease. This is caused by the buffer saturation and, consequently, decreased buffer power at high [Ca2+]i (FIGURE 10Ad). Thus, a given rate of Ca2+ pumping results in a faster fall of [Ca2+]i. Such saturation of Ca2+ buffers has been suggested to account for the initial rapid decay of the caffeine-evoked Ca2+ transient in ventricular myocytes (197). As evident from the normalized traces (FIGURE 10Ab) and especially the instantaneous rate constants, the buffer with the highest affinity (Kd = 0.25 µM) causes a faster initial decay of the [Ca2+]i transients than do the others, because of the greater binding of Ca2+ and consequent decrease of buffer power at the peak (FIGURE 10Ad). As [Ca2+]i approaches resting levels, buffer power of the high-affinity buffer increases and the Ca2+ transient decays more slowly. In other words, for a given amount of Ca2+ added to the cytoplasm, increasing the affinity of the buffer causes stronger buffer saturation and produces a [Ca2+]i decay that crosses over with the lower-affinity cases (see sects. 6.2.2–6.2.5 for other practical consequences in cardiac muscle).
FIGURE 10B demonstrates the predictions for increasing the amount of added Ca2+. Doubling the Ca2+ release into the cytosol results in a larger Ca2+ transient (FIGURE 10Ba), which initially decays much faster because of buffer saturation (FIGURE 10Bb). Once again, it is clear from the plot of instantaneous rate constants (FIGURE 10Bc) that the decays are not single exponentials. However, this might not be apparent in real, noisy data, and it would be all too easy to assume that the larger transient decays more quickly because of enhanced Ca2+ pumping as opposed to buffer saturation.
4.4. The Effects of Cooperative Buffers
As shown above, for noncooperative buffers buffer power decreases as [Ca2+]i increases. In contrast, for a cooperative buffer, buffer power will first increase but then decrease once [Ca2+]i rises above an optimal concentration (FIGURE 5). FIGURE 11 demonstrates that this may have marked effects on Ca2+ signaling. The black traces show a simulation (similar to that of FIGURE 10A) for a noncooperative buffer (n = 1). The red trace shows the prediction for a cooperative buffer (n = 2). Here the peak value of [Ca2+]i is greater than the level at which maximum buffering occurs, resulting in an increase of buffer power. As [Ca2+]i falls toward the level at which peak buffer power occurs, buffer power increases further and this is accompanied by a decrease of the instantaneous rate constant of decay of [Ca2+]i. However, as [Ca2+]i decays below the level of peak buffering, buffer power decreases and the rate constant of decay will increase. Therefore, in contrast to a simple buffer, where for large Ca2+ transients the instantaneous rate constant of decay is expected to decrease as [Ca2+]i falls, a slowing followed by acceleration should be observed. It should, however, be noted that the kinetics of decay of the Ca transient will also be affected by the dependence of Ca2+ pumping on [Ca2+]i.
FIGURE 11.
The effect of a cooperative buffer on the decay of intracellular calcium concentration [Ca2+]i. Model similar to FIGURE 10. The buffer had a Kapp of 0.5 µM. Traces show the predicted effects of values of Hill coefficient (n) of 1 (black solid lines) and 2 (red dashed lines). A: [Ca2+]i. B: normalized [Ca2+]i. C: instantaneous rate constant of decay. D: buffer power.
4.5. The Influence of Buffering on the Diffusion of Ca2+
Diffusion of Ca2+, i.e., the ability of the ion to move along its concentration gradient, is affected by interactions of solvated Ca2+ with the components in the medium. A striking physiological example of the ability of Ca2+ buffers to enhance diffusion is that of the calbindin proteins, which facilitate Ca2+ diffusion across epithelial cell layers (200) (see sect. 10). Often, the bulk of cytoplasmic Ca2+ buffering is from fixed structures, i.e., cytoskeleton, membranes, contractile proteins, and pumps on organelles, which will not contribute to the diffusive flux of Ca2+ (173, 201). Therefore, even if mobile cytoplasmic Ca2+ buffers only make a small contribution to the total cellular buffer power, they can have a large role in diffusion (174, 202–204). The exact contribution will depend on the relative diffusion coefficient of the Ca2+ bound forms and on their buffering power (see Eq. 29). An estimate for the diffusion coefficient (D) can be obtained from the known dependence of D on molecular weight (M) as described by a simplified form of the Stokes–Einstein equation (D ∝ 1/Mr); r can be measured empirically and is 0.33–0.40 over a wide range of M (205, 206). In this review, we use a value of 0.33.
Kinetic considerations aside, intracellular Ca2+ buffers will affect Ca2+ diffusion in two distinct ways: 1) for any cytosolic influx/efflux event, buffers will reduce the free Ca2+ concentration gradient generated and therefore the rate of intracellular diffusion and 2) the reduction in diffusive flux of Ca2+ due to a lower free Ca2+ concentration gradient may be offset by a contribution to diffusion by the Ca2+-bound form of the buffer; the magnitude of this component will depend on the value of the diffusion coefficient of the complex and the concentration gradient of the bound Ca2+ (178, 202, 204).
Specifically, the flux of free Ca2+ due to diffusion is described by the Fick equation.
(21) |
where DCa is the diffusion coefficient of free Ca2+ and dCa/dx is its concentration gradient.
The flux of bound Ca2+ is similarly
(22) |
where DCaB is the diffusion coefficient of Ca2+ bound to buffer and dCaB/dx is its concentration gradient.
The total flux is
(23) |
Henceforth, we assume that Ca2+ binding and dissociation are fast enough such that the local bound and free Ca2+ are in equilibrium; therefore dCaB/dx can be replaced by the product of dCa/dx (the free Ca2+ gradient) and dCaB/dCa (the buffer power), resulting in
(24) |
which can be rewritten as follows:
(25) |
Dividing Eq. 25 by Eq. 21 gives the relative contribution made by the flux of bound compared to free Ca2+ to diffusion as
(26) |
This means that buffers make the biggest fractional contribution to Ca2+ diffusion when they also buffer Ca2+ most strongly (βi,max). For a simple noncooperative buffer, where the binding is described by a single dissociation constant (Kd), we can substitute from Eq. 8 for βi(Ca) to give
(27) |
More generally, including cooperative buffers, we substitute from Eq. 16:
(28) |
Hence, noncooperative buffers contribute most to Ca2+ diffusion at low [Ca2+] (Eq. 27). As [Ca2+] reaches high levels, the buffer will become saturated [β(Ca) = 0], there will be no gradient of CaB, and therefore no contribution to net diffusion. In contrast, for cooperative buffers, their contribution to Ca2+ diffusion will be low at both very low and high [Ca2+] and greatest at a value of [Ca2+] given by Eq. 18. As regards the dependence on Kd, for noncooperative buffers their contribution to diffusion will be largest when Kd = [Ca2+]i (Eq. 13) (207), and a similar condition holds for cooperative buffers (Eq. 20).
Equations 21–28 describe the local fluxes contributed by individual Ca2+-binding components of the cytosol. When confronted with the problem of solving the spatiotemporal pattern of diffusion of several interacting components, this can, of course, be accomplished by numerical integration (see sect. 4.7). For certain limiting cases, however, simple analytical solutions can be derived, which often provide a more intuitive understanding of the processes involved.
Thus, for the case of small changes in [Ca2+] around low values, where the buffering power of all Ca2+-binding species is constant, the “apparent diffusion coefficient” for Ca2+ (Dapp) can be calculated as
(29) |
where the sums extend over all Ca2+-binding species (178, 202). The use of κ in Eq. 29 reflects the original papers, but the equation could equally well be written in terms of βi (sect. 3).
This equation provides an intuitive understanding in the sense that in the numerator the individual terms of the sum specify contributions of the respective buffers to Ca2+ mobility, normalized by the sum of Ca2+ binding ratios of all species, irrespective of whether they are mobile or stationary.
TABLE 3 gives values for some important mobile intracellular buffers. The respective relative diffusion constant (Di/DCa), appearing in Eqs. 26–29, is given in the table for each buffer. One major contributor to Dapp is ATP, which is present at 5–8 mM in virtually all cell types (34, 203). ATP binds Ca2+, Mg2+, and H+ and, despite its relatively low Ca affinity (Kd of ∼1 mM), because of its high concentration it can nevertheless bind significant amounts of intracellular Ca2+ (sect. 2.4.2). This and its low molecular weight mean that the diffusive flux carried by Ca-ATP will be greater than that due to free Ca2+ (see TABLE 3 and Eq. 27). This contribution will be larger in cytoplasmic regions close to channels/exchangers where Ca2+ influx can increase the local [Ca2+] to higher levels, which will saturate higher-affinity buffers. Striated muscle cells also contain high concentrations (up to 20 mM) of histidyl dipeptides (HDPs) such as carnosine (211, 212), and, subject to the concerns about Ca2+ binding at physiological pH (sect. 2.4.1), these may also contribute to Ca2+ diffusion.
Table 3.
Predicted effects of various mobile Ca2+ buffers on diffusion of Ca2+
Mol Weight, g | Relative Diffusion Constant | Conc, µM | Kapp, µM | Hill Slope | Binding Sites | Ca-Bound, µM | β(Ca) | Relative Flux | |
---|---|---|---|---|---|---|---|---|---|
Free calcium | 40 | 1.00 | 0.1 | 1 | 1 | ||||
Calmodulin | 16,700 | 0.14 | 20a | 2.7/12.7b | 1.8/1.9 | 4 | 0.05 | 0.49 | 0.09 |
ATP | 525 | 0.43 | 7,000c | 2,496d | 1 | 1 | 0.28 | 2.80 | 1.20 |
Carnosine | 226 | 0.56 | 10,000e | 1,350e | 1 | 1 | 0.74 | 7.41 | 4.18 |
CB-D9k | 10,000 | 0.16 | 60f | 0.21g | 1.2 | 2 | 34.9 | 148.6 | 48.0 |
CB-D28k | 28,000 | 0.12 | 60f | 0.31h | 1.1 | 4 | 53.7 | 114.6 | 52.8 |
EGTA | 380 | 0.48 | 100 | 0.18i | 1 | 1 | 35.7 | 230 | 109.2 |
Fura-2 | 636 | 0.40 | 100 | 0.3j | 1 | 1 | 25.0 | 188 | 75.3 |
GCAMP | 50,000 | 0.10 | 10 | 0.32k | 3 | 4 | 11.8 | 86.2 | 32.8 |
From left to right: molecular (atomic) weight (MW); diffusion constant relative to Ca2+, calculated as (40/MW)0.33; concentration of buffer; Kapp; Hill slope; no. of Ca2+ binding sites per molecule; concentration of bound Ca2+; buffer power (from Eq. 8 or 16); diffusion flux relative to that of Ca2+, calculated at free Ca2+ of 0.1 µM (from Eq. 27 or 28). N.B. The concentration for carnosine applies to striated muscle and that for calbindinD9K (CB-D9k) and CalbindinD28K (CB-D28k) to Ca2+-transporting epithelia, the concentrations for EGTA, fura-2, and GCAMP represent typical values under experimental conditions. The CaM data represent both the COOH- and NH2-terminal sites. All compounds bind 1 Ca2+ per molecule except: CaM, 4; CB-D9K, 2; CB-D28K, and 4; GCAMP, 4. The values come from Refs. 208 (a), 187 (b), 34 (c), 33 (d), 49 (e), 209 (f), 186 (g), 30 (h), 24 (i), 25 (j), and 210 (k). Note that the table shows a value of 3 for the Hill coefficient of GCAMP but a range of Kds is found for different variants from 1 to 3 (210). These values were used to calculate Kapp. For noncooperative buffers, this was the apparent Kd at pH 7.0 and 1 mM Mg2+; for cooperative buffers, it was calculated as in TABLE 2. Buffers listed in italics are synthetic molecules used in research.
The low intracellular concentration and large molecular weight of Ca2+-signaling molecules such as annexin and CaM mean that their contributions to diffusive flux are small (TABLE 3). For CaM, the inferred diffusion coefficient (20–30 µm2/s) (203, 204) is approximately what would be anticipated from its molecular weight (∼17,000). The calbindins (CB-D9k and CB-D28k) at nominal cytoplasmic concentrations of 300 µM make a significant contribution to diffusive flux. Interestingly, the lower diffusion coefficient of the larger 28k form is offset by being able to bind four Ca2+ compared to the two Ca2+ binding sites on the 9k form (see TABLE 3). It is useful to compare the calbindins with CaM. The molecular weights and therefore diffusion constants are of the same order. The much greater contribution to Ca2+ diffusion of the calbindins arises from two factors: 1) The concentration of the calbindins in epithelia is three times greater than that of CaM. 2) More importantly, the higher affinity of the calbindins means that at lower values of [Ca2+]i the buffer power of calbindin, and therefore its contribution to diffusion, is much greater. Finally, the effects of Ca2+ indicators to promote intracellular Ca2+ diffusion are illustrated by the inclusion of fura-2 and the protein indicator GCaMP in TABLE 3 (note the very high relative fluxes of EGTA and fura-2, which at the assumed concentrations dominate diffusion; see also sect. 4.6).
A mechanism through which a relatively immobile Ca2+ buffer can enhance net Ca2+ diffusion has been suggested to apply in the case of the intra-SR buffer CSQ (213). CSQ polymers within the SR may mediate enhanced Ca2+ diffusion from a distal region of the polymer to a region next to the SR Ca2+ channel via the “jumping” of Ca2+ ions from binding site to site without going through the aqueous phase, which in a linear polymer would form a “Ca2+ wire.” This theory is based on the well-known physical process of diffusion on an adsorptive surface (214) and may enhance net diffusion by 10- to 100-fold. There is, however, no direct evidence that this process features in intra-SR Ca2+ diffusion.
4.5.1. Linkage between diffusion of calcium and protons.
Buffer-mediated diffusion of Ca2+ may also be linked to that of protons, since many buffers can bind both ions. This was originally proposed for calbindin-mediated transport, where it was suggested that protons might facilitate Ca2+ binding resulting in a coupled cotransport of Ca2+ and protons (207). However, there is no evidence to suggest that protons increase Ca2+ binding, and, indeed, the opposite is more likely, as ∼1 H+ is released when Ca2+ binds to an EF-hand domain (see sect. 2.6.1.2). It is therefore likely that calbindin-mediated Ca2+-facilitated diffusion results in countertransport of protons. More recently, competition of H+ and Ca2+ for binding to ATP and carnosine has been suggested to produce a countertransport of Ca2+ and protons in cardiac cells. Acidifying a region of a cell resulted in an increase of [Ca2+]i in that region. This was explained by the buffers shuttling protons away from this region and returning with Ca2+ bound (49).
4.6. Effects of Exogenous Buffers on Ca2+ Signaling
The use of small-molecule intracellular Ca2+ indicators will increase Ca2+ buffering and thus decrease the amplitude of [Ca2+]i transients and alter their kinetics (174, 215) (sect. 4.1). The magnitude of the effect will depend on both the concentration and Kd of the indicator. An important requirement for [Ca2+]i fluorimetry, therefore, is to keep the concentration of indicators low, preferably much lower than that of endogenous buffers. The low molecular weight of fura-2 and its high buffer power at low [Ca2+]i (as a consequence of its high affinity) require extremely low concentrations of this indicator for unperturbed measurement of [Ca2+]i signals (204). The low Kd means that this problem lessens at higher [Ca2+]i as buffer power falls. Equally, however, the accuracy of measuring [Ca2+]i falls at higher [Ca2+]i as the indicator saturates. If the indicator is loaded via a patch pipette, then its concentration is known. Most studies, however, use the membrane-permeant acetoxymethyl (AM) ester form, and the cytosolic concentration is therefore unknown. TABLE 3 shows the effects of a typical 100 µM dye concentration. At 0.1 µM [Ca2+]i, there is a significant contribution to buffer power. Lower-affinity indicators will produce less buffering. However, as [Ca2+]i increases, the contribution to buffering of low-affinity indicators will be greater than that of high-affinity indicators (FIGURE 3).
Exogenously applied slow buffers have been used for a variety of purposes in physiological experiments. In muscle, Ca2+ release from the SR has been studied as Ca2+ “spikes” by using a fast fluorescent Ca2+ indicator in combination with a higher concentration of the slow Ca2+ buffer EGTA (198). The slow buffer keeps [Ca2+]i low so the fast indicator will be mainly in the low-fluorescence state. When Ca2+ is released, [Ca2+]i will briefly and locally reach high levels and thereby be detected by the fluorescent indicator before Ca2+ is bound to EGTA. The amplitude of the spike of [Ca2+]i then gives a measure of the release flux (198, 216, 217). Similarly, so-called “puffs” of local elevated [Ca2+]i, released from the endoplasmic reticulum by IP3, have been studied in various cell types (218–220). Local domains of elevated [Ca2+] in the vicinity of Ca2+ channel clusters were analyzed at ribbon synapses in auditory hair cells by both Ca2+ imaging and computational approaches (221–223). Ca2+ signals due to the opening of single voltage-gated Ca2+ channels could be resolved by total internal reflection fluorescence (TIRF) microscopy (224). A careful balance of fast low-affinity indicator dyes and slow high-affinity buffers (usually EGTA) is required for optimum resolution (see sect. 4.7.2 for a detailed discussion of local [Ca2+]i domains).
The buffering effect of Ca2+ indicators can also be used to advantage. In the extreme case of a large excess of indicator, changes in [Ca2+]i will be very small, but since the indicator binds essentially all the Ca2+ changes in fluorescence at the Ca2+-dependent wavelength of an indicator dye will be proportional to fluxes and can be used for quantitative determination of Ca2+ fluxes through ion channels (173, 225–227).
Exogenous buffers will also affect Ca2+ diffusion. As mentioned above, mobile endogenous buffers contribute only ∼20% of the total cytoplasmic buffer power, so added Ca2+ indicators will increase the relative concentration of mobile buffers within the cytosol and therefore alter physiological Ca2+ gradients and fluxes (174, 203, 204). Although generally present at concentrations (∼50 µM) lower than those of endogenous buffers, these indicators, mostly derivatives of BAPTA, have relatively high affinity (Kd 0.3–1 µM) and therefore their contribution to buffer power and thence Ca2+ diffusion may increase to levels that match or exceed that of endogenous mobile buffers, thereby increasing the diffusion of Ca2+ and attenuating intracellular Ca2+ gradients. This is analogous to the very large influence of low-molecular weight metabolites such as ATP (sect. 2.4.2) on Dapp of Ca2+ (Eq. 29) in cells that contain only low levels of endogenous mobile CBPs. Therefore, and as indicated in TABLE 3, only by ensuring low cytoplasmic indicator concentrations (<10 µM) or using indicators with a low diffusion coefficient can experimenters hope to preserve physiological intracellular Ca2+ gradients while measuring intracellular Ca2+ signals with high-affinity Ca2+ dyes. Genetically encoded Ca2+ indicators of the GCaMP family (228), which are derivatives of GFP, have a MW of ∼ 50,000 and therefore diffuse considerably more slowly (TABLE 3). Nevertheless, at least at resting [Ca2+]i, they are predicted to make a significant contribution to Ca2+ buffering and to interfere with physiological processes (215, 229). In practice, there is a compromise between avoiding excessive additional mobile buffer, while having adequate signal to noise. Efforts are being made to develop more sensitive and brighter protein-based probes (230, 231), which, because of their high MW, may preserve the endogenous diffusion properties better than conventional Ca2+ indicators. An alternative approach is to use dextran-coupled dyes (220, 232). Furthermore, Ca2+ indicators can be targeted to structures within the cell (233, 234), thereby becoming immobile. However, as immobile buffers they may still retard Ca2+ diffusion according to studies of diffusion in cytoplasm (235) because of their contribution in the denominator of Eq. 29.
Studies on the influence of Ca2+-indicator dyes and other small Ca2+-binding molecules on Dapp of Ca2+ have shown that this very much depends on the presence of endogenous buffers. With a MW of ∼400, the diffusion coefficient of indicators would be ∼50% of that of free Ca2+ (200–220 µm2/s). Thus, one would expect Dapp to be near 100 µm2/s if the indicator dye were the dominating buffer. If, on the other hand, an immobile low-affinity buffer with a buffering power of ∼20 (236) were present as the only buffer, Dapp would be expected to be around 200/20 = 10 µm2/s. Measured or inferred values are often in between these values. In rod photoreceptor cells, for instance, Dapp was found to be only ∼15 µm2/s (236), pointing toward a small contribution of the mobile indicator or else a lower buffering power of the fixed buffer. A detailed Ca2+ imaging and modeling study on atrial myocytes (203) concluded that because of the presence of immobile buffers the apparent diffusion coefficient would be as low as 4.25 μm2/s if diffusion were not accelerated by the presence of indicator dyes and ATP. An imaging study of Ca2+ diffusion within nerve axons (204) measured Dapp for a range of different indicator concentrations and extrapolated to zero indicator. This yielded an upper bound for Dapp of <16 µm2/s (see also sect. 4.7.3). Estimates of Dapp in skeletal muscle (237) as well as cardiac muscle (238) were in a similar range of 15–30 µm2/s. Such low values are only obtained, however, if care is taken to use as low indicator concentrations as possible, unless endogenous buffers dominate. Otherwise, the relatively mobile indicators speed up Ca2+ diffusion. The study of Gabso et al. (204), for instance, concluded that Dapp is increased by more than a factor of 3 by 100 µM fura-2. In contrast, the low-affinity indicator CaGreen 5N increases Dapp by only twofold when used at even twice the concentration (200 µM). This was calculated for the case of nerve axons of cultured Aplysia neurons, which have a low abundance of endogenous buffers. Higher concentrations of endogenous mobile and immobile buffers reduce the influence of Ca2+-indicator dyes on Dapp. Furthermore, immobile buffers may be heterogeneously distributed in subregions of the cell cytoplasm in some of which large Ca2+ gradients are generated. A good example is in muscle, where the major Ca2+ buffer, TnC, is only found in the myofilament compartment.
4.7. Modeling of Spatiotemporal Gradients
Quantitative solutions to the combined problem of binding of Ca2+ to several buffers and diffusion in a realistic geometry can only be obtained by numerical computation. Early work by Roberts (239) on the [Ca2+]i patterns generated by Ca2+ influx through arrays of Ca2+ channels in saccular hair cells provided important insights into the interaction between Ca2+ channels and Ca2+-activated K+ channels. This sparked major efforts for the development of software platforms to handle such calculations. Both stochastic tools such as MCell (https://mcell.org/) (240) and deterministic approaches, e.g., CalC (https://web.njit.edu/~matveev/calc.html) (241) have become available. These tools and custom-made software have been used to simulate signals generated by arrays of voltage-dependent Ca2+ channels (222, 242–244) and so-called Ca2+ sparks or Ca2+ puffs (218–220, 245–247). Such calculations provide a complete picture of the complex features, which arise by the combination of buffering and diffusion in a specific geometric environment. They may include the kinetics of Ca2+ binding to and dissociation from several ligands. However, images and graphs produced by these software tools are often snapshots that convey limited intuitive understanding. Therefore, it is worthwhile considering approximate solutions to the problem, which are valid for certain limiting scenarios (178). One of these, the so-called “single compartment model” (SCM) considers the case that the structure of interest is small enough that diffusional equilibration in a given compartment is fast on the timescale of interest. Thus, the SCM considers homogeneous concentrations throughout, and mobility of buffers is not an issue. This assumption was made for many of the equations and features discussed so far.
Another limiting case is the so-called “rapid buffer approximation” (RBA), which considers timescales that are long enough that all binding/dissociation reactions are at equilibrium, while there may still be diffusional gradients within the structure of interest. Equation 29, which allows one to calculate an apparent diffusion coefficient, Dapp, is an example for the rapid buffer approximation. With the additional assumption of linearity of the buffer, the so-called “linear buffer approximation” (LBA), predictions such as those by Eq. 29 become very simple but are restricted to [Ca2+]i smaller than the Kd of the ligand with the highest affinity.
4.7.1. Approximations for special cases.
Here we describe the assumptions and the use of approximate calculations for the special cases listed above.
4.7.1.1. the single compartment model.
For small structures on the order of a few micrometers, such as small cell bodies and longitudinally homogeneous dendrites, [Ca2+] gradients equilibrate on the timescale of 20 ms according to the Stokes–Einstein equation, assuming Dapp = 200 µm2/s. The single compartment model (SCM) is typically applied to describe time courses of small global [Ca2+]i transients elicited by action potentials (APs) in presynaptic boutons, dendritic spines, small cell bodies of neurons, and endocrine cells, such as adrenal chromaffin cells. Although the latter have diameters up to 15 µm, a detailed quantitative analysis of [Ca2+]i gradients found that such gradients elicited by short episodes of Ca2+ influx largely dissipated within 25 ms (163). The SCM is also applicable to striated muscle cells made up of many identical sarcomeres, each of which can be modeled as a single compartment. Together with the assumption of linearity and rapid equilibration (see the LBA and RBA discussed below), the SCM predicts that [Ca2+]i transients have a rapid rise with an amplitude proportional to the total amount of Ca2+ influx and an exponential decay according to
(30) |
with
(31) |
where qCa is the charge carried by the Ca2+ influx, F is the Faraday constant, and v is the accessible volume of the compartment (174). The sum of the denominator in Eq. 31 extends over all Ca2+ buffers. The time constant τCa is given by
(32) |
with γ representing the relationship between the Ca2+ extrusion (pump) mechanism and [Ca2+]i (Refs. 174, 248; see also a graphical representation of these equations in FIGURE 3). It is readily seen that the time integral of [Ca2+]i(t), the product of amplitude Δ[Ca2+]i and time constant τCa, is independent of buffering (qCa/(2F·v·γ).
4.7.1.2. the rapid buffer approximation.
Most Ca2+ buffers of interest have relatively fast binding kinetics (sect. 2.1), with the exception of PV, a CBP with high expression in skeletal muscle and certain neuron types (sects. 7.2 and 9.2), and the exogenous buffers EGTA and EDTA. For studying cellular properties on the millisecond to second timescale, it is, therefore, often convenient to assume that [Ca2+]i is at equilibrium with all buffers (201, 249–254). This approximation misses, of course, some very interesting kinetic features caused by slow buffers, but it simplifies computations enormously (178). In particular in combination with the LBA, discussed below, some important features of buffering and diffusion can be understood intuitively on the basis of very simple equations, such as Eq. 30–32. But the rapid buffer approximation (RBA) can also be extended by including buffer saturation, which expands its range of applications (202). Furthermore, extensions for the case of buffers with two binding sites have been described (188, 241).
4.7.1.3. the linear buffer approximation.
As stated by Eq. 8, the buffer power β of a given ligand with dissociation constant Kd is nearly constant for [Ca2+]i ≪ Kd. For small [Ca2+]i transients around resting values and in the range of a few hundred nanomolar, neglecting buffer saturation may be acceptable and proves helpful in understanding the changes in [Ca2+]i dynamics induced by the presence of buffers. This implies that changes in [Ca2+]i are proportional to changes in total [Ca2+] (178, 255–257). Furthermore, Ca2+ signals caused by various processes add linearly—as long as the resulting summed response stays well below the Kd of the buffer with the highest affinity. Exploiting these properties, the aggregate signal and expected neurotransmitter release has been calculated for various arrays of Ca2+ channels, which becomes particularly simple in the presence of millimolar concentrations of EGTA (Refs. 255, 256, 258, 259; see also sect. 4.7.2 on local domains). As noted above, the linear buffer approximation (LBA) is reasonable for the small changes of [Ca2+]i that typically occur in neurons. For example, from Eq. 8, assuming a buffer Kd of 0.75 µM, the buffer power at a peak [Ca2+] of 200 nM will be 71% of that at 50 nM resting [Ca2+]i. In contrast, with the same Kd, the buffer power at the very much higher peak of a cardiac Ca2+ transient (1 µM) will be only 24% of that at a resting [Ca2+]i of 100 nM, and the approximation would be problematic (see also FIGURE 3).
The LBA, together with the RBA, leads to simple equations like Eq. 29 for the apparent diffusion coefficient. It was used to describe the diffusional spread in cylindrical (178, 204) and spherical (163) structures. Together with RBA and applied to a single compartment, it leads to a simple description (Eqs. 30–32) of exponentially decaying small [Ca2+]i transients, as frequently observed in neurons.
4.7.2. Local domains.
Many Ca2+-dependent processes are triggered very locally by single-molecule events. Prominent examples are the release of neurotransmitters upon the opening of one or only few voltage-dependent Ca2+ channels and the generation of “sparks” of [Ca2+]i in muscle due to opening of RyR-operated channels in the SR. Numerical simulations have shown that such local domains, often called “microdomains” or “nanodomains,” of elevated [Ca2+]i rise within microseconds, stay elevated as long as Ca2+ channels are open, and decay equally rapidly after channel closure (178, 239, 246, 256, 257, 260–266). Mobile buffers are particularly efficient in shaping such local domains, since diffusion across the small dimensions of these domains is very rapid, such that when buffer binds Ca2+ it is rapidly replaced by free buffer. A particularly simple scenario emerges in the presence of slow buffers, such as EGTA (256, 262, 263) or PV (267). Because of their slow Ca2+ binding and rapid diffusional replacement they can “penetrate” the local domain without significant changes in their free concentration, [B]0. Where [B]0 is high enough, such that its local value does not change appreciably with respect to the bulk [B]0, the buffer constitutes a spatially and temporally uniform sink, leading to a differential equation with a particularly simple solution for Δ[Ca2+](r), the steady-state increment in [Ca2+]i at distance r from the channel mouth (262):
(33) |
where iCa is the single-channel calcium current, F is the Faraday constant, and λ is the length constant of the domain, which is given by
(34) |
Here, kon is the apparent Ca2+ binding rate constant of the buffer (178, 262). The time constant, τ, of rise and decay of the local domain is given by
(35) |
These equations hold for the case when the slow buffer is the dominating one. They do not consider stationary buffers. In the presence of stationary, very fast-binding low-affinity buffers, for which the RBA and the LBA apply, a steady-state solution for free [Ca2+]i very similar to that in the absence of stationary buffers is obtained. The characteristic time constant, however, has to be increased by a factor of 1 + κs (268), where κs is the sum of values for all such stationary buffers. It should be noted that in general diffusion is slowed down in the presence of stationary buffers. Nevertheless, the diffusion coefficient appearing in Eqs. 33 and 34 is that of free Ca2+, since the equations describe the steady state for which fixed buffers are in equilibrium with free Ca2+. Therefore, the rates that represent binding and dissociation of Ca2+ to/from fixed buffers in the respective differential equations cancel each other. Values for λ and τ × (κs + 1) for various concentrations of EGTA and PV, assuming κs = 21, are given in TABLE 4 (248). These are meant to provide an “order-of-magnitude” idea about the extent and dynamics of local domains in the presence of a dominating slow buffer.
Table 4.
The dependence of length constant and time constant of local [Ca2+] domains
Buffer Conc, mM | λEGTA, µm | τEGTA, ms | λPV, µm | τPV, ms |
---|---|---|---|---|
1 | 0.22 | 5.02 | 0.27 | 7.10 |
2 | 0.16 | 2.51 | 0.19 | 3.55 |
5 | 0.10 | 1.00 | 0.12 | 1.42 |
10 | 0.07 | 0.50 | 0.084 | 0.71 |
20 | 0.05 | 0.25 | 0.060 | 0.35 |
Length constant (λ) and time constant (τ) for EGTA and parvalbumin (PV) were calculated, using Eqs. 34 and 35 under the following assumptions: diffusion coefficient of free Ca2+ (DCa) = 220 µm2/s; apparent association rate constants kon,app = 4.38 µM−1·s−1 for EGTA (248) and kon,app = 3.1 µM−1·s−1 for PV at a Mg2+ concentration of 1 mM. For the calculation of kon,app, we assumed 1 binding site per molecule to be relevant for the buildup and decay of the local intracellular calcium concentration ([Ca2+]i) domains with kon = 103 µM−1·s−1 (27). Time constants were further multiplied by a factor of 22 to include the effects of a fast, immobile buffer with a buffering power of 21 (248).
Equation 33 provides intuitive insight: The term represents the solution in the absence of buffers and reflects diffusion of Ca2+ from a point source in homogeneous medium. The effect of the slow buffer at steady state is exclusively contained in the length constant, where it appears as the so-called “buffer product” kon × [B]0. The diffusion coefficient of the slow buffer does not appear in these equations, since it is assumed that diffusion is fast enough to secure a constant concentration of free buffer within the domain. These approximations do not depend on the RBA and LBA but, importantly, on the assumption regarding nondepletion of free slow buffer. Clearly, the slow buffer is not at equilibrium with [Ca2+]i within that domain. Nevertheless, local domains in the presence of sufficiently high concentrations of free slow buffer superimpose linearly as long as their combined Ca2+ load does not deplete free buffer (see also sect. 9.4).
The validity of these conclusions about local domains has been confirmed by solving the underlying equations analytically (269) and by numerical computation (270). In these calculations, several buffer species were included and it was shown that very close to the site of entry Ca2+ ions are first captured by the buffer with the highest buffer product (kon × [B]0). This was ATP, which was assumed to be present at a free concentration of 0.17 mM. However, as a low-affinity buffer, ATP rapidly releases Ca2+ again. During cycles of binding and unbinding from ATP, buffers with lower buffer product can capture Ca2+. If they are of higher affinity, they will retain the Ca2+ ions for longer times, while moving away from the Ca2+ source. In the end, the buffer species with the highest affinity will carry the bulk of Ca2+, according to equilibrium conditions. The resulting spatial profile of the local domain decays multiexponentially, unless the buffer with the highest affinity also has the highest buffer product. Numbers given in TABLE 4 assume that this is the case for EGTA and PV, respectively.
The influence of small organic molecules with low affinity but higher buffer product on the properties of local domains has largely been neglected in the literature so far, despite their important role in shaping these domains and likely consequences for physiological processes (see sect. 9.4 and Ref. 60 on the effect of gluconate on neurotransmitter release).
4.7.3. Diffusion in a cylinder.
Elongated cells or subcellular compartments such as axons, dendrites, cardiac myocytes, and muscle fibers often have narrow diameters such that diffusional gradients along their cross section are small because of rapid equilibration. However, gradients may be very pronounced longitudinally. In this case, the spread of Ca2+ along the longitudinal axis can be conveniently described by simple analytical expressions, Additionally, if both rapid buffer approximation (RBA) and linearity of buffers (LBA) are assumed, the transport of Ca2+ along the cylinder will follow the classical one-dimensional diffusion equation with Dapp according to Eq. 29 (178, 202, 204). If a leak of Ca2+ across the cylinder wall is considered, the same simplifying assumptions (RBA and LBA) lead to an equation analogous to that for propagation of voltage in a leaky cable (268). Using this simplified formalism, Gabso et al. (204) attempted to measure Dapp in largely intact axons of the snail Aplysia. They locally injected, via a microelectrode, boluses of Ca2+ into uniform lengths of axons, which had previously been loaded with defined concentrations of indicator dyes. Analyzing images of Ca2+ spread as a function of position and time after injection, they found that the local [Ca2+]i elevation had a Gaussian shape, which spread and decayed rapidly. According to theory, the square of the half-width of that Gaussian should be proportional to the time since the Ca2+ injection and the apparent diffusion coefficient, Dapp. This allowed evaluation of buffering power and Dapp for various concentrations of the indicator dye. To obtain estimates for the case of an unperturbed cell, the results were extrapolated to zero indicator dye. This yielded values for Dapp (<16 µm2/s) and for κs (<60). Unfortunately, only upper bounds could be provided in this analysis, since the extrapolation depended very much on the measurements with the lowest indicator concentrations, which were very noisy. The finding that Dapp is much smaller than DCa points toward the presence of fixed buffers that retard diffusion. On the other hand, the upper bound (16 µm2/s) is higher than the lowest possible value, DCa/(κs + 1) = 223 µm2/s/61 ≈ 3.7 µm2/s, which would be expected if all endogenous buffers were immobile. Thus, it can be concluded that mobile buffers must also be present.
In the light of the above considerations, we now turn to the role of buffers in specific instances of physiology and pathology. Most of the quantitative work has been carried out in nerve and muscle, but, although there are tissue-specific differences in buffering, the general principles will apply throughout.
5. THE RED BLOOD CELL
Red blood cells differ from all other cell types in that their total Ca2+ content is unmeasurably low. In human red blood cells the Ca2+ content extractable by ionophores and chelator treatments amounts to <1 μmol/(L packed cells) (271). Most mammalian red blood cells contain no organelles, making it simpler to measure their cytoplasmic buffering than is the case in many other cell types. An early study measured Ca2+ binding at relatively high levels of Ca2+ (20 µM to 1.4 mM) in suspensions of freeze-thawed red blood cells. A Kd of ∼0.3 mM was reported (272). Subsequent work measured Ca2+ buffering in lysed red blood cells (thereby preserving any contribution from cellular contents) and found a linear dependence of bound on free Ca2+, over the range 0.1 to 3 mM, with bound and free Ca2+ being approximately equal. When the cell constituents were removed by dialysis, all the bound Ca2+ was lost, suggesting that cell membranes do not contribute appreciably to buffering (273). Because these studies were performed at Ca2+ levels much greater than normal cytoplasmic, higher-affinity buffers would have been saturated and their contribution would have been overlooked.
Buffering was measured in intact red blood cells by controlling intracellular free Ca2+ with an ionophore (A23187) while measuring total intracellular Ca2+ with 45Ca (274). Again, total Ca2+ was proportional to free with a buffer power of only 2–3. A subsequent study (275) measured buffering at [Ca2+]i levels as low as 100 nM and again found a linear (nonsaturable) binding, attributed to binding to hemoglobin and other proteins, with bound Ca2+ equal to ∼35% of free, a very low buffer power. In addition, there was evidence of a saturable buffer with Kd of ∼8 µM and maximum capacity of ∼100 µM, thereby contributing a buffer power of ∼12 at 100 nM [Ca2+]i. A tentative suggestion for the identity of this component was calpromotin, now known as Peroxiredoxin 2, a major antioxidant (276). CaM was also considered, but its concentration is too low to account for more than a small fraction. It should, however, be noted that the saturable buffer was identified by subtracting the nonsaturable component from the total, and, as pointed out by the authors, this limited the precision of the estimation. Importantly, no evidence was found for high-affinity buffers with Kd values below 1 µM. This contrasts with the tissues reviewed in subsequent sections, where the major buffer power is provided by buffers with such low Kds.
The low Ca2+ buffering of the red blood cell has been suggested to be physiologically important as it means that small Ca2+ fluxes can produce large changes of [Ca2+]i. Given the very limited metabolic reserves of this cell, minimizing these fluxes, and the consequent demand for ATP to pump Ca2+ out of the cell, is important for the economy of the cell. Human red cells have a programmed circulatory life span of ∼120 days (277). As the cells deform when traversing capillaries, the mechanosensitive PIEZO1 channels of the cell membrane (278, 279) become transiently activated, allowing brief episodes of Ca2+ influx. The low calcium buffering power of the cells allows this minimal Ca2+ influx to elevate [Ca2+]i sufficiently to activate Ca2+-sensitive K+ channels, leading to progressive KCl and water loss. This explains the physiological mechanism behind the progressive increase in the density of aging red cells in the circulation and highlights the important role of low Ca2+ buffering by minimizing the magnitude and maximizing the speed of the leak-restorative Ca2+ fluxes by the Ca2+ pump during each capillary transit (280–282).
The low buffer power of the red blood cell contrasts with the much higher values (by a factor of 10–100 fold) in the cells and tissues considered in subsequent sections. A major aim of these sections is to identify the substances responsible for this difference of buffering and the physiological roles of the greater buffering.
6. CARDIAC MUSCLE
Each heartbeat is activated by a systolic increase of [Ca2+]i, derived from Ca2+ entry via the L-type Ca2+ channel and by Ca2+ release from the sarcoplasmic reticulum through the ryanodine receptor (RyR). The amplitude of this free Ca2+ transient is of the order of 1 µM and results from an increase of total cytoplasmic [Ca2+] of ∼100 µM. Changes of the size of the Ca2+ transient are the major factor regulating contraction strength (see Ref. 81 for review). For the heart to work as a pump it must also relax during diastole; Ca2+ is lowered back to resting levels via a combination of uptake back into the SR via SERCA and pumping out of the cell (largely via Na/Ca exchange, NCX). Ca2+ buffers will therefore affect both the amplitude and rate of decay of the Ca2+ transient.
6.1. Measurements of Buffer Power in Cardiac Muscle
There are many direct measurements of Ca2+ buffering in intact myocytes that can be compared with the measured concentrations of potential buffers. Initial estimates were obtained from Ca2+ titration of isolated myofilaments indicating that these alone could contribute a buffering power of at least 10 (283). Similar studies on cardiac homogenates gave a value of 70 (284), and work on permeabilized myocytes showed that cytoplasmic Ca2+ buffering could be represented by a buffer with a Kd of 0.42 µM at a concentration of 78 µM as well as a lower-affinity component (285). The higher-affinity component would contribute a buffer power of ∼120 at 0.1 µM [Ca2+]i. Subsequent studies have measured cytoplasmic buffering in intact cells by comparing changes of [Ca2+]i measured with a fluorescent indicator to those of total Ca2+ calculated from the integral of a membrane current. One approach uses the L-type Ca2+ current activated by depolarizing pulses (286, 287). A related technique applies caffeine to release Ca2+ from the SR. The Ca2+ is then pumped out of the cell largely by the electrogenic Na/Ca exchange (NCX), and integrating the NCX current gives a measure of the change of total Ca2+ (254). This method requires correction for that fraction of Ca2+ that is transported by the electroneutral PMCA. In rat ventricular myocytes, the L-type Ca2+ current method provided a Kd of 0.96 µM and a Bmax of 123 µM (287) whereas the NCX method gave 0.49 µM and 149 µM, respectively. Various sources of errors need to be considered when comparing these values. First, the cytoplasmic volume must be measured to convert the measured total change of Ca2+ to a concentration. Second, the accuracy depends on that of the measurement of [Ca2+]i. There are at least two potential problems: 1) Calibrating the measured fluorescence signals in terms of absolute [Ca2+]i requires knowledge of the minimum and maximum fluorescence as well as the Kd of the indicator dye (36), and these values can be affected by the cytoplasmic environment (288–292). Although this is a particular problem for single-wavelength indicators such as the Fluo family, it is not trivial even for ratiometric indicators. Indeed, the Fluo indicators have the advantage that fluorescence is essentially zero in the absence of Ca2+, meaning that calibration only requires knowledge of the maximum fluorescence and Kd. The literature shows a range of values assumed for even the most used indicators (288–292). 2) Accurate calculation of the Kd of the buffer requires measurements over a wide range of [Ca2+]i, certainly much greater than the Kd value. That this is not always possible complicates, for example, establishing whether buffering is cooperative. Indeed, many cardiac buffer curves can be fit reasonably well by a linear regression. It would also be useful to make measurements at values of [Ca2+]i below the normal resting level. These issues should not, however, affect measurement of the buffer power over the range of [Ca2+]i studied, which implies that such measurements are probably most useful for comparing buffering between two conditions. Absolute values need to be treated with caution, a problem when trying to compare with the measured concentrations of cellular CBPs.
[Ca2+]i in the bulk cytoplasm in cardiac muscle is generally in the range 0.1 to 1 µM. The major buffers in cardiac muscle are listed in TABLE 5. This follows from previous compilations (81, 294). For simplicity, we ignore the modest contributions from the sarcolemmal binding sites. The top five rows of TABLE 5 show the faster buffers and the lower two rows the slower (see sect. 6.1.3). At 0.1 µM [Ca2+]i, the total buffer power provided by the faster buffers is predicted to be ∼140, in reasonable agreement with the experimental values reviewed above. Importantly, the bulk of the buffering is provided by immobile buffers, with the mobile buffers (calmodulin, carnosine, and ATP) contributing <10%.
The effects of experimentally altering Ca2+ buffering in cardiac muscle are shown in FIGURE 12 (199). Here, buffering was first increased by adding a photolabile Ca2+ chelator. Buffering was then decreased by photolysis. As expected from FIGURE 7, the fast buffer nitr-5 decreases the amplitude and slows the kinetics of the Ca transient (FIGURE 12A). In contrast (FIGURE 12B), the slow buffer NP-EGTA results in a biphasic decay (cf. FIGURE 8 and FIGURE 9).
6.1.1. Contribution of TnC and SERCA.
As shown in TABLE 5, TnC is expected to be the major buffer at diastolic levels of [Ca2+]i. As discussed in sect. 2.5.1.1, TnC has two classes of Ca2+ binding sites: 1) the regulatory Ca2+ site, which has a Kd of ∼0.6 µM, and 2) a pair of nonspecific “Ca2+/Mg2+” sites with a much lower Kd for both Ca2+ (0.003 µM) and Mg2+ (3.3 µM) (discussed in sect. 6.1.3). At diastolic Ca2+ levels, these sites bind significant amounts of Ca2+ and Mg2+. Experimental studies have investigated the effects of increasing the binding affinity of the regulatory sites of TnC with a single amino acid substitution (L48Q). With an adenoviral technique, ∼20% of the TnC in a rat heart was replaced. As expected, this increased myocyte contraction but did not alter the rate of decay of the [Ca2+]i transient (295), suggesting little effect on buffering. There are two possible explanations. 1) If TnC accounts for slightly less than 50% of total buffering (TABLE 5), then 20% replacement will decrease total buffering by only 10%, resulting in only a small effect on the decay of the [Ca2+]i transient. 2) A decrease of Kd will increase buffering at low and decrease it at high [Ca2+]i (Eq. 14), leading to a mixed effect on relaxation. Although both factors will presumably contribute, the importance of the former is supported by a subsequent study expressing a higher concentration of the same mutation in mice myocytes (50% replacement) resulting in slowing of the decay of the [Ca2+]i transient (296). A recent study has emphasized the importance of TnC in buffering by comparing buffering and Ca2+ handling in myocytes from the right and left ventricle of rat hearts. This showed a lower buffer power in the right ventricle that was associated with, and suggested to result from, lower expression of TnC (297). The other major contribution to buffering is expected to come from binding to SERCA (94). Direct evidence for a role of SERCA in buffering comes from the observation that inhibiting it with thapsigargin decreases buffer power by 16% in mouse ventricular myocytes (298).
6.1.1.1. cooperative buffering.
The cooperative nature of Ca2+ binding to some buffers has not generally been considered in cardiac Ca2+ buffering. SERCA binds Ca2+ cooperatively, and this would be expected to produce a bimodal dependence on [Ca2+]i of the buffer power contributed by fast buffers, with buffer power being low at both low and high [Ca2+]i (TABLE 5). This contrasts with the simple hyperbolic buffer curves shown in the literature, where buffer power decreases with increasing [Ca2+]i (180, 254, 287, 297, 299, 300). It is therefore possible that TABLE 5 overestimates the contribution of cooperative buffers. Equally, the experimental studies may have insufficient accuracy to resolve sigmoidal binding curves. A limitation of these studies is the lack of data below the normal resting [Ca2+]i of ∼100 nM, and it will be important to obtain this. As pointed out in sect. 3.3, despite having only one regulatory site TnC may also show cooperativity due to interactions along the thin filament. Such effects would be expected to be different between isotonic and isometric contractions (sect. 6.2.3) and, given the lack of detailed information, are ignored in TABLE 5.
6.1.2. The role of other fast buffers.
TABLE 5 shows that the buffer power provided by both SERCA and TnC falls as [Ca2+]i increases to micromolar levels. During Ca2+ release, the Ca2+ concentration in the dyads may reach levels of up to 100 µM (55). Even at 10 µM [Ca2+]i, the combined contribution of TnC and SERCA is predicted to produce a buffer power of only 0.5. This raises the question as to what, if anything, provides buffering at this high [Ca2+]i. TABLE 5 suggests that total fast buffer power is ∼10, contributed largely by CaM, ATP, and carnosine. In other words, these buffers, which make only a modest fractional contribution to buffering at diastolic and systolic levels, may be important during Ca2+ release. Not only will they buffer Ca2+, but ATP and carnosine will accelerate its diffusion away from the release sites (TABLE 3). However, as mentioned in sect. 2.4.1, there is controversy as to whether carnosine buffers Ca2+ appreciably under physiological conditions. More work is required to investigate carnosine’s buffering role and the extent to which this may be altered, for example, in heart failure and other disease.
6.1.3. The role of “slow” buffers.
Above, we have ignored the slower buffers, including the nonspecific or Ca2+/Mg2+ sites on TnC as well as Ca2+ binding to myosin. The nonspecific sites on TnC are present at twice the concentration of the regulatory sites and have a high affinity for Ca2+ with an apparent Kd for Ca2+ (in the presence of 1 mM Mg2+) of ∼100 nM (78, 79, 301). This means that they are ∼50% occupied by Ca2+ and contribute a buffer power of ∼350 at a diastolic level of Ca2+ of 100 nM. However, the relevance of this buffer power will depend on the kinetics and, specifically, on whether appreciable binding occurs during a [Ca2+]i transient. This is illustrated by the difference between the instantaneous and steady-state values of buffer power in TABLE 5. Two experimental studies have provided very different measurements as exemplified by values for koff of 0.33 s−1 (78) and 0.032 s−1 (79). This 10-fold difference in Ca2+ unbinding kinetics results in different predictions for the time course of Ca2+ occupancy of the sites as shown in FIGURE 13 (78, 79, 301). The faster kinetics predict some beat-to-beat change of bound Ca2+ and therefore buffering during the Ca2+ transient. In contrast, the slower kinetics result in essentially no beat-to-beat changes but a gradual increase of bound Ca2+ at higher frequencies. Either kinetic may lead to significant problems in the interpretation of experimental studies. For example, maneuvers that increase the time-averaged level of [Ca2+]i will result in a slow loading of these sites (82). This will occur when stimulation is resumed after a quiescent period (FIGURE 13) or during the application of many agents that increase systolic [Ca2+]i. On cessation of stimulation, the Ca2+ gradually dissociating from these slow buffers is expected to keep both cytoplasmic and SR Ca2+ concentrations elevated for a period. The subsequent gradual decrease of [Ca2+]i might erroneously be ascribed to movements of Ca2+ across SR or surface membranes. We are unaware of work investigating this possibility. The other major Ca2+ binding site on the myofilaments, which can also bind Mg2+, is on myosin, with a high apparent Kd for Ca2+ of ∼9 µM. As shown in FIGURE 13, Ca2+ binding to myosin is also slow, and the high Kd predicts that the amount of Ca2+ bound, even during repetitive stimulation, is modest and is unlikely to make a major contribution to total buffer power.
FIGURE 13.
Simulation of Ca2+ binding to slow sites in cardiac muscle. The model simulates the effects of cytoplasmic Ca2+ transients. Traces show (from top to bottom) intracellular calcium concentration ([Ca2+]i); Ca2+ bound to the regulatory site of Troponin-C (TnC); Ca2+ bound to the nonspecific sites on TnC, comparing rate constants taken from Refs. 78, 301 (solid) and Ref. 79 (dashed); Ca2+ binding to myosin. Note the different vertical scales for TnC and myosin binding. The model shows the effect of (from left to right) a single stimulus; 1-Hz stimulation; and 2-Hz stimulation. The free Ca2+ transient was modeled with an instantaneous increase from 0.1 to 0.8 µM followed by an exponential decay (time constant 0.2 s).
6.1.4. SR buffers.
The above section has concentrated on the cytoplasmic buffers. The major Ca2+ buffer in the SR is calsequestrin, CSQ2. As well as acting as a buffer, it also interacts with triadin and junctin in regulating the RyR (302). The Ca2+ buffering properties of the SR have been measured in rabbit ventricular myocytes by measuring free SR [Ca2+] with Fluo-5N and estimating changes of total SR [Ca2+] from the corresponding changes of total cytoplasmic Ca2+ (as assessed from changes of [Ca2+]i and cytoplasmic buffering). The buffering could be represented by a single buffer present at a concentration of 2.5 mM with a Kd of ∼0.6 mM (Ref. 303; see also Ref. 293). Even at very low [Ca2+], this would give a buffer power of only ∼4. With a loaded SR (typical free [Ca2+] of ∼1 mM) there will be 1.6 mM bound Ca2+. Two points are worth emphasizing: 1) The buffer power of the SR is much less than that of the cytoplasm. 2) The low buffer power means that complete removal of the major buffer (CSQ2) only decreases total Ca2+ by ∼60% and significant Ca2+ release from the SR can occur in the absence of CSQ2. Indeed, the CSQ2 knockout (KO) mouse shows normal cardiac contractility, and the loss of CSQ2 is compensated for by an increase of SR volume (304). As discussed in sects. 3.3.1 and 7.5, in skeletal muscle the binding of Ca2+ is cooperative and promoted by polymerization of CSQ2. In contrast, there is no evidence for cooperativity in the cardiac measurements (303). Although this may simply reflect a lack of sufficiently precise data, it is also possible that the low concentration of CSQ2 decreases the probability of polymerization. Further work is required to look for such cooperativity. Increasing the amount of CSQ2 prolongs Ca2+ release from the SR, an effect attributed to maintaining free SR content (305). Some mutations have been shown to decrease Ca2+ binding capacity (306). A decrease in CSQ2 concentration or mutations results in abnormal Ca2+ release from the SR, and this is one of the causes of the inherited arrhythmia syndrome CPVT (catecholaminergic polymorphic ventricular tachycardia) (307). At least in transgenic mouse models, removal of CSQ2 leads to a compensatory increase of another ER/SR Ca2+-binding protein, calreticulin (308). Calsequestrin is discussed more extensively in the context of skeletal muscle (sect. 7.5).
6.2. Factors Affecting Cardiac Calcium Buffering
Cardiac contraction is controlled by the amplitude of the systolic Ca2+ transient, which is generally assumed to be regulated by the size of the Ca2+ fluxes into and out of the cytoplasm. An important, and unresolved, issue is the extent to which changes of the amplitude of the [Ca2+]i transient seen in both physiology and disease may also result from changes of Ca2+ buffering as opposed to those of Ca2+ fluxes. In other words, is Ca2+ buffering constant over time? We now consider some of the factors that may alter buffering. These include changes of [Ca2+]i as well as of the properties of the buffers.
6.2.1. Calcium concentration.
As discussed above (sect. 3.1) and also previously (82, 309, 310), for noncooperative buffers the available buffer power is affected by the level of [Ca2+]i, tending to decrease as [Ca2+]i increases (see TABLE 5). This will have two consequences: 1) An increase of diastolic [Ca2+]i will decrease buffer power. This raises the possibility that the positive inotropic effects of maneuvers that increase diastolic [Ca2+]i such as increased stimulation rate (311–313) and the application of cardiac glycosides (314) include a contribution from decreased Ca2+ buffering. Direct experimental work is required to assess this. The decreased buffer power at elevated [Ca2+]i also complicates comparing buffer power between different animals, for example in the context of heart failure versus control. It is important to exclude the possibility that changes of buffer power are simply a consequence of differences of diastolic [Ca2+]i as opposed to alterations of the buffer properties. This is not helped by the relative paucity of absolute measurements of diastolic [Ca2+]i in different species (for review see Ref. 315). 2) More generally, Ca2+ concentration-dependent changes in Ca2+ buffer power are predicted to contribute to shaping the time course of [Ca2+]i transients as suggested in sect. 4.3 and FIGURE 10 and FIGURE 11. For example, the fact that the rate constant of decay of the Ca2+ transient increases with the amplitude of the transient is at least in part due to decreased buffering at increased [Ca2+]i (316). Another example is the initial rapid phase of the decay of the [Ca2+]i transient, which can be explained by the decreased buffer power at elevated [Ca2+]i since a given rate of total Ca2+ removal by SERCA produces a larger fall of [Ca2+]i (197).
6.2.2. Phosphorylation of buffers.
A further question is whether the properties of cardiac cytoplasmic Ca2+ buffers can vary at fixed [Ca2+]i. The apparent Kds for Ca2+ of the two major buffers (SERCA and TnC) are regulated by phosphorylation [of phospholamban and troponin I (TnI), respectively], raising the possibility that this will also affect buffering. An experimental study, however, found no effect of beta adrenergic stimulation on Ca2+ buffering (298), and it was suggested that this resulted from the fact that, while phosphorylation of phospholamban increases the Ca2+ affinity of SERCA phosphorylation of troponin decreases its affinity, with the net result being no change. It should, however, be noted that previous work has not explicitly considered that the effects of changes of buffer Kd are expected to depend on the level of [Ca2+]i (sect. 3.2). It is possible that the lack of change of buffer power could reflect a smearing of opposite effects above and below the value of [Ca2+]i where changing Kd has no effect. Precise measurements of buffer power over a broader range of [Ca2+]i are required.
6.2.3. Effects of muscle length.
One special feature of TnC is that its location on the myofilaments makes it potentially sensitive to force development, with an increase of force increasing the affinity of Ca2+ binding. Direct measurements of [Ca2+]i in cat papillary muscles showed that rapid stretches and releases resulted in transient decreases and increases of [Ca2+]i consistent with Ca2+ being taken up by or released from TnC, respectively (317, 318). A maintained increase of muscle length was found to accelerate the initial rate of decay of the [Ca2+]i transient and slow down the final phase, with both effects attributed to increased affinity of Ca2+ binding to troponin (319, 320). This is consistent with the analysis in sect. 3.2 and the simulations of FIGURE 10 illustrating that increased buffer affinity decreases buffering at higher [Ca2+]i but increases it at lower [Ca2+]i. The former effect will accelerate the initial decay and the latter prolong the final phase (see also Ref. 321). Stretch has previously been reported to increase Ca2+ release from the SR (322). A recent study showed that the rise of [Ca2+]i was increased by the actomyosin inhibitor blebbistatin (sect. 6.2.5), suggesting that, in the absence of blebbistatin, stretch increases the binding affinity of TnC, thereby attenuating the rise of [Ca2+]i (323).
6.2.4. Effects of pH on cardiac Ca2+ buffering.
As mentioned in sect. 2.5.1.2, changes of pH may significantly affect Ca2+ binding to buffers, either because of direct competition between protons and Ca2+ for binding sites or via effects on the tertiary structure of the buffers. An example is the decrease of Ca2+ binding to troponin produced by acidification (324). This interaction provides a potential mechanism linking changes of intracellular pH to those of [Ca2+]i. It is important to consider the relative magnitude of the effects on intracellular pH and [Ca2+]i. For example, a displacement of 20 µM total Ca2+ from TnC and its one-to-one replacement by protons will elevate [Ca2+]i by 200 nM (given a Ca2+ buffer power of 100). In contrast, the typical intracellular pH buffer power is 30 mM per pH unit (325), so the absorption of 20 µM protons would be expected to change intracellular pH by 0.02/30 ≈ 0.0007 pH units. Although the change of [Ca2+]i would be easily measurable, that of pH would not. An acid-induced increase of [Ca2+]i has been measured in cardiac muscle (93, 194). However, as pointed out previously (82, 326) and in sect. 4.1, buffers cannot change the steady-state level of [Ca2+]i and one would expect that the Ca2+ ions released would be pumped out of the cell, thereby restoring [Ca2+]i. In this context, it is worth noting that effects of protons on various membrane channels including acid-sensing ion channels (ASICs) and TRPV channels have been shown to contribute to changes of [Ca2+]i in rat ventricular myocytes (327) and that protons also affect NCX activity (328). A further complication arises from the finding that changes of intracellular pH of the order of 0.4 pH units do not produce measurable changes of buffer power (329). Again, this may be a consequence of the fact that the direction of the effect of a given change of Kd on buffer power is opposite at high and low [Ca2+]i (see sect. 3.2). A resolution of this issue will require further studies measuring buffer power as a function of both [Ca2+]i and pH.
6.2.5. Effects of modulators of actin-myosin interactions.
Given the importance of TnC to buffering, it would not be surprising if modulating Ca2+ binding directly or via interaction with other sarcomeric proteins affected buffering. Compounds such as EMD57033 are contractile sensitizers that increase TnC’s apparent affinity for Ca2+. In mouse ventricular myocytes, EMD57033 had no effect on Bmax but decreased the buffer Kd (180). A recent study investigated its effects on Ca2+ buffering in cardiac myocytes derived from human induced pluripotent stem cells (iPSCs). Again, there was no effect on Bmax, and the Kd decreased from ∼0.47 to 0.25 µM (300). This was accompanied by a slowing of the decay of the [Ca2+]i transient. The authors also calculated the rate of decrease of total Ca2+ from the measured buffering, which was unaffected by the drug, consistent with the conclusion that the slowing of decay of free Ca2+ is due to buffering. Strictly speaking (see sect. 3.2), one would expect the decrease of Kd to only increase buffer power at [Ca2+]i below the geometric mean of the Kds (here 0.34 µM), with decreased buffering and therefore a faster decay of [Ca2+]i above this level. It would therefore be interesting to study the effect of EMD57033 over a wider range of [Ca2+]i. The effects of such contractile sensitizers are of more than academic interest, as members of this group such as levosimendan (330) and omecamtiv mecarbil (331) have been developed for use in heart failure.
In contrast, other drugs decrease actin-myosin interactions, and one example is the myosin ATPase activity inhibitor blebbistatin (332). This can decrease the apparent affinity for Ca2+ to activate force (333, 334). One might therefore expect that it would lower the affinity for Ca2+ buffering by TnC, but we are unaware of any direct measurements. Whereas blebbistatin is simply an experimental tool, another inhibitor, mavacamten (see Ref. 335 for review), has been developed as a treatment for hypertrophic cardiomyopathy (336). As mentioned in sect. 6.4, mavacamtem decreases the affinity for Ca2+ activating contraction (337), and it is therefore also important to characterize its effects on Ca2+ buffering.
6.3. Ca2+ Buffers and Ca2+ Diffusion in Cardiac Muscle
As discussed in sect. 4.5, mobile Ca2+ buffers accelerate the diffusion of Ca2+. The major cardiac Ca2+ buffers, SERCA and TnC, are immobile, and the best-characterized mobile buffers are CaM and ATP. As we have seen above, the contribution of CaM to diffusion of Ca2+ is negligible, particularly since the majority of CaM is bound. With the values of TABLE 3, the flux of Ca2+ carried by Ca-ATP will be ∼1.4 times that of free Ca2+ over a wide range of [Ca2+]i, thereby making a significant contribution. It is also worth noting that there is up to 20 mM carnosine and other histidyl dipeptides in cardiac muscle (211). As mentioned in sect. 2.4.1, there is controversy as to their degree of Ca2+ binding, but, if significant, these will also contribute greatly to Ca2+ diffusion.
What effect will such Ca2+ diffusion mediated by mobile buffers have? As discussed below for skeletal muscle (sect. 7.6), diffusion is required for Ca2+ ions to move from the sarcoplasmic reticulum release sites to the TnC on the myofilaments (34, 338), and therefore the mobile buffers, in particular ATP, will accelerate activation of the myofilaments. We are unaware of any experimental evidence in cardiac muscle, and it will be important to obtain this. This effect of mobile buffers will promote cardiac contractility. Set against this, however, is the fact that excitation-contraction coupling depends on local Ca2+ release, observed as Ca2+ sparks. Larger [Ca2+]i transients result from the summation of closely spaced individual Ca2+ sparks (339). These sparks are spatially independent, and this is important for controlling cardiac contractility in a stable way (340). Increasing Ca2+ diffusion will potentially increase the spatial extent of Ca2+ sparks and remove this independence. One consequence of loss of spatial independence of sparks is the generation of propagating Ca2+ waves. These are not seen normally but occur under conditions where Ca2+ spark frequency (341) and leak from the SR are increased, because of either an increase of SR Ca2+ content or alterations in the properties of the RyR (for review see Ref. 342). Modeling has suggested that such waves can propagate by a “fire-diffuse-fire” mechanism. Importantly, the speed of propagation of the waves is proportional to the apparent diffusion constant for Ca2+ (343) and would therefore be expected to be increased by mobile buffers. In particular, at least for IP3-induced Ca2+ waves, for a given buffer concentration calculations suggest that the more mobile the buffer, the faster the wave propagation velocity (344). Another modeling study predicted that Ca2+ waves are more likely to appear at low cytoplasmic buffer concentrations and that, at a given buffer concentration, the higher the Kd for the buffer the more likely are waves (345). At least at low [Ca2+]i, high Kd and low buffer concentration will both decrease buffer power, suggesting that increased buffering decreases wave occurrence. It would be useful to extend this modeling to consider mobile buffers.
6.4. Ca2+ Buffering in Cardiac Disease
Several studies have examined whether Ca2+ buffering changes in cardiac disease. One example is atrial fibrillation, the most common cardiac arrhythmia (346). When this was simulated with rapid electrical pacing in rabbits, a large (2- to 3-fold) increase of buffer power was observed that was attributed to decreased phosphorylation of TnI decreasing the Kd for Ca2+ binding to TnC (299). This was accompanied by (and suggested to cause) a failure of the Ca2+ release to propagate from the periphery to the interior of the cell. A threefold decrease of Kd would increase buffer power by the same factor at low [Ca2+]i (sect. 3.1). At higher [Ca2+]i, however, buffer power would increase by less and would be expected to decrease above a [Ca2+]i given by the geometric mean of the Kds in control and rapid pacing. This effect on propagation could be mimicked by the addition of BAPTA to increase buffering. Addition of EGTA similarly stops propagation in normal feline atrial myocytes (347). It should be noted that the waves occur because atrial myocytes from small animal species contain relatively few transverse tubules and therefore Ca2+ release initially occurs at the periphery where the SR and surface membrane are in contact (348). However, atrial myocytes from larger species, including human, have a more complete network of t tubules, and Ca2+ release also occurs in the center of the cell, making excitation-contraction coupling less dependent on Ca2+ waves (349–351). This may impact on the broader relevance of this decreased wave propagation mechanism in atrial fibrillation. A decrease of buffer power was seen in a sheep model of atrial fibrillation, where it was suggested to facilitate the spread of Ca2+ release and thereby maintain fibrillation (352). Also in sheep atrium, rapid pacing to induce heart failure decreased buffer power (353). In contrast, a recent study found no difference in Ca2+ buffering between myocytes taken from human atrium with or without postoperative atrial fibrillation (354). The origin of these disparate findings is unclear and may be in part due to the different species and models studied. Further work is required to resolve this and also the possibility that changes of diastolic [Ca2+]i may have contributed.
Many studies have shown that heart failure is associated with increased Ca2+ sensitivity of the contractile machinery, possibly as a result of decreased phosphorylation of troponin I (334, 355–357) or myosin binding protein C (358, 359). It is therefore surprising that direct measurements have found that ventricular Ca2+ buffering is unaffected by heart failure (360, 361). Two factors need to be considered: 1) Increased Ca2+ sensitivity does not necessarily imply increased affinity for Ca2+, as it is possible that events downstream from changes in Ca2+ binding are augmented, giving an apparent increase of Ca2+ affinity. 2) An increase of Ca2+ affinity of buffers will increase buffering at low [Ca2+]i but decrease it at higher (sect. 3.2). Measuring buffer power over the whole range of [Ca2+]i will average out these changes and result in no apparent change of buffer power. More detailed measurements are therefore required.
In humans, hypertrophic cardiomyopathy (HCM) frequently results from mutations in thin filament proteins including troponin and tropomyosin (362). Mutations in these and other proteins are also associated with dilated cardiomyopathy (DCM) (363). Many of the HCM mutations have been shown to increase the affinity of Ca2+ binding, whereas those producing DCM decrease it (364). Troponin mutations associated with HCM slow the decay and decrease the amplitude of the [Ca2+]i transient (180, 365). These effects could result from an increase of buffer power due to the increased affinity, although, as pointed out previously (180), buffer power would only be expected to increase at low [Ca2+]i (in the range of [Ca2+]i below the geometric mean of the Kds). The myosin inhibitor mavacamtem decreased the apparent Ca2+ affinity of such HCM mutations and reversed the slowing of the decay of the [Ca2+]i transient (337), and it is therefore possible that these effects also result from normalization of buffering. However, buffering was not measured. The increase of buffering and therefore of total Ca2+ in HCM has been suggested to increase Ca2+ loading of the SR and thence increased arrhythmogenic Ca2+ release during diastole, partly accounting for the increased arrhythmia burden (180). The link between increased Ca2+ binding and arrhythmias is supported by the fact that those mutations that have the greatest effect on Ca2+ binding are also the most arrhythmogenic (333). A modeling study has, however, questioned whether increasing Ca2+ sensitivity does promote arrhythmias in human HCM (366). Increased buffering also leads to shortening and triangulation of the AP (333), findings that were reproduced in both work on cardiac myocytes derived from human iPSCs (367) and computer simulations (368). It was suggested that the AP shortening results from the increased Ca2+ buffering decreasing the amplitude of the Ca2+ transient and thence the inward NCX current, which contributes to maintaining the normal AP plateau (367). Such electrophysiological alterations can also be arrhythmogenic. A significant factor that predisposes to arrhythmias is that of electrical alternans (a condition in which the AP duration alternates on a beat-to-beat basis). This is often accompanied by alternation in the amplitude of contraction and the underlying Ca2+ transient (for recent review see Ref. 369) and is a particular issue when the alternation is inhomogeneous throughout the ventricle (370). It is therefore interesting to note that HCM mutants can increase the probability of such electrical alternans (367). Modeling has also indicated that changes of Ca2+ affinity would be expected to affect the likelihood of alternans (371), with an increase of Ca2+ sensitivity increasing alternans in the atrium (368).
The study referred to in sect. 6.2.5 on iPSC-derived cardiomyocytes compared control cells with those derived from patients with a troponin T mutation leading to DCM. The mutation was associated with increased affinity of Ca2+ buffering (Kd = 0.32 µM in mutant vs. 0.40 µM in control), with no effect on maximum buffer power (300). It is worth noting (see sect. 3.2) that with these values of Kd buffer power will only be increased in a range of [Ca2+]i below ∼0.35 µM. At higher [Ca2+]i, the mutant troponin will have a lower buffer power. The decreased buffer Kd was accompanied by increased occurrence of alternans, and this link was reinforced by showing that the myofilament Ca2+ sensitizer EMD57033 increased both buffering and alternans. This contrasts with the suggestion (above) that DCM mutations have lower affinity. It may be accounted for by the decreased phosphorylation of troponin I leading to increased affinity of Ca binding found in iPSC-derived cardiomyocytes (372). Nevertheless, this change of buffering was associated with increased occurrence of alternans, suggesting a link between altered Ca2+ handling and the arrhythmias seen in DCM.
7. SKELETAL MUSCLE
The intracellular Ca2+ signals in skeletal muscle control body movements lasting from <1 s to many minutes. Excitation-contraction (EC) coupling is the process whereby an action potential (AP) lasting only a few milliseconds generates a twitch with a duration of hundreds of milliseconds. Most movements are initiated by a short volley of APs (20–50 Hz) in a motor neuron innervating a group of skeletal muscle fibers, resulting in a train of skeletal muscle APs and tetanic contraction (373). There are two main subtypes of skeletal muscle fibers, fast and slow twitch. This distinction is based on the kinetics of contraction, and different muscles have predominantly one or the other fiber type. Muscles involved with posture are generally slow twitch and are more energetically efficient than the predominantly fast twitch involved in rapid limb movements (374, 375). A skeletal muscle AP triggers a pulse of Ca2+ release from the SR of ∼1-ms duration at rates of 200 µM/ms (34, 376) (for review see Ref. 84). SERCA-mediated Ca2+ uptake then returns the released Ca2+ back to the SR over ∼100 ms. The timescale of the release event is comparable to the half-time of Ca2+ binding to TnC, ∼1.5 ms (84), and therefore the Ca2+-TnC interaction does not reach equilibrium during a twitch (see also FIGURE 14 and Ref. 377). As demonstrated by [Ca2+]i measurements and computational models, the kinetics and relative affinities of the cytoplasmic buffers determine the rapid and sustained contractile responses to the [Ca2+]i signal (83, 84). It is important to note that the levels of [Ca2+]i reached in skeletal muscle [up to ∼20 µM (84)] are considerably greater than those in cardiac muscle.
FIGURE 14.
The role of buffers in skeletal muscle. A: changes of free and bound Ca2+ during a twitch in a mouse fast-twitch fiber. Records show (from top to bottom) rate of release of Ca2+ from sarcoplasmic reticulum (SR); changes of total calcium concentration (Δ[CaT]); [Ca-parvalbumin (PV)]; [Ca-Troponin-C (TnC)]; [Ca-ATP]; Ca2+ bound to the indicator (furaptra) (CaD); and intracellular [Ca2+] ([Ca2+]i). Note that [Ca2+]i decays more quickly than total; [Ca-TnC] falls while [Ca-PV] is still increasing. B: changes during tetanic stimulation. Traces show (from top to bottom) rate of release of Ca2+; total [Ca2+]; [Ca-TnC]; and [Ca2+]i. Note that the release flux and increase of total [Ca2+] on the second stimulus is much less than on the first, but the change of [Ca2+]i is almost comparable. Figure taken from Ref. 377, with permission from the Journal of Physiology.
7.1. Skeletal Muscle Ca2+ Buffers
Several methods have been used to obtain direct measurements of Ca2+ buffering in skeletal muscle. An early approach used frog skeletal muscle fibers with the ends cut to allow control of the intracellular environment. Buffer power was increased by raising the concentration of the Ca2+ indicator Antipyrylazo III (APIII) (378). As expected, the higher the concentration of indicator, the slower the decay of [Ca2+]i. Extrapolating this relationship to zero indicator suggested that ∼25% of the cytoplasmic Ca2+ was free, with the remainder bound to endogenous buffers, giving a rather low value of ∼4 for buffer power. The authors pointed out that the accuracy of this method was limited by the accuracy of the estimate of the Ca2+ affinity of the indicator, although it is not clear that this could reconcile this low buffer power with the higher ones discussed below. In later work using the same technique, a higher buffer power of 16 was obtained (379), and this was suggested to have a significant contribution from TnC. Another study (on intact fibers) used the Ca2+ indicator Arsenazo III at high concentrations such that the bulk of the Ca2+ release from the SR would be bound to the indicator (225). This allowed a minimum estimate of the Ca2+ release from the SR of ∼100 µM. With a lower indicator concentration, the rise of free [Ca2+]i was found to be ∼4 µM. This would suggest a buffer power of ∼25, but, as mentioned, the estimate of SR content is a minimum estimate so that the actual buffer power could have been up to ∼37.
Most research into the effects of Ca2+ buffers on Ca2+ signaling and contraction in skeletal muscle has used concentrations and affinities of the buffers found in biochemical studies (e.g., Ref. 78). The main intracellular Ca2+ buffers for fast- and slow-twitch fibers are listed in TABLE 6. Such lists have been used in studies designed to investigate the interrelationship between Ca2+ fluxes from the sarcoplasmic reticulum and the resulting contraction (83, 84, 225, 376). As well as giving information about the steady-state contributions to buffer power of the various buffers, TABLE 6 also gives values for the average buffer power on increasing [Ca2+]i from a resting value of 50 nM to either 2.0 or 20 µM. Two values are given, “instantaneous” and “steady state.” The former assumes that the fast buffers such as TnC and SERCA will have reached equilibrium with [Ca2+]I, whereas the latter (higher) value assumes that all buffers are equilibrated. Note that the contribution of the slower buffers will be much greater during a tetanus compared to a single twitch.
The concentrations of the Ca2+ binding sites on the major fast buffers (TnC and SERCA) are greater than in cardiac muscle (TABLE 5). In contrast, the buffer powers reviewed above in skeletal muscle are lower than in cardiac muscle. This is partly a consequence of the much higher levels of [Ca2+]i in skeletal muscle and the consequent buffer saturation. For example, TABLE 6 shows that, in fast-twitch fibers, the buffer contribution produced by the sum of SERCA and the regulatory sites on TnC is 455 at 50 nM [Ca2+]i but only 0.7 at 20 µM [Ca2+]i. Another study, on frog skeletal muscle, used depolarizing pulses to activate the L-type Ca2+ current. Ca2+ release from the SR was inhibited with EGTA so that the observed increase of [Ca2+]i could be assumed to originate from the Ca2+ current. The change of total Ca2+ (calculated from the integral of the current) was then compared with the measured free [Ca2+]i. With this approach, a high buffer power of ∼150–190 was calculated (380). This study was performed over a lower range of [Ca2+]i (<1 µM) than occurs during twitches and tetani, and this (TABLE 6) would contribute to a larger buffer power. Two complications should be noted. 1) There will be a contribution of ∼14 from the Ca2+ indicator (again APIII), and 2) in these experiments the muscle also contained 1 mM EGTA, which will make a major contribution to buffering. As mentioned above, there is still controversy about the buffer role of carnosine and other HDPs under physiological conditions. As previously reviewed (381), carnosine, the major member of the HDP family, has many biological roles including acting as an antioxidant and intracellular pH buffer. In skeletal muscle, it sensitizes contraction to activation by Ca2+ (382) but may also have a role as a mobile calcium buffer analogous to that of ATP (338).
7.2. The Role of Parvalbumin
The cytoplasmic CBP parvalbumin is an example of a pure buffer: a protein with apparently no role in skeletal muscle other than to shape the intracellular Ca2+ signal and thereby influence function (61). It also plays an important role as a buffer in many neurons (see sect. 9.2). The PVs are a group of small acidic proteins (MW 11,000–12,000) with three EF-hand motifs, two of which can bind Ca2+ at cytoplasmic Ca2+ concentrations There are two major forms, α-PV and β-PV, arising from separate genes; the PVβ group includes the protein oncomodulin (see Ref. 383 for review). In the α form of PV (present in mammalian skeletal muscle), the two EF-hand motifs have similar binding affinities, are Ca2+/Mg2+ sites, and do not show significant cooperativity (29, 83, 384). PV is present at higher concentrations in smaller compared with larger mammals and, in a given species, in fast- compared with slow-twitch fibers (385). The conversion by high-intensity exercise of fast- to slow-twitch fibers is accompanied by a decrease of PV concentration (386). In humans, only very low levels of PV are expressed in both fiber types, and it is only found in the intrafusal fibers (muscle spindles); the reason for this difference is not known (385). Increase of PV concentration by direct gene transfer accelerates mechanical relaxation (387).
The EF hands of PV bind Mg2+ ions in competition with Ca2+. The dissociation constant of Mg2+ is ∼104 times greater than that for Ca2+, but since cytoplasmic free Mg2+ levels are ∼104 times higher than Ca2+, PV binds comparable amounts of the two ions, leaving only a small amount of PV unbound even in a relaxed muscle. Using the values of TABLE 6 for fast-twitch fibers, at rest ([Ca2+]i = 0.05 µM) 68% of PV will have Mg2+ bound, 26% is Ca2+ bound, and 6% remains unbound. This unbound fraction will contribute to rapid buffering; the time course of the larger subsequent buffering by Ca2+ binding to the Mg2+ sites will depend on the rate constant of dissociation of Mg2+ [koff,Mg = 3 s−1 (34)] as the subsequent Ca2+ association rate constant is rapid [kon,Ca = 42 µM−1·s−1, equivalent to a pseudo-first-order rate constant of 42 s−1 at 1 µM Ca2+ (83)]. Despite being a slow buffer, the buffer capacity of PV is significant, particularly in the case of some fast-twitch muscles. This is illustrated in TABLE 6, in the form of the calculation of an “instantaneous” (in) versus “steady-state” (ss) buffer power for mouse skeletal muscle fibers.
The role of PV in mechanical relaxation depends on the fact that it binds Ca2+ slowly via the Mg2+ exchange mechanism such that binding continues when [Ca2+]i decreases and Ca2+ dissociates from TnC (87). The relaxation of [Ca2+]i and contraction therefore reflects a sequential mechanism: Ca2+ is initially bound to TnC and is then taken up by PV before being removed by SERCA into the SR (see Ref. 388 for review). Binding of Ca2+ to PV allows [Ca2+]i to decrease quickly at the end of the SR release phase as the slow displacement of Mg2+ from PV provides a buffer in addition to SR Ca2+ uptake and therefore accelerates the decrease in cytoplasmic Ca2+ (83, 376). As shown in FIGURE 14A, as Ca2+ dissociates from TnC, it binds to PV (377). Under experimental conditions, PV can promote full mechanical relaxation, albeit at a slow rate, even when SERCA is virtually completely inhibited (389). This can also be seen in the simulation of FIGURE 9D. This additional boost to Ca2+ removal will only occur at the end of a twitch or brief tetanus; longer tetani will cause increased Ca2+ binding to PV as it displaces bound Mg2+, leading to saturation of PV (83, 390, 391, 564). This explains why PV is expressed at higher concentrations in fast-twitch muscle in smaller mammals (see sect. 7.3). For example, in rat skeletal muscle the faster relaxation of a twitch compared to that after a long tetanus is more prominent in fast- compared with slow-twitch fibers (390), consistent with the presence of PV in the former but not the latter (385).
Adding EGTA, a slow buffer, can also accelerate the rate of [Ca2+]i decay (392). A subsequent study characterized [Ca2+]i transients accompanying tetani in mouse fast- and slow-twitch muscle fibers. Slow-twitch muscle fibers show a stimulus-to-stimulus increase of [Ca2+]i during the initial phase of the tetanus. Such a staircase is absent in fast-twitch fibers. Addition of the exogenous buffer EGTA to a slow-twitch fiber virtually abolished this staircase and made the [Ca2+]i time course resemble that of a fast-twitch fiber (393). Although alternative mechanisms exist to explain this data, these findings are consistent with the role of PV in shaping the fast-twitch Ca2+ transient, and measurements and simulations confirm this mechanism (377).
The high concentration of PV in fast-twitch fibers of small mammals enhances relaxation in an energetically efficient manner. A decrease of [Ca2+]i produced by binding to PV does not consume ATP (see sect. 4.2). ATP is required subsequently as SERCA eventually lowers [Ca2+]i, leading to Ca2+ dissociation from PV. However, the peak SERCA rate is much lower than would be the case in the absence of PV (FIGURE 9 and Ref. 394). This role of PV comes at the price of a higher buffer power at resting [Ca2+]i (0.05 µM; TABLE 6), which would in part explain the higher Ca2+-storage capacity of the SR in fast-twitch fibers as larger Ca2+ release is required to achieve comparable loading of the contractile proteins (225). This distinction between rapid and slow effects of PV is also relevant to neurons (see sects. 9.4–9.6). Related to this, the lower temperature dependence of PV kinetics compared to those of SERCA increases the relative contribution of PV to relaxation at lower temperatures (388, 395). As pointed out in sect. 7.7, modification of PV to increase the affinity for Mg2+ would decrease the concentration of free PV and therefore the buffering during Ca2+ release. It would, however, also decrease the ability to buffer Ca2+ during relaxation, and it is possible that the existing values represent an optimum balance between effects during Ca2+ release and relaxation. This is supported by recent data from an inducible mouse PV knockout that showed increased peak twitch and tetanic Ca2+ acutely in fast-twitch fibers (396) such that short-term fatigue was reduced, despite the slower rate of the early phase of decay of the intracellular Ca2+ signal at the end of a tetanus while longer-term fatigue was unaffected. This work contrasts with earlier work that showed greater long-term fatigue in a constitutive PV knockout model and parallel increases in mitochondrial volume and elevated resting [Ca2+]i (397).
There are several significant cytoplasmic Ca2+/Mg2+ buffers in skeletal muscle, but PV shows the greatest difference between the instantaneous and steady-state buffer power at Ca2+ concentrations normally achieved during tetani. As shown in TABLE 6, in fast-twitch fibers the steady-state contributions of the TnC nonspecific (Ca2+/Mg2+) sites and of myosin are much less than provided by PV. In slow-twitch fibers (TABLE 6) these sites may play a larger role, but we can find no experimental data testing this.
Finally, an interesting example of the importance of PV is provided by the woodpecker, which drums its beak against a tree at a rate of almost 20 s−1. The neck muscles responsible for this drumming have much higher PV concentrations than both other skeletal muscle in the woodpecker and in the neck muscles of nondrumming species (398).
7.3. Differences in Buffering between Fast-Twitch and Slow-Twitch Fibers
The calculated buffer power differs considerably between fiber types because of differences in expression of proteins associated with Ca2+ signaling and contraction. (399). TABLE 6 lists the levels of the major cellular buffers found in small mammals such as mice (83, 84) and shows a much higher buffer power in fast- compared to slow-twitch fibers at both 0.05 and 2.0 µM [Ca2+]i. The greater concentration of SERCA in fast-twitch fibers corresponds to a larger SR Ca2+ store in this fiber type and the higher number of SR Ca2+ release sites compared with slow-twitch fibers (84). This arrangement generates a larger Ca2+ release during E-C coupling and along with the cooperative binding of Ca2+ at two sites on TnC (compared to only a single site in slow twitch), generating a faster rate of contraction in fast fibers. As discussed in sect. 7.2, the major contributor to intracellular buffering at [Ca2+]i values normally achieved during tetani (up to 20 µM) is PV, which is expressed at ∼10-fold higher levels in fast-twitch compared with slow-twitch fibers (385). The Ca2+ binding and buffer powers due to PV for the two fiber types are shown in FIGURE 15 and illustrate the minimal additional Ca2+ binding and Ca2+ buffering attributable to PV both immediately after SR Ca2+ release (instantaneous) and during sustained (>200 ms) tetani (steady state) in slow-twitch muscle fibers. This contrasts with the very large additional binding and the associated buffer power that operates in mouse fast-twitch muscle with 10-fold higher concentrations of PV (see TABLE 6).
FIGURE 15.
Theoretical parvalbumin (PV) buffer and Ca2+ binding curves in mouse skeletal muscle. A and B: buffer power [β(ΔCa)] and the change in total Ca2+ binding to PV assuming a resting (initial) Ca2+ of 50 nM. Black lines (dashed and solid) show the buffer power for a range of changes in free Ca2+ (Δfree[Ca2+]) that is normally experienced during a tetanus. The dashed black lines represent the initial buffer power (before dissociation of Mg2+ from PV), corresponding to the initial 20–30 ms [“instantaneous” values, in βi(ΔCa)]. The solid black lines show steady-state values calculated assuming the equilibrium of Ca2+ and Mg2+ with PV that would be approached during tetani lasting >200 ms [“steady-state” values, ss βi(ΔCa)]. Red lines and right axis represent the concentration of Ca2+ bound to PV at the instantaneous time point (in ΔBound [Ca2+]) and that during the steady state (ss ΔBound [Ca2+]).
7.4. The Role of Buffer Saturation in Tetanic Contraction
Another example of the physiological consequences of buffer kinetics for muscle function involves the Ca2+ binding events at the start of a tetanic contraction. As shown in FIGURE 14B, the initial SR Ca2+ release from the first in a train of action potentials (APs) raises total cytoplasmic Ca2+ by ∼350 µM, resulting in an increase of free [Ca2+]i of ∼20 µM. Free and total [Ca2+] as well as Ca2+ bound to TnC then begin to decay as a result of SR reaccumulation by SERCA; however, [Ca2+]i is still elevated at the time of the next stimulus (∼20 ms later). Since the SR Ca2+ content is decreased (not shown), this AP results in a smaller increase of total cytoplasmic Ca2+. However, this smaller SR Ca2+ release will be added on top of an already elevated [Ca2+]i with the majority of high-affinity Ca2+ binding sites being occupied. This occupancy reduces the cytoplasmic Ca2+ buffer power considerably, and although the second SR Ca2+ release is much smaller than the first, it is still sufficient to raise [Ca2+]i to a peak value almost as high as that produced by the first AP, maintaining an almost maximal Ca2+ binding to TnC (84, 179, 400). It should be noted that the slow kinetics of PV means that it does not substantially influence the time course of [Ca2+]i during such high-frequency stimulation. The interval between consecutive APs is critical. Too short intervals will attenuate APs and SR Ca2+ release. Too long intervals will allow the free Ca2+ concentration to decay sufficiently to decrease Ca2+ binding to TnC, and therefore contraction will not be maintained (179). The net effect is that a train of APs in skeletal muscle generates an immediate (<10 ms) increase in contractile activity that is sustained at close to maximal levels for the duration of the tetanic train. This effect of buffer saturation is analogous to that seen for buffer saturation-induced synaptic facilitation (sect. 9.5).
This principle also explains the benefit of doublet or triplet firing patterns recorded commonly in motor neurons of fast mammalian skeletal muscle fibers (373, 401). The first two or three APs are separated by shorter intervals (e.g., 5 ms) compared with the intervals in the rest of the train (e.g., 30 ms) (402). The doublet/triplet pattern of motor neuron firing ensures a rapid step change in contractile activity and sustained force response during a brief tetanic response (179) by sustaining [Ca2+]i for the initial 10–15 ms.
It has been suggested that cytoplasmic Ca2+ buffering may have significant effects on the level of [Ca2+]i reached during a tetanus (see Ref. 400 for review). For example, acidosis decreases force while increasing [Ca2+]i, and both effects can be attributed, at least in part, to decreased binding of Ca2+ to troponin (403). In this context, it is worth emphasizing an important difference between these effects and what would be expected in cardiac muscle. Ca2+ fluxes across the sarcolemma are small in skeletal muscle, so a decrease of Ca2+ binding to buffers would be expected to produce an increase of [Ca2+]i that will be maintained for longer than would be the case in the heart. Another mechanism by which altered buffering may be involved in fatigue has been proposed. The increase of cytoplasmic phosphate concentration, as a result of breakdown of phosphocreatine, has been suggested to increase the SR phosphate concentration, leading to precipitation of Ca phosphate in the SR and a reduction of Ca2+ release (404, 405). It should, however, be noted that changes of metabolites have a variety of other effects on Ca2+ handling (see Ref. 406 for review).
7.5. The Role of Calsequestrin
The concentration of free Ca2+ in the SR ([Ca2+]SR) of skeletal muscle is ∼300–600 µM (112, 407, with considerable Ca2+ also bound to CSQ (for review see Ref. 213). The major isoform in skeletal muscle is CSQ1 (408). The properties of SR Ca2+ buffering have been studied in intact frog skeletal muscle fibers in which the ends were cut to allow incorporation of indicators into the cytoplasm and SR. A high concentration of EGTA was added for two reasons: 1) to suppress changes of cytoplasmic [Ca2+]i, allowing the added Ca2+ indicator (tetramethylmurexide) to measure changes of [Ca2+]SR uncontaminated by those of cytoplasmic [Ca2+]i (409), and 2) Ca2+ released from the SR will bind to cytoplasmic EGTA releasing protons. The resulting change of intracellular pH gives a measure of the change of total SR Ca2+ concentration (256). This method gave a value (in frog skeletal muscle) of 670 µM for free and 17.1 mM for total SR Ca2+ with ∼30 mM CSQ. It was concluded that ∼95% of the Ca2+ released from the SR came from that bound to CSQ as opposed to free Ca2+. The SR buffering could be fit with a Hill coefficient of 3.0, indicating pronounced cooperativity of Ca2+ binding (cf. sect. 3.3). A subsequent study measured SR Ca2+ buffering in mouse skeletal muscle by comparing changes of free SR Ca2+ with those of calculated total Ca2+ assessed from changes of [Ca2+]i measured with a cytoplasmic indicator (410). This provided an average SR buffer power of 157. Measurements in myocytes of CSQ knockout mice showed that buffer power was reduced to 25%, indicating that CSQ is the major SR Ca2+ buffer. The SR Ca2+ buffer curve could be fit with a Hill coefficient of 3.52, again indicating cooperative binding. As discussed in sect. 3.3, the cooperativity of Ca2+ binding has been associated with polymerization of CSQ. This was investigated in intact skeletal muscle by using fluorescence recovery after bleaching (FRAP) to measure the diffusion of CSQ in the SR. Depleting the SR of Ca2+ increased the diffusion coefficient, indicating depolymerization (411). The cooperative nature of Ca2+ binding to CSQ has been suggested to have important physiological consequences (411, 412).
It is not simple to relate the biophysical properties of SR Ca2+ buffering to the known biochemical properties of CSQ (410). During Ca2+ release from the SR, the decrease of free [Ca2+] from ∼1 mM to 50 µM will cause depolymerization of the CSQ filament and a reduction in the available binding sites and therefore potentially buffer power (20, 132). Whereas the on and off rate constants of Ca2+ binding to the individual chelation sites are relatively rapid, the kinetics of the polymerization-depolymerization processes, which would determine the overall buffer kinetics, are unknown. Estimates of dimerization kinetics have been derived from the analysis of records of [Ca2+]SR signals from smooth muscle (189). More direct measurements are required, particularly since it is uncertain what degree of CSQ polymerization occurs within the SR lumen of different muscle types. EM images of the terminal cisternae of skeletal muscle show extensive branching networks of CSQ suggesting a high degree of polymerization (213, 413). However, the estimate of the maximum [Ca2+]SR in skeletal and cardiac muscle is ∼1 mM. Based on biochemical measurements, this concentration would only result in CSQ dimers and not higher-order polymers (132, 133). It is, of course, possible that the Ca2+ sensitivity of polymerization is affected by cellular constituents absent in biochemical studies. Furthermore, the law of mass action would suggest that the higher CSQ concentration found in fast-twitch SR compared to that used in the biochemical studies would favor polymer formation. The degree of polymerization may also be expected to be affected by the CSQ concentration, something that has yet to be studied in detail.
If CSQ did not buffer Ca2+ cooperatively, it would have its lowest buffer power at high [Ca2+]SR (FIGURE 3). Release of a given amount of Ca2+ would result in a large decrease of [Ca2+]SR, thereby increasing buffer power and decreasing the release flux. In contrast, the high buffer power of the cooperative CSQ buffer maintains [Ca2+]SR and thence Ca2+ release (213). Direct measurements of [Ca2+]SR have shown the occurrence of abrupt rises of [Ca2+]SR attributed to depolymerization of CSQ and therefore release of bound Ca2+ (411, 414). These papers also advanced an explanation for muscle fatigue with the idea that a decrease of [Ca2+]SR would (because of the cooperative nature of Ca2+ buffering) decrease the buffer power and thereby the Ca2+ release. In this context it is noteworthy that mutations in CSQ that alter polymerization and reduce Ca2+ binding have been associated with disease, including malignant hyperthermia and vacuole aggregate myopathy (415–417).
7.6. Diffusion of Ca2+ in Skeletal Muscle
The release of Ca2+ from the terminal cisternae of the SR results in a high local, cytoplasmic Ca2+ signal (376). To enable contraction, this Ca2+ must diffuse from the terminal cisternae of the SR, located near the Z lines (418), to bind to TnC molecules, located along most of the sarcomere, at a distance of up to 1 µm away. Modeling studies suggested that Ca2+ diffusion is accelerated by binding to ATP (34). An experimental study using frog skeletal muscle fibers compared the rise time of [Ca2+]i at the release sites (Z line) with that in the middle of the sarcomere (M line). The latter lagged by 2–3 ms, and these relative kinetics could be explained by a model including diffusion of CaATP (338). The situation may be different in mammalian skeletal muscle, where the Ca2+ release triads reside at the A-I border as opposed to the Z line (419), placing them closer to the TnC and thereby decreasing the diffusion distance and the need for acceleration of Ca2+ diffusion by mobile buffers. Note that skeletal muscle contains significant amounts of the mobile buffer carnosine (212) (cf. sect. 2.4.1). Subject to current uncertainties about its Ca2+ binding, this may also aid Ca2+ diffusion.
7.7. The Role of Buffers in Skeletal vs. Cardiac Muscle
The above sections emphasize an important difference between the role of buffers in these two types of striated muscle. There are four important roles of Ca2+ buffers in skeletal muscle: 1) SR Ca2+ buffering by CSQ limits the fall of Ca2+ that occurs during Ca2+ release from the SR. 2) Mobile buffers help transfer Ca2+ ions from the SR release sites to the myofibrils. 3) Buffering by PV accelerates relaxation. 4) Buffer saturation increases the rise of [Ca2+]i during a tetanus. In contrast, in cardiac muscle only the first two of these occur. Furthermore, the degree of SR Ca2+ buffering by CSQ is considerably greater in skeletal compared with cardiac muscle, perhaps reflecting a higher concentration of CSQ (409). Functionally, the difference may reflect the fact that depletion of the cardiac SR on a single beat does not matter as the SR is refilled before the next beat. Indeed, the decrease of free Ca2+ may contribute to closing of RyRs and thence termination of Ca2+ release (420). In contrast, the SR will gradually deplete during a tetanus in skeletal muscle, and the higher CSQ content will minimize this.
Other differences between buffering in cardiac versus skeletal muscle do not result from the differences in the buffers expressed but rather from the time course and concentration range of [Ca2+]i. As discussed above, the higher peak [Ca2+]i in skeletal muscle will result in decreased buffer power due to saturation, and the prolonged rise of [Ca2+]i during a tetanus will increase contributions from slow buffers.
The role of PV to accelerate relaxation in skeletal muscle contrasts with the fact that Ca2+ buffers slow relaxation of cardiac muscle. The explanation of this paradox is the slow kinetics of PV, which works in parallel with SERCA to reduce [Ca2+]i in fast-twitch skeletal fibers. In particular, the kinetics of PV are sufficiently slow that it continues to take up Ca2+ even when [Ca2+]i is declining. There is no equivalently slow buffer in cardiac muscle or slow-twitch skeletal fibers. In this context, it has been suggested that expression of a PV-like buffer in cardiomyocytes might be a useful therapeutic approach in heart failure as it could accelerate relaxation without decreasing contraction. PV binds Ca2+ too slowly to interfere appreciably with peak [Ca2+]i in systole but would bind significant amounts of Ca2+ during diastolic relaxation (421, 422). It has further been shown that β-PV performs better in this respect because it binds Mg2+ better than α-PV (423). This will increase the concentration of the Mg2+-bound and decrease that of the free form of the buffer, thereby decreasing the rapid component of buffering and allowing it to accelerate relaxation with less effect on the amplitude of the systolic [Ca2+]i transient and thence contractility. A similar improvement has been reported for single amino acid mutations in the α-PV EF-hand domain (424, 425).
8. SMOOTH MUSCLE
Several studies have measured the Ca2+ buffer power in smooth muscle by using the method described in sect. 6.1 in which cells are depolarized to activate L-type Ca2+ currents. The measured Ca2+ current and inferred total Ca2+ can then be compared with the rise of [Ca2+]i measured with an indicator dye. The resulting value of buffer power includes contributions from both the endogenous buffers and the indicator used to measure [Ca2+]i. A value of 82 was reported for toad stomach (426). A similar approach in rat portal vein myocytes gave a buffering power of 170, which, corrected for the contribution of the Ca2+ indicator, provided an endogenous buffering power of 114 (427). A smaller buffer power of 30–40 was reported for guinea pig urinary bladder smooth muscle cells (428). In guinea pig colonic smooth muscle cells, a much higher buffer power (∼400) was measured. This was reduced to 250 by the addition of ryanodine, suggesting that some of this apparent buffer power represented Ca2+ uptake into the SR (429), suggesting the lower value as the better estimate. In guinea pig coronary vascular smooth muscle cells, a buffer power of 300 was obtained, which corresponded to ∼150 when corrected for the contribution by the fluorescent indicator (430). A different study using the added buffer method (173) (sect. 9.1) provided a buffer power of 46 (431). Further research is required to clarify whether the wide range of reported buffer power values reflects real differences between the various types of smooth muscle as opposed to being a consequence of methodological differences. It is, however, worth noting that the lowest buffer power mentioned above was obtained with changes of [Ca2+]i of ∼4 µM (428) whereas the other studies typically involved maximum [Ca2+]i of 250–800 nM. It is therefore possible (cf. sect. 3.1) that the lowest buffer power values might result from buffer saturation. Given this, it is challenging to model buffering. A computer model of smooth muscle Ca2+ movements used a lumped buffer with a total concentration of 230 µM and Kd of 1 µM, equivalent to a rather high buffer power at low [Ca2+]i of 230 µM/1 µM = 230 (432).
What is the origin of the measured Ca2+ buffering? One potential contributor is CaM, which is present at concentrations of 34 and 40 µM in guinea pig taenia coli (433) and cultured bovine tracheal cells (434), respectively. Taking 40 µM and the Kapps listed in TABLE 5 (187), one calculates a buffer power of 4 at 100 nM [Ca2+]i. This value is lower than most of the buffer powers reported above. The affinity for Ca2+ binding increases when CaM is bound to proteins (435), and this would be expected to increase buffer power. A study using electron probe microanalysis found that the increase of total cytoplasmic Ca2+ accompanying a contracture produced by the combination of KCl-induced depolarization and norepinephrine was ∼235 µM (436). The authors pointed out that this was considerably greater than what might be expected to bind to CaM. They therefore suggested that Ca2+ must also bind to myosin and possibly other proteins. It has, however, been questioned whether myosin binds significant amounts of Ca2+ in the presence of cytoplasmic Mg2+ concentrations (see Ref. 437 for review). The reader will note that the studies referred to in this paragraph are mainly >30 years old. In their review published in 1986 (437), Sommerville and Hartshorne note, “A contemporary problem is to identify the various components of the Ca2+ buffer system.” It is our opinion that this problem is still “contemporary.”
A recent paper reported expression of calreticulin in endothelial cells forming myoendothelial junctions, which are important in coordinating calcium signaling between endothelium and smooth muscle. Much of the calreticulin expression was not associated with other ER markers, suggesting that it is located somewhere other than in the ER (438). Further work is required to establish whether this is present in the cytoplasm.
9. NEURONS AND NEUROENDOCRINE CELLS
As in muscle cells, Ca2+ ions function as the principal second messengers coupling action potential (AP) firing to diverse intracellular events, from activation of cytosolic signaling pathways to gene transcription. In addition, Ca2+ serves a dual neuron-specific function at chemical synapses: Ca2+ triggers synaptic vesicle fusion at presynaptic active zones, and Ca2+ is a pivotal signaling molecule controlling synaptic plasticity. The spatiotemporal characteristics of neuronal [Ca2+]i transients are determined by Ca2+ influx, extrusion, and buffering. Ca2+ influx into somata, dendrites, and axon terminals occurs primarily in response to membrane depolarizations opening voltage-gated Ca2+ channels (VGCCs). In some neurons, Ca2+ influx through Ca2+-permeable ligand-gated channels and/or Ca2+ release from internal stores represent two additional routes to increase [Ca2+]i. AP-induced Ca2+ influx is generally brief, lasting at most a few milliseconds. At glutamatergic synapses equipped with postsynaptic NMDA channels, Ca2+ influx can last longer owing to the slow deactivation kinetics of these channels (227, 439). Nevertheless, global cytosolic [Ca2+]i transients that remain after the collapse of local Ca2+ domains following channel closure can outlast the duration of Ca2+ influx by one or two orders of magnitude. These global cytosolic [Ca2+]i transients are shaped by Ca2+ extrusion and buffering, which differ between different neuron types and different subcellular compartments and which temporally and spatially confine Ca2+ signals. The manner in which neuronal [Ca2+]i transients are affected by Ca2+ buffers is determined by three parameters: 1) their concentration in the cytosol, 2) their affinity and binding kinetics for Ca2+ and Mg2+ ions, and 3) their mobility in the cytosol.
9.1. Measuring Cytosolic Ca2+ Buffering in Neurons and Neuroendocrine Cells
Much work characterizing neuronal buffers has used the “added buffer approach” as originally applied to adrenal chromaffin cells (27, 174) and later extended to nerve terminals (248, 440) and dendrites of hippocampal neurons (59, 201, 441, 442). In these experiments, a whole cell recording is established on the compartment of interest, with a patch pipette containing a ratiometric Ca2+ indicator, usually fura-2 or one of its low-affinity analogs at a known concentration and therefore value. Either single APs are elicited or short depolarizing pulses are applied under voltage clamp at various time points while the cell is loaded with the indicator dye (FIGURE 16) (248, 440, 443). Both the amplitudes (Δ[Ca2+]i) as well as the decay time constants τCa,i of the resulting [Ca2+]i transients are evaluated. According to the simple theory presented in Eqs. 31 and 32 (27, 174), plots of 1/A or τ against the amount of “added buffer power” allow the calculation of values for A and τ in the absence of exogenous buffers by extrapolation to zero added buffer (y-axis intercept, FIGURE 16Bb). The negative x-axis intercept of such plots is an estimate for the endogenous buffer power (1 + ; FIGURE 16Bb). This method is therefore a refinement of the approach using different concentrations of Antipyrylazo III in skeletal muscle (sect. 7.1 and Ref. 378). Unlike the techniques presented above for cardiac and smooth muscle (sects. 6.1 and 8), where buffer power is estimated by comparing the movement of Ca2+ into the cytoplasm with the change of [Ca2+]i, the added buffer approach does not require knowledge of the cytoplasmic volume. However, it does not easily provide information about buffer type, its concentration, or affinity. In fact, several mobile and immobile buffer species may often contribute to the obtained buffer power estimates. The added buffer approach is typically used with small and compact cells such as adrenal chromaffin cells or with subcellular compartments of neurons such as somata or nerve terminals, for which it can be assumed that spatial [Ca2+]i gradients collapse over a time span much shorter than typical values for the decay constants of [Ca2+]i transients (>10 ms). The method is also suitable for cylindrical structures such as axons and dendrites under careful selection of uniform diameter, such that longitudinal gradients are minor. FIGURE 16 illustrates its use to estimate buffer power in neuronal cell bodies and nerve terminals. When it is applied to cerebellar Purkinje cells, their high endogenous buffer power results in dialysis with fura-2 having only small effects on measured [Ca2+]i transients (FIGURE 16A). In contrast, much larger effects are observed (FIGURE 16, B AND C) in calyx of Held terminals because of their much lower buffer power.
As originally described (174), the results of the added buffer approach are somewhat ambiguous since it is not clear which types of endogenous buffer are being assayed. All fixed buffers are expected to be included, but mobile buffers may be washed out while the indicator dye is infused (FIGURE 17) (444). Typical buffers, with diffusion coefficients about four times smaller than that of fura-2 (TABLE 3), should be retained at the early time points of the whole cell recording and are therefore probably included. Small metabolites, however (sect. 2.4), may be washed out quickly. During extended recording periods, diffusional equilibration between pipette and cytoplasm is achieved also for larger buffer molecules such as calbindin-D28k, resulting in nearly complete washout of that mobile buffer with a time constant of ∼10 min (FIGURE 17E). At the same time gluconate−, often chosen as the main anion of the pipette solutions in these types of experiments, is infused together with the indicator. The gluconate anion binds Ca2+ weakly (60) and may well compensate for the loss of endogenous counterparts. Therefore, estimates of buffering power contributed by small low-affinity Ca2+ binders and determined by the original added buffer approach cannot be considered reliable. Although their contribution to the absolute value of buffering power is expected to be small, their influence on Ca2+ diffusion and on local domains may be substantial (see sects. 4.5. and 4.7.2). To remove this uncertainty and to focus on low-mobility and stationary buffers, Matthews, Schoch, and Dietrich (59) proposed a modification of the procedure: Instead of measuring changes in [Ca2+]i during a single dye-loading experiment, several whole cell recordings are performed with different concentrations of indicator dyes in the pipette. [Ca2+]i transients are elicited after dye loading is complete and mobile Ca2+ buffers are expected to have been washed out. Therefore, the plots of A−1 and τ plots versus of the indicator, as described above, should report the buffering power of immobile and slowly mobile buffers, depending on the geometry of the cell under study and the waiting period before measurement. When applied to the calyx of Held nerve terminal, this method yielded an endogenous Ca2+ buffer power of 21 (248), whereas somewhat higher estimates (40–46) were obtained with the original procedure (440, 445).
FIGURE 17.
Loss of the endogenous mobile Ca2+ buffer CB-D28k during whole cell recordings in hippocampal dentate gyrus granule neurons (GCs). A–D: several GCs were recorded and filled with the fluorescent dye Lucifer yellow (LY) during whole cell recording episodes lasting for 30 s (A), 6 min (B), 12 min (C), or 38 min (D). Subsequently, hippocampal slices were fixed and stained for CB-D28k. Arrows indicate recorded neurons identified by their LY fluorescence. E: time course of CB-D28k loss during whole cell recordings. The CB-D28k immunofluorescence of each recorded GC was normalized to that of neighboring GCs. Modified from Ref. 444, with permission from Journal of Neuroscience.
The same problem was addressed in studies on bovine chromaffin cells (446) by “balanced dye loading.” In this approach, individual cells are preloaded with a certain concentration of fura-2, either by AM-ester loading or by dye preloading, during a brief whole cell recording episode. Subsequently, a perforated-patch recording is established, this time with a concentration of fura-2 in the pipette, which is expected to match the concentration inside the cell under study. Total Ca2+ buffer power, including mobile and immobile buffers, is estimated from total Ca2+ influx and changes in free [Ca2+]i during short depolarizing voltage-clamp pulses. Once a stationary situation has been established, the perforated patch is ruptured, resulting in a whole cell configuration. Subsequent changes in buffering power are followed by repetitive application of depolarizing pulses. Experiments are considered successful if dye loading is “balanced,” meaning that total fura-2 fluorescence did not change rapidly after rupture. In that case, later changes in Ca2+ buffer power are interpreted as changes in endogenous buffers. No such changes were observed in some cells after “balanced loading,” indicative of little or no contribution to Ca2+ buffering from mobile buffers, which would have washed out. However, on a longer timescale about half of the cells displayed a partial drop in buffering power compatible with the loss of proteins of ∼7,000–20,000 molecular weight, which before their washout contributed ∼25% of the total buffering power of 40.
All versions of the added buffer approach discussed so far depend on the approximations discussed in sect. 4.7.1, employing Eqs. 30–32 for the analysis of exponentially decaying [Ca2+]i transients. When studying slow buffers, such as PV, which cause biphasic decays of [Ca2+]i transients, more complex stimulus patterns and analysis procedures need to be invoked (27). Various other methods for measuring Ca2+ buffer power, and also for titrating endogenous buffers, have been described (251, 447, 448), similar to methods typically used when studying muscle cells (sects. 6.1 and 8). With photolysis of caged Ca2+ compounds, Kd, Ca2+ binding rate constant, and Ca2+ binding ratio of the endogenous fixed buffer in adrenal chromaffin cells were estimated as ∼100 µM, 5.17 × 108 M−1·s−1, and 40, respectively (449). Altogether, values of buffer power are typically in the range of 15–300; maximum values of ∼2,000 for Purkinje cell somata and dendrites were found. For a compilation of Ca2+ buffer powers and affinities of endogenous Ca2+ ligands see Refs. 173, 201. Thus, the fraction of free Ca2+ in the cytosol varies typically between ∼0.3% and ∼7% of total [Ca2+]i at equilibrium but can be as low as 0.05%. The buffering power of fixed buffers is small but nevertheless the dominating one in many types of neurons. Higher buffering powers are primarily achieved by a higher abundance of endogenous mobile Ca2+ binding proteins and confer special characteristics of short-term synaptic plasticity and electrical excitability (see below).
9.2. Neuronal Expression of EF-Hand Domain Ca2+-Binding Proteins
In the mammalian brain, the predominant mobile Ca2+ buffers found in neurons are parvalbumin (PV/α-PV), calbindin-D28K (CB-D28k), and calretinin (CR), which are members of the superfamily of EF-hand domain CBPs (sect. 2.5.1). Neurons typically express only one of these three proteins, but coexpression of combinations of two of the three EF-hand domain CBPs is also observed. Two other relevant molecules that bind Ca2+ rapidly and are present at high cytoplasmic concentrations in neurons, and therefore shape the spatiotemporal characteristics of local and/or global Ca2+ transients, are ATP (450) and CaM (451). CaM binds Ca2+ with faster kinetics than any of the three CBPs CR, CB-D28k, or PV (187).
The commonly applied criterion (sect. 2.5.1) that pure Ca2+ buffer proteins do not undergo marked conformational changes upon Ca2+ binding and do not interact with other proteins in a Ca2+-dependent way holds for PV among the family of EF-hand domain CBPs (565). In contrast, a substantial conformational change is observed upon Ca2+ binding to CB-D28k (31). The apparent diffusion coefficient for CB-D28k in the cytosol depends on the subcellular compartment, and an interaction of CB-D28k with myo-inositol monophosphatase has been identified in spines and dendrites of cerebellar Purkinje neurons, indicating that CB-D28k is both a Ca2+ buffer and a Ca2+-signaling molecule (452). A similar picture emerges with respect to CR: CR and P/Q-type (CaV2.1) VGCCs coimmunoprecipitate from mouse cerebellum homogenate. In HEK293T cells, coexpression of CR attenuates Ca2+-dependent inactivation (CDI) and augments Ca2+-dependent facilitation (CDF) of heterologously expressed P/Q-type VGCCs via a direct interaction with the α12.1 subunit (453). It remains to be established whether a modulation of VGCCs by CR as observed in HEK293T cells also occurs in neurons.
The expression of EF-hand domain CBPs in different brain regions has been studied extensively at the cellular and subcellular levels and in various species, once specific antibodies became available. It is the topic of several excellent reviews (454–460), and here we only summarize some key findings. PV is predominantly, but not exclusively, expressed in inhibitory interneurons of the brain. Many of these are fast-spiking GABAergic interneurons, including fast-spiking basket and chandelier cells of the neocortex and the hippocampus, fast-spiking striatal interneurons, two classes of GABAergic interneurons of the molecular layer of the cerebellum, stellate and basket cells, and many interneurons of the thalamus. PV is absent from pyramidal neurons of the neocortex and hippocampus, but it is expressed in large amounts in the output neurons of the cerebellum, the Purkinje cells, which additionally coexpress CB-D28k. PV is heavily present in somata and neuropil of the rodent auditory system, for example in auditory nerve fibers, and neurons of the spiral ganglia, the cochlear nucleus, and the inferior colliculus (461).
Prominent PV immunoreactivity is found in three types of large calyciform synaptic terminals: GABAergic terminals of interneurons in the thalamus (462) and glutamatergic endbulb of Held and calyx of Held terminals in the auditory brain stem (461). In subsets of the latter two types of auditory terminals coexpression of presynaptic CR but not CB-D28k was found (463, 464). Interestingly, the onset of PV expression in numerous classes of neurons of the auditory brain stem of mice and rats occurs relatively late and coincides approximately with the onset of hearing (461, 463).
CB-D28k is expressed in a variety of neurons, mostly interneurons, and the adult pattern of expression is in general established at birth. In the rat neocortex, CB-D28k-immunoreactive cells are predominantly found in the upper layers II and III (457). In the hippocampus, CB-D28k-immunoreactive interneurons are found in all subdivisions (465). CB-D28k is also a major Ca2+ buffer of hippocampal dentate gyrus granule cells and CA1 pyramidal neurons, whereas it is absent from CA3 pyramidal cells (444, 466). In the cerebellum, Purkinje cells are the only neurons expressing CB-D28k (455).
Similar to CB-D28k, CR is expressed in various interneurons of the neocortex, most abundantly in the upper layers II and III, and in specific interneurons of the hippocampus (467, 468). However, coexpression of CB-D28k and CR in individual neurons is rare (469). In the cerebellum, CR is expressed in the glutamatergic granule cells, which relay mossy fiber inputs to Purkinje cell dendrites via the parallel fiber pathway (455).
Expression of CBPs within the vertebrate retina varies among species. In the rodent, photoreceptors and bipolar cells lack expression of CBPs whereas some classes of horizontal, amacrine, and ganglion cells express one or more of CB-D28K, CR, and PV (470–472). In the rat cochlea, in situ hybridization shows exclusive expression of β-PV (oncomodulin) in outer hair cells (OHCs) whereas inner hair cells (IHCs) express both α-PV and β-PV (473). The developmental profile of PV isoform, CB-D28k, and CR expression revealed a transient β-PV expression during IHC development while the expression of CR and α-PV slightly increases. The sum of the CBP concentrations decreases in IHCs but increases in OHCs during cochlear maturation (474).
The selective expression of CBPs in distinct neuron populations of different brain areas suggests that specific functional properties of CBPs confer specific physiological properties to those neurons. This finding has been exploited experimentally to identify and/or selectively manipulate the respective neurons in situ. The availability of Cre-driver mouse lines for both PV and CR allows targeting PV- and CR-expressing neurons for genetic manipulations, including the expression of fluorescent markers, fluorophores for Ca2+-imaging, and channelrhodopsin and its derivatives enabling stimulation or silencing of specific neuron populations (475, 476).
9.3. Cytosolic Concentration of Ca2+-Binding Proteins
For some neurons and subcellular compartments, cytosolic concentrations of CBPs have been estimated either by using calibrated immunohistochemistry or via functional assays during which neurons lacking a certain CBP were loaded with recombinant protein to restore normal function.
Hippocampal dentate gyrus granule cells (hDG GCs) contain ∼40 µM CB-D28k, corresponding to ∼160 µM Ca2+ binding sites, as estimated by performing postrecording immunohistochemistry following whole cell dialysis with known concentrations of recombinant CB-D28k and comparing immunofluorescence intensities to that in neighboring, unperturbed neurons (444). Cytosolic CB-D28k concentrations for hippocampal CA3 stratum radiatum interneurons and CA1 pyramidal neurons were similar, with ∼47 µM and ∼45 µM, respectively (444). With a similar approach, a cytosolic PV concentration of ∼12 µM was estimated for hippocampal dentate gyrus basket cells (hDG BCs) (267). The PV concentration was variable among individual hDG BCs but similar in somata and boutons of a given cell. For cerebellar basket cell (cBC) somata, a substantially higher mean PV concentration of ∼565 µM was reported, which was also less variable among individual cBCs (267).
In other studies, the concentrations of CBPs were estimated from densities of gold particles with electron microscopic postembedding immunogold procedures. Quantification of particle densities allows for a statistical comparison of the relative levels of CBPs in somata, dendrites, dendritic spines, axons, and axon terminals. Using this approach, Kosaka et al. (477) found significantly higher levels of PV immunoreactivity in axons and axon terminals of Purkinje cells and basket cells than in their respective somata and dendrites. In contrast, CB-D28k immunoreactivity was more similar in somata, dendrites, and spines of Purkinje cells. Estimates for absolute PV concentrations were obtained by comparison to calibration curves deduced from quantitative immunogold analyses of standard PV samples. Estimated PV concentrations were 50–100 µM for Purkinje cell somata and dendrites as well as for interneuron somata and 1 mM or more in axons and axon terminals of Purkinje cells and cBCs (477), corresponding to 100–200 µM and 2 mM Ca2+ binding sites, respectively. Estimates for PV and CB-D28k concentrations in Purkinje cell somata were obtained more recently from calibrated immunogold tissue counts, with 116 µM and 208 µM for PV and CB-D28k, respectively (474), corresponding to 232 µM and 832 µM Ca2+ binding sites. In postnatal day 26 rats with fully developed hearing, cochlear inner hair cells contained one-tenth of the amount of CBPs of outer hair cells. In these latter cells, the cell body contained β-PV and CB-D28k at high levels equivalent to 5 mM Ca2+ binding sites. In contrast, the concentration of Ca2+ binding sites in inner hair cells was ∼0.5 mM and was dominated by α-PV (474).
An alternative approach for estimating CBP concentrations is based on functional assays during which normal neuronal function in neurons lacking a certain CBP, either due to genetic ablation or because of washout during whole cell dialysis, is restored by supplying exogenous buffer via the patch pipette. With such a “rescue approach,” a concentration of 1.2 mM CR was estimated for frog saccular hair cells, corresponding to ∼6 mM Ca2+ binding sites. The criterion for rescue was the concentration of exogenous CR required to restore the voltage dependence of activation of Ca2+-sensitive potassium channels to the level of that measured in perforated-patch recordings (478) (FIGURE 18A). In voltage-clamped cochlear inner hair cells (IHCs) of young (postnatal day 14–23) constitutive triple KO (α-PV−/−CB-D28k−/−CR−/−) mice, a relationship between Ca2+ influx duration and exocytosis similar, but not identical, to that of wild type could be restored by buffer infusion via the patch pipette (223). The required mobile buffer concentrations were equivalent to ∼1 mM synthetic Ca2+ binding sites, half of them with kinetics as fast as BAPTA, the remainder with properties like EGTA.
FIGURE 18.
Estimating the concentration of endogenous buffers by functional rescue. A: exogenous calretinin (CR) in the pipette solution mimics the native Ca2+ buffer in suppressing the Ca2+-activated potassium channel (KCa) current in frog saccular hair cells. KCa currents normalized to the amplitude at –15 mV are plotted against membrane potential (Vm). The voltage dependence of KCa is similar to that seen in perforated-patch recordings if the pipette solution for whole cell recordings contains 1.2 mM CR. Modified from Ref. 478, with permission from Nature Neuroscience. B: exogenous parvalbumin (PV) in the pipette solution mimics the native Ca2+ buffer in accelerating the decay of action potential (AP)-triggered intracellular calcium concentration ([Ca2+]i) transients. Average waveforms of AP-evoked [Ca2+]i transients during presynaptic whole cell recordings in calyx of Held terminals obtained with pipette solutions supplemented with either 100 µM fura-6F, 100 µM fura-6F + 100 µM recombinant PV, 100 µM fura-6F + 100 µM EGTA, or 50 µM fura-2. The slow Ca2+ buffer EGTA and recombinant PV restore the fast decay of AP-evoked [Ca2+]i transients. Modified from Ref. 479, with permission from Journal of Neuroscience.
Presynaptic AP-induced [Ca2+]i transients in calyx of Held terminals decay with a biphasic time course when the intracellular medium is only minimally perturbed by brief dye preloading. When such transients are measured during standard whole cell recordings without added Ca2+ buffers, a slower monoexponential decay is observed. Ca2+ transient amplitudes differ only marginally between the two conditions. This finding is consistent with a washout of a mobile Ca2+ buffer with slow binding kinetics such as PV during whole cell recording. In agreement with this notion, AP-induced presynaptic [Ca2+]i transients decayed slowly in calyx terminals of PV-deficient (PV−/−) mice. A fast decay of AP-evoked [Ca2+]i transients was restored when whole cell recordings were performed with 50–100 µM of the slow buffers EGTA or PV added to the presynaptic pipette solution, indicating that unperturbed terminals probably contain an equivalent amount of PV (479) (FIGURE 18B).
9.4. Local [Ca2+]i Domains Triggering Synaptic Vesicle Fusion
Local Ca2+ domains that are established when VGCCs open and rapidly collapse after channel closure dominate fast Ca2+ signaling. They may trigger synaptic vesicle fusion and also activate Ca2+-sensitive potassium channels, which contribute to AP repolarization.
Immobile endogenous Ca2+ buffers retard diffusion and thereby prolong the time until local domains of elevated [Ca2+]i reach steady state. Mobile endogenous buffers, on the other hand, accelerate the redistribution of Ca2+ ions and thereby reduce both the amplitude of changes of local [Ca2+]i and their spatial extent. The importance of the binding rate constant kon as a major factor determining the effect of exogenous buffers on local Ca2+ domains was first considered in a study on Ca2+-sensitive BK potassium channels in adrenal chromaffin cells (480). It was observed that EGTA is much less effective than BAPTA in blocking the activation of these channels, even though the buffers have similar equilibrium dissociation constants Kd (TABLE 1) and consequently similar steady-state buffering power (see Eqs. 33–35 and TABLE 4 for order-of-magnitude estimates of the spatial extent and kinetics of local Ca2+ domains). The size of the local domain depends mainly on the product of the binding rate constant kon and the concentration of free, mobile chelator, [B]0.
Thus, anions that are present at high concentration in the cytosol may have a strong effect on local domains, even if their equilibrium buffering power is very small, e.g., because of a very fast off-rate constant. For example, gluconate−, which is often infused as the main anion in patch-clamp experiments together with ATP2−, was found to increase the Ca2+ influx necessary for eliciting a given amount of transmitter release by a factor of 2.7, when comparing experiments performed with different intracellular solutions containing either gluconate− and ATP2− or methanesulfonate−, lacking Ca2+ binding and no ATP2− (60). Other small anions, present in unperturbed cells, as discussed in sect. 2.4.4, may have similar effects on local domains.
From the perspective of intracellular Ca2+ sensors, the terms “nanodomain” and “microdomain” are frequently used to describe the tightness of coupling between sites of Ca2+ entry and the sensor. However, these terms are not precisely defined with respect to spatial distances, and the distinction is primarily based on the experimentally determined effectiveness of the slow and fast buffers EGTA and BAPTA, respectively, in uncoupling the Ca2+-triggered process under study from Ca2+ influx. Synapses in which transmitter release is sensitive to EGTA (indicative of microdomain coupling) are the calyx of Held (481) and synapses between layer 5 cortical pyramidal neurons (482) in young rats. Synapses in which release is largely insensitive to EGTA (nanodomain coupling) are the mature calyx of Held (483), inner hair cells of the cochlea (484), GABAergic hippocampal dentate gyrus BC→GC synapses (485), glutamatergic cerebellar GC→PC synapses (486), and rod bipolar cell→amacrine cell synapses of the retina (487). Numerical simulations suggest mean coupling distance of ∼100 nm and <30 nm for microdomain and nanodomain coupling, respectively (243, 445, 450, 485, 486, 488).
To base the determination of coupling distance between VGCCs and docked synaptic vesicles exclusively on the differential effectiveness of EGTA and BAPTA may be an oversimplification because processes upstream of synaptic vesicle fusion can also be Ca2+ dependent, for example the resupply of fusion-competent synaptic vesicles (489–492) or the balance between distinct priming states (492). Therefore, reduced synaptic strength after prolonged presynaptic whole cell dialysis with EGTA-containing pipette solution may result not only from uncoupling of the Ca2+ sensor for vesicle fusion but also from reduced availability of fusion-competent vesicles unless resting [Ca2+]i is guaranteed to remain unchanged while changing the concentration of free chelator. This can be achieved by maintaining a fixed ratio of Ca2+-bound to free buffer in the recording pipettes while increasing total amount of buffer.
Relative contributions of the fast CBPs CB-D28k and CaM to buffering of AP-evoked [Ca2+]i transients at presynaptic active zones were studied by numerical simulations (451). These predicted that each buffer contributes to the reduction of AP-evoked local [Ca2+]i transients and resulting decrease of synaptic vesicle fusion probability. At an assumed VGCC-to-synaptic vesicle distance of 40 nm, CB-D28k caused ∼50% reduction of the synaptic vesicle fusion probability relative to control simulations without CB-D28k and CaM. CaM had a stronger inhibitory effect of ∼80% reduction, and addition of CB-D28k on top of CaM caused only a minor further reduction (∼85%). The reduction of AP-evoked fusion at synapses that contain both CB-D28k and CaM is mainly caused by fast Ca2+ binding to the N- and C-lobes of CaM. CB-D28k plays only secondary roles (451). At a resting [Ca2+]i of 50 nM, >99.8% of CaM C-lobes are in the Ca2+-free apo-state.
The slow Ca2+ buffer PV has Ca2+ binding kinetics similar to EGTA. Because of its relatively high affinity for Mg2+, the majority of PV is Mg2+ bound and little PV is free at physiological cytosolic [Mg2+]i (sect. 4.2). Competition with bound Mg2+ slows the binding of Ca2+ to PV in response to a cytosolic [Ca2+]i increase. However, if the total cytosolic PV concentration is high, sufficient amounts of PV are not Mg2+ bound but free and can therefore act as a fast Ca2+ buffer. For example, for cerebellar BC→PC synapses, which contain on average >0.5 mM PV, 5% of the PV is free at rest, 73% is bound to Mg2+, and 22% is bound to Ca2+, when assuming resting concentrations of 40 nM free Ca2+ and 400 µM free Mg2+ (267). Thus, there is ∼30 µM of free PV available that can rapidly bind Ca2+ and affect peak local domain [Ca2+]i. When free PV is depleted in the local domain, it is replenished both by diffusion of free PV from the periphery and from a large reservoir of Mg2+-bound PV, which, however, has to shed Mg2+ before being able to bind Ca2+. The Mg2+ binding of PV, therefore, represents a mechanism for generating new buffer (called “metabuffering”) (267), albeit on a slower timescale. This distinction between fast buffering by free PV and slower buffering following dissociation of Mg2+ is identical to that discussed above for skeletal muscle (sect. 7.2).
9.5. Modulation of Short-Term Plasticity by Ca2+ Buffers
During repetitive activation, the strength of synapses can transiently increase or decrease, resulting in synaptic facilitation or depression. Facilitating and depressing mechanisms are likely to operate simultaneously at many synapses, and the balance between the two defines magnitude and time course of changes in synaptic strength during stimulus trains. Local and global [Ca2+]i signaling plays a key role in the presynaptic mechanisms of short-term plasticity (STP). Ca2+ buffers can affect presynaptic STP in two ways: 1) buffers shape amplitude and spatial profile of local Ca2+ domains triggering transmitter release, and 2) buffers determine the time course of global [Ca2+]i changes, which occur in nerve terminals after diffusional equilibration after AP firing. These signals, also called “residual [Ca2+]i changes,” are crucially involved in regulating STP.
The relationship between transmitter release rates and [Ca2+]i is highly nonlinear, and the spatial profile of local [Ca2+]i domains that build up in the vicinity of open presynaptic VGCCs after AP arrival is affected by Ca2+ buffers. Provided that these buffers bind Ca2+ fast enough, they are able to intercept incoming Ca2+ ions before binding to the Ca2+ sensor for transmitter release (see sect. 9.4). Thereby, Ca2+ buffers determine the [Ca2+]i seen by the Ca2+ sensor and control synaptic vesicle fusion probability. During repetitive presynaptic AP firing, the concentration of free Ca2+ buffers may decrease if Ca2+ does not completely dissociate during interstimulus intervals or else if global [Ca2+]i is transiently elevated. Such buffer saturation can lead to release facilitation because synaptic vesicles experience incrementally higher local [Ca2+]i during repetitive stimulation (178, 493). Such a mechanism has been proposed to underlie facilitation at cortical synapses between multipolar bursting (MB) interneurons and pyramidal neurons (177) (FIGURE 19) and is analogous to the buffer saturation underpinning tetanic contraction in skeletal muscle (sect. 7.4). MB interneurons are CB-D28k positive, and MB→CA3 pyramidal cell synapses show pronounced paired-pulse facilitation (PPF). Washout of CB-D28k from MB interneurons during prolonged whole cell recordings increased the amplitude of the first responses and reduced PPF at MB→pyramidal cell synapses. Recordings in synapses of CB-D28k−/− mice showed a similar pattern. CB-D28k loading into MB interneurons of CB-D28k−/− mice via the recording pipette restored wild-type-like synaptic amplitudes and PPF (177). Taken together, these observations demonstrate that rapid Ca2+ binding to CB-D28k is able to reduce [Ca2+]i within local domains that trigger vesicle fusion, and thus reduces synaptic strength.
FIGURE 19.
Ca2+ buffer saturation contributes to paired-pulse facilitation (PPF). A and B: opposite effects of changing Ca2+ influx on PPF in wild-type (MF, top) and CB-D28k-deficient (MF CB-KO, bottom) hippocampal MF→CA3 pyramidal cell synapses. A: excitatory postsynaptic currents (EPSCs) recorded in CA3 pyramidal cells during paired-pulse stimulation (interval 100 ms) of mossy fibers. B: paired-pulse ratios (EPSC2/EPSC1) normalized to those measured in 2 mM external calcium concentration ([Ca2+]o). In wild-type MF→CA3 synapses (□), increasing external [Ca2+] increases facilitation, whereas decreasing ([Ca2+]o reduces facilitation. In CB-D28k-deficient MF→CA3 synapses (●), the opposite was observed. Normalized PPFs measured at Schaffer collateral (SC)→CA1 synapses (■) are shown for comparison. Right graph illustrates a similar dependence of EPSC1 on ([Ca2+]o in all 3 synapses. Modified from Ref. 177, with permission from Neuron. C–E: a fast endogenous buffer attenuates transmitter release during the first action potential (AP). Buffer saturation during the first AP leads to higher local intracellular [Ca2+] ([Ca2+]i) during the second AP and thereby produces PPF at MF→CA3 synapses. Upon washout of endogenous buffer, substitution with the slow buffer EGTA alters PPF whereas the fast buffer BAPTA restores PPF. C: schematic illustration of recording configurations. Presynaptic APs were elicited via stimulation in bouton-attached (left) or whole bouton (right) configuration with either EGTA or BAPTA in the pipette. The whole bouton configuration leads to rapid washout of mobile endogenous buffers from the presynaptic MF terminal. D: EPSCs recorded in a CA3 neuron while the MF bouton was first stimulated in the bouton-attached configuration (left) and subsequently in the whole bouton configuration with 100 µM EGTA in the stimulation pipette (right). E: similar experiment as shown in D but with 300 µM BAPTA in the pipette. Modified from Ref. 494, with permission from Science.
Buffer saturation also contributes to PPF and frequency facilitation at hippocampal mossy fiber (hMF)→CA3 pyramidal cell synapses (177, 494). Synaptic facilitation counteracts a potential reduction in synaptic strength caused by the consumption of synaptic vesicles during repetitive AP firing. In contrast to many other types of synapses, which show increased synaptic facilitation when lowering external [Ca2+] to attenuate release probability and thereby prevent synaptic vesicle exhaustion, the magnitudes of both PPF and frequency facilitation are decreased in these synapses at lower external [Ca2+] (177). This finding is consistent with reduced buffer saturation due to reduced presynaptic Ca2+ influx (FIGURE 19, A AND B). Dialysis of hippocampal mossy fiber boutons (hMFBs) with pipette solution containing 0.1 mM EGTA strongly augmented initial excitatory postsynaptic currents (EPSCs) and decreased paired-pulse ratios, whereas dialyzing 0.3 mM BAPTA into hMFBs restored amplitudes and facilitation to a pattern similar to synaptic responses generated by stimulating unperturbed hMBs (494), consistent with the loss of a mobile fast binding endogenous Ca2+ buffer during whole cell dialysis of the hMFB terminal (FIGURE 19, D AND E).
GABAergic synapses between cerebellar interneurons and Purkinje cells (PCs) show paired-pulse depression (PPD) in response to paired stimuli delivered at intervals between 30 and 300 ms (495). In PV−/− mice, in synapses as well as in whole cell recordings of connected interneuron-PC pairs, the same stimulus pattern induced PPF (267, 495). Wild-type-like short-term plasticity could be restored by loading recombinant PV into presynaptic interneurons (267). Likewise, dialysis of presynaptic interneurons with 1 mM of the slow Ca2+ buffer EGTA rescued PPD in synapses of PV−/− mice. EGTA and PV do not reduce markedly the peak of AP-induced [Ca2+]i transients, but both accelerate their initial rate of decay. Thus, the acceleration of the Ca2+ decay is likely to reduce the residual [Ca2+]i and hence to attenuate facilitation (495) (FIGURE 20A).
FIGURE 20.
The slow endogenous buffer parvalbumin (PV) affects the time course of presynaptic global intracellular calcium concentration ([Ca2+]i) transients and thereby regulates synaptic short-term plasticity. A: absence of paired-pulse depression at GABAergic synapses between interneurons and Purkinje cells (PCs) of PV knockout (KO) mice. Inhibitory postsynaptic currents (IPSCs) in response to extracellular stimulations of GABAergic interneurons with paired stimuli (30-ms interval) were recorded in Purkinje cells of a PV+/+ mouse (top) and a PV−/− mouse (middle). Bottom: average paired-pulse ratios [PPR = excitatory postsynaptic current (EPSC)2/EPSC1] are shown as a function of interstimulus interval for PV+/+ (circles) and PV−/− (diamonds) mice. Modified from Ref. 495, with permission from Proceedings of the National Academy of Sciences USA. B: slower decay of PPF at glutamatergic calyx of Held synapses of PV−/− mice. EPSCs were recorded in response to pulse pairs at 3 different interstimulus intervals (Δt) in a PV−/− calyx of Held synapse at reduced (0.6 mM) external [Ca2+] (top). PPRs plotted as a function of Δt were fitted by an exponential function with a time constant of 77 ms (middle). Paired-pulse facilitation in a calyx synapse from a wild-type (PV+/+) mouse decayed notably faster, with a time constant of 17 ms (bottom). Modified from Ref. 479, with permission from Journal of Neuroscience.
Similar observations were made at calyx of Held synapses, which show pronounced PPF at low-release probability conditions (0.6 mM external [Ca2+]). PPF decays with a time constant of 17 ms in synapses of wild-type mice, whereas genetic ablation of PV expression increases the time constant of decay to 77 ms. PV+/+ and PV−/− calyx of Held synapses show a similar magnitude of PPF when probed at an interstimulus interval of only 4 ms. However, at an interstimulus interval of 50 ms, PPF had nearly completely decayed in the wild type, whereas it was reduced to only half in PV−/− calyx synapses (479) (FIGURE 20B). These observations are consistent with the notion that the slow Ca2+ buffer PV only marginally reduces the peak of residual [Ca2+]i transients, because of its slow action. Nevertheless, PV binds Ca2+ fast enough to strongly accelerate their initial decay. In contrast to the case of calyx of Held synapses, genetic elimination of PV has only minimal effects on synaptic transmission at hippocampal BC→GC synapses, consistent with its low concentration in hBCs (267).
Cerebellar Purkinje cells (PCs), which express large amounts of CB-D28k and PV, are reciprocally connected via inhibitory synapses that show PPF during high-frequency activation. Surprisingly, PPF is not affected by the absence of either CB-D28k or PV at these GABAergic recurrent PC→PC synapses. Likewise, PPF measured in experiments on wild-type pairs of connected PCs that were dialyzed for 60–70 min with 10 mM EGTA remains unaltered compared with controls. Thus, PPF at PC→PC synapses is largely independent of the major endogenous Ca2+ buffers CB-D28k and PV and of the decay of residual [Ca2+]i. Instead, it was proposed that PPF results from long-lived Ca2+-bound states of the sensor for transmitter release (496).
9.6. Modulation of Delayed Asynchronous Transmitter Release by Ca2+ Buffers
In many synapses, two kinetically distinct components of AP-evoked transmitter release can be distinguished: 1) synchronous release that is temporally tightly coupled to the arrival of the presynaptic AP and 2) delayed asynchronous release that can last up to hundreds of milliseconds (497, 498). The relative contributions of these components to the total release depend on synapse type and can change with repetitive synapse activation during which asynchronous release typically builds up with increasing frequency and duration of stimulus trains.
At synapses between interneurons of the cerebellar molecular layer (MLIs), the time course of global [Ca2+]i determines the time course of delayed asynchronous release observed after stimulus trains (250). After 50-Hz stimulation, a barrage of asynchronous release events was observed that can last for up to ∼2.5 s after cessation of stimulation. This delayed release was strongly attenuated in PV−/− mice. This is consistent with the postulated effect of the slow buffer PV on global [Ca2+]i transients. In the presence of PV, the initial [Ca2+]i decay is sped up because of Ca2+ binding. The subsequent unloading of Ca2+ from PV during the decline of global [Ca2+]i generates a slowly decaying component of the [Ca2+]i transient that supports delayed asynchronous release during a time window between 0.5 and 2.5 s after cessation of stimulation. In PV−/− mice, the slow [Ca2+]i transient component is absent, and delayed asynchronous release is strongly reduced.
A similar scenario was described for hippocampal mossy fiber to CA3 pyramidal cell (hMF→CA3) synapses, where brief 20-Hz EPSC trains are followed by delayed asynchronous release lasting for ∼1–2 s (499). [Ca2+]i measurements in mossy fiber boutons revealed a slowly decaying time course of presynaptic global Ca2+. Treatment of the synapses with the membrane-permeable slow buffer EGTA-AM decreased the amplitude of delayed asynchronous release but prolonged its duration, reminiscent of what was observed in cerebellar MLI→MLI synapses of PV-expressing in comparison to PV-lacking mice (250).
Presynaptic terminals of glutamatergic endbulb→bushy cell synapses in the mammalian antero-ventral cochlear nucleus invariably express PV, but only a subset coexpress CR. Interestingly, postsynaptic neurons contacted by CR/PV-coexpressing terminals show lower rates of asynchronous release compared with those that are contacted by terminals void of presynaptic CR immunoreactivity (464). At glutamatergic calyx of Held→MNTB synapses, the magnitude of delayed asynchronous release observed after high-frequency stimulus trains strongly decreases during postnatal maturation. Developmental decrease in asynchronous release coincides with an upregulation of presynaptic CR expression, but virtually all calyces are PV immunoreactive already shortly after hearing onset (postnatal day 12) (463). Neither endbulb nor calyx terminals express CB-D28k. Whether this developmental downregulation of asynchronous release is related to differences in CR expression remains to be established. Nevertheless, these findings can be understood on the basis that CR is a fast buffer, which reduces the peak of the [Ca2+]i transient. During the [Ca2+]i decay, free and Ca2+-bound forms of CR are at equilibrium, such that Eqs. 30–32 hold. Although these equations predict that the area under the [Ca2+]i transient is unchanged in the presence of the buffer (FIGURE 7D), its amplitude is reduced (FIGURE 7, A AND C). For a cooperative process, such as Ca2+-triggered synaptic vesicle fusion, this implies a reduction of the overall effect of the [Ca2+]i transient.
Among hippocampal dentate gyrus inhibitory interneuron→principal neuron synapses, PV-expressing interneurons provide temporally precise inhibition, which is tightly synchronized to the timing of presynaptic AP firing, whereas cholecystokinin (CCK)-expressing interneurons release GABA in a less synchronized manner and exhibit prominent asynchronous release for a few hundreds of milliseconds after stimulus trains. CCK-expressing interneurons seem to lack expression of PV, CB-D28k, and CR. It is therefore conceivable that the differential expression of PV contributes to the mechanisms responsible for the ∼15-fold difference in the ratio of synchronous versus asynchronous release at these two synapses by accelerating the initial decay of AP-induced [Ca2+]i transients in PV-expressing interneurons and thereby reducing the number of asynchronously released quanta immediately following presynaptic APs (500).
9.7. Modulation of Excitability, AP Firing, and Network Activity
Neuronal membrane excitability and discharge properties are determined by an interplay of various voltage- and/or Ca2+-sensitive ion conductances. The amount of Ca2+ entering per AP is controlled by the AP duration, which needs to be precisely regulated to limit Ca2+ influx especially during repetitive AP firing. Many cortical neurons, including hippocampal granule cells (hGCs), express large-conductance Ca2+-sensitive potassium channels (BK channels), which contribute to AP termination. When recorded with pipettes containing 10 mM BAPTA, AP durations in hGCs increase similarly as observed after application of BK channel blockers, indicating that BAPTA uncouples BK channels from Ca2+ influx. Intracellular application of 10 mM EGTA does not affect AP kinetics, because BK channels are spatially tightly coupled to VGCCs. By quantitatively analyzing the extent of AP prolongation by different intracellular BAPTA concentrations, the mean diffusional distance for Ca2+ ions from VGCCs to BK channels was estimated to be ∼13 nm in hGCs (501). Such tight colocalization of the two types of channels together with low Ca2+ sensitivity of BK channels confines the activation of BK channels to short, well-timed episodes, which is required for fast AP repolarization. It also decouples channel activation from changes in global [Ca2+]i levels.
CR is the main cytosolic Ca2+ buffer in cerebellar granule cells (cGCs), which provide the major excitatory input to Purkinje cells via their parallel fibers. Recordings in cGCs of CR−/− mice showed that CR-deficient cGCs exhibit briefer APs compared with cGCs of wild-type mice. They also generate repetitive spike discharge with a steeper rise in firing frequency with increasing current injections. Addition of 0.15 mM BAPTA to the patch pipette solution restored normal excitability levels in CR-deficient cGCs, indicating that 0.15 mM BAPTA effectively mimics the contribution of CR to the cytosolic Ca2+-buffering power in cGCs of wild-type mice. A minimal single compartment-based model of cGC AP firing, which considers voltage- and Ca2+-sensitive conductances as well as Ca2+ buffering and extrusion, supports the conclusion that larger and faster-decaying AP-induced [Ca2+]i transients in CR-deficient cGCs are a consequence of reduced Ca2+ buffering, which enhances activation of Ca2+-sensitive BK potassium channels, speeds up AP repolarization, and thus produces shorter APs that enable faster discharge rates (502).
A similar modeling approach was used to study the role of the slow buffer PV for regulating the discharge properties of striatal fast-spiking (FS) interneurons (503). A conductance-based model that also includes Ca2+-sensitive small-conductance (SK) potassium channels and the presence of a Ca2+ buffer reproduces average firing frequencies and spike adaptation as observed during whole cell recordings in striatal FS interneurons in response to current injection. Higher concentrations of PV lead to elevated [Ca2+]i between consecutive AP-induced [Ca2+]i transients when Ca2+ dissociates from PV and thereby facilitate activation of the SK current. This increases the duration of the afterhyperpolarization (AHP) following each AP and thereby delays the next AP and reduces the firing frequency. Thus, variable concentrations of PV in the cytoplasm of striatal FS interneurons can modulate their intrinsic excitability and may potentially alter striatal information processing (503).
Neurons of the reticular thalamic nucleus (RTN) express high levels of PV and are characterized by low-threshold voltage-activated (LVA) Ca2+ currents. Based on the firing patterns observed in extracellular in vivo recordings, four types of neurons can be distinguished in the RTN: irregularly firing, medium bursting, long bursting, and tonically firing. Neurons of the medium-bursting type are more frequently observed than those of the long-bursting type in PV−/− mice. The generation of AP bursts involves Ca2+ influx thorough LVA VGCCs that subsequently activates Ca2+-dependent SK potassium channels. It is possible that a lack of PV in RTN neurons alters SK channel activation following Ca2+ influx and thereby affects their firing properties (504).
Cerebellar Purkinje cells generate two types of membrane discharges: simple and complex spikes. Simple spikes occur spontaneously or are triggered synaptically by parallel fiber input, whereas complex spikes are driven by climbing fiber input. Extracellular recordings in Purkinje cells from cerebella of adult calretinin-deficient (CR−/−) mice revealed a strongly enhanced spontaneous simple spike firing rate compared with wild-type mice, whereas mean spontaneous firing rates of complex spikes were unaltered. The duration of complex spikes was reduced in CR−/− PCs, as was the duration of pauses in simple spike firing following spontaneous complex spikes (505). A reduced complex spike duration may decrease Ca2+ influx and thence activation of Ca2+-sensitive potassium channels, leading to shorter pauses in simple spike firing that normally follows the complex spike (505). Motor coordination is impaired in CR−/− mice consistent with the role of the cerebellum in motor control and motor learning (506). Genetic ablation of both CR and CB-D28k (CR−/−CB-D28k −/−) induces 160-Hz local field potential oscillations in the cerebellar cortex of alert mice, with PCs firing simple spikes phase-locked to the oscillations. Since the intrinsic excitability of PCs is unaltered in CR−/−CB-D28k −/− mice but oscillations reversibly disappear when gap junctions or either GABAA or NMDA receptors are blocked, these 160-Hz oscillations are likely to emerge at the network level, demonstrating that changes in intracellular Ca2+ buffering in specific neuron types can alter network dynamics (507).
9.8. Modulation of Ca2+ Signaling in Dendritic Shafts and Spines
Dendritic Ca2+ signaling and its modulation by Ca2+ buffers has been studied experimentally by dye loading into dendrites of neocortical and hippocampal pyramidal neurons, cerebellar Purkinje cells and various interneurons, as well as by numerical simulations using deterministic or stochastic approaches. Dendritic [Ca2+]i transients in pyramidal neurons evoked by backpropagating APs have amplitudes of several hundreds of nanomolar and decay rapidly with time constants generally <100 ms at physiological temperature (441, 508, 509) (FIGURE 21A). Neocortical and hippocampal CA3 pyramidal neurons lack expression of PV, CR, and CB-D28K. Hippocampal CA1 pyramidal neurons are also void of PV and CR but contain ∼45 µM CB-D28K (444). For these neurons, estimates for the endogenous Ca2+ buffer power in dendrites ranged between 170 and 200, indicating that <1% of the total Ca2+ entering per AP remains free. During AP trains, [Ca2+]i increases to a steady-state level that depends linearly on the firing frequency. Thus, the dendritic [Ca2+]i level linearly encodes the frequency of APs in pyramidal neurons (441). Dendritic Ca2+ transients elicited by backpropagating APs in bitufted interneurons in layer 2/3 of the somatosensory cortex have a mean amplitude of ∼140 nM and decay with a slow time constant of ∼200 ms, consistent with their higher endogenous Ca2+ buffer power (∼285) compared with that of pyramidal neurons (511). [Ca2+]i transients measured in dendrites of hippocampal dentate gyrus basket cells (hDG BCs) show similar kinetics, with a mean decay time constant of ∼200 ms, but have smaller amplitudes, with a mean of only ∼40 nM. The endogenous Ca2+ buffer power in hippocampal dentate gyrus BC dendrites is ∼200. In these neurons, Ca2+ buffering power was estimated both during Ca2+ indicator loading and under steady-state conditions up to 20 min after establishing whole cell configuration, when mobile buffers are expected to be largely washed out. The similarity of the buffering power estimates obtained under these two conditions suggests that the Ca2+ buffering power in proximal apical dendrites of BCs is primarily determined by buffers that are resistant to washout (512). Similarly to pyramidal cells, dendritic [Ca2+]i transients in these two classes of cortical and hippocampal interneurons summate linearly during short AP bursts (511, 512) whereas dendrites of midbrain dopamine neurons of the substantia nigra exhibit a supralinear summation of single AP-evoked Ca2+ transients (513).
AP-induced [Ca2+]i transients in dendritic spines and shafts of Purkinje cells (PCs), which contain high concentrations of PV and CB-D28k, last considerably longer than those in pyramidal cell dendrites (510, 514) (FIGURE 21B). In both compartments, [Ca2+]i transients decay double exponentially with fast and slow time constants of ∼20–30 and ∼300–400 ms, respectively (510). Whereas amplitudes of the slowly decaying component are comparable for [Ca2+]i transients measured in dendritic shafts and spines, amplitudes of the fast-decaying components are considerably greater in spines, which largely accounts for the higher total amplitudes of their [Ca2+]i transients. To reproduce the rapid initial decay of [Ca2+]i transients in spines in numerical simulations, diffusional coupling had to be assumed, allowing free Ca2+ as well as all Ca2+-bound and free buffer species to diffuse between both compartments. The simulations further suggest that 1) neither CB-D28k nor PV saturates in spines or shafts during climbing fiber-evoked Ca2+ transients, 2) kon of CB’s medium-affinity binding is fast enough to reduce [Ca2+]i transient peaks, and 3) slow Ca2+ binding to PV and CB-D28k leads to biphasic decay of [Ca2+]i transients in dendritic shafts (510).
In cerebellar Purkinje cells, the functional consequences of genetic removal of one or both mobile Ca2+ buffers PV and CB-D28k were studied. Kinetic analysis of [Ca2+]i transients in PC dendrites of CB-D28k−/− mice (515, 516) and in spines and dendritic shafts of PCs of PV−/− and PV−/−CB-D28k−/− mice (510) shows that Ca2+ buffers contribute to sculpting amplitude and decay of [Ca2+]i transients. In CB-D28k−/− PCs, resting [Ca2+]i is unaltered but peak amplitudes of synaptically evoked dendritic [Ca2+]i transients associated with complex spikes are enhanced on average by >80% compared with wild-type PCs. This is mainly due to an increase of the fast, but not the slow, decay component of [Ca2+]i transients (515, 516). In PV−/− PCs, [Ca2+]i transients reach the same peak amplitudes as in wild-type animals, but the biphasic nature of the decay is less pronounced. When both PV and CB-D28k are removed (PV−/−CB-D28k−/−), peak amplitudes of [Ca2+]i transients are about two times higher than those in wild-type PCs, and their decay is nearly monophasic (510).
9.9. Ca2+ Buffers in Neurological and Neurodegenerative Diseases
It was originally considered that Ca2+ binding proteins may exert a neuroprotective function by protecting cells against damaging effects of excessively high [Ca2+]i during periods of strong activity (517). Distinct populations of neurons are selectively affected by different neurodegenerative diseases, which could be facilitated by a weakening of their Ca2+ buffering power because of intrinsically low or diminished expression of CBPs. Differential expression of CBPs in neuronal subpopulation could thus contribute to defining their susceptibility to disease processes. For example, motoneuron populations lost early in amyotrophic lateral sclerosis (ALS) lack CB-D28K and PV expression, whereas those damaged late or infrequently express higher levels of these CBPs (518). On the other hand, a substantial loss of neostriatal neurons containing CB-D28K has been reported for postmortem brain specimens from patients diagnosed with Huntington’s disease, indicating a selective vulnerability of these neuron populations (519). In general, the concept of CBPs primarily serving neuroprotective functions has received only limited experimental support. Rather, CBPs are essential components in neuronal Ca2+ homeostasis and signaling that primarily regulate the timing and spatial extent of Ca2+ signals. Deficiencies in one or more CBPs lead to distinct alterations in neuronal excitability and/or synaptic physiology that possibly also penetrate to the behavioral level.
Mice with genetic ablation of PV, CB-D28k, or CR expression are generally healthy and fertile and show no gross abnormalities in brain morphology (520). For example, CB-D28k−/− mice develop normally and have a normal anatomy and synaptic connectivity with no signs of cell death. However, CB-D28k-deficient mice show motor deficits. They develop a graded cerebellar ataxia, which is correlated with marked changes in the amplitude and kinetics of synaptically mediated [Ca2+]i transients, and display aberrant network activity (515). In CR−/− mice, gross development and many morphological, biochemical, and behavioral characteristics were found to be unaffected, but these animals show a motor discoordination that worsens dramatically with age (505, 521). In addition, long-term potentiation (LTP) but not basal synaptic transmission is impaired in CR−/− mice at synapses between the perforant pathway and granule cells in the dentate gyrus connecting entorhinal cortex with the hippocampal formation. Normal LTP can be restored in these animals by application of the GABAA receptor antagonist bicuculline, suggesting that in CR−/− dentate gyrus an excess of GABA release interferes with LTP induction. Thus, expression of CR contributes to the control of synaptic plasticity in mouse dentate gyrus by indirectly regulating the activity of GABAergic interneurons (522). Despite such roles of Ca2+ buffers in neuronal signaling, a complete absence of α-PV, CB-D28k, and CR has little impact on cochlear function and hearing when comparing distortion-product otoacoustic emissions and auditory brain stem responses between constitutive triple KO and wild-type mice (223). PV−/− and PV+/− mice show social behavior deficits with an autism spectrum disorder (ADS)-like phenotype including impairments in communication and repetitive and stereotyped patterns of behavior (523). This seems to relate to an absent or reduced PV expression in PV−/− and PV+/− mice, respectively, rather than to a loss of fast-spiking GABAergic interneurons (523).
We are unaware of human monogenic disorders causally linked to mutations in the CB-D28k, PV, or CR genes. However, altered expression of the CBPs PV, CB-D28k, or CR has been reported in the context of neurological diseases including ADS, dementia, epilepsy, and ataxia (524, 525). For example, a decrease in the levels of CB-D28K has been found in the cortex of brains of Alzheimer disease (AD) patients (526, 527). Vulnerability to AD also extends to PV-containing interneurons. A decrease in PV immunostaining was reported for parts of the entorhinal cortex when AD neuropathological markers are present. As the density of pathological markers in the entorhinal cortex becomes greater and more widespread, the decrease of PV immunostaining encompasses additional layers, even though some changes that are observed in PV-expressing interneurons may be linked to the fate of the projection neurons on which they synapse (528). AD disease-sensitive PV-containing inhibitory interneurons were further identified in the perirhinal cortex of AD patients (529). In the temporal cortex of AD patients the number of PV-immunoreactive somata was unchanged, but in layer II a decreased density of terminals of PV-expressing chandelier cells was observed, suggesting that not the interneurons themselves but only their terminals are decreased (530).
These findings raise the possibility that dysregulation of Ca2+ buffering power in certain neuron populations may contribute to the etiopathology. It is, however, challenging to differentiate between selective loss of neurons expressing a certain CBP and decreased protein expression. For example, for caldendrin, a member of the neuronal calcium-binding protein (nCaBP) family with rapid Ca2+-binding kinetics and high abundance in postsynaptic density of spine synapses, it was reported that in postmortem brains of subjects with chronic schizophrenia the number of caldendrin-immunoreactive neurons is significantly reduced in the left dorsolateral prefrontal cortex (531). Despite the reduced number of immunoreactive neurons, absolute caldendrin protein levels were elevated (532). A reduction of mRNAs encoding PV was found in the prefrontal cortex of subjects with schizophrenia. This was primarily due to a reduction in neuronal PV mRNA expression rather than a decreased density of PV mRNA-positive neurons. In contrast, the same measures of CR mRNA expression were not altered in schizophrenia (533).
10. EPITHELIA
Transport of Ca2+ across epithelia is important for the body’s Ca2+ homeostasis (for review see Ref. 534). To summarize, Ca2+ is absorbed by the intestine and then excreted from the body via the kidneys. Although the kidney may be a site of net Ca2+ loss from the body, it is important to remember that there is a balance between glomerular filtration and subsequent reabsorption, so Ca2+ (re)absorption occurs in both kidney and intestine. There are two routes for transepithelial Ca2+ transport. Some can occur via a paracellular route, through tight junctions between cells (535). Another important route involves transcellular transport with Ca2+ present in the lumen, entering the cell, along its electrochemical gradient, before being actively transported (by NCX and/or PMCA) at the basolateral membrane into the extracellular fluid and thence the blood.
Early work on rat duodenum found that calcium absorption comprised two components as characterized by their response to elevating luminal [Ca2+] and vitamin D. One was a saturating function of luminal [Ca2+] and was stimulated by vitamin D, whereas the other was proportional to [Ca2+] and independent of vitamin D (536, 537). The transport through the former pathway was found to be proportional to the concentration of a vitamin D-induced CBP (536, 538, 539). This CBP was subsequently shown to be comprised of the two calbindins (CBs) CB-D9k and CB-D28k (for review see Ref. 540). CB-D9k has two Ca2+ binding sites, and CB-D28k has four (30, 186) (see TABLES 2 and 3). In general, tissues only express one of the two CBs, for example CB-D9k in the duodenum and CB-D28k in the kidney.
A major challenge for the transport mechanism described above is that, to move between the apical and basolateral membranes, Ca2+ must traverse the cytoplasm, where [Ca2+]i is very low. It was calculated that the diffusion rate of free Ca2+ would be ∼70 times smaller than the experimentally measured transepithelial flux (541). Therefore the role of CBs is to allow rapid diffusion of Ca2+ across the cytoplasm while maintaining [Ca2+]i low. As pointed out previously (174, 200, 209) (see also sect. 4.5), there are two opposing factors. 1) Since CB has a larger MW than Ca2+, the diffusion rate of an individual Ca2+ bound to CB will be much slower, in proportion to the 0.33th power of the ratio of molecular weights, than that of a free Ca2+ ion. Calcium has an atomic weight of 40, so Ca2+ bound to CB-D9k will diffuse at a speed of about (40/9000)0.33 ≈ 16% of that of free Ca2+. For Ca2+ bound to CB-D28k, the corresponding value is ∼12%. This effect is greatly outweighed by the fact that the concentration of Ca2+ bound to CB is much higher than that of free Ca2+ so the net effect is a marked increase in the rate of Ca2+ diffusion (≈50-fold with the values of TABLE 3). It should also be noted that free Ca2+ ions will bind to fixed charges in the cell, slowing their diffusion. Charge neutralization by binding to CB will thereby accelerate diffusion. We could not find measurements of the contribution of fixed buffers in epithelia, but, given the high concentration of CBs, it is probable that Ca2+-transporting epithelia represent a case where the majority of buffering comes from mobile rather than fixed buffers.
Evidence suggests that CB-D9k plays the major role as the CBP in the intestine. A similar mechanism involving CB is also important for renal reabsorption of Ca2+ in the distal convoluted tubule (DCT), but here CB-D28k is the major player. Indeed, the appearance of CB in urine has been proposed as a biomarker for damage to the distal nephron (542). PV is also expressed in the cells of the DCT (543). Knockout of PV resulted in diuresis. PV also appeared to act as a Ca2+ buffer and thereby reduced the amplitude of intracellular Ca2+ signals, suggesting that the effects of PV are as an intracellular Ca2+ buffer modulating Ca2+ signaling as opposed to transport (544). It has also been shown that most of the PV in the kidney is located in the early DCT, a region where there is little Ca2+ transport (543, 545). In other sections we have discussed the importance of the kinetics of PV in shaping Ca2+ signals. Given the slow kinetics of changes of [Ca2+]i in the DCT (544), it is unclear whether these kinetics are relevant. It is, of course, possible that fast changes of [Ca2+]i in a subcompartment of the cytoplasm are affected by PV.
Work showing that CB-D9k is localized to the basolateral membrane in the rat distal nephron has also led to the suggestion that it may regulate Ca2+ transport as well as buffering (543). The entry of Ca2+ along its electrochemical gradient from lumen into cell occurs in many epithelia through TRPV5 channels, with the Ca2+ subsequently binding to CB and being transported across the cell. It has been shown that the expressions of CB and TRPV5 are coordinated (546). As well as being present in the cytoplasm, a fraction of the CB is localized on the apical membrane, and this apical fraction is decreased in TRPV5 knockout mice. It has been suggested that Ca2+ buffering by CB in the vicinity of the TRPV5 channel increases influx through this channel.
Mg2+ is also reabsorbed by the DCT, and it has been suggested that it enters the cell from the lumen via a TRPM6/7 channel (547). That work showed that the TRPM channel was located close to either PV or CB-D28k, and it was suggested that binding of Mg2+ to buffers may be important in transepithelial Mg2+ transport (548). As pointed out in sect. 2.2, many Ca2+ buffers can also bind Mg2+. The major Mg2+ buffer is probably ATP. The relative molecular weights (atomic mass of Mg2+ is 24 and MW of ATP is 500) mean that the bound form diffuses at (24/500)0.33 = 0.37 of the free. Given a free Mg2+ of 1 mM, a total ATP concentration of 5 mM and Kd of 100 µM for Mg2+ binding to ATP, we obtain from Eq. 27 that the flux of bound will be only 12% of that at the free. The contribution to Mg2+ diffusion of protein buffers, with higher molecular weights and lower concentrations than ATP, will be even smaller. Consistent with this, knockout of CB-D28k does not affect Mg2+ balance (549).
11. IMMUNE CELLS
Early studies using the Ca2+ indicator quin-2 found that the intrinsic buffering in human neutrophils could be represented by ∼0.76 mM of a buffer with Kd of 0.55 µM (550), equivalent to a buffer power of ∼1,000 at low [Ca2+]i. Neutrophils and monocytes were then shown to contain CBPs of the S100 family (551). The major components are migration inhibitory factor-related proteins 8 and 14 (MRP8 and MRP14), more commonly known, respectively, as S100A8 and S100A9, which form a dimer known as calprotectin. Two such dimers can form a tetramer, particularly at elevated [Ca2+], and this tetramerization is essential for biological function (552). Calprotectin can also bind Mn2+ and Zn2+, and it has been shown that release of calprotectin into the extracellular fluid has an antibacterial action by chelating these ions (553–555). Mn2+ is required for the function of bacterial superoxide dismutase as well as a variety of other bacterial enzymes, and Zn2+ is required for several bacterial enzymes (556). It appears that Ca2+ binding to the EF hands on calprotectin increases the affinity for Zn2+. Consequently, the affinity for Zn2+ and Mn2+ will be much greater outside, where Ca2+ is higher, than inside the cell, thereby allowing calprotectin to avidly bind Mn2+ or Zn2+ once it is released into the extracellular fluid where the bacteria are present (557, 558). The sequestration of extracellular Mn2+ and Zn2+ will therefore deprive bacteria of these essential metal ions, thereby preventing their growth (556).
Activated macrophages have been shown to release β-PV (oncomodulin), and this stimulates the outgrowth of axons from retinal ganglion cells (559). This effect could not be mimicked by other CBPs, suggesting that it results from an action of oncomodulin other than Ca2+ binding. Subsequent studies have also shown that oncomodulin released from neutrophils has similar effects (560).
A very different role of Ca2+ buffers has been suggested as a defense against West Nile virus. Infection of a cell by this virus has been shown to increase [Ca2+]i (561). Subsequent work found that CB-D28k decreased viral replication in cultured cells (562). A study with another virus (Borna disease virus) found that infection increased expression of CB-D28k in submucous and myenteric neurons (563). Although the effects of CB on viral invasion may be due to effects on [Ca2+]i, direct measurements are required to confirm this.
12. CONCLUSIONS
The physiological properties of Ca2+ buffers have been studied in a variety of tissues. Their biochemical parameters such as ion specificity, binding kinetics, affinity, and cooperativity of binding generate a range of properties required for optimal cell function. Endogenous and exogenous Ca2+ buffers shape local and global [Ca2+]i transients in predictable ways in different cell compartments. Together with Ca2+ influx and extrusion, the specific properties of Ca2+-binding molecules enable cells to encode activity using the universal second messenger Ca2+ and relay signals to specific intracellular pathways. By shaping and confining local [Ca2+]i domains, Ca2+ buffering allows local signaling at spatially separate sites to coexist with global Ca2+ actions.
Some Ca2+-binding proteins serve a dual role as detectors of Ca2+ signals and Ca2+ buffers, whereas others act mainly as buffers shaping the spatiotemporal properties of intracellular Ca2+ signals. Ca2+ buffers can sculpt intracellular Ca2+ signaling in many ways. They can accelerate or retard the diffusion of Ca2+ in the cell, sharpen the spatial profile of local Ca2+ domains that build up during transmembrane Ca2+ flux, and thereby contribute to isolating local from global Ca2+ signaling pathways.
Each Ca2+ buffer can be used in several cell types. For example, the calbindins facilitate both epithelial Ca2+ transport and neuronal signaling, and the slow kinetics of parvalbumin are used to selectively accelerate the decay of Ca2+ signals in both neurons and skeletal muscle without interfering with their rise and amplitude. There are, however, tissue-dependent differences in buffering. Cardiac muscle is at one extreme, with most of the buffering provided by the nondiffusible buffers TnC and SERCA, essential for excitation-contraction coupling, with only a small contribution from molecules whose major function is to buffer. Skeletal muscle is broadly similar but also makes use of the diffusible buffer parvalbumin. In contrast, in neurons a high cytosolic Ca2+ buffering power is often achieved by expressing high amounts of one or two species of mobile Ca2+ buffer proteins. In general, neurons with low Ca2+ buffering power exhibit relatively fast decays of global [Ca2+]i transients, whereas those with high buffer power, as a consequence of expression of high amounts of Ca2+ binding proteins, exhibit slower or biphasic global [Ca2+]i transients. In neurons, this type of multiphasic kinetics of [Ca2+]i transients allows cells to shape responses to action potentials in cell-specific ways and thereby produce a variety of short-term plasticity patterns. Slow tails of [Ca2+]i transients build up during repetitive activity and may support asynchronous transmitter release and enable various forms of synaptic plasticity at intermediate timescales. Understanding of the quantitative effects of buffering has advanced less for many other tissues than is the case in nerve and muscle, and more research is required. For now, we will have to be satisfied with knowing which buffers are present and applying the basic principles learned from work on excitable cells.
A major area for future work concerns the role of changes of Ca2+ buffering in disease. Although there is evidence for such a link in cardiac muscle, it remains to be seen what the situation is in other tissues.
GRANTS
D.E. was funded by a British Heart Foundation Chair (CH/2000004/12801) and Grant FS/CRTF/21/24140. E.N.’s and H.T.’s work is supported by the German Research Foundation (DFG), Collaborative Research Center 1286 “Quantitative Synaptology.” E.N. acknowledges funding by the DFG under Germany’s Excellence Strategy-EXC 2067/1-390729940. G.S. was funded by British Heart Foundation Grant PG/19/55/34545.
DISCLOSURES
No conflicts of interest, financial or otherwise, are declared by the authors.
AUTHOR CONTRIBUTIONS
D.E., E.N., H.T., and G.S. conceived and designed research; D.E., H.T., and G.S. prepared figures; D.E., E.N., H.T., and G.S. drafted manuscript; D.E., E.N., H.T., and G.S. edited and revised manuscript; D.E., E.N., H.T., and G.S. approved final version of manuscript.
ACKNOWLEDGMENTS
We are indebted to the following colleagues for helpful discussion: Jorge Amich, Stephen Baylor, Rene Bindels, Francis Burton, Virgilio Lew, Kun-Han Lin, Elizabeth Murphy, and Eduardo Rios. The graphical abstract was produced with BioRender.
REFERENCES
- 1. Steinhardt R, Zucker R, Schatten G. Intracellular calcium release at fertilization in the sea urchin egg. Dev Biol 58: 185–196, 1977. doi: 10.1016/0012-1606(77)90084-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2. Szalai G, Krishnamurthy R, Hajnóczky G. Apoptosis driven by IP3-linked mitochondrial calcium signals. EMBO J 18: 6349–6361, 1999. doi: 10.1093/emboj/18.22.6349. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3. Berridge MJ, Lipp P, Bootman MD. The versatility and universality of calcium signalling. Nat Rev Mol Cell Biol 1: 11–21, 2000. doi: 10.1038/35036035. [DOI] [PubMed] [Google Scholar]
- 4. Case RM, Eisner D, Gurney A, Jones O, Muallem S, Verkhratsky A. Evolution of calcium homeostasis: from birth of the first cell to an omnipresent signalling system. Cell Calcium 42: 345–350, 2007. doi: 10.1016/j.ceca.2007.05.001. [DOI] [PubMed] [Google Scholar]
- 5. Petersen OH, Gerasimenko JV, Gerasimenko OV, Gryshchenko O, Peng S. The roles of calcium and ATP in the physiology and pathology of the exocrine pancreas. Physiol Rev 101: 1691–1744, 2021. doi: 10.1152/physrev.00003.2021. [DOI] [PubMed] [Google Scholar]
- 6. Stafford N, Wilson C, Oceandy D, Neyses L, Cartwright EJ. The plasma membrane calcium ATPases and their role as major new players in human disease. Physiol Rev 97: 1089–1125, 2017. doi: 10.1152/physrev.00028.2016. [DOI] [PubMed] [Google Scholar]
- 7. Enomoto M, Nishikawa T, Siddiqui N, Chung S, Ikura M, Stathopulos PB. From stores to sinks: structural mechanisms of cytosolic calcium regulation. Adv Exp Med Biol 981: 215–251, 2017. doi: 10.1007/978-3-319-55858-5_10. [DOI] [PubMed] [Google Scholar]
- 8. Berridge MJ. The inositol trisphosphate/calcium signaling pathway in health and disease. Physiol Rev 96: 1261–1296, 2016. doi: 10.1152/physrev.00006.2016. [DOI] [PubMed] [Google Scholar]
- 9. Eisner DA, Caldwell JL, Kistamás K, Trafford AW. Calcium and excitation-contraction coupling in the heart. Circ Res 121: 181–195, 2017. doi: 10.1161/CIRCRESAHA.117.310230. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10. Woll KA, Van Petegem F. Calcium-release channels: structure and function of IP3 receptors and ryanodine receptors. Physiol Rev 102: 209–268, 2022. doi: 10.1152/physrev.00033.2020. [DOI] [PubMed] [Google Scholar]
- 11. Koppel M, Spiro K. Ober die Wirkung von Moderatoren (Puffern) bei der Verschiebung des Säure-Basengleichgewichtes in biologischen Fliissigkeiten. Biochem Ztschr 65: 409–439, 1914. [Google Scholar]
- 12. Roos A, Boron WF. The buffer value of weak acids and bases: origin of the concept, and first mathematical derivation and application to physico-chemical systems. The work of M. Koppel and K. Spiro (1914). Respir Physiol 40: 1–32, 1980. doi: 10.1016/0034-5687(80)90002-x. [DOI] [PubMed] [Google Scholar]
- 13. Van Slyke DD. On the measurement of buffer values and on the relationship of buffer value to the dissociation constant of the buffer and the concentration and reaction of the buffer solution. J Biol Chem 52: 525–570, 1922. doi: 10.1016/S0021-9258(18)85845-8. [DOI] [Google Scholar]
- 14. Hodgkin AL, Keynes RD. Movements of labelled calcium in squid giant axons. J Physiol 138: 253–281, 1957. doi: 10.1113/jphysiol.1957.sp005850. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15. Brinley FJ. Calcium buffering in squid axons. Annu Rev Biophys Bioeng 7: 363–392, 1978. doi: 10.1146/annurev.bb.07.060178.002051. [DOI] [PubMed] [Google Scholar]
- 16. Hewish NA, Neilson GW, Enderby JE. Environment of Ca2+ ions in aqueous solvent. Nature 297: 138–139, 1982. doi: 10.1038/297138a0. [DOI] [Google Scholar]
- 17. Gifford JL, Walsh MP, Vogel HJ. Structures and metal-ion-binding properties of the Ca2+-binding helix-loop-helix EF-hand motifs. Biochem J 405: 199–221, 2007. doi: 10.1042/BJ20070255. [DOI] [PubMed] [Google Scholar]
- 18. Dudev T, Lim C. Principles governing Mg, Ca, and Zn binding and selectivity in proteins. Chem Rev 103: 773–788, 2003. doi: 10.1021/cr020467n. [DOI] [PubMed] [Google Scholar]
- 19. Dochia M, Stănescu M, Constantin C. Calcium content indicator of scouring efficiency. Fibres Text East Eur 21: 22–25, 2013. [Google Scholar]
- 20. Park H, Wu S, Dunker AK, Kang C. Polymerization of calsequestrin. Implications for Ca2+ regulation. J Biol Chem 278: 16176–16182, 2003. doi: 10.1074/jbc.M300120200. [DOI] [PubMed] [Google Scholar]
- 21. Eigen M. Fast elementary steps in chemical reaction mechanisms. Pure Appl Chem 6: 97–116, 1963. doi: 10.1351/pac196306010097. [DOI] [Google Scholar]
- 22. Kula RJ, Reed GH. Nuclear magnetic resonance investigation of ligand exchange kinetics in the calcium(II)-EDTA system. Anal Chem 38: 697–701, 1966. doi: 10.1021/ac60238a007. [DOI] [Google Scholar]
- 23. Smith GL, Miller DJ. Potentiometric measurements of stoichiometric and apparent affinity constants of EGTA for protons and divalent ions including calcium. Biochim Biophys Acta 839: 287–299, 1985. doi: 10.1016/0304-4165(85)90011-x. [DOI] [PubMed] [Google Scholar]
- 24. Mirti P. Kinetics of ligand exchange and dissociation reactions of the calcium(II)-EGTA complex investigated by the NMR technique. J Inorg Nucl Chem 41: 323–330, 1979. doi: 10.1016/0022-1902(79)80141-4. [DOI] [Google Scholar]
- 25. Naraghi M. T-jump study of calcium binding kinetics of calcium chelators. Cell Calcium 22: 255–268, 1997. doi: 10.1016/s0143-4160(97)90064-6. [DOI] [PubMed] [Google Scholar]
- 26. White HD. Kinetic mechanism of calcium binding to whiting parvalbumin. Biochemistry 27: 3357–3365, 1988. doi: 10.1021/bi00409a036. [DOI] [PubMed] [Google Scholar]
- 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 525: 419–432, 2000. doi: 10.1111/j.1469-7793.2000.t01-2-00419.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28. Ogawa Y, Tanokura M. Steady-state properties of calcium binding to parvalbumins from bullfrog skeletal muscle: effects of Mg2+, pH, ionic strength, and temperature. J Biochem 99: 73–80, 1986. doi: 10.1093/oxfordjournals.jbchem.a135481. [DOI] [PubMed] [Google Scholar]
- 29. Eberhard M, Erne P. Calcium and magnesium binding to rat parvalbumin. Eur J Biochem 222: 21–26, 1994. doi: 10.1111/j.1432-1033.1994.tb18836.x. [DOI] [PubMed] [Google Scholar]
- 30. Nagerl UV, Novo D, Mody I, Vergara JL. Binding kinetics of calbindin-D28k determined by flash photolysis of caged Ca2+. Biophys J 79: 3009–3018, 2000. doi: 10.1016/S0006-3495(00)76537-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31. Berggård T, Miron S, Onnerfjord P, Thulin E, Akerfeldt KS, Enghild JJ, Akke M, Linse S. Calbindin D28k exhibits properties characteristic of a Ca2+ sensor. J Biol Chem 277: 16662–16672, 2002. doi: 10.1074/jbc.M200415200. [DOI] [PubMed] [Google Scholar]
- 32. Davis JP, Rall JA, Reiser PJ, Smillie LB, Tikunova SB. Engineering competitive magnesium binding into the first EF-hand of skeletal troponin C. J Biol Chem 277: 49716–49726, 2002. doi: 10.1074/jbc.M208488200. [DOI] [PubMed] [Google Scholar]
- 33. Martell AE, Smith RM. Critical Stability Constants. New York: Plenum, 1974. [Google Scholar]
- 34. Baylor SM, Hollingworth S. Model of sarcomeric Ca2+ movements, including ATP Ca2+ binding and diffusion, during activation of frog skeletal muscle. J Gen Physiol 112: 297–316, 1998. doi: 10.1085/jgp.112.3.297. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35. Tsien RY. New calcium indicators and buffers with high selectivity against magnesium and protons: design, synthesis, and properties of prototype structures. Biochemistry 19: 2396–2404, 1980. doi: 10.1021/bi00552a018. [DOI] [PubMed] [Google Scholar]
- 36. Grynkiewicz G, Poenie M, Tsien RY. A new generation of Ca2+ indicators with greatly improved fluorescence properties. J Biol Chem 260: 3440–3450, 1985. doi: 10.1016/S0021-9258(19)83641-4. [DOI] [PubMed] [Google Scholar]
- 37. Portzehl H, Caldwell PC, Rueegg JC. The dependence of contraction and relaxation of muscle fibres from the crab maia squinado on the internal concentration of free calcium ions. Biochim Biophys Acta 79: 581–591, 1964. doi: 10.1016/0926-6577(64)90224-4. [DOI] [PubMed] [Google Scholar]
- 38. Blaustein MP, Hodgkin AL. The effect of cyanide on the efflux of calcium from squid axons. J Physiol 200: 497–527, 1969. doi: 10.1113/jphysiol.1969.sp008704. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39. Gurney AM, Tsien RY, Lester HA. Activation of a potassium current by rapid photochemically generated step increases of intracellular calcium in rat sympathetic neurons. Proc Natl Acad Sci USA 84: 3496–3500, 1987. doi: 10.1073/pnas.84.10.3496. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40. Ellis-Davies GC, Kaplan JH. Nitrophenyl-EGTA, a photolabile chelator that selectively binds Ca2+ with high affinity and releases it rapidly upon photolysis. Proc Natl Acad Sci USA 91: 187–191, 1994. doi: 10.1073/pnas.91.1.187. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41. Ellis-Davies GC, Kaplan JH, Barsotti RJ. Laser photolysis of caged calcium: rates of calcium release by nitrophenyl-EGTA and DM-nitrophen. Biophys J 70: 1006–1016, 1996. doi: 10.1016/S0006-3495(96)79644-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42. Mulligan IP, Ashley CC. Rapid relaxation of single frog skeletal muscle fibres following laser flash photolysis of the caged calcium chelator, diazo-2. FEBS Lett 255: 196–200, 1989. doi: 10.1016/0014-5793(89)81090-7. [DOI] [PubMed] [Google Scholar]
- 43. Bers DM. A simple method for the accurate determination of free [Ca] in Ca-EGTA solutions. Am J Physiol Cell Physiol 242: C404–C408, 1982. doi: 10.1152/ajpcell.1982.242.5.C404. [DOI] [PubMed] [Google Scholar]
- 44. Miller DJ, Smith GL. EGTA purity and the buffering of calcium ions in physiological solutions. Am J Physiol Cell Physiol 246: C160–C166, 1984. doi: 10.1152/ajpcell.1984.246.1.C160. [DOI] [PubMed] [Google Scholar]
- 45. Choppin GR, Shanbhag PM. Binding of calcium by humic acid. J Inorg Nucl Chem 43: 921–922, 1981. doi: 10.1016/0022-1902(81)80150-9. [DOI] [Google Scholar]
- 46. Tipping E, Hurley MA. A unifying model of cation binding by humic substances. Geochim Cosmochim Acta 56: 3627–3641, 1992. doi: 10.1016/0016-7037(92)90158-F. [DOI] [Google Scholar]
- 47. Tang N, Skibsted LH. Calcium binding to amino acids and small glycine peptides in aqueous solution: toward peptide design for better calcium bioavailability. J Agric Food Chem 64: 4376–4389, 2016. doi: 10.1021/acs.jafc.6b01534. [DOI] [PubMed] [Google Scholar]
- 48. Guéguen L, Pointillart A. The bioavailability of dietary calcium. J Am Coll Nutr 19: 119s–136s, 2000. doi: 10.1080/07315724.2000.10718083. [DOI] [PubMed] [Google Scholar]
- 49. Swietach P, Youm JB, Saegusa N, Leem CH, Spitzer KW, Vaughan-Jones RD. Coupled Ca2+/H+ transport by cytoplasmic buffers regulates local Ca2+ and H+ ion signaling. Proc Natl Acad Sci USA 110: E2064–E2073, 2013. doi: 10.1073/pnas.1222433110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50. Lenz GR, Martell AE. Metal complexes of carnosine. Biochemistry 3: 750–753, 1964. doi: 10.1021/bi00894a002. [DOI] [PubMed] [Google Scholar]
- 51. Baran EJ. Metal complexes of carnosine. Biochemistry (Mosc) 65: 789–797, 2000. [PubMed] [Google Scholar]
- 52. Abate C, Cassone G, Cordaro M, Giuffre O, Mollica-Nardo V, Ponterio RC, Saija F, Sponer J, Trusso S, Foti C. Understanding the behaviour of carnosine in aqueous solution: an experimental and quantum-based computational investigation on acid-base properties and complexation mechanisms with Ca2+ and Mg2+. New J Chem 45: 20352–20364, 2021. doi: 10.1039/D1NJ04094D. [DOI] [Google Scholar]
- 53. Melcrová A, Pokorna S, Pullanchery S, Kohagen M, Jurkiewicz P, Hof M, Jungwirth P, Cremer PS, Cwiklik L. The complex nature of calcium cation interactions with phospholipid bilayers. Sci Rep 6: 38035, 2016. doi: 10.1038/srep38035. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54. Post JA, Langer GA. Sarcolemmal calcium binding sites in heart: I. Molecular origin in “gas-dissected” sarcolemma. J Membr Biol 129: 49–57, 1992. doi: 10.1007/BF00232054. [DOI] [PubMed] [Google Scholar]
- 55. Shannon TR, Wang F, Puglisi J, Weber C, Bers DM. A mathematical treatment of integrated Ca dynamics within the ventricular myocyte. Biophys J 87: 3351–3371, 2004. doi: 10.1529/biophysj.104.047449. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56. Bers DM, Philipson KD, Peskoff A. Calcium at the surface of cardiac plasma membrane vesicles: cation binding, surface charge screening, and Na-Ca exchange. J Membr Biol 85: 251–261, 1985. doi: 10.1007/BF01871520. [DOI] [PubMed] [Google Scholar]
- 57. Dudev T, Lim C. Competition between protein ligands and cytoplasmic inorganic anions for the metal cation: a DFT/CDM study. J Am Chem Soc 128: 10541–10548, 2006. doi: 10.1021/ja063111s. [DOI] [PubMed] [Google Scholar]
- 58. Lenz RA, Pitler TA, Alger BE. High intracellular Cl- concentrations depress G-protein-modulated ionic conductances. J Neurosci 17: 6133–6141, 1997. doi: 10.1523/JNEUROSCI.17-16-06133.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59. Matthews EA, Schoch S, Dietrich D. Tuning local calcium availability: cell-type-specific immobile calcium buffer capacity in hippocampal neurons. J Neurosci 33: 14431–14445, 2013. doi: 10.1523/JNEUROSCI.4118-12.2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60. Woehler A, Lin KH, Neher E. Calcium-buffering effects of gluconate and nucleotides, as determined by a novel fluorimetric titration method. J Physiol 592: 4863–4875, 2014. doi: 10.1113/jphysiol.2014.281097. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61. Schwaller B. Cytosolic Ca2+ buffers. Cold Spring Harb Perspect Biol 2: a004051, 2010. doi: 10.1101/cshperspect.a004051. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62. Schwaller B. Calretinin: from a “simple” Ca2+ buffer to a multifunctional protein implicated in many biological processes. Front Neuroanat 8: 3, 2014. doi: 10.3389/fnana.2014.00003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63. Elíes J, Yáñez M, Pereira TM, Gil-Longo J, MacDougall DA, Campos-Toimil M. An update to calcium binding proteins. Adv Exp Med Biol 1131: 183–213, 2020. doi: 10.1007/978-3-030-12457-1_8. [DOI] [PubMed] [Google Scholar]
- 64. Schwaller B. Cytosolic Ca2+ buffers are inherently Ca2+ signal modulators. Cold Spring Harb Perspect Biol 12: a035543, 2020. doi: 10.1101/cshperspect.a035543. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65. Saroff HA, Lewis MS. The binding of calcium ions to serum albumin. J Phys Chem 67: 1211–1216, 1963. doi: 10.1021/j100800a011. [DOI] [Google Scholar]
- 66. Schwaller B. The continuing disappearance of “pure” Ca2+ buffers. Cell Mol Life Sci 66: 275–300, 2009. doi: 10.1007/s00018-008-8564-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67. Murakoshi H, Shin ME, Parra-Bueno P, Szatmari EM, Shibata AC, Yasuda R. Kinetics of endogenous CaMKII required for synaptic plasticity revealed by optogenetic kinase inhibitor. Neuron 94: 37–47.e5, 2017. doi: 10.1016/j.neuron.2017.02.036. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68. Gonzalez LL, Garrie K, Turner MD. Role of S100 proteins in health and disease. Biochim Biophys Acta Mol Cell Res 1867: 118677, 2020. doi: 10.1016/j.bbamcr.2020.118677. [DOI] [PubMed] [Google Scholar]
- 69. Gilston BA, Skaar EP, Chazin WJ. Binding of transition metals to S100 proteins. Sci China Life Sci 59: 792–801, 2016. doi: 10.1007/s11427-016-5088-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70. Kretsinger RH, Nockolds CE. Carp muscle calcium-binding protein: II. Structure determination and general description. J Biol Chem 248: 3313–3326, 1973. doi: 10.1016/S0021-9258(19)44043-X. [DOI] [PubMed] [Google Scholar]
- 71. Kawasaki H, Kretsinger RH. Structural and functional diversity of EF-hand proteins: Evolutionary perspectives. Protein Sci 26: 1898–1920, 2017. doi: 10.1002/pro.3233. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72. Evenäs J, Malmendal A, Thulin E, Carlström G, Forsén S. Ca2+ binding and conformational changes in a calmodulin domain. Biochemistry 37: 13744–13754, 1998. doi: 10.1021/bi9806448. [DOI] [PubMed] [Google Scholar]
- 73. Mäler L, Blankenship J, Rance M, Chazin WJ. Site-site communication in the EF-hand Ca2+-binding protein calbindin D9k. Nat Struct Biol 7: 245–250, 2000. doi: 10.1038/73369. [DOI] [PubMed] [Google Scholar]
- 74. Grabarek Z. Insights into modulation of calcium signaling by magnesium in calmodulin, troponin C and related EF-hand proteins. Biochim Biophys Acta 1813: 913–921, 2011. doi: 10.1016/j.bbamcr.2011.01.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75. Malmendal A, Linse S, Evenäs J, Forsén S, Drakenberg T. Battle for the EF-hands: magnesium-calcium interference in calmodulin. Biochemistry 38: 11844–11850, 1999. doi: 10.1021/bi9909288. [DOI] [PubMed] [Google Scholar]
- 76. Rayani K, Seffernick J, Li AY, Davis JP, Spuches AM, Van Petegem F, Solaro RJ, Lindert S, Tibbits GF. Binding of calcium and magnesium to human cardiac troponin C. J Biol Chem 296: 100350, 2021. doi: 10.1016/j.jbc.2021.100350. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77. Martin SR, Masino L, Bayley PM. Enhancement by Mg2+ of domain specificity in Ca2+-dependent interactions of calmodulin with target sequences. Protein Sci 9: 2477–2488, 2000. doi: 10.1110/ps.9.12.2477. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78. Robertson SP, Johnson JD, Potter JD. The time-course of Ca2+ exchange with calmodulin, troponin, parvalbumin, and myosin in response to transient increases in Ca2+. Biophys J 34: 559–569, 1981. doi: 10.1016/S0006-3495(81)84868-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79. Pan BS, Solaro RJ. Calcium-binding properties of troponin C in detergent-skinned heart muscle fibers. J Biol Chem 262: 7839–7849, 1987. doi: 10.1016/S0021-9258(18)47644-2. [DOI] [PubMed] [Google Scholar]
- 80. Shannon TR, Ginsburg KS, Bers DM. Reverse mode of the sarcoplasmic reticulum calcium pump and load-dependent cytosolic calcium decline in voltage-clamped cardiac ventricular myocytes. Biophys J 78: 322–333, 2000. doi: 10.1016/S0006-3495(00)76595-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81. Bers DM. Excitation-Contraction Coupling and Cardiac Contractile Force. Dordrecht, The Netherlands: Kluwer Academic, 2001. [Google Scholar]
- 82. Smith GL, Eisner DA. Calcium buffering in the heart in health and disease. Circulation 139: 2358–2371, 2019. doi: 10.1161/CIRCULATIONAHA.118.039329. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83. Baylor SM, Hollingworth S. Simulation of Ca2+ movements within the sarcomere of fast-twitch mouse fibers stimulated by action potentials. J Gen Physiol 130: 283–302, 2007. doi: 10.1085/jgp.200709827. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84. Baylor SM, Hollingworth S. Intracellular calcium movements during excitation-contraction coupling in mammalian slow-twitch and fast-twitch muscle fibers. J Gen Physiol 139: 261–272, 2012. doi: 10.1085/jgp.201210773. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85. Song Q, Saucerman JJ, Bossuyt J, Bers DM. Differential integration of Ca2+-calmodulin signal in intact ventricular myocytes at low and high affinity Ca2+-calmodulin targets. J Biol Chem 283: 31531–31540, 2008. doi: 10.1074/jbc.M804902200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86. Romano DR, Pharris MC, Patel NM, Kinzer-Ursem TL. Competitive tuning: competition's role in setting the frequency-dependence of Ca2+-dependent proteins. PLoS Comput Biol 13: e1005820, 2017. doi: 10.1371/journal.pcbi.1005820. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87. Gillis JM, Thomason D, Lefèvre J, Kretsinger RH. Parvalbumins and muscle relaxation: a computer simulation study. J Muscle Res Cell Motil 3: 377–398, 1982. doi: 10.1007/BF00712090. [DOI] [PubMed] [Google Scholar]
- 88. Kretsinger RH, Schaffer JE. Calcium | calcium-modulated proteins (EF-hand). In: Encyclopedia of Biological Chemistry III (3rd ed.), edited by Jez J. Oxford: Elsevier, 2021, p. 630–636. doi: 10.1016/B978-0-12-819460-7.00153-5. [DOI] [Google Scholar]
- 89. Honoré B, Vorum H. The CREC family, a novel family of multiple EF-hand, low-affinity Ca2+-binding proteins localised to the secretory pathway of mammalian cells. FEBS Lett 466: 11–18, 2000. doi: 10.1016/s0014-5793(99)01780-9. [DOI] [PubMed] [Google Scholar]
- 90. Li MX, Hwang PM. Structure and function of cardiac troponin C (TNNC1): implications for heart failure, cardiomyopathies, and troponin modulating drugs. Gene 571: 153–166, 2015. doi: 10.1016/j.gene.2015.07.074. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91. Kesvatera T, Jönsson B, Telling A, Tõugu V, Vija H, Thulin E, Linse S. Calbindin D9k: a protein optimized for calcium binding at neutral pH. Biochemistry 40: 15334–15340, 2001. doi: 10.1021/bi0114022. [DOI] [PubMed] [Google Scholar]
- 92. Johnson RA, Fulcher LM, Vang K, Palmer CD, Grossoehme NE, Spuches AM. In depth, thermodynamic analysis of Ca2+ binding to human cardiac troponin C: extracting buffer-independent binding parameters. Biochim Biophys Acta Proteins Proteom 1867: 359–366, 2019. doi: 10.1016/j.bbapap.2019.01.004. [DOI] [PubMed] [Google Scholar]
- 93. Bers DM, Ellis D. Intracellular calcium and sodium activity in sheep heart Purkinje fibres. Effect of changes of external sodium and intracellular pH. Pflugers Arch 393: 171–178, 1982. doi: 10.1007/BF00582941. [DOI] [PubMed] [Google Scholar]
- 94. Higgins ER, Cannell MB, Sneyd J. A buffering SERCA pump in models of calcium dynamics. Biophys J 91: 151–163, 2006. doi: 10.1529/biophysj.105.075747. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95. Chen J, Sitsel A, Benoy V, Sepúlveda MR, Vangheluwe P. Primary active Ca2+ transport systems in health and disease. Cold Spring Harb Perspect Biol 12: a035113, 2020. doi: 10.1101/cshperspect.a035113. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96. Toyoshima C, Nomura H. Structural changes in the calcium pump accompanying the dissociation of calcium. Nature 418: 605–611, 2002. doi: 10.1038/nature00944. [DOI] [PubMed] [Google Scholar]
- 97. Gong D, Chi X, Ren K, Huang G, Zhou G, Yan N, Lei J, Zhou Q. Structure of the human plasma membrane Ca2+-ATPase 1 in complex with its obligatory subunit neuroplastin. Nat Commun 9: 3623, 2018. doi: 10.1038/s41467-018-06075-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98. Inesi G, Kurzmack M, Coan C, Lewis DE. Cooperative calcium binding and ATPase activation in sarcoplasmic reticulum vesicles. J Biol Chem 255: 3025–3031, 1980. doi: 10.1016/S0021-9258(19)85846-5. [DOI] [PubMed] [Google Scholar]
- 99. Koivumäki JT, Takalo J, Korhonen T, Tavi P, Weckström M. Modelling sarcoplasmic reticulum calcium ATPase and its regulation in cardiac myocytes. Philos Trans A Math Phys Eng Sci 367: 2181–2202, 2009. doi: 10.1098/rsta.2008.0304. [DOI] [PubMed] [Google Scholar]
- 100. Smith GL, Duncan AM, Neary P, Bruce L, Burton FL. Pi inhibits the SR Ca2+ pump and stimulates pump-mediated Ca2+ leak in rabbit cardiac myocytes. Am J Physiol Heart Circ Physiol 279: H577–H585, 2000. doi: 10.1152/ajpheart.2000.279.2.H577. [DOI] [PubMed] [Google Scholar]
- 101. Teucher N, Prestle J, Seidler T, Currie S, Elliott EB, Reynolds DF, Schott P, Wagner S, Kogler H, Inesi G, Bers DM, Hasenfuss G, Smith GL. Excessive sarcoplasmic/endoplasmic reticulum Ca2+-ATPase expression causes increased sarcoplasmic reticulum Ca2+ uptake but decreases myocyte shortening. Circulation 110: 3553–3559, 2004. doi: 10.1161/01.CIR.0000145161.48545.B3. [DOI] [PubMed] [Google Scholar]
- 102. Niggli V, Adunyah ES, Penniston JT, Carafoli E. Purified (Ca2+-Mg2+)-ATPase of the erythrocyte membrane. Reconstitution and effect of calmodulin and phospholipids. J Biol Chem 256: 395–401, 1981. doi: 10.1016/S0021-9258(19)70149-5. [DOI] [PubMed] [Google Scholar]
- 103. Go CK, Hooper R, Aronson MR, Schultz B, Cangoz T, Nemani N, Zhang Y, Madesh M, Soboloff J. The Ca2+ export pump PMCA clears near-membrane Ca2+ to facilitate store-operated Ca2+ entry and NFAT activation. Sci Signal 12: eaaw2627, 2019. doi: 10.1126/scisignal.aaw2627. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104. Schultz J, Milpetz F, Bork P, Ponting CP. SMART, a simple modular architecture research tool: identification of signaling domains. Proc Natl Acad Sci USA 95: 5857–5864, 1998. doi: 10.1073/pnas.95.11.5857. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 105. Nalefski EA, Falke JJ. The C2 domain calcium-binding motif: structural and functional diversity. Protein Sci 5: 2375–2390, 1996. doi: 10.1002/pro.5560051201. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 106. Sugita S, Shin OH, Han W, Lao Y, Südhof TC. Synaptotagmins form a hierarchy of exocytotic Ca2+ sensors with distinct Ca2+ affinities. EMBO J 21: 270–280, 2002. doi: 10.1093/emboj/21.3.270. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 107. Torrecillas A, Laynez J, Menéndez M, Corbalán-García S, Gómez-Fernández JC. Calorimetric study of the interaction of the C2 domains of classical protein kinase C isoenzymes with Ca2+ and phospholipids. Biochemistry 43: 11727–11739, 2004. doi: 10.1021/bi0489659. [DOI] [PubMed] [Google Scholar]
- 108. Pinheiro PS, Houy S, Sørensen JB. C2-domain containing calcium sensors in neuroendocrine secretion. J Neurochem 139: 943–958, 2016. doi: 10.1111/jnc.13865. [DOI] [PubMed] [Google Scholar]
- 109. Südhof TC. Calcium control of neurotransmitter release. Cold Spring Harb Perspect Biol 4: a011353, 2012. doi: 10.1101/cshperspect.a011353. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 110. Huber R, Schneider M, Mayr I, Römisch J, Paques EP. The calcium binding sites in human annexin V by crystal structure analysis at 2.0 A resolution. Implications for membrane binding and calcium channel activity. FEBS Lett 275: 15–21, 1990. doi: 10.1016/0014-5793(90)81428-q. [DOI] [PubMed] [Google Scholar]
- 111. Jost M, Thiel C, Weber K, Gerke V. Mapping of three unique Ca2+-binding sites in human annexin II. Eur J Biochem 207: 923–930, 1992. doi: 10.1111/j.1432-1033.1992.tb17125.x. [DOI] [PubMed] [Google Scholar]
- 112. Rudolf R, Magalhães PJ, Pozzan T. Direct in vivo monitoring of sarcoplasmic reticulum Ca2+ and cytosolic cAMP dynamics in mouse skeletal muscle. J Cell Biol 173: 187–193, 2006. doi: 10.1083/jcb.200601160. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 113. Miyawaki A, Llopis J, Heim R, McCaffery JM, Adams JA, Ikura M, Tsien RY. Fluorescent indicators for Ca2+ based on green fluorescent proteins and calmodulin. Nature 388: 882–887, 1997. doi: 10.1038/42264. [DOI] [PubMed] [Google Scholar]
- 114. Arnaudeau S, Frieden M, Nakamura K, Castelbou C, Michalak M, Demaurex N. Calreticulin differentially modulates calcium uptake and release in the endoplasmic reticulum and mitochondria. J Biol Chem 277: 46696–46705, 2002. doi: 10.1074/jbc.M202395200. [DOI] [PubMed] [Google Scholar]
- 115. Shannon TR, Guo T, Bers DM. Ca2+ scraps: local depletions of free [Ca2+] in cardiac sarcoplasmic reticulum during contractions leave substantial Ca2+ reserve. Circ Res 93: 40–45, 2003. doi: 10.1161/01.RES.0000079967.11815.19. [DOI] [PubMed] [Google Scholar]
- 116. Ozawa M, Muramatsu T. Reticulocalbin, a novel endoplasmic reticulum resident Ca2+-binding protein with multiple EF-hand motifs and a carboxyl-terminal HDEL sequence. J Biol Chem 268: 699–705, 1993. doi: 10.1016/S0021-9258(18)54208-3. [DOI] [PubMed] [Google Scholar]
- 117. Weis K, Griffiths G, Lamond AI. The endoplasmic reticulum calcium-binding protein of 55 kDa is a novel EF-hand protein retained in the endoplasmic reticulum by a carboxyl-terminal His-Asp-Glu-Leu motif. J Biol Chem 269: 19142–19150, 1994. doi: 10.1016/S0021-9258(17)32286-X. [DOI] [PubMed] [Google Scholar]
- 118. Honoré B. The rapidly expanding CREC protein family: members, localization, function, and role in disease. Bioessays 31: 262–277, 2009. doi: 10.1002/bies.200800186. [DOI] [PubMed] [Google Scholar]
- 119. Yabe D, Nakamura T, Kanazawa N, Tashiro K, Honjo T. Calumenin, a Ca2+-binding protein retained in the endoplasmic reticulum with a novel carboxyl-terminal sequence, HDEF. J Biol Chem 272: 18232–18239, 1997. doi: 10.1074/jbc.272.29.18232. [DOI] [PubMed] [Google Scholar]
- 120. Vorum H, Liu X, Madsen P, Rasmussen HH, Honoré B. Molecular cloning of a cDNA encoding human calumenin, expression in Escherichia coli and analysis of its Ca2+-binding activity. Biochim Biophys Acta 1386: 121–131, 1998. doi: 10.1016/s0167-4838(98)00089-2. [DOI] [PubMed] [Google Scholar]
- 121. Mazzorana M, Hussain R, Sorensen T. Ca-dependent folding of human calumenin. PLoS One 11: e0151547, 2016. doi: 10.1371/journal.pone.0151547. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 122. Jung DH, Mo SH, Kim DH. Calumenin, a multiple EF-hands Ca2+-binding protein, interacts with ryanodine receptor-1 in rabbit skeletal sarcoplasmic reticulum. Biochem Biophys Res Commun 343: 34–42, 2006. doi: 10.1016/j.bbrc.2006.02.115. [DOI] [PubMed] [Google Scholar]
- 123. Sahoo SK, Kim T, Kang GB, Lee JG, Eom SH, Kim DH. Characterization of calumenin-SERCA2 interaction in mouse cardiac sarcoplasmic reticulum. J Biol Chem 284: 31109–31121, 2009. doi: 10.1074/jbc.M109.031989. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 124. Lunz V, Romanin C, Frischauf I. STIM1 activation of Orai1. Cell Calcium 77: 29–38, 2019. doi: 10.1016/j.ceca.2018.11.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 125. Michalak M, Groenendyk J, Szabo E, Gold LI, Opas M. Calreticulin, a multi-process calcium-buffering chaperone of the endoplasmic reticulum. Biochem J 417: 651–666, 2009. doi: 10.1042/BJ20081847. [DOI] [PubMed] [Google Scholar]
- 126. Baksh S, Michalak M. Expression of calreticulin in Escherichia coli and identification of its Ca2+ binding domains. J Biol Chem 266: 21458–21465, 1991. doi: 10.1016/S0021-9258(18)54661-5. [DOI] [PubMed] [Google Scholar]
- 127. Corbett EF, Michalak M. Calcium, a signaling molecule in the endoplasmic reticulum? Trends Biochem Sci 25: 307–311, 2000. doi: 10.1016/s0968-0004(00)01588-7. [DOI] [PubMed] [Google Scholar]
- 128. Corbett EF, Oikawa K, Francois P, Tessier DC, Kay C, Bergeron JJ, Thomas DY, Krause KH, Michalak M. Ca2+ regulation of interactions between endoplasmic reticulum chaperones. J Biol Chem 274: 6203–6211, 1999. doi: 10.1074/jbc.274.10.6203. [DOI] [PubMed] [Google Scholar]
- 129. Vassilakos A, Michalak M, Lehrman MA, Williams DB. Oligosaccharide binding characteristics of the molecular chaperones calnexin and calreticulin. Biochemistry 37: 3480–3490, 1998. doi: 10.1021/bi972465g. [DOI] [PubMed] [Google Scholar]
- 130. Kozlov G, Gehring K. Calnexin cycle - structural features of the ER chaperone system. FEBS J 287: 4322–4340, 2020. doi: 10.1111/febs.15330. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 131. Tjoelker LW, Seyfried CE, Eddy RL Jr, Byers MG, Shows TB, Calderon J, Schreiber RB, Gray PW. Human, mouse, and rat calnexin cDNA cloning: identification of potential calcium binding motifs and gene localization to human chromosome 5. Biochemistry 33: 3229–3236, 1994. doi: 10.1021/bi00177a013. [DOI] [PubMed] [Google Scholar]
- 132. Sanchez EJ, Lewis KM, Danna BR, Kang C. High-capacity Ca2+ binding of human skeletal calsequestrin. J Biol Chem 287: 11592–11601, 2012. doi: 10.1074/jbc.M111.335075. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 133. Park H, Park IY, Kim E, Youn B, Fields K, Dunker AK, Kang C. Comparing skeletal and cardiac calsequestrin structures and their calcium binding: a proposed mechanism for coupled calcium binding and protein polymerization. J Biol Chem 279: 18026–18033, 2004. doi: 10.1074/jbc.M311553200. [DOI] [PubMed] [Google Scholar]
- 134. Sztretye M, Yi J, Figueroa L, Zhou J, Royer L, Allen P, Brum G, Ríos E. Measurement of RyR permeability reveals a role of calsequestrin in termination of SR Ca2+ release in skeletal muscle. J Gen Physiol 138: 231–247, 2011. doi: 10.1085/jgp.201010592. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 135. Terentyev D, Kubalova Z, Valle G, Nori A, Vedamoorthyrao S, Terentyeva R, Viatchenko-Karpinski S, Bers DM, Williams SC, Volpe P, Gyorke S. Modulation of SR Ca release by luminal Ca and calsequestrin in cardiac myocytes: effects of CASQ2 mutations linked to sudden cardiac death. Biophys J 95: 2037–2048, 2008. doi: 10.1529/biophysj.107.128249. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 136. Stevens SC, Terentyev D, Kalyanasundaram A, Periasamy M, Györke S. Intra-sarcoplasmic reticulum Ca2+ oscillations are driven by dynamic regulation of ryanodine receptor function by luminal Ca2+ in cardiomyocytes. J Physiol 587: 4863–4872, 2009. doi: 10.1113/jphysiol.2009.175547. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 137. Krause KH, Milos M, Luan-Rilliet Y, Lew DP, Cox JA. Thermodynamics of cation binding to rabbit skeletal muscle calsequestrin. Evidence for distinct Ca2+- and Mg2+-binding sites. J Biol Chem 266: 9453–9459, 1991. doi: 10.1016/S0021-9258(18)92842-5. [DOI] [PubMed] [Google Scholar]
- 138. Titus EW, Deiter FH, Shi C, Wojciak J, Scheinman M, Jura N, Deo RC. The structure of a calsequestrin filament reveals mechanisms of familial arrhythmia. Nat Struct Mol Biol 27: 1142–1151, 2020. doi: 10.1038/s41594-020-0510-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 139. Prins D, Michalak M. Organellar calcium buffers. Cold Spring Harb Perspect Biol 3: a004069, 2011. doi: 10.1101/cshperspect.a004069. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 140. McCormack JG, Denton RM. The effects of calcium ions and adenine nucleotides on the activity of pig heart 2-oxoglutarate dehydrogenase complex. Biochem J 180: 533–544, 1979. doi: 10.1042/bj1800533. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 141. Williams GS, Boyman L, Lederer WJ. Mitochondrial calcium and the regulation of metabolism in the heart. J Mol Cell Cardiol 78: 35–45, 2015. doi: 10.1016/j.yjmcc.2014.10.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 142. Lu X, Ginsburg KS, Kettlewell S, Bossuyt J, Smith GL, Bers DM. Measuring local gradients of intramitochondrial [Ca2+] in cardiac myocytes during sarcoplasmic reticulum Ca2+ release. Circ Res 112: 424–431, 2013. doi: 10.1161/CIRCRESAHA.111.300501. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 143. Boyman L, Chikando AC, Williams GS, Khairallah RJ, Kettlewell S, Ward CW, Smith GL, Kao JP, Lederer WJ. Calcium movement in cardiac mitochondria. Biophys J 107: 1289–1301, 2014. doi: 10.1016/j.bpj.2014.07.045. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 144. Dedkova EN, Blatter LA. Calcium signaling in cardiac mitochondria. J Mol Cell Cardiol 58: 125–133, 2013. doi: 10.1016/j.yjmcc.2012.12.021. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 145. Drago I, De Stefani D, Rizzuto R, Pozzan T. Mitochondrial Ca2+ uptake contributes to buffering cytoplasmic Ca2+ peaks in cardiomyocytes. Proc Natl Acad Sci USA 109: 12986–12991, 2012. doi: 10.1073/pnas.1210718109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 146. Maack C, Cortassa S, Aon MA, Ganesan AN, Liu T, O’Rourke B. Elevated cytosolic Na+ decreases mitochondrial Ca2+ uptake during excitation-contraction coupling and impairs energetic adaptation in cardiac myocytes. Circ Res 99: 172–182, 2006. doi: 10.1161/01.RES.0000232546.92777.05. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 147. Kettlewell S, Cabrero P, Nicklin SA, Dow JA, Davies S, Smith GL. Changes of intra-mitochondrial Ca2+ in adult ventricular cardiomyocytes examined using a novel fluorescent Ca2+ indicator targeted to mitochondria. J Mol Cell Cardiol 46: 891–901, 2009. doi: 10.1016/j.yjmcc.2009.02.016. [DOI] [PubMed] [Google Scholar]
- 148. Sedova M, Dedkova EN, Blatter LA. Integration of rapid cytosolic Ca2+ signals by mitochondria in cat ventricular myocytes. Am J Physiol Cell Physiol 291: C840–C850, 2006. doi: 10.1152/ajpcell.00619.2005. [DOI] [PubMed] [Google Scholar]
- 149. Palty R, Silverman WF, Hershfinkel M, Caporale T, Sensi SL, Parnis J, Nolte C, Fishman D, Shoshan-Barmatz V, Herrmann S, Khananshvili D, Sekler I. NCLX is an essential component of mitochondrial Na+/Ca2+ exchange. Proc Natl Acad Sci USA 107: 436–441, 2010. doi: 10.1073/pnas.0908099107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 150. Hansford RG, Castro F. Intramitochondrial and extramitochondrial free calcium ion concentrations of suspensions of heart mitochondria with very low, plausibly physiological, contents of total calcium. J Bioenerg Biomembr 14: 361–376, 1982. doi: 10.1007/BF00743064. [DOI] [PubMed] [Google Scholar]
- 151. Coll KE, Joseph SK, Corkey BE, Williamson JR. Determination of the matrix free Ca2+ concentration and kinetics of Ca2+ efflux in liver and heart mitochondria. J Biol Chem 257: 8696–8704, 1982. doi: 10.1016/S0021-9258(18)34184-X. [DOI] [PubMed] [Google Scholar]
- 152. Chalmers S, Nicholls DG. The relationship between free and total calcium concentrations in the matrix of liver and brain mitochondria. J Biol Chem 278: 19062–19070, 2003. doi: 10.1074/jbc.M212661200. [DOI] [PubMed] [Google Scholar]
- 153. Wolf SG, Mutsafi Y, Dadosh T, Ilani T, Lansky Z, Horowitz B, Rubin S, Elbaum M, Fass D. 3D visualization of mitochondrial solid-phase calcium stores in whole cells. Elife 6: e29929, 2017. doi: 10.7554/eLife.29929. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 154. Blomeyer CA, Bazil JN, Stowe DF, Pradhan RK, Dash RK, Camara AK. Dynamic buffering of mitochondrial Ca2+ during Ca2+ uptake and Na+-induced Ca2+ release. J Bioenerg Biomembr 45: 189–202, 2013. doi: 10.1007/s10863-012-9483-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 155. Bazil JN, Blomeyer CA, Pradhan RK, Camara AK, Dash RK. Modeling the calcium sequestration system in isolated guinea pig cardiac mitochondria. J Bioenerg Biomembr 45: 177–188, 2013. doi: 10.1007/s10863-012-9488-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 156. Wei AC, Liu T, Winslow RL, O'Rourke B. Dynamics of matrix-free Ca2+ in cardiac mitochondria: two components of Ca2+ uptake and role of phosphate buffering. J Gen Physiol 139: 465–478, 2012. doi: 10.1085/jgp.201210784. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 157. Andrienko TN, Picht E, Bers DM. Mitochondrial free calcium regulation during sarcoplasmic reticulum calcium release in rat cardiac myocytes. J Mol Cell Cardiol 46: 1027–1036, 2009. doi: 10.1016/j.yjmcc.2009.03.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 158. Bassani JW, Bassani RA, Bers DM. Relaxation in rabbit and rat cardiac cells: species-dependent differences in cellular mechanisms. J Physiol 476: 279–293, 1994. doi: 10.1113/jphysiol.1994.sp020130. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 159. Nicholls DG, Chalmers S. The integration of mitochondrial calcium transport and storage. J Bioenerg Biomembr 36: 277–281, 2004. doi: 10.1023/B:JOBB.0000041753.52832.f3. [DOI] [PubMed] [Google Scholar]
- 160. Bootman MD, Berridge MJ, Lipp P. Cooking with calcium: the recipes for composing global signals from elementary events. Cell 91: 367–373, 1997. doi: 10.1016/s0092-8674(00)80420-1. [DOI] [PubMed] [Google Scholar]
- 161. Yang Z, Kirton HM, MacDougall DA, Boyle JP, Deuchars J, Frater B, Ponnambalam S, Hardy ME, White E, Calaghan SC, Peers C, Steele DS. The Golgi apparatus is a functionally distinct Ca2+ store regulated by the PKA and Epac branches of the beta1-adrenergic signaling pathway. Sci Signal 8: ra101, 2015. doi: 10.1126/scisignal.aaa7677. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 162. Fox JL, Burgstahler AD, Nathanson MH. Mechanism of long-range Ca2+ signalling in the nucleus of isolated rat hepatocytes. Biochem J 326: 491–495, 1997. doi: 10.1042/bj3260491. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 163. Naraghi M, Müller TH, Neher E. Two-dimensional determination of the cellular Ca2+ binding in bovine chromaffin cells. Biophys J 75: 1635–1647, 1998. doi: 10.1016/S0006-3495(98)77606-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 164. Yarom R, Hall TA, Peters PD. Calcium in myonuclei: electron microprobe X-ray analysis. Experientia 31: 154–157, 1975. doi: 10.1007/BF01990677. [DOI] [PubMed] [Google Scholar]
- 165. Dobi A, Agoston D. Submillimolar levels of calcium regulates DNA structure at the dinucleotide repeat (TG/AC)n. Proc Natl Acad Sci USA 95: 5981–5986, 1998. doi: 10.1073/pnas.95.11.5981. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 166. Phengchat R, Takata H, Morii K, Inada N, Murakoshi H, Uchiyama S, Fukui K. Calcium ions function as a booster of chromosome condensation. Sci Rep 6: 38281, 2016. doi: 10.1038/srep38281. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 167. Echevarría W, Leite MF, Guerra MT, Zipfel WR, Nathanson MH. Regulation of calcium signals in the nucleus by a nucleoplasmic reticulum. Nat Cell Biol 5: 440–446, 2003. doi: 10.1038/ncb980. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 168. Malhas A, Goulbourne C, Vaux DJ. The nucleoplasmic reticulum: form and function. Trends Cell Biol 21: 362–373, 2011. doi: 10.1016/j.tcb.2011.03.008. [DOI] [PubMed] [Google Scholar]
- 169. Mandinova A, Atar D, Schäfer BW, Spiess M, Aebi U, Heizmann CW. Distinct subcellular localization of calcium binding S100 proteins in human smooth muscle cells and their relocation in response to rises in intracellular calcium. J Cell Sci 111: 2043–2054, 1998. doi: 10.1242/jcs.111.14.2043. [DOI] [PubMed] [Google Scholar]
- 170. Mauceri D, Hagenston AM, Schramm K, Weiss U, Bading H. Nuclear calcium buffering capacity shapes neuronal architecture. J Biol Chem, 2015. doi: 10.1074/jbc.M115.654962. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 171. Gilchrist JS, Czubryt MP, Pierce GN. Calcium and calcium-binding proteins in the nucleus. Mol Cell Biochem 135: 79–88, 1994. doi: 10.1007/BF00925963. [DOI] [PubMed] [Google Scholar]
- 172. Balakier H, Dziak E, Sojecki A, Librach C, Michalak M, Opas M. Calcium-binding proteins and calcium-release channels in human maturing oocytes, pronuclear zygotes and early preimplantation embryos. Hum Reprod 17: 2938–2947, 2002. doi: 10.1093/humrep/17.11.2938. [DOI] [PubMed] [Google Scholar]
- 173. Neher E. The use of fura-2 for estimating Ca buffers and Ca fluxes. Neuropharmacology 34: 1423–1442, 1995. doi: 10.1016/0028-3908(95)00144-u. [DOI] [PubMed] [Google Scholar]
- 174. Neher E, Augustine GJ. Calcium gradients and buffers in bovine chromaffin cells. J Physiol 450: 273–301, 1992. doi: 10.1113/jphysiol.1992.sp019127. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 175. Klingauf J, Neher E. Modeling buffered Ca2+ diffusion near the membrane: implications for secretion in neuroendocrine cells. Biophys J 72: 674–690, 1997. doi: 10.1016/s0006-3495(97)78704-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 176. Neher E. Vesicle pools and Ca2+ microdomains: new tools for understanding their roles in neurotransmitter release. Neuron 20: 389–399, 1998. doi: 10.1016/s0896-6273(00)80983-6. [DOI] [PubMed] [Google Scholar]
- 177. Blatow M, Caputi A, Burnashev N, Monyer H, Rozov A. Ca2+ buffer saturation underlies paired pulse facilitation in calbindin-D28k-containing terminals. Neuron 38: 79–88, 2003. doi: 10.1016/s0896-6273(03)00196-x. [DOI] [PubMed] [Google Scholar]
- 178. Neher E. Usefulness and limitations of linear approximations to the understanding of Ca++ signals. Cell Calcium 24: 345–357, 1998. doi: 10.1016/s0143-4160(98)90058-6. [DOI] [PubMed] [Google Scholar]
- 179. Bakker AJ, Cully TR, Wingate CD, Barclay CJ, Launikonis BS. Doublet stimulation increases Ca2+ binding to troponin C to ensure rapid force development in skeletal muscle. J Gen Physiol 149: 323–334, 2017. doi: 10.1085/jgp.201611727. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 180. Schober T, Huke S, Venkataraman R, Gryshchenko O, Kryshtal D, Hwang HS, Baudenbacher FJ, Knollmann BC. Myofilament Ca sensitization increases cytosolic Ca binding affinity, alters intracellular Ca homeostasis, and causes pause-dependent Ca-triggered arrhythmia. Circ Res 111: 170–179, 2012. doi: 10.1161/CIRCRESAHA.112.270041. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 181. Tikunova SB, Davis JP. Designing calcium-sensitizing mutations in the regulatory domain of cardiac troponin C. J Biol Chem 279: 35341–35352, 2004. doi: 10.1074/jbc.M405413200. [DOI] [PubMed] [Google Scholar]
- 182. Kreutziger KL, Piroddi N, McMichael JT, Tesi C, Poggesi C, Regnier M. Calcium binding kinetics of troponin C strongly modulate cooperative activation and tension kinetics in cardiac muscle. J Mol Cell Cardiol 50: 165–174, 2011. doi: 10.1016/j.yjmcc.2010.10.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 183. Weiss JN. The Hill equation revisited: uses and misuses. FASEB J 11: 835–841, 1997.doi: 10.1096/fasebj.11.11.9285481. [DOI] [PubMed] [Google Scholar]
- 184. Hill AV. The possible effects of the aggregation of the molecules of haemoglobin on its dissociation curve. Proceedings of The Physiological Society: January 22, 1910. J Physiol 40: iv–vii, 1910. [Google Scholar]
- 185. Faas GC, Schwaller B, Vergara JL, Mody I. Resolving the fast kinetics of cooperative binding: Ca2+ buffering by calretinin. PLoS Biol 5: e311, 2007. doi: 10.1371/journal.pbio.0050311. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 186. Martin SR, Linse S, Johansson C, Bayley PM, Forsén S. Protein surface charges and Ca2+ binding to individual sites in calbindin D9k: stopped-flow studies. Biochemistry 29: 4188–4193, 1990. doi: 10.1021/bi00469a023. [DOI] [PubMed] [Google Scholar]
- 187. Faas GC, Raghavachari S, Lisman JE, Mody I. Calmodulin as a direct detector of Ca2+ signals. Nat Neurosci 14: 301–304, 2011. doi: 10.1038/nn.2746. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 188. Matveev V. Extension of rapid buffering approximation to Ca2+ buffers with two binding sites. Biophys J 114: 1204–1215, 2018. doi: 10.1016/j.bpj.2018.01.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 189. Sánchez-Gómez L, Guerrero-Hernández A, Santillán M. Polymerization of sarcoplasmic-reticulum calcium-binding proteins might explain observed reticulum kinetics-on-demand behavior. J Theor Biol 482: 109986, 2019. doi: 10.1016/j.jtbi.2019.08.017. [DOI] [PubMed] [Google Scholar]
- 190. Li MX, Gagné SM, Tsuda S, Kay CM, Smillie LB, Sykes BD. Calcium binding to the regulatory N-domain of skeletal muscle troponin C occurs in a stepwise manner. Biochemistry 34: 8330–8340, 1995. doi: 10.1021/bi00026a014. [DOI] [PubMed] [Google Scholar]
- 191. Davis JP, Norman C, Kobayashi T, Solaro RJ, Swartz DR, Tikunova SB. Effects of thin and thick filament proteins on calcium binding and exchange with cardiac troponin C. Biophys J 92: 3195–3206, 2007. doi: 10.1529/biophysj.106.095406. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 192. Dargan SL, Schwaller B, Parker I. Spatiotemporal patterning of IP3-mediated Ca2+ signals in Xenopus oocytes by Ca2+-binding proteins. J Physiol 556: 447–461, 2004. doi: 10.1113/jphysiol.2003.059204. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 193. Schwiening CJ, Thomas RC. Relationship between intracellular calcium and its muffling measured by calcium iontophoresis in snail neurones. J Physiol 491: 621–633, 1996. doi: 10.1113/jphysiol.1996.sp021244. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 194. Allen DG, Eisner DA, Orchard CH. Factors influencing free intracellular calcium concentration in quiescent ferret ventricular muscle. J Physiol 350: 615–630, 1984. doi: 10.1113/jphysiol.1984.sp015221. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 195. Friel DD, Tsien RW. A caffeine- and ryanodine-sensitive Ca2+ store in bullfrog sympathetic neurones modulates effects of Ca2+ entry on [Ca2+]i. J Physiol 450: 217–246, 1992. doi: 10.1113/jphysiol.1992.sp019125. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 196. Ríos E. The cell boundary theorem: a simple law of the control of cytosolic calcium concentration. J Physiol Sci 60: 81–84, 2010. doi: 10.1007/s12576-009-0069-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 197. Díaz ME, Trafford AW, Eisner DA. The role of intracellular Ca buffers in determining the shape of the systolic Ca transient in cardiac ventricular myocytes. Pflugers Arch 442: 96–100, 2001. doi: 10.1007/s004240000509. [DOI] [PubMed] [Google Scholar]
- 198. Song LS, Sham JS, Stern MD, Lakatta EG, Cheng H. Direct measurement of SR release flux by tracking ‘Ca2+ spikes’ in rat cardiac myocytes. J Physiol 512: 677–691, 1998. doi: 10.1111/j.1469-7793.1998.677bd.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 199. Díaz ME, Trafford AW, Eisner DA. The effects of exogenous calcium buffers on the systolic calcium transient in rat ventricular myocytes. Biophys J 80: 1915–1925, 2001. doi: 10.1016/S0006-3495(01)76161-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 200. Bronner F. Renal calcium transport: mechanisms and regulation—an overview. Am J Physiol Renal Physiol 257: F707–F711, 1989. doi: 10.1152/ajprenal.1989.257.5.F707. [DOI] [PubMed] [Google Scholar]
- 201. Matthews EA, Dietrich D. Buffer mobility and the regulation of neuronal calcium domains. Front Cell Neurosci 9: 48, 2015.doi: 10.3389/fncel.2015.00048. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 202. Wagner J, Keizer J. Effects of rapid buffers on Ca2+ diffusion and Ca2+ oscillations. Biophys J 67: 447–456, 1994. doi: 10.1016/S0006-3495(94)80500-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 203. Michailova A, DelPrincipe F, Egger M, Niggli E. Spatiotemporal features of Ca2+ buffering and diffusion in atrial cardiac myocytes with inhibited sarcoplasmic reticulum. Biophys J 83: 3134–3151, 2002. doi: 10.1016/S0006-3495(02)75317-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 204. Gabso M, Neher E, Spira ME. Low mobility of the Ca2+ buffers in axons of cultured Aplysia neurons. Neuron 18: 473–481, 1997. doi: 10.1016/s0896-6273(00)81247-7. [DOI] [PubMed] [Google Scholar]
- 205. Gribbon P, Hardingham TE. Macromolecular diffusion of biological polymers measured by confocal fluorescence recovery after photobleaching. Biophys J 75: 1032–1039, 1998. doi: 10.1016/S0006-3495(98)77592-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 206. Valencia DP, González F. Understanding the linear correlation between diffusion coefficient and molecular weight. A model to estimate diffusion coefficients in acetonitrile solutions. Electrochem Commun 13: 129–132, 2011. doi: 10.1016/j.elecom.2010.11.032. [DOI] [Google Scholar]
- 207. Stein WD. Facilitated diffusion of calcium across the rat intestinal epithelial cell. J Nutr 122: 651–656, 1992. doi: 10.1093/jn/122.suppl_3.651. [DOI] [PubMed] [Google Scholar]
- 208. Saucerman JJ, Bers DM. Calmodulin mediates differential sensitivity of CaMKII and calcineurin to local Ca2+ in cardiac myocytes. Biophys J 95: 4597–4612, 2008. doi: 10.1529/biophysj.108.128728. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 209. Feher JJ, Fullmer CS, Wasserman RH. Role of facilitated diffusion of calcium by calbindin in intestinal calcium absorption. Am J Physiol Cell Physiol 262: C517–C526, 1992. doi: 10.1152/ajpcell.1992.262.2.C517. [DOI] [PubMed] [Google Scholar]
- 210. Badura A, Sun XR, Giovannucci A, Lynch LA, Wang SS. Fast calcium sensor proteins for monitoring neural activity. Neurophotonics 1: 025008, 2014. doi: 10.1117/1.NPh.1.2.025008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 211. O’Dowd JJ, Robins DJ, Miller DJ. Detection, characterisation, and quantification of carnosine and other histidyl derivatives in cardiac and skeletal muscle. Biochim Biophys Acta 967: 241–249, 1988. doi: 10.1016/0304-4165(88)90015-3. [DOI] [PubMed] [Google Scholar]
- 212. Baguet A, Everaert I, Achten E, Thomis M, Derave W. The influence of sex, age and heritability on human skeletal muscle carnosine content. Amino Acids 43: 13–20, 2012. doi: 10.1007/s00726-011-1197-3. [DOI] [PubMed] [Google Scholar]
- 213. Royer L, Ríos E. Deconstructing calsequestrin. Complex buffering in the calcium store of skeletal muscle. J Physiol 587: 3101–3111, 2009. doi: 10.1113/jphysiol.2009.171934. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 214. Naumovets AG, Vedula YS. Surface diffusion of adsorbates. Surf Sci Rep 4: 365–434, 1985. doi: 10.1016/0167-5729(85)90007-X. [DOI] [Google Scholar]
- 215. McMahon SM, Jackson MB. An inconvenient truth: calcium sensors are calcium buffers. Trends Neurosci 41: 880–884, 2018. doi: 10.1016/j.tins.2018.09.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 216. Callamaras N, Parker I. Phasic characteristic of elementary Ca2+ release sites underlies quantal responses to IP3. EMBO J 19: 3608–3617, 2000. doi: 10.1093/emboj/19.14.3608. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 217. Schuhmeier RP, Melzer W. Voltage-dependent Ca2+ fluxes in skeletal myotubes determined using a removal model analysis. J Gen Physiol 123: 33–51, 2004. doi: 10.1085/jgp.200308908. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 218. Lock JT, Parker I, Smith IF. A comparison of fluorescent Ca2+ indicators for imaging local Ca2+ signals in cultured cells. Cell Calcium 58: 638–648, 2015. doi: 10.1016/j.ceca.2015.10.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 219. Smith IF, Parker I. Imaging the quantal substructure of single IP3R channel activity during Ca2+ puffs in intact mammalian cells. Proc Natl Acad Sci USA 106: 6404–6409, 2009. doi: 10.1073/pnas.0810799106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 220. Swaminathan D, Dickinson GD, Demuro A, Parker I. Noise analysis of cytosolic calcium image data. Cell Calcium 86: 102152, 2020. doi: 10.1016/j.ceca.2019.102152. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 221. Frank T, Khimich D, Neef A, Moser T. Mechanisms contributing to synaptic Ca2+ signals and their heterogeneity in hair cells. Proc Natl Acad Sci USA 106: 4483–4488, 2009. doi: 10.1073/pnas.0813213106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 222. Neef J, Urban NT, Ohn TL, Frank T, Jean P, Hell SW, Willig KI, Moser T. Quantitative optical nanophysiology of Ca2+ signaling at inner hair cell active zones. Nat Commun 9: 290, 2018. doi: 10.1038/s41467-017-02612-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 223. Pangršič T, Gabrielaitis M, Michanski S, Schwaller B, Wolf F, Strenzke N, Moser T. EF-hand protein Ca2+ buffers regulate Ca2+ influx and exocytosis in sensory hair cells. Proc Natl Acad Sci USA 112: E1028–E1037, 2015. doi: 10.1073/pnas.1416424112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 224. Demuro A, Parker I. Imaging single-channel calcium microdomains by total internal reflection microscopy. Biol Res 37: 675–679, 2004. doi: 10.4067/s0716-97602004000400025. [DOI] [PubMed] [Google Scholar]
- 225. Baylor SM, Chandler WK, Marshall MW. Sarcoplasmic reticulum calcium release in frog skeletal muscle fibres estimated from Arsenazo III calcium transients. J Physiol 344: 625–666, 1983. doi: 10.1113/jphysiol.1983.sp014959. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 226. Schneggenburger R, Zhou Z, Konnerth A, Neher E. Fractional contribution of calcium to the cation current through glutamate receptor channels. Neuron 11: 133–143, 1993. doi: 10.1016/0896-6273(93)90277-x. [DOI] [PubMed] [Google Scholar]
- 227. Bollmann JH, Helmchen F, Borst JG, Sakmann B. Postsynaptic Ca2+ influx mediated by three different pathways during synaptic transmission at a calyx-type synapse. J Neurosci 18: 10409–10419, 1998. doi: 10.1523/JNEUROSCI.18-24-10409.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 228. Nakai J, Ohkura M, Imoto K. A high signal-to-noise Ca2+ probe composed of a single green fluorescent protein. Nat Biotechnol 19: 137–141, 2001. doi: 10.1038/84397. [DOI] [PubMed] [Google Scholar]
- 229. Singh M, Lujan B, Renden R. Presynaptic GCaMP expression decreases vesicle release probability at the calyx of Held. Synapse 72: e22040, 2018. doi: 10.1002/syn.22040. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 230. Dana H, Sun Y, Mohar B, Hulse BK, Kerlin AM, Hasseman JP, Tsegaye G, Tsang A, Wong A, Patel R, Macklin JJ, Chen Y, Konnerth A, Jayaraman V, Looger LL, Schreiter ER, Svoboda K, Kim DS. High-performance calcium sensors for imaging activity in neuronal populations and microcompartments. Nat Methods 16: 649–657, 2019. doi: 10.1038/s41592-019-0435-6. [DOI] [PubMed] [Google Scholar]
- 231. Inoue M. Genetically encoded calcium indicators to probe complex brain circuit dynamics in vivo. Neurosci Res 169: 2–8, 2021. doi: 10.1016/j.neures.2020.05.013. [DOI] [PubMed] [Google Scholar]
- 232. Horne JH, Meyer T. Characterization of a dextran-based bifunctional calcium indicator immobilized in cells by the enzymatic addition of isoprenoid lipids. Cell Calcium 25: 1–7, 1999. doi: 10.1054/ceca.1998.0006. [DOI] [PubMed] [Google Scholar]
- 233. Despa S, Shui B, Bossuyt J, Lang D, Kotlikoff MI, Bers DM. Junctional cleft [Ca2+]i measurements using novel cleft-targeted Ca2+ sensors. Circ Res 115: 339–347, 2014. doi: 10.1161/CIRCRESAHA.115.303582. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 234. Sparrow AJ, Sievert K, Patel S, Chang YF, Broyles CN, Brook FA, Watkins H, Geeves MA, Redwood CS, Robinson P, Daniels MJ. Measurement of myofilament-localized calcium dynamics in adult cardiomyocytes and the effect of hypertrophic cardiomyopathy mutations. Circ Res 124: 1228–1239, 2019. doi: 10.1161/CIRCRESAHA.118.314600. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 235. Kekenes-Huskey PM, Scott CE, Atalay S. Quantifying the influence of the crowded cytoplasm on small molecule diffusion. J Phys Chem B 120: 8696–8706, 2016. doi: 10.1021/acs.jpcb.6b03887. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 236. Nakatani K, Chen C, Koutalos Y. Calcium diffusion coefficient in rod photoreceptor outer segments. Biophys J 82: 728–739, 2002. doi: 10.1016/S0006-3495(02)75435-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 237. Harkins AB, Kurebayashi N, Baylor SM. Resting myoplasmic free calcium in frog skeletal muscle fibers estimated with fluo-3. Biophys J 65: 865–881, 1993. doi: 10.1016/S0006-3495(93)81112-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 238. Blatter LA, Wier WG. Intracellular diffusion, binding, and compartmentalization of the fluorescent calcium indicators indo-1 and fura-2. Biophys J 58: 1491–1499, 1990. doi: 10.1016/S0006-3495(90)82494-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 239. Roberts WM. Spatial calcium buffering in saccular hair cells. Nature 363: 74–76, 1993. doi: 10.1038/363074a0. [DOI] [PubMed] [Google Scholar]
- 240. Kerr RA, Bartol TM, Kaminsky B, Dittrich M, Chang JC, Baden SB, Sejnowski TJ, Stiles JR. Fast Monte Carlo simulation methods for biological reaction-diffusion systems in solution and on surfaces. SIAM J Sci Comput 30: 3126, 2008. doi: 10.1137/070692017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 241. Chen Y, Matveev V. Stationary Ca2+ nanodomains in the presence of buffers with two binding sites. Biophys J 120: 1942–1956, 2021. doi: 10.1016/j.bpj.2021.03.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 242. Laghaei R, Meriney SD. Microphysiological modeling of the structure and function of neuromuscular transmitter release sites. Front Synaptic Neurosci 14: 917285, 2022. doi: 10.3389/fnsyn.2022.917285. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 243. Nakamura Y, Harada H, Kamasawa N, Matsui K, Rothman JS, Shigemoto R, Silver RA, DiGregorio DA, Takahashi T. Nanoscale distribution of presynaptic Ca2+ channels and its impact on vesicular release during development. Neuron 85: 145–158, 2015. doi: 10.1016/j.neuron.2014.11.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 244. Rebola N, Reva M, Kirizs T, Szoboszlay M, Lőrincz A, Moneron G, Nusser Z, DiGregorio DA. Distinct nanoscale calcium channel and synaptic vesicle topographies contribute to the diversity of synaptic function. Neuron 104: 693–710.e9, 2019. doi: 10.1016/j.neuron.2019.08.014. [DOI] [PubMed] [Google Scholar]
- 245. Basnayake K, Mazaud D, Bemelmans A, Rouach N, Korkotian E, Holcman D. Fast calcium transients in dendritic spines driven by extreme statistics. PLoS Biol 17: e2006202, 2019. doi: 10.1371/journal.pbio.2006202. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 246. Cheng H, Lederer WJ. Calcium sparks. Physiol Rev 88: 1491–1545, 2008. doi: 10.1152/physrev.00030.2007. [DOI] [PubMed] [Google Scholar]
- 247. Friedhoff VN, Antunes G, Falcke M, Simões de Souza FM. Stochastic reaction-diffusion modeling of calcium dynamics in 3D dendritic spines of Purkinje cells. Biophys J 120: 2112–2123, 2021. doi: 10.1016/j.bpj.2021.03.027. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 248. Lin KH, Taschenberger H, Neher E. Dynamics of volume-averaged intracellular Ca2+ in a rat CNS nerve terminal during single and repetitive voltage-clamp depolarizations. J Physiol 595: 3219–3236, 2017. doi: 10.1113/JP272773. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 249. Bornschein G, Schmidt H. Synaptotagmin Ca2+ sensors and their spatial coupling to presynaptic Cav channels in central cortical synapses. Front Mol Neurosci 11: 494, 2018. doi: 10.3389/fnmol.2018.00494. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 250. Collin T, Chat M, Lucas MG, Moreno H, Racay P, Schwaller B, Marty A, Llano I. Developmental changes in parvalbumin regulate presynaptic Ca2+ signaling. J Neurosci 25: 96–107, 2005. doi: 10.1523/JNEUROSCI.3748-04.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 251. McMahon SM, Chang C-W, Jackson MB. Multiple cytosolic calcium buffers in posterior pituitary nerve terminals. J Gen Physiol 147: 243–254, 2016. doi: 10.1085/jgp.201511525. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 252. Schmidt H, Brown EB, Schwaller B, Eilers J. Diffusional mobility of parvalbumin in spiny dendrites of cerebellar Purkinje neurons quantified by fluorescence recovery after photobleaching. Biophys J 84: 2599–2608, 2003. doi: 10.1016/S0006-3495(03)75065-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 253. Clarke JD, Caldwell JL, Pearman CM, Eisner DA, Trafford AW, Dibb KM. Increased Ca buffering underpins remodelling of Ca2+ handling in old sheep atrial myocytes. J Physiol 595: 6263–6279, 2017. doi: 10.1113/JP274053. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 254. Trafford AW, Díaz ME, Eisner DA. A novel, rapid and reversible method to measure Ca buffering and time-course of total sarcoplasmic reticulum Ca content in cardiac ventricular myocytes. Pflugers Arch 437: 501–503, 1999. doi: 10.1007/s004240050808. [DOI] [PubMed] [Google Scholar]
- 255. Fogelson AL, Zucker RS. Presynaptic calcium diffusion from various arrays of single channels. Implications for transmitter release and synaptic facilitation. Biophys J 48: 1003–1017, 1985. doi: 10.1016/S0006-3495(85)83863-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 256. Pape PC, Jong DS, Chandler WK. Calcium release and its voltage dependence in frog cut muscle fibers equilibrated with 20 mM EGTA. J Gen Physiol 106: 259–336, 1995. doi: 10.1085/jgp.106.2.259. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 257. Stern MD. Buffering of calcium in the vicinity of a channel pore. Cell Calcium 13: 183–192, 1992. doi: 10.1016/0143-4160(92)90046-u. [DOI] [PubMed] [Google Scholar]
- 258. Bertram R, Smith GD, Sherman A. Modeling study of the effects of overlapping Ca2+ microdomains on neurotransmitter release. Biophys J 76: 735–750, 1999. doi: 10.1016/S0006-3495(99)77240-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 259. Trommershäuser J, Schneggenburger R, Zippelius A, Neher E. Heterogeneous presynaptic release probabilities: functional relevance for short-term plasticity. Biophys J 84: 1563–1579, 2003. doi: 10.1016/S0006-3495(03)74967-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 260. Bers DM, Peskoff A. Diffusion around a cardiac calcium channel and the role of surface bound calcium. Biophys J 59: 703–721, 1991. doi: 10.1016/S0006-3495(91)82284-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 261. Soeller C, Cannell MB. Numerical simulation of local calcium movements during L-type calcium channel gating in the cardiac diad. Biophys J 73: 97–111, 1997. doi: 10.1016/S0006-3495(97)78051-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 262. Neher E. Concentration profiles of intracellular calcium in the presence of a diffusable chelator. In: Calcium electrogenesis and neuronal functioning, edited by Heinemann U, Klee M, Neher E, Singer W. Berlin: Springer, 1986, p. 80–96. [Google Scholar]
- 263. Pape PC, Jong DS, Chandler WK. Effects of partial sarcoplasmic reticulum calcium depletion on calcium release in frog cut muscle fibers equilibrated with 20 mM EGTA. J Gen Physiol 112: 263–295, 1998. doi: 10.1085/jgp.112.3.263. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 264. Xu T, Ashery U, Burgoyne RD, Neher E. Early requirement for alpha-SNAP and NSF in the secretory cascade in chromaffin cells. EMBO J 18: 3293–3304, 1999. doi: 10.1093/emboj/18.12.3293. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 265. Yamada WM, Zucker RS. Time course of transmitter release calculated from simulations of a calcium diffusion model. Biophys J 61: 671–682, 1992. doi: 10.1016/S0006-3495(92)81872-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 266. Zucker RS, Fogelson AL. Relationship between transmitter release and presynaptic calcium influx when calcium enters through discrete channels. Proc Natl Acad Sci USA 83: 3032–3036, 1986. doi: 10.1073/pnas.83.9.3032. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 267. Eggermann E, Jonas P. How the ‘slow’ Ca2+ buffer parvalbumin affects transmitter release in nanodomain-coupling regimes. Nat Neurosci 15: 20–22, 2011. doi: 10.1038/nn.3002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 268. Zador A, Koch C. Linearized models of calcium dynamics: formal equivalence to the cable equation. J Neurosci 14: 4705–4715, 1994. doi: 10.1523/JNEUROSCI.14-08-04705.1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 269. Naraghi M, Neher E. Linearized buffered Ca2+ diffusion in microdomains and its implications for calculation of [Ca2+] at the mouth of a calcium channel. J Neurosci 17: 6961–6973, 1997. doi: 10.1523/JNEUROSCI.17-18-06961.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 270. Smith GD, Wagner J, Keizer J. Validity of the rapid buffering approximation near a point source of calcium ions. Biophys J 70: 2527–2539, 1996. doi: 10.1016/S0006-3495(96)79824-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 271. Bookchin RM, Lew VL. Progressive inhibition of the Ca pump and Ca:Ca exchange in sickle red cells. Nature 284: 561–563, 1980. doi: 10.1038/284561a0. [DOI] [PubMed] [Google Scholar]
- 272. Gent WL, Trounce JR, Walser M. The binding of calcium ion by the human erythrocyte membrane. Arch Biochem Biophys 105: 582–589, 1964. doi: 10.1016/0003-9861(64)90054-2. [DOI] [PubMed] [Google Scholar]
- 273. Schatzmann HJ. Dependence on calcium concentration and stoichiometry of the calcium pump in human red cells. J Physiol 235: 551–569, 1973. doi: 10.1113/jphysiol.1973.sp010403. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 274. Ferreira HG, Lew VL. Use of ionophore A23187 to measure cytoplasmic Ca buffering and activation of the Ca pump by internal Ca. Nature 259: 47–49, 1976. doi: 10.1038/259047a0. [DOI] [PubMed] [Google Scholar]
- 275. Tiffert T, Lew VL. Cytoplasmic calcium buffers in intact human red cells. J Physiol 500: 139–154, 1997. doi: 10.1113/jphysiol.1997.sp022005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 276. Low FM, Hampton MB, Winterbourn CC. Peroxiredoxin 2 and peroxide metabolism in the erythrocyte. Antioxid Redox Signal 10: 1621–1630, 2008. doi: 10.1089/ars.2008.2081. [DOI] [PubMed] [Google Scholar]
- 277. Clark MR. Senescence of red blood cells: progress and problems. Physiol Rev 68: 503–554, 1988. doi: 10.1152/physrev.1988.68.2.503. [DOI] [PubMed] [Google Scholar]
- 278. Cahalan SM, Lukacs V, Ranade SS, Chien S, Bandell M, Patapoutian A. Piezo1 links mechanical forces to red blood cell volume. Elife 4: e07370, 2015. doi: 10.7554/eLife.07370. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 279. Lew VL, Tiffert T. On the mechanism of human red blood cell longevity: roles of calcium, the sodium pump, PIEZO1, and Gardos channels. Front Physiol 8: 977, 2017. doi: 10.3389/fphys.2017.00977. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 280. Rogers S, Lew VL. Up-down biphasic volume response of human red blood cells to PIEZO1 activation during capillary transits. PLoS Comput Biol 17: e1008706, 2021. doi: 10.1371/journal.pcbi.1008706. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 281. Rogers S, Lew VL. PIEZO1 and the mechanism of the long circulatory longevity of human red blood cells. PLoS Comput Biol 17: e1008496, 2021. doi: 10.1371/journal.pcbi.1008496. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 282. Lew VL. The circulatory dynamics of human red blood cell homeostasis: oxy-deoxy and PIEZO1-triggered changes. Biophys J 122: 484–495, 2023. doi: 10.1016/j.bpj.2022.12.038. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 283. Solaro RJ, Wise RM, Shiner JS, Briggs FN. Calcium requirements for cardiac myofibrillar activation. Circ Res 34: 525–530, 1974. doi: 10.1161/01.res.34.4.525. [DOI] [PubMed] [Google Scholar]
- 284. Pierce GN, Philipson KD, Langer GA. Passive calcium-buffering capacity of a rabbit ventricular homogenate preparation. Am J Physiol Cell Physiol 249: C248–C255, 1985. doi: 10.1152/ajpcell.1985.249.3.C248. [DOI] [PubMed] [Google Scholar]
- 285. Hove-Madsen L, Bers DM. Passive Ca buffering and SR Ca uptake in permeabilized rabbit ventricular myocytes. Am J Physiol Cell Physiol 264: C677–C86, 1993. doi: 10.1152/ajpcell.1993.264.3.C677. [DOI] [PubMed] [Google Scholar]
- 286. Sipido KR, Wier WG. Flux of Ca2+ across the sarcoplasmic reticulum of guinea-pig cardiac cells during excitation-contraction coupling. J Physiol 435: 605–630, 1991. doi: 10.1113/jphysiol.1991.sp018528. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 287. Berlin JR, Bassani JW, Bers DM. Intrinsic cytosolic calcium buffering properties of single rat cardiac myocytes. Biophys J 67: 1775–1787, 1994. doi: 10.1016/S0006-3495(94)80652-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 288. Konishi M, Olson A, Hollingworth S, Baylor SM. Myoplasmic binding of fura-2 investigated by steady-state fluorescence and absorbance measurements. Biophys J 54: 1089–1104, 1988. doi: 10.1016/S0006-3495(88)83045-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 289. Klein MG, Simon BJ, Szucs G, Schneider MF. Simultaneous recording of calcium transients in skeletal muscle using high- and low-affinity calcium indicators. Biophys J 53: 971–988, 1988. doi: 10.1016/S0006-3495(88)83178-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 290. Poenie M. Alteration of intracellular Fura-2 fluorescence by viscosity: a simple correction. Cell Calcium 11: 85–91, 1990. doi: 10.1016/0143-4160(90)90062-y. [DOI] [PubMed] [Google Scholar]
- 291. Popov EG, Gavrilov Y, Pozin E, Gabbasov ZA. Multiwavelength method for measuring concentration of free cytosolic calcium using the fluorescent probe indo-1. Arch Biochem Biophys 261: 91–96, 1988. doi: 10.1016/0003-9861(88)90107-5. [DOI] [PubMed] [Google Scholar]
- 292. Hove-Madsen L, Bers DM. Indo-1 binding to protein in permeabilized ventricular myocytes alters its spectral and Ca binding properties. Biophys J 63: 89–97, 1992. doi: 10.1016/S0006-3495(92)81597-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 293. Swietach P, Spitzer KW, Vaughan-Jones RD. Ca2+-mobility in the sarcoplasmic reticulum of ventricular myocytes is low. Biophys J 95: 1412–1427, 2008. doi: 10.1529/biophysj.108.130385. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 294. Fabiato A. Calcium-induced release of calcium from the cardiac sarcoplasmic reticulum. Am J Physiol Cell Physiol 245: C1–C14, 1983. doi: 10.1152/ajpcell.1983.245.1.C1. [DOI] [PubMed] [Google Scholar]
- 295. Feest ER, Steven Korte F, Tu AY, Dai J, Razumova MV, Murry CE, Regnier M. Thin filament incorporation of an engineered cardiac troponin C variant (L48Q) enhances contractility in intact cardiomyocytes from healthy and infarcted hearts. J Mol Cell Cardiol 72: 219–227, 2014. doi: 10.1016/j.yjmcc.2014.03.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 296. Davis J, Davis LC, Correll RN, Makarewich CA, Schwanekamp JA, Moussavi-Harami F, Wang D, York AJ, Wu H, Houser SR, Seidman CE, Seidman JG, Regnier M, Metzger JM, Wu JC, Molkentin JD. A tension-based model distinguishes hypertrophic versus dilated cardiomyopathy. Cell 165: 1147–1159, 2016. doi: 10.1016/j.cell.2016.04.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 297. Jeon YK, Kwon JW, Jang J, Choi SW, Woo J, Cho SH, Yu BI, Chun YS, Youm JB, Zhang YH, Kim SJ. Lower troponin expression in the right ventricle of rats explains interventricular differences in E-C coupling. J Gen Physiol 154: e202112949, 2022. doi: 10.1085/jgp.202112949. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 298. Briston SJ, Dibb KM, Solaro RJ, Eisner DA, Trafford AW. Balanced changes in Ca buffering by SERCA and troponin contribute to Ca handling during β-adrenergic stimulation in cardiac myocytes. Cardiovasc Res 104: 347–354, 2014. doi: 10.1093/cvr/cvu201. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 299. Greiser M, Kerfant BG, Williams GS, Voigt N, Harks E, Dibb KM, Giese A, Meszaros J, Verheule S, Ravens U, Allessie MA, Gammie JS, van der Velden J, Lederer WJ, Dobrev D, Schotten U. Tachycardia-induced silencing of subcellular Ca2+ signaling in atrial myocytes. J Clin Invest 124: 4759–4772, 2014. doi: 10.1172/JCI70102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 300. Jung P, Seibertz F, Fakuade FE, Ignatyeva N, Sampathkumar S, Ritter M, Li H, Mason FE, Ebert A, Voigt N. Increased cytosolic calcium buffering contributes to a cellular arrhythmogenic substrate in iPSC-cardiomyocytes from patients with dilated cardiomyopathy. Basic Res Cardiol 117: 5, 2022. doi: 10.1007/s00395-022-00912-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 301. Holroyde MJ, Robertson SP, Johnson JD, Solaro RJ, Potter JD. The calcium and magnesium binding sites on cardiac troponin and their role in the regulation of myofibrillar adenosine triphosphatase. J Biol Chem 255: 11688–11693, 1980. doi: 10.1016/S0021-9258(19)70187-2. [DOI] [PubMed] [Google Scholar]
- 302. Györke I, Hester N, Jones LR, Györke S. The role of calsequestrin, triadin, and junctin in conferring cardiac ryanodine receptor responsiveness to luminal calcium. Biophys J 86: 2121–2128, 2004. doi: 10.1016/S0006-3495(04)74271-X. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 303. Guo T, Ai X, Shannon TR, Pogwizd SM, Bers DM. Intra-sarcoplasmic reticulum free [Ca2+] and buffering in arrhythmogenic failing rabbit heart. Circ Res 101: 802–810, 2007. doi: 10.1161/CIRCRESAHA.107.152140. [DOI] [PubMed] [Google Scholar]
- 304. Knollmann BC, Chopra N, Hlaing T, Akin B, Yang T, Ettensohn K, Knollmann BE, Horton KD, Weissman NJ, Holinstat I, Zhang W, Roden DM, Jones LR, Franzini-Armstrong C, Pfeifer K. Casq2 deletion causes sarcoplasmic reticulum volume increase, premature Ca2+ release, and catecholaminergic polymorphic ventricular tachycardia. J Clin Invest 116: 2510–2520, 2006. doi: 10.1172/JCI29128. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 305. Terentyev D, Viatchenko-Karpinski S, Györke I, Volpe P, Williams SC, Györke S. Calsequestrin determines the functional size and stability of cardiac intracellular calcium stores: Mechanism for hereditary arrhythmia. Proc Natl Acad Sci USA 100: 11759–11764, 2003. doi: 10.1073/pnas.1932318100. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 306. Kim E, Youn B, Kemper L, Campbell C, Milting H, Varsanyi M, Kang C. Characterization of human cardiac calsequestrin and its deleterious mutants. J Mol Biol 373: 1047–1057, 2007. doi: 10.1016/j.jmb.2007.08.055. [DOI] [PubMed] [Google Scholar]
- 307. Postma AV, Denjoy I, Hoorntje TM, Lupoglazoff JM, Da Costa A, Sebillon P, Mannens MM, Wilde AA, Guicheney P. Absence of calsequestrin 2 causes severe forms of catecholaminergic polymorphic ventricular tachycardia. Circ Res 91: e21–e26, 2002. doi: 10.1161/01.res.0000038886.18992.6b. [DOI] [PubMed] [Google Scholar]
- 308. Song L, Alcalai R, Arad M, Wolf CM, Toka O, Conner DA, Berul CI, Eldar M, Seidman CE, Seidman JG. Calsequestrin 2 (CASQ2) mutations increase expression of calreticulin and ryanodine receptors, causing catecholaminergic polymorphic ventricular tachycardia. J Clin Invest 117: 1814–1823, 2007. doi: 10.1172/JCI31080. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 309. MacQuaide N, Dempster J, Smith GL. Measurement and modeling of Ca2+ waves in isolated rabbit ventricular cardiomyocytes. Biophys J 93: 2581–2595, 2007. doi: 10.1529/biophysj.106.102293. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 310. Gattoni S, Røe ÅT, Frisk M, Louch WE, Niederer SA, Smith NP. The calcium–frequency response in the rat ventricular myocyte: an experimental and modelling study. J Physiol 594: 4193–4224, 2016. doi: 10.1113/JP272011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 311. Layland J, Kentish JC. Positive force- and [Ca2+]i-frequency relationships in rat ventricular trabeculae at physiological frequencies. Am J Physiol Heart Circ Physiol 276: H9–H18, 1999. doi: 10.1152/ajpheart.1999.276.1.H9. [DOI] [PubMed] [Google Scholar]
- 312. Antoons G, Mubagwa K, Nevelsteen I, Sipido KR. Mechanisms underlying the frequency dependence of contraction and [Ca2+]i transients in mouse ventricular myocytes. J Physiol 543: 889–898, 2002. doi: 10.1113/jphysiol.2002.025619. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 313. Sankaranarayanan R, Kistamás K, Greensmith DJ, Venetucci LA, Eisner DA. Systolic [Ca2+]i regulates diastolic levels in rat ventricular myocytes. J Physiol 595: 5545–5555, 2017. doi: 10.1113/JP274366. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 314. Wier WG, Hess P. Excitation-contraction coupling in cardiac Purkinje fibers. Effects of cardiotonic steroids on the intracellular [Ca2+] transient, membrane potential, and contraction. J Gen Physiol 83: 395–415, 1984. doi: 10.1085/jgp.83.3.395. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 315. Eisner DA, Caldwell JL, Trafford AW, Hutchings DC. The control of diastolic calcium in the heart: basic mechanisms and functional implications. Circ Res 126: 395–412, 2020. doi: 10.1161/CIRCRESAHA.119.315891. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 316. Bers DM, Berlin JR. Kinetics of [Ca]i decline in cardiac myocytes depend on peak [Ca]i. Am J Physiol Cell Physiol 268: C271–C277, 1995. doi: 10.1152/ajpcell.1995.268.1.C271. [DOI] [PubMed] [Google Scholar]
- 317. Allen DG, Kurihara S. The effects of muscle length on intracellular calcium transients in mammalian cardiac muscle. J Physiol 327: 79–94, 1982. doi: 10.1113/jphysiol.1982.sp014221. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 318. Kurihara S, Komukai K. Tension-dependent changes of the intracellular Ca2+ transients in ferret ventricular muscles. J Physiol 489: 617–625, 1995. doi: 10.1113/jphysiol.1995.sp021077. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 319. Backx PH, Ter Keurs HE. Fluorescent properties of rat cardiac trabeculae microinjected with fura-2 salt. Am J Physiol Heart Circ Physiol 264: H1098–H1110, 1993. doi: 10.1152/ajpheart.1993.264.4.H1098. [DOI] [PubMed] [Google Scholar]
- 320. Kentish JC, Wrzosek A. Changes in force and cytosolic Ca2+ concentration after length changes in isolated rat ventricular trabeculae. J Physiol 506: 431–444, 1998. doi: 10.1111/j.1469-7793.1998.431bw.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 321. Lookin O. The use of Ca-transient to evaluate Ca2+ utilization by myofilaments in living cardiac muscle. Clin Exp Pharmacol Physiol 47: 1824–1833, 2020. doi: 10.1111/1440-1681.13376. [DOI] [PubMed] [Google Scholar]
- 322. Prosser BL, Ward CW, Lederer WJ. X-ROS signaling: rapid mechano-chemo transduction in heart. Science 333: 1440–1445, 2011. doi: 10.1126/science.1202768. [DOI] [PubMed] [Google Scholar]
- 323. Limbu S, Prosser BL, Lederer WJ, Ward CW, Jafri MS. X-ROS signaling depends on length-dependent calcium buffering by troponin. Cells 10: 1189, 2021. doi: 10.3390/cells10051189. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 324. Blanchard EM, Solaro RJ. Inhibition of the activation and troponin calcium binding of dog cardiac myofibrils by acidic pH. Circ Res 55: 382–391, 1984. doi: 10.1161/01.res.55.3.382. [DOI] [PubMed] [Google Scholar]
- 325. Vaughan-Jones RD, Wu ML. pH dependence of intrinsic H+ buffering power in the sheep cardiac Purkinje fibre. J Physiol 425: 429–448, 1990. doi: 10.1113/jphysiol.1990.sp018112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 326. Cairns SP, Westerblad H, Allen DG. Changes in myoplasmic pH and calcium concentration during exposure to lactate in isolated rat ventricular myocytes. J Physiol 464: 561–574, 1993. doi: 10.1113/jphysiol.1993.sp019651. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 327. Hu YL, Mi X, Huang C, Wang HF, Song JR, Shu Q, Ni L, Chen JG, Wang F, Hu ZL. Multiple H+ sensors mediate the extracellular acidification-induced [Ca2+]i elevation in cultured rat ventricular cardiomyocytes. Sci Rep 7: 44951, 2017. doi: 10.1038/srep44951. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 328. Boyman L, Hagen BM, Giladi M, Hiller R, Lederer WJ, Khananshvili D. Proton-sensing Ca2+ binding domains regulate the cardiac Na+/Ca2+ exchanger. J Biol Chem 286: 28811–28820, 2011. doi: 10.1074/jbc.M110.214106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 329. Choi HS, Trafford AW, Orchard CH, Eisner DA. The effect of acidosis on systolic Ca2+ and sarcoplasmic reticulum calcium content in isolated rat ventricular myocytes. J Physiol 529: 661–668, 2000. doi: 10.1111/j.1469-7793.2000.00661.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 330. Papp Z, Agostoni P, Alvarez J, Bettex D, Bouchez S, Brito D, et al. Levosimendan efficacy and safety: 20 years of SIMDAX in clinical use. J Cardiovasc Pharmacol 76: 4–22, 2020. doi: 10.1097/FJC.0000000000000859. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 331. Malik FI, Hartman JJ, Elias KA, Morgan BP, Rodriguez H, Brejc K, et al. Cardiac myosin activation: a potential therapeutic approach for systolic heart failure. Science 331: 1439–1443, 2011. doi: 10.1126/science.1200113. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 332. Dou Y, Arlock P, Arner A. Blebbistatin specifically inhibits actin-myosin interaction in mouse cardiac muscle. Am J Physiol Cell Physiol 293: C1148–C1153, 2007. doi: 10.1152/ajpcell.00551.2006. [DOI] [PubMed] [Google Scholar]
- 333. Baudenbacher F, Schober T, Pinto JR, Sidorov VY, Hilliard F, Solaro RJ, Potter JD, Knollmann BC. Myofilament Ca2+ sensitization causes susceptibility to cardiac arrhythmia in mice. J Clin Invest 118: 3893–3903, 2008. doi: 10.1172/JCI36642. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 334. Venkataraman R, Baldo MP, Hwang HS, Veltri T, Pinto JR, Baudenbacher FJ, Knollmann BC. Myofilament calcium de-sensitization and contractile uncoupling prevent pause-triggered ventricular tachycardia in mouse hearts with chronic myocardial infarction. J Mol Cell Cardiol 60: 8–15, 2013. doi: 10.1016/j.yjmcc.2013.03.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 335. Lehman SJ, Crocini C, Leinwand LA. Targeting the sarcomere in inherited cardiomyopathies. Nat Rev Cardiol 19: 353–363, 2022. doi: 10.1038/s41569-022-00682-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 336. Olivotto I, Oreziak A, Barriales-Villa R, Abraham TP, Masri A, Garcia-Pavia P, Saberi S, Lakdawala NK, Wheeler MT, Owens A, Kubanek M, Wojakowski W, Jensen MK, Gimeno-Blanes J, Afshar K, Myers J, Hegde SM, Solomon SD, Sehnert AJ, Zhang D, Li W, Bhattacharya M, Edelberg JM, Waldman CB, Lester SJ, Wang A, Ho CY, Jacoby D; EXPLORER-HCM study investigators. Mavacamten for treatment of symptomatic obstructive hypertrophic cardiomyopathy (EXPLORER-HCM): a randomised, double-blind, placebo-controlled, phase 3 trial. Lancet 396: 759–769, 2020. doi: 10.1016/S0140-6736(20)31792-X. [DOI] [PubMed] [Google Scholar]
- 337. Sparrow AJ, Watkins H, Daniels MJ, Redwood C, Robinson P. Mavacamten rescues increased myofilament calcium sensitivity and dysregulation of Ca2+ flux caused by thin filament hypertrophic cardiomyopathy mutations. Am J Physiol Heart Circ Physiol 318: H715–H722, 2020. doi: 10.1152/ajpheart.00023.2020. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 338. Hollingworth S, Soeller C, Baylor SM, Cannell MB. Sarcomeric Ca2+ gradients during activation of frog skeletal muscle fibres imaged with confocal and two-photon microscopy. J Physiol 526: 551–560, 2000. doi: 10.1111/j.1469-7793.2000.t01-1-00551.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 339. Cheng H, Lederer WJ, Cannell MB. Calcium sparks: elementary events underlying excitation-contraction coupling in heart muscle. Science 262: 740–744, 1993. doi: 10.1126/science.8235594. [DOI] [PubMed] [Google Scholar]
- 340. Cannell MB, Cheng H, Lederer WJ. The control of calcium release in heart muscle. Science 268: 1045–1049, 1995. doi: 10.1126/science.7754384. [DOI] [PubMed] [Google Scholar]
- 341. Cheng H, Lederer MR, Lederer WJ, Cannell MB. Calcium sparks and [Ca2+]i waves in cardiac myocytes. Am J Physiol Cell Physiol 270: C148–C159, 1996. doi: 10.1152/ajpcell.1996.270.1.C148. [DOI] [PubMed] [Google Scholar]
- 342. Venetucci LA, Trafford AW, O’Neill SC, Eisner DA. The sarcoplasmic reticulum and arrhythmogenic calcium release. Cardiovasc Res 77: 285–292, 2008. doi: 10.1093/cvr/cvm009. [DOI] [PubMed] [Google Scholar]
- 343. Keizer J, Smith GD, Ponce-Dawson S, Pearson JE. Saltatory propagation of Ca2+ waves by Ca2+ sparks. Biophys J 75: 595–600, 1998. doi: 10.1016/S0006-3495(98)77550-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 344. Jafri MS, Keizer J. On the roles of Ca2+ diffusion, Ca2+ buffers, and the endoplasmic reticulum in IP3-induced Ca2+ waves. Biophys J 69: 2139–2153, 1995. doi: 10.1016/S0006-3495(95)80088-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 345. Marchena M, Echebarria B, Shiferaw Y, Alvarez-Lacalle E. Buffering and total calcium levels determine the presence of oscillatory regimes in cardiac cells. PLoS Comput Biol 16: e1007728, 2020. doi: 10.1371/journal.pcbi.1007728. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 346. Chugh SS, Havmoeller R, Narayanan K, Singh D, Rienstra M, Benjamin EJ, Gillum RF, Kim YH, McAnulty JH, Zheng ZJ, Forouzanfar MH, Naghavi M, Mensah GA, Ezzati M, Murray CJ. Worldwide epidemiology of atrial fibrillation. Circulation 129: 837–847, 2014. doi: 10.1161/CIRCULATIONAHA.113.005119. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 347. Sheehan KA, Blatter LA. Regulation of junctional and non-junctional sarcoplasmic reticulum calcium release in excitation-contraction coupling in cat atrial myocytes. J Physiol 546: 119–135, 2003. doi: 10.1113/jphysiol.2002.026963. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 348. Hüser J, Lipsius SL, Blatter LA. Calcium gradients during excitation-contraction coupling in cat atrial myocytes. J Physiol 494: 641–651, 1996. doi: 10.1113/jphysiol.1996.sp021521. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 349. Dibb KM, Clarke JD, Horn MA, Richards MA, Graham HK, Eisner DA, Trafford AW. Characterization of an extensive transverse tubular network in sheep atrial myocytes and its depletion in heart failure. Circ Heart Fail 2: 482–489, 2009. doi: 10.1161/CIRCHEARTFAILURE.109.852228. [DOI] [PubMed] [Google Scholar]
- 350. Lenaerts I, Bito V, Heinzel FR, Driesen RB, Holemans P, D’hooge J, Heidbüchel H, Sipido KR, Willems R. Ultrastructural and functional remodeling of the coupling between Ca2+ influx and sarcoplasmic reticulum Ca2+ release in right atrial myocytes from experimental persistent atrial fibrillation. Circ Res 105: 876–885, 2009. doi: 10.1161/CIRCRESAHA.109.206276. [DOI] [PubMed] [Google Scholar]
- 351. Richards MA, Clarke JD, Saravanan P, Voigt N, Dobrev D, Eisner DA, Trafford AW, Dibb KM. Transverse (t-) tubules are a common feature in large mammalian atrial myocytes including human. Am J Physiol Heart Circ Physiol 301: H1996–H2005, 2011. doi: 10.1152/ajpheart.00284.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 352. Macquaide N, Tuan HT, Hotta J, Sempels W, Lenaerts I, Holemans P, Hofkens J, Jafri MS, Willems R, Sipido KR. Ryanodine receptor cluster fragmentation and redistribution in persistent atrial fibrillation enhance calcium release. Cardiovasc Res 108: 387–398, 2015. doi: 10.1093/cvr/cvv231. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 353. Clarke JD, Caldwell JL, Horn MA, Bode EF, Richards MA, Hall MC, Graham HK, Briston SJ, Greensmith DJ, Eisner DA, Dibb KM, Trafford AW. Perturbed atrial calcium handling in an ovine model of heart failure: potential roles for reductions in the L-type calcium current. J Mol Cell Cardiol 79: 169–179, 2015. doi: 10.1016/j.yjmcc.2014.11.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 354. Fakuade FE, Steckmeister V, Seibertz F, Gronwald J, Kestel S, Menzel J, Pronto JR, Taha K, Haghighi F, Kensah G, Pearman CM, Wiedmann F, Teske AJ, Schmidt C, Dibb KM, El-Essawi A, Danner BC, Baraki H, Schwappach B, Kutschka I, Mason FE, Voigt N. Altered atrial cytosolic calcium handling contributes to the development of postoperative atrial fibrillation. Cardiovasc Res 117: 1790–1801, 2021. doi: 10.1093/cvr/cvaa162. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 355. Wolff MR, Buck SH, Stoker SW, Greaser ML, Mentzer RM. Myofibrillar calcium sensitivity of isometric tension is increased in human dilated cardiomyopathies: role of altered beta-adrenergically mediated protein phosphorylation. J Clin Invest 98: 167–176, 1996. doi: 10.1172/JCI118762. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 356. Wolff MR, Whitesell LF, Moss RL. Calcium sensitivity of isometric tension is increased in canine experimental heart failure. Circ Res 76: 781–789, 1995. doi: 10.1161/01.res.76.5.781. [DOI] [PubMed] [Google Scholar]
- 357. de Waard MC, van der Velden J, Bito V, Ozdemir S, Biesmans L, Boontje NM, Dekkers DH, Schoonderwoerd K, Schuurbiers HC, de Crom R, Stienen GJ, Sipido KR, Lamers JM, Duncker DJ. Early exercise training normalizes myofilament function and attenuates left ventricular pump dysfunction in mice with a large myocardial infarction. Circ Res 100: 1079–1088, 2007. doi: 10.1161/01.RES.0000262655.16373.37. [DOI] [PubMed] [Google Scholar]
- 358. Hegemann N, Primessnig U, Bode D, Wakula P, Beindorff N, Klopfleisch R, Michalick L, Grune J, Hohendanner F, Messroghli D, Pieske B, Kuebler WM, Heinzel FR. Right-ventricular dysfunction in HFpEF is linked to altered cardiomyocyte Ca2+ homeostasis and myofilament sensitivity. ESC Heart Fail 8: 3130–3144, 2021. doi: 10.1002/ehf2.13419. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 359. Moss RL, Fitzsimons DP, Ralphe JC. Cardiac MyBP-C regulates the rate and force of contraction in mammalian myocardium. Circ Res 116: 183–192, 2015. doi: 10.1161/CIRCRESAHA.116.300561. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 360. Hobai IA, O’Rourke B. Enhanced Ca2+-activated Na+-Ca2+ exchange activity in canine pacing-induced heart failure. Circ Res 87: 690–698, 2000. doi: 10.1161/01.res.87.8.690. [DOI] [PubMed] [Google Scholar]
- 361. Briston SJ, Caldwell JL, Horn MA, Clarke JD, Richards MA, Greensmith DJ, Graham HK, Hall MC, Eisner DA, Dibb KM, Trafford AW. Impaired β-adrenergic responsiveness accentuates dysfunctional excitation-contraction coupling in an ovine model of tachypacing-induced heart failure. J Physiol 589: 1367–1382, 2011. doi: 10.1113/jphysiol.2010.203984. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 362. Fatkin D, Graham RM. Molecular mechanisms of inherited cardiomyopathies. Physiol Rev 82: 945–980, 2002. doi: 10.1152/physrev.00012.2002. [DOI] [PubMed] [Google Scholar]
- 363. McNally EM, Mestroni L. Dilated cardiomyopathy: genetic determinants and mechanisms. Circ Res 121: 731–748, 2017. doi: 10.1161/CIRCRESAHA.116.309396. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 364. Robinson P, Griffiths PJ, Watkins H, Redwood CS. Dilated and hypertrophic cardiomyopathy mutations in troponin and alpha-tropomyosin have opposing effects on the calcium affinity of cardiac thin filaments. Circ Res 101: 1266–1273, 2007. doi: 10.1161/CIRCRESAHA.107.156380. [DOI] [PubMed] [Google Scholar]
- 365. Li Y, Zhang L, Jean-Charles PY, Nan C, Chen G, Tian J, Jin JP, Gelb IJ, Huang X. Dose-dependent diastolic dysfunction and early death in a mouse model with cardiac troponin mutations. J Mol Cell Cardiol 62: 227–236, 2013. doi: 10.1016/j.yjmcc.2013.06.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 366. Zile MA, Trayanova NA. Myofilament protein dynamics modulate EAD formation in human hypertrophic cardiomyopathy. Prog Biophys Mol Biol 130: 418–428, 2017. doi: 10.1016/j.pbiomolbio.2017.06.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 367. Wang L, Kryshtal DO, Kim K, Parikh S, Cadar AG, Bersell KR, He H, Pinto JR, Knollmann BC. Myofilament calcium-buffering dependent action potential triangulation in human-induced pluripotent stem cell model of hypertrophic cardiomyopathy. J Am Coll Cardiol 70: 2600–2602, 2017. doi: 10.1016/j.jacc.2017.09.033. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 368. Zile MA, Trayanova NA. Increased thin filament activation enhances alternans in human chronic atrial fibrillation. Am J Physiol Heart Circ Physiol 315: H1453–H1462, 2018. doi: 10.1152/ajpheart.00658.2017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 369. Qu Z, Weiss JN. Cardiac alternans: from bedside to bench and back. Circ Res 132: 127–149, 2023. doi: 10.1161/CIRCRESAHA.122.321668. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 370. Pastore JM, Girouard SD, Laurita KR, Akar FG, Rosenbaum DS. Mechanism linking T-wave alternans to the genesis of cardiac fibrillation. Circulation 99: 1385–1394, 1999. doi: 10.1161/01.cir.99.10.1385. [DOI] [PubMed] [Google Scholar]
- 371. Zile MA, Trayanova NA. Rate-dependent force, intracellular calcium, and action potential voltage alternans are modulated by sarcomere length and heart failure induced-remodeling of thin filament regulation in human heart failure: a myocyte modeling study. Prog Biophys Mol Biol 120: 270–280, 2016. doi: 10.1016/j.pbiomolbio.2015.12.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 372. Dai Y, Amenov A, Ignatyeva N, Koschinski A, Xu H, Soong PL, Tiburcy M, Linke WA, Zaccolo M, Hasenfuss G, Zimmermann WH, Ebert A. Troponin destabilization impairs sarcomere-cytoskeleton interactions in iPSC-derived cardiomyocytes from dilated cardiomyopathy patients. Sci Rep 10: 209, 2020. doi: 10.1038/s41598-019-56597-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 373. Hennig R, Lømo T. Gradation of force output in normal fast and slow muscles of the rat. Acta Physiol Scand 130: 133–142, 1987. doi: 10.1111/j.1748-1716.1987.tb08119.x. [DOI] [PubMed] [Google Scholar]
- 374. Denny-Brown DE, Sherrington CS. The histological features of striped muscle in relation to its functional activity. Proc R Soc Lond Ser B 104: 371–411, 1929. [Google Scholar]
- 375. Tavi P, Westerblad H. The role of in vivo Ca2+ signals acting on Ca2+-calmodulin-dependent proteins for skeletal muscle plasticity. J Physiol 589: 5021–5031, 2011. doi: 10.1113/jphysiol.2011.212860. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 376. Cannell MB, Allen DG. Model of calcium movements during activation in the sarcomere of frog skeletal muscle. Biophys J 45: 913–925, 1984. doi: 10.1016/S0006-3495(84)84238-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 377. Baylor SM, Hollingworth S. Sarcoplasmic reticulum calcium release compared in slow-twitch and fast-twitch fibres of mouse muscle. J Physiol 551: 125–138, 2003. doi: 10.1113/jphysiol.2003.041608. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 378. Kovacs L, Rios E, Schneider MF. Measurement and modification of free calcium transients in frog skeletal muscle fibres by a metallochromic indicator dye. J Physiol 343: 161–196, 1983. doi: 10.1113/jphysiol.1983.sp014887. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 379. Melzer W, Ríos E, Schneider MF. The removal of myoplasmic free calcium following calcium release in frog skeletal muscle. J Physiol 372: 261–292, 1986. doi: 10.1113/jphysiol.1986.sp016008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 380. García J, Amador M, Stefani E. Relationship between myoplasmic calcium transients and calcium currents in frog skeletal muscle. J Gen Physiol 94: 973–986, 1989. doi: 10.1085/jgp.94.6.973. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 381. Boldyrev AA, Aldini G, Derave W. Physiology and pathophysiology of carnosine. Physiol Rev 93: 1803–1845, 2013. doi: 10.1152/physrev.00039.2012. [DOI] [PubMed] [Google Scholar]
- 382. Lamont C, Miller DJ. Calcium sensitizing action of carnosine and other endogenous imidazoles in chemically skinned striated muscle. J Physiol 454: 421–434, 1992. doi: 10.1113/jphysiol.1992.sp019271. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 383. Pauls TL, Cox JA, Berchtold MW. The Ca2+-binding proteins parvalbumin and oncomodulin and their genes: new structural and functional findings. Biochim Biophys Acta 1306: 39–54, 1996. doi: 10.1016/0167-4781(95)00221-9. [DOI] [PubMed] [Google Scholar]
- 384. Henzl MT, Larson JD, Agah S. Estimation of parvalbumin Ca2+- and Mg2+-binding constants by global least-squares analysis of isothermal titration calorimetry data. Anal Biochem 319: 216–233, 2003. doi: 10.1016/s0003-2697(03)00288-4. [DOI] [PubMed] [Google Scholar]
- 385. Heizmann CW, Berchtold MW, Rowlerson AM. Correlation of parvalbumin concentration with relaxation speed in mammalian muscles. Proc Natl Acad Sci USA 79: 7243–7247, 1982. doi: 10.1073/pnas.79.23.7243. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 386. Green HJ, Klug GA, Reichmann H, Seedorf U, Wiehrer W, Pette D. Exercise-induced fibre type transitions with regard to myosin, parvalbumin, and sarcoplasmic reticulum in muscles of the rat. Pflugers Arch 400: 432–438, 1984. doi: 10.1007/BF00587545. [DOI] [PubMed] [Google Scholar]
- 387. Müntener M, Käser L, Weber J, Berchtold MW. Increase of skeletal muscle relaxation speed by direct injection of parvalbumin cDNA. Proc Natl Acad Sci USA 92: 6504–6508, 1995. doi: 10.1073/pnas.92.14.6504. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 388. Rall J. Role of parvalbumin in skeletal muscle relaxation. Physiology 11: 249–255, 1996. doi: 10.1152/physiologyonline.1996.11.6.249. [DOI] [Google Scholar]
- 389. Jiang Y, Johnson JD, Rall JA. Parvalbumin relaxes frog skeletal muscle when sarcoplasmic reticulum Ca2+-ATPase is inhibited. Am J Physiol Cell Physiol 270: C411–C417, 1996. doi: 10.1152/ajpcell.1996.270.2.C411. [DOI] [PubMed] [Google Scholar]
- 390. Carroll SL, Klein MG, Schneider MF. Decay of calcium transients after electrical stimulation in rat fast- and slow-twitch skeletal muscle fibres. J Physiol 501: 573–588, 1997. doi: 10.1111/j.1469-7793.1997.573bm.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 391. Hou TT, Johnson JD, Rall JA. Parvalbumin content and Ca2+ and Mg2+ dissociation rates correlated with changes in relaxation rate of frog muscle fibres. J Physiol 441: 285–304, 1991. doi: 10.1113/jphysiol.1991.sp018752. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 392. Johnson JD, Jiang Y, Flynn M. Modulation of Ca2+ transients and tension by intracellular EGTA in intact frog muscle fibers. Am J Physiol Cell Physiol 272: C1437–C1444, 1997. doi: 10.1152/ajpcell.1997.272.5.C1437. [DOI] [PubMed] [Google Scholar]
- 393. Calderón JC, Bolaños P, Caputo C. Tetanic Ca2+ transient differences between slow- and fast-twitch mouse skeletal muscle fibres: a comprehensive experimental approach. J Muscle Res Cell Motil 35: 279–293, 2014. doi: 10.1007/s10974-014-9388-7. [DOI] [PubMed] [Google Scholar]
- 394. Coutu P, Metzger JM. Genetic manipulation of calcium-handling proteins in cardiac myocytes. II. Mathematical modeling studies. Am J Physiol Heart Circ Physiol 288: H613–H631, 2005. doi: 10.1152/ajpheart.00425.2004. [DOI] [PubMed] [Google Scholar]
- 395. Hou TT, Johnson JD, Rall JA. Effect of temperature on relaxation rate and Ca2+, Mg2+ dissociation rates from parvalbumin of frog muscle fibres. J Physiol 449: 399–410, 1992. doi: 10.1113/jphysiol.1992.sp019092. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 396. Nogueira L, Gilmore NK, Hogan MC. Role of parvalbumin in fatigue-induced changes in force and cytosolic calcium transients in intact single mouse myofibers. J Appl Physiol (1985) 132: 1041–1053, 2022. doi: 10.1152/japplphysiol.00861.2021. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 397. Chen G, Carroll S, Racay P, Dick J, Pette D, Traub I, Vrbova G, Eggli P, Celio M, Schwaller B. Deficiency in parvalbumin increases fatigue resistance in fast-twitch muscle and upregulates mitochondria. Am J Physiol Cell Physiol 281: C114–C122, 2001. doi: 10.1152/ajpcell.2001.281.1.C114. [DOI] [PubMed] [Google Scholar]
- 398. Schuppe ER, Petersen JO, Fuxjager MJ. Woodpecker drumming behavior is linked to the elevated expression of genes that encode calcium handling proteins in the neck musculature. J Exp Biol 221, 2018. doi: 10.1242/jeb.180190. [DOI] [PubMed] [Google Scholar]
- 399. Föhr UG, Weber BR, Müntener M, Staudenmann W, Hughes GJ, Frutiger S, Banville D, Schäfer BW, Heizmann CW. Human alpha and beta parvalbumins. Structure and tissue-specific expression. Eur J Biochem 215: 719–727, 1993. doi: 10.1111/j.1432-1033.1993.tb18084.x. [DOI] [PubMed] [Google Scholar]
- 400. Westerblad H, Allen DG. Mechanisms underlying changes of tetanic [Ca2+]i and force in skeletal muscle. Acta Physiol Scand 156: 407–416, 1996. doi: 10.1046/j.1365-201X.1996.196000.x. [DOI] [PubMed] [Google Scholar]
- 401. Desmedt JE, Godaux E. Ballistic contractions in man: characteristic recruitment pattern of single motor units of the tibialis anterior muscle. J Physiol 264: 673–693, 1977. doi: 10.1113/jphysiol.1977.sp011689. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 402. Hennig R, Lømo T. Firing patterns of motor units in normal rats. Nature 314: 164–166, 1985. doi: 10.1038/314164a0. [DOI] [PubMed] [Google Scholar]
- 403. Westerblad H, Allen DG. The influence of intracellular pH on contraction, relaxation and [Ca2+]i in intact single fibres from mouse muscle. J Physiol 466: 611–628, 1993. [PMC free article] [PubMed] [Google Scholar]
- 404. Duke AM, Steele DS. Interdependent effects of inorganic phosphate and creatine phosphate on sarcoplasmic reticulum Ca2+ regulation in mechanically skinned rat skeletal muscle. J Physiol 531: 729–742, 2001. doi: 10.1111/j.1469-7793.2001.0729h.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 405. Dutka TL, Cole L, Lamb GD. Calcium phosphate precipitation in the sarcoplasmic reticulum reduces action potential-mediated Ca2+ release in mammalian skeletal muscle. Am J Physiol Cell Physiol 289: C1502–C1512, 2005. doi: 10.1152/ajpcell.00273.2005. [DOI] [PubMed] [Google Scholar]
- 406. Steele DS, Duke AM. Metabolic factors contributing to altered Ca2+ regulation in skeletal muscle fatigue. Acta Physiol Scand 179: 39–48, 2003. doi: 10.1046/j.1365-201X.2003.01169.x. [DOI] [PubMed] [Google Scholar]
- 407. Fénelon K, Lamboley CR, Carrier N, Pape PC. Calcium buffering properties of sarcoplasmic reticulum and calcium-induced Ca2+ release during the quasi-steady level of release in twitch fibers from frog skeletal muscle. J Gen Physiol 140: 403–419, 2012. doi: 10.1085/jgp.201110730. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 408. Fliegel L, Ohnishi M, Carpenter MR, Khanna VK, Reithmeier RA, MacLennan DH. Amino acid sequence of rabbit fast-twitch skeletal muscle calsequestrin deduced from cDNA and peptide sequencing. Proc Natl Acad Sci USA 84: 1167–1171, 1987. doi: 10.1073/pnas.84.5.1167. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 409. Pape PC, Fénelon K, Lamboley CR, Stachura D. Role of calsequestrin evaluated from changes in free and total calcium concentrations in the sarcoplasmic reticulum of frog cut skeletal muscle fibres. J Physiol 581: 319–367, 2007. doi: 10.1113/jphysiol.2006.126474. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 410. Manno C, Sztretye M, Figueroa L, Allen PD, Ríos E. Dynamic measurement of the calcium buffering properties of the sarcoplasmic reticulum in mouse skeletal muscle. J Physiol 591: 423–442, 2013. doi: 10.1113/jphysiol.2012.243444. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 411. Manno C, Figueroa LC, Gillespie D, Fitts R, Kang C, Franzini-Armstrong C, Rios E. Calsequestrin depolymerizes when calcium is depleted in the sarcoplasmic reticulum of working muscle. Proc Natl Acad Sci USA 114: E638–E647, 2017. doi: 10.1073/pnas.1620265114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 412. Royer L, Sztretye M, Manno C, Pouvreau S, Zhou J, Knollmann BC, Protasi F, Allen PD, Ríos E. Paradoxical buffering of calcium by calsequestrin demonstrated for the calcium store of skeletal muscle. J Gen Physiol 136: 325–338, 2010. doi: 10.1085/jgp.201010454. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 413. Perni S, Close M, Franzini-Armstrong C. Novel details of calsequestrin gel conformation in situ. J Biol Chem 288: 31358–31362, 2013. doi: 10.1074/jbc.M113.507749. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 414. Launikonis BS, Zhou J, Royer L, Shannon TR, Brum G, Ríos E. Depletion “skraps” and dynamic buffering inside the cellular calcium store. Proc Natl Acad Sci USA 103: 2982–2987, 2006. doi: 10.1073/pnas.0511252103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 415. Rossi D, Vezzani B, Galli L, Paolini C, Toniolo L, Pierantozzi E, Spinozzi S, Barone V, Pegoraro E, Bello L, Cenacchi G, Vattemi G, Tomelleri G, Ricci G, Siciliano G, Protasi F, Reggiani C, Sorrentino V. A mutation in the CASQ1 gene causes a vacuolar myopathy with accumulation of sarcoplasmic reticulum protein aggregates. Hum Mutat 35: 1163–1170, 2014. doi: 10.1002/humu.22631. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 416. Kraeva N, Zvaritch E, Frodis W, Sizova O, Kraev A, MacLennan DH, Riazi S. CASQ1 gene is an unlikely candidate for malignant hyperthermia susceptibility in the North American population. Anesthesiology 118: 344–349, 2013. doi: 10.1097/01.anes.0000530185.78660.da. [DOI] [PubMed] [Google Scholar]
- 417. Lewis KM, Ronish LA, Ríos E, Kang C. Characterization of two human skeletal calsequestrin mutants implicated in malignant hyperthermia and vacuolar aggregate myopathy. J Biol Chem 290: 28665–28674, 2015. doi: 10.1074/jbc.M115.686261. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 418. Peachey LD. The sarcoplasmic reticulum and transverse tubules of the frog’s sartorius. J Cell Biol 25: 209–231, 1965. doi: 10.1083/jcb.25.3.209. [DOI] [PubMed] [Google Scholar]
- 419. Franzini-Armstrong C, Ferguson DG, Champ C. Discrimination between fast- and slow-twitch fibres of guinea pig skeletal muscle using the relative surface density of junctional transverse tubule membrane. J Muscle Res Cell Motil 9: 403–414, 1988. doi: 10.1007/BF01774067. [DOI] [PubMed] [Google Scholar]
- 420. Cannell MB, Kong CHT, Imtiaz MS, Laver DR. Control of sarcoplasmic reticulum Ca2+ release by stochastic RyR gating within a 3D model of the cardiac dyad and importance of induction decay for CICR termination. Biophys J 104: 2149–2159, 2013. doi: 10.1016/j.bpj.2013.03.058. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 421. Asp ML, Martindale JJ, Heinis FI, Wang W, Metzger JM. Calcium mishandling in diastolic dysfunction: mechanisms and potential therapies. Biochim Biophys Acta 1833: 895–900, 2013. doi: 10.1016/j.bbamcr.2012.09.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 422. Hirsch JC, Borton AR, Albayya FP, Russell MW, Ohye RG, Metzger JM. Comparative analysis of parvalbumin and SERCA2a cardiac myocyte gene transfer in a large animal model of diastolic dysfunction. Am J Physiol Heart Circ Physiol 286: H2314–H2321, 2004. doi: 10.1152/ajpheart.01137.2003. [DOI] [PubMed] [Google Scholar]
- 423. Rodenbaugh DW, Wang W, Davis J, Edwards T, Potter JD, Metzger JM. Parvalbumin isoforms differentially accelerate cardiac myocyte relaxation kinetics in an animal model of diastolic dysfunction. Am J Physiol Heart Circ Physiol 293: H1705–H1713, 2007. doi: 10.1152/ajpheart.00232.2007. [DOI] [PubMed] [Google Scholar]
- 424. Coutu P, Metzger JM. Genetic manipulation of calcium-handling proteins in cardiac myocytes. I. Experimental studies. Am J Physiol Heart Circ Physiol 288: H601–H612, 2005. doi: 10.1152/ajpheart.00424.2004. [DOI] [PubMed] [Google Scholar]
- 425. Asp ML, Sjaastad FV, Siddiqui JK, Davis JP, Metzger JM. Effects of modified parvalbumin EF-hand motifs on cardiac myocyte contractile function. Biophys J 110: 2094–2105, 2016. doi: 10.1016/j.bpj.2016.03.037. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 426. Guerrero A, Singer JJ, Fay FS. Simultaneous measurement of Ca2+ release and influx into smooth muscle cells in response to caffeine. A novel approach for calculating the fraction of current carried by calcium. J Gen Physiol 104: 395–422, 1994. doi: 10.1085/jgp.104.2.395. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 427. Kamishima T, McCarron JG. Depolarization-evoked increases in cytosolic calcium concentration in isolated smooth muscle cells of rat portal vein. J Physiol 492: 61–74, 1996. doi: 10.1113/jphysiol.1996.sp021289. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 428. Daub B, Ganitkevich VY. An estimate of rapid cytoplasmic calcium buffering in a single smooth muscle cell. Cell Calcium 27: 3–13, 2000. doi: 10.1054/ceca.1999.0084. [DOI] [PubMed] [Google Scholar]
- 429. Bradley KN, Craig JW, Muir TC, McCarron JG. The sarcoplasmic reticulum and sarcolemma together form a passive Ca2+ trap in colonic smooth muscle. Cell Calcium 36: 29–41, 2004. doi: 10.1016/j.ceca.2003.11.008. [DOI] [PubMed] [Google Scholar]
- 430. Ganitkevich VY, Isenberg G. Efficacy of peak Ca2+ currents (ICa) as trigger of sarcoplasmic reticulum Ca2+ release in myocytes from the guinea-pig coronary artery. J Physiol 484: 287–306, 1995. doi: 10.1113/jphysiol.1995.sp020665. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 431. Ganitkevich VY. The amount of acetylcholine mobilisable Ca2+ in single smooth muscle cells measured with the exogenous cytoplasmic Ca2+ buffer, Indo-1. Cell Calcium 20: 483–492, 1996. doi: 10.1016/s0143-4160(96)90090-1. [DOI] [PubMed] [Google Scholar]
- 432. Kargacin G, Fay FS. Ca2+ movement in smooth muscle cells studied with one- and two-dimensional diffusion models. Biophys J 60: 1088–1100, 1991. doi: 10.1016/S0006-3495(91)82145-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 433. Rüegg JC, Pfitzer G, Zimmer M, Hofmann F. The calmodulin fraction responsible for contraction in an intestinal smooth muscle. FEBS Lett 170: 383–386, 1984. doi: 10.1016/0014-5793(84)81349-6. [DOI] [PubMed] [Google Scholar]
- 434. Tansey MG, Luby-Phelps K, Kamm KE, Stull JT. Ca2+-dependent phosphorylation of myosin light chain kinase decreases the Ca2+ sensitivity of light chain phosphorylation within smooth muscle cells. J Biol Chem 269: 9912–9920, 1994. doi: 10.1016/S0021-9258(17)36969-7. [DOI] [PubMed] [Google Scholar]
- 435. Olwin BB, Storm DR. Calcium binding to complexes of calmodulin and calmodulin binding proteins. Biochemistry 24: 8081–8086, 1985. doi: 10.1021/bi00348a037. [DOI] [PubMed] [Google Scholar]
- 436. Bond M, Shuman H, Somlyo AP, Somlyo AV. Total cytoplasmic calcium in relaxed and maximally contracted rabbit portal vein smooth muscle. J Physiol 357: 185–201, 1984. doi: 10.1113/jphysiol.1984.sp015496. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 437. Sommerville LE, Hartshorne DJ. Intracellular calcium and smooth muscle contraction. Cell Calcium 7: 353–364, 1986. doi: 10.1016/0143-4160(86)90038-2. [DOI] [PubMed] [Google Scholar]
- 438. Biwer LA, Good ME, Hong K, Patel RK, Agrawal N, Looft-Wilson R, Sonkusare SK, Isakson BE. Non-endoplasmic reticulum-based calr (Calreticulin) can coordinate heterocellular calcium signaling and vascular function. Arterioscler Thromb Vasc Biol 38: 120–130, 2018. doi: 10.1161/ATVBAHA.117.309886. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 439. Koh DS, Geiger JR, Jonas P, Sakmann B. Ca2+-permeable AMPA and NMDA receptor channels in basket cells of rat hippocampal dentate gyrus. J Physiol 485: 383–402, 1995. doi: 10.1113/jphysiol.1995.sp020737. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 440. Helmchen F, Borst JG, Sakmann B. Calcium dynamics associated with a single action potential in a CNS presynaptic terminal. Biophys J 72: 1458–1471, 1997. doi: 10.1016/S0006-3495(97)78792-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 441. Helmchen F, Imoto K, Sakmann B. Ca2+ buffering and action potential-evoked Ca2+ signaling in dendrites of pyramidal neurons. Biophys J 70: 1069–1081, 1996. doi: 10.1016/S0006-3495(96)79653-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 442. Lee SH, Rosenmund C, Schwaller B, Neher E. Differences in Ca2+ buffering properties between excitatory and inhibitory hippocampal neurons from the rat. J Physiol 525: 405–418, 2000. doi: 10.1111/j.1469-7793.2000.t01-3-00405.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 443. Fierro L, Llano I. High endogenous calcium buffering in Purkinje cells from rat cerebellar slices. J Physiol 496: 617–625, 1996. doi: 10.1113/jphysiol.1996.sp021713. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 444. Müller A, Kukley M, Stausberg P, Beck H, Müller W, Dietrich D. Endogenous Ca2+ buffer concentration and Ca2+ microdomains in hippocampal neurons. J Neurosci 25: 558–565, 2005. doi: 10.1523/JNEUROSCI.3799-04.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 445. Babai N, Kochubey O, Keller D, Schneggenburger R. An alien divalent ion reveals a major role for Ca2+ buffering in controlling slow transmitter release. J Neurosci 34: 12622–12635, 2014. doi: 10.1523/JNEUROSCI.1990-14.2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 446. Zhou Z, Neher E. Mobile and immobile calcium buffers in bovine adrenal chromaffin cells. J Physiol 469: 245–273, 1993. doi: 10.1113/jphysiol.1993.sp019813. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 447. McMahon SM, Jackson MB. In situ Ca2+ titration in the fluorometric study of intracellular Ca2+ binding. Cell Calcium 56: 504–512, 2014. doi: 10.1016/j.ceca.2014.10.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 448. Jackson MB, Redman SJ. Calcium dynamics, buffering, and buffer saturation in the boutons of dentate granule-cell axons in the hilus. J Neurosci 23: 1612–1621, 2003. doi: 10.1523/JNEUROSCI.23-05-01612.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 449. Xu T, Naraghi M, Kang H, Neher E. Kinetic studies of Ca2+ binding and Ca2+ clearance in the cytosol of adrenal chromaffin cells. Biophys J 73: 532–545, 1997. doi: 10.1016/S0006-3495(97)78091-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 450. Meinrenken CJ, Borst JG, Sakmann B. Calcium secretion coupling at calyx of Held governed by nonuniform channel-vesicle topography. J Neurosci 22: 1648–1667, 2002. doi: 10.1523/JNEUROSCI.22-05-01648.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 451. Timofeeva Y, Volynski K. Calmodulin as a major calcium buffer shaping vesicular release and short-term synaptic plasticity: facilitation through buffer dislocation. Front Cell Neurosci 9: 239, 2015. doi: 10.3389/fncel.2015.00239. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 452. Schmidt H, Schwaller B, Eilers J. Calbindin D28k targets myo-inositol monophosphatase in spines and dendrites of cerebellar Purkinje neurons. Proc Natl Acad Sci USA 102: 5850–5855, 2005. doi: 10.1073/pnas.0407855102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 453. Christel CJ, Schaer R, Wang S, Henzi T, Kreiner L, Grabs D, Schwaller B, Lee A. Calretinin regulates Ca2+-dependent inactivation and facilitation of Cav2.1 Ca2+ channels through a direct interaction with the alpha12.1 subunit. J Biol Chem 287: 39766–39775, 2012. doi: 10.1074/jbc.M112.406363. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 454. Markram H, Toledo-Rodriguez M, Wang Y, Gupta A, Silberberg G, Wu C. Interneurons of the neocortical inhibitory system. Nat Rev Neurosci 5: 793–807, 2004. doi: 10.1038/nrn1519. [DOI] [PubMed] [Google Scholar]
- 455. Bastianelli E. Distribution of calcium-binding proteins in the cerebellum. Cerebellum 2: 242–262, 2003. doi: 10.1080/14734220310022289. [DOI] [PubMed] [Google Scholar]
- 456. Tremblay R, Lee S, Rudy B. GABAergic Interneurons in the neocortex: from cellular properties to circuits. Neuron 91: 260–292, 2016. doi: 10.1016/j.neuron.2016.06.033. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 457. Hof PR, Glezer II, Condé F, Flagg RA, Rubin MB, Nimchinsky EA, Vogt Weisenhorn DM. Cellular distribution of the calcium-binding proteins parvalbumin, calbindin, and calretinin in the neocortex of mammals: phylogenetic and developmental patterns. J Chem Neuroanat 16: 77–116, 1999. doi: 10.1016/s0891-0618(98)00065-9. [DOI] [PubMed] [Google Scholar]
- 458. Celio MR. Calbindin D-28k and parvalbumin in the rat nervous system. Neuroscience 35: 375–475, 1990. doi: 10.1016/0306-4522(90)90091-h. [DOI] [PubMed] [Google Scholar]
- 459. Girard F, Venail J, Schwaller B, Celio MR. The EF-hand Ca2+-binding protein super-family: a genome-wide analysis of gene expression patterns in the adult mouse brain. Neuroscience 294: 116–155, 2015. doi: 10.1016/j.neuroscience.2015.02.018. [DOI] [PubMed] [Google Scholar]
- 460. Freund TF, Buzsáki G. Interneurons of the hippocampus. Hippocampus 6: 347–470, 1996. doi:. [DOI] [PubMed] [Google Scholar]
- 461. Lohmann C, Friauf E. Distribution of the calcium-binding proteins parvalbumin and calretinin in the auditory brainstem of adult and developing rats. J Comp Neurol 367: 90–109, 1996. doi:. [DOI] [PubMed] [Google Scholar]
- 462. Csillik B, Mihály A, Krisztin-Péva B, Chadaide Z, Samsam M, Knyihár-Csillik E, Fenyo R. GABAergic parvalbumin-immunoreactive large calyciform presynaptic complexes in the reticular nucleus of the rat thalamus. J Chem Neuroanat 30: 17–26, 2005. doi: 10.1016/j.jchemneu.2005.03.010. [DOI] [PubMed] [Google Scholar]
- 463. Felmy F, Schneggenburger R. Developmental expression of the Ca2+-binding proteins calretinin and parvalbumin at the calyx of Held of rats and mice. Eur J Neurosci 20: 1473–1482, 2004. doi: 10.1111/j.1460-9568.2004.03604.x. [DOI] [PubMed] [Google Scholar]
- 464. Zhang C, Wang M, Lin S, Xie R. Calretinin-expressing synapses show improved synaptic efficacy with reduced asynchronous release during high-rate activity. J Neurosci 42: 2729–2742, 2022. doi: 10.1523/JNEUROSCI.1773-21.2022. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 465. Gulyás AI, Freund TF. Pyramidal cell dendrites are the primary targets of calbindin D28k-immunoreactive interneurons in the hippocampus. Hippocampus 6: 525–534, 1996. doi:. [DOI] [PubMed] [Google Scholar]
- 466. Baimbridge KG, Miller JJ. Immunohistochemical localization of calcium-binding protein in the cerebellum, hippocampal formation and olfactory bulb of the rat. Brain Res 245: 223–229, 1982. doi: 10.1016/0006-8993(82)90804-6. [DOI] [PubMed] [Google Scholar]
- 467. Miettinen R, Gulyás AI, Baimbridge KG, Jacobowitz DM, Freund TF. Calretinin is present in non-pyramidal cells of the rat hippocampus–II. Co-existence with other calcium binding proteins and GABA. Neuroscience 48: 29–43, 1992. doi: 10.1016/0306-4522(92)90335-y. [DOI] [PubMed] [Google Scholar]
- 468. Blasco-Ibáñez JM, Freund TF. Distribution, ultrastructure, and connectivity of calretinin-immunoreactive mossy cells of the mouse dentate gyrus. Hippocampus 7: 307–320, 1997. doi:. [DOI] [PubMed] [Google Scholar]
- 469. Schwaller B, Brückner G, Celio MR, Härtig W. A polyclonal goat antiserum against the calcium-binding protein calretinin is a versatile tool for various immunochemical techniques. J Neurosci Methods 92: 137–144, 1999. doi: 10.1016/s0165-0270(99)00106-5. [DOI] [PubMed] [Google Scholar]
- 470. Sanna PP, Keyser KT, Battenberg E, Bloom FE. Parvalbumin immunoreactivity in the rat retina. Neurosci Lett 118: 136–139, 1990. doi: 10.1016/0304-3940(90)90267-d. [DOI] [PubMed] [Google Scholar]
- 471. Sanna PP, Keyser KT, Celio MR, Karten HJ, Bloom FE. Distribution of parvalbumin immunoreactivity in the vertebrate retina. Brain Res 600: 141–150, 1993. doi: 10.1016/0006-8993(93)90412-g. [DOI] [PubMed] [Google Scholar]
- 472. Pasteels B, Rogers J, Blachier F, Pochet R. Calbindin and calretinin localization in retina from different species. Vis Neurosci 5: 1–16, 1990. doi: 10.1017/s0952523800000031. [DOI] [PubMed] [Google Scholar]
- 473. Yang D, Thalmann I, Thalmann R, Simmons DD. Expression of alpha and beta parvalbumin is differentially regulated in the rat organ of Corti during development. J Neurobiol 58: 479–492, 2004. doi: 10.1002/neu.10289. [DOI] [PubMed] [Google Scholar]
- 474. Hackney CM, Mahendrasingam S, Penn A, Fettiplace R. The concentrations of calcium buffering proteins in mammalian cochlear hair cells. J Neurosci 25: 7867–7875, 2005. doi: 10.1523/JNEUROSCI.1196-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 475. Boyden ES, Zhang F, Bamberg E, Nagel G, Deisseroth K. Millisecond-timescale, genetically targeted optical control of neural activity. Nat Neurosci 8: 1263–1268, 2005. doi: 10.1038/nn1525. [DOI] [PubMed] [Google Scholar]
- 476. Madisen L, Mao T, Koch H, Zhuo JM, Berenyi A, Fujisawa S, Hsu YW, Garcia AJ 3rd, Gu X, Zanella S, Kidney J, Gu H, Mao Y, Hooks BM, Boyden ES, Buzsáki G, Ramirez JM, Jones AR, Svoboda K, Han X, Turner EE, Zeng H. A toolbox of Cre-dependent optogenetic transgenic mice for light-induced activation and silencing. Nat Neurosci 15: 793–802, 2012. doi: 10.1038/nn.3078. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 477. Kosaka T, Kosaka K, Nakayama T, Hunziker W, Heizmann CW. Axons and axon terminals of cerebellar Purkinje cells and basket cells have higher levels of parvalbumin immunoreactivity than somata and dendrites: quantitative analysis by immunogold labeling. Exp Brain Res 93: 483–491, 1993. doi: 10.1007/BF00229363. [DOI] [PubMed] [Google Scholar]
- 478. Edmonds B, Reyes R, Schwaller B, Roberts WM. Calretinin modifies presynaptic calcium signaling in frog saccular hair cells. Nat Neurosci 3: 786–790, 2000. doi: 10.1038/77687. [DOI] [PubMed] [Google Scholar]
- 479. Müller M, Felmy F, Schwaller B, Schneggenburger R. Parvalbumin is a mobile presynaptic Ca2+ buffer in the calyx of Held that accelerates the decay of Ca2+ and short-term facilitation. J Neurosci 27: 2261–2271, 2007. doi: 10.1523/JNEUROSCI.5582-06.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 480. Marty A, Neher E. Potassium channels in cultured bovine adrenal chromaffin cells. J Physiol 367: 117–141, 1985. doi: 10.1113/jphysiol.1985.sp015817. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 481. Borst JG, Sakmann B. Calcium influx and transmitter release in a fast CNS synapse. Nature 383: 431–434, 1996. doi: 10.1038/383431a0. [DOI] [PubMed] [Google Scholar]
- 482. Ohana O, Sakmann B. Transmitter release modulation in nerve terminals of rat neocortical pyramidal cells by intracellular calcium buffers. J Physiol 513: 135–148, 1998. doi: 10.1111/j.1469-7793.1998.135by.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 483. Fedchyshyn MJ, Wang LY. Developmental transformation of the release modality at the calyx of Held synapse. J Neurosci 25: 4131–4140, 2005. doi: 10.1523/JNEUROSCI.0350-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 484. Moser T, Beutner D. Kinetics of exocytosis and endocytosis at the cochlear inner hair cell afferent synapse of the mouse. Proc Natl Acad Sci USA 97: 883–888, 2000. doi: 10.1073/pnas.97.2.883. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 485. Bucurenciu I, Kulik A, Schwaller B, Frotscher M, Jonas P. Nanodomain coupling between Ca2+ channels and Ca2+ sensors promotes fast and efficient transmitter release at a cortical GABAergic synapse. Neuron 57: 536–545, 2008. doi: 10.1016/j.neuron.2007.12.026. [DOI] [PubMed] [Google Scholar]
- 486. Schmidt H, Brachtendorf S, Arendt O, Hallermann S, Ishiyama S, Bornschein G, Gall D, Schiffmann SN, Heckmann M, Eilers J. Nanodomain coupling at an excitatory cortical synapse. Curr Biol 23: 244–249, 2013. doi: 10.1016/j.cub.2012.12.007. [DOI] [PubMed] [Google Scholar]
- 487. Singer JH, Diamond JS. Sustained Ca2+ entry elicits transient postsynaptic currents at a retinal ribbon synapse. J Neurosci 23: 10923–10933, 2003. doi: 10.1523/JNEUROSCI.23-34-10923.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 488. Adler EM, Augustine GJ, Duffy SN, Charlton MP. Alien intracellular calcium chelators attenuate neurotransmitter release at the squid giant synapse. J Neurosci 11: 1496–1507, 1991. doi: 10.1523/JNEUROSCI.11-06-01496.1991. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 489. Hosoi N, Sakaba T, Neher E. Quantitative analysis of calcium-dependent vesicle recruitment and its functional role at the calyx of Held synapse. J Neurosci 27: 14286–14298, 2007. doi: 10.1523/JNEUROSCI.4122-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 490. Dittman JS, Regehr WG. Calcium dependence and recovery kinetics of presynaptic depression at the climbing fiber to Purkinje cell synapse. J Neurosci 18: 6147–6162, 1998. doi: 10.1523/JNEUROSCI.18-16-06147.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 491. Wang LY, Kaczmarek LK. High-frequency firing helps replenish the readily releasable pool of synaptic vesicles. Nature 394: 384–388, 1998. doi: 10.1038/28645. [DOI] [PubMed] [Google Scholar]
- 492. Lin KH, Taschenberger H, Neher E. A sequential two-step priming scheme reproduces diversity in synaptic strength and short-term plasticity. Proc Natl Acad Sci USA 119: e2207987119, 2022. doi: 10.1073/pnas.2207987119. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 493. Felmy F, Neher E, Schneggenburger R. Probing the intracellular calcium sensitivity of transmitter release during synaptic facilitation. Neuron 37: 801–811, 2003. doi: 10.1016/s0896-6273(03)00085-0. [DOI] [PubMed] [Google Scholar]
- 494. Vyleta NP, Jonas P. Loose coupling between Ca2+ channels and release sensors at a plastic hippocampal synapse. Science 343: 665–670, 2014. doi: 10.1126/science.1244811. [DOI] [PubMed] [Google Scholar]
- 495. Caillard O, Moreno H, Schwaller B, Llano I, Celio MR, Marty A. Role of the calcium-binding protein parvalbumin in short-term synaptic plasticity. Proc Natl Acad Sci USA 97: 13372–13377, 2000. doi: 10.1073/pnas.230362997. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 496. Bornschein G, Arendt O, Hallermann S, Brachtendorf S, Eilers J, Schmidt H. Paired-pulse facilitation at recurrent Purkinje neuron synapses is independent of calbindin and parvalbumin during high-frequency activation. J Physiol 591: 3355–3370, 2013. doi: 10.1113/jphysiol.2013.254128. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 497. Goda Y, Stevens CF. Two components of transmitter release at a central synapse. Proc Natl Acad Sci USA 91: 12942–12946, 1994. doi: 10.1073/pnas.91.26.12942. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 498. Rozov A, Bolshakov AP, Valiullina-Rakhmatullina F. The ever-growing puzzle of asynchronous release. Front Cell Neurosci 13: 28, 2019. doi: 10.3389/fncel.2019.00028. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 499. Chamberland S, Timofeeva Y, Evstratova A, Norman CA, Volynski K, Tóth K. Slow-decaying presynaptic calcium dynamics gate long-lasting asynchronous release at the hippocampal mossy fiber to CA3 pyramidal cell synapse. Synapse 74: e22178, 2020. doi: 10.1002/syn.22178. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 500. Hefft S, Jonas P. Asynchronous GABA release generates long-lasting inhibition at a hippocampal interneuron-principal neuron synapse. Nat Neurosci 8: 1319–1328, 2005. doi: 10.1038/nn1542. [DOI] [PubMed] [Google Scholar]
- 501. Müller A, Kukley M, Uebachs M, Beck H, Dietrich D. Nanodomains of single Ca2+ channels contribute to action potential repolarization in cortical neurons. J Neurosci 27: 483–495, 2007. doi: 10.1523/JNEUROSCI.3816-06.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 502. Gall D, Roussel C, Susa I, D’Angelo E, Rossi P, Bearzatto B, Galas MC, Blum D, Schurmans S, Schiffmann SN. Altered neuronal excitability in cerebellar granule cells of mice lacking calretinin. J Neurosci 23: 9320–9327, 2003. doi: 10.1523/JNEUROSCI.23-28-09320.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 503. Bischop DP, Orduz D, Lambot L, Schiffmann SN, Gall D. Control of neuronal excitability by calcium binding proteins: a new mathematical model for striatal fast-spiking interneurons. Front Mol Neurosci 5: 78, 2012. doi: 10.3389/fnmol.2012.00078. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 504. Albéri L, Lintas A, Kretz R, Schwaller B, Villa AEP. The calcium-binding protein parvalbumin modulates the firing 1 properties of the reticular thalamic nucleus bursting neurons. J Neurophysiol 109: 2827–2841, 2013. doi: 10.1152/jn.00375.2012. [DOI] [PubMed] [Google Scholar]
- 505. Schiffmann SN, Cheron G, Lohof A, d’Alcantara P, Meyer M, Parmentier M, Schurmans S. Impaired motor coordination and Purkinje cell excitability in mice lacking calretinin. Proc Natl Acad Sci USA 96: 5257–5262, 1999. doi: 10.1073/pnas.96.9.5257. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 506. Welsh JP, Lang EJ, Suglhara I, Llinás R. Dynamic organization of motor control within the olivocerebellar system. Nature 374: 453–457, 1995. doi: 10.1038/374453a0. [DOI] [PubMed] [Google Scholar]
- 507. Cheron G, Gall D, Servais L, Dan B, Maex R, Schiffmann SN. Inactivation of calcium-binding protein genes induces 160 Hz oscillations in the cerebellar cortex of alert mice. J Neurosci 24: 434–441, 2004. doi: 10.1523/JNEUROSCI.3197-03.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 508. Markram H, Helm PJ, Sakmann B. Dendritic calcium transients evoked by single back-propagating action potentials in rat neocortical pyramidal neurons. J Physiol 485: 1–20, 1995. doi: 10.1113/jphysiol.1995.sp020708. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 509. Yuste R, Denk W. Dendritic spines as basic functional units of neuronal integration. Nature 375: 682–684, 1995. doi: 10.1038/375682a0. [DOI] [PubMed] [Google Scholar]
- 510. Schmidt H, Stiefel KM, Racay P, Schwaller B, Eilers J. Mutational analysis of dendritic Ca2+ kinetics in rodent Purkinje cells: role of parvalbumin and calbindin D28k. J Physiol 551: 13–32, 2003. doi: 10.1113/jphysiol.2002.035824. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 511. Kaiser KM, Zilberter Y, Sakmann B. Back-propagating action potentials mediate calcium signalling in dendrites of bitufted interneurons in layer 2/3 of rat somatosensory cortex. J Physiol 535: 17–31, 2001. doi: 10.1111/j.1469-7793.2001.t01-1-00017.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 512. Aponte Y, Bischofberger J, Jonas P. Efficient Ca2+ buffering in fast-spiking basket cells of rat hippocampus. J Physiol 586: 2061–2075, 2008. doi: 10.1113/jphysiol.2007.147298. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 513. Hage TA, Khaliq ZM. Tonic firing rate controls dendritic Ca2+ signaling and synaptic gain in substantia nigra dopamine neurons. J Neurosci 35: 5823–5836, 2015. doi: 10.1523/JNEUROSCI.3904-14.2015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 514. Eilers J, Callewaert G, Armstrong C, Konnerth A. Calcium signaling in a narrow somatic submembrane shell during synaptic activity in cerebellar Purkinje neurons. Proc Natl Acad Sci USA 92: 10272–10276, 1995. doi: 10.1073/pnas.92.22.10272. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 515. Airaksinen MS, Eilers J, Garaschuk O, Thoenen H, Konnerth A, Meyer M. Ataxia and altered dendritic calcium signaling in mice carrying a targeted null mutation of the calbindin D28k gene. Proc Natl Acad Sci USA 94: 1488–1493, 1997. doi: 10.1073/pnas.94.4.1488. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 516. Barski JJ, Hartmann J, Rose CR, Hoebeek F, Mörl K, Noll-Hussong M, De Zeeuw CI, Konnerth A, Meyer M. Calbindin in cerebellar Purkinje cells is a critical determinant of the precision of motor coordination. J Neurosci 23: 3469–3477, 2003. doi: 10.1523/JNEUROSCI.23-08-03469.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 517. Baimbridge KG, Celio MR, Rogers JH. Calcium-binding proteins in the nervous system. Trends Neurosci 15: 303–308, 1992. doi: 10.1016/0166-2236(92)90081-i. [DOI] [PubMed] [Google Scholar]
- 518. Alexianu ME, Ho BK, Mohamed AH, La Bella V, Smith RG, Appel SH. The role of calcium-binding proteins in selective motoneuron vulnerability in amyotrophic lateral sclerosis. Ann Neurol 36: 846–858, 1994. doi: 10.1002/ana.410360608. [DOI] [PubMed] [Google Scholar]
- 519. Seto-Ohshima A, Emson PC, Lawson E, Mountjoy CQ, Carrasco LH. Loss of matrix calcium-binding protein-containing neurons in Huntington’s disease. Lancet 1: 1252–1255, 1988. doi: 10.1016/s0140-6736(88)92073-9. [DOI] [PubMed] [Google Scholar]
- 520. Schwaller B, Meyer M, Schiffmann S. ‘New’ functions for ‘old’ proteins: the role of the calcium-binding proteins calbindin D-28k, calretinin and parvalbumin, in cerebellar physiology. Studies with knockout mice. Cerebellum 1: 241–258, 2002. doi: 10.1080/147342202320883551. [DOI] [PubMed] [Google Scholar]
- 521. Cheron G, Schurmans S, Lohof A, d’Alcantara P, Meyer M, Draye JP, Parmentier M, Schiffmann SN. Electrophysiological behavior of Purkinje cells and motor coordination in calretinin knock-out mice. Prog Brain Res 124: 299–308, 2000. doi: 10.1016/S0079-6123(00)24024-7. [DOI] [PubMed] [Google Scholar]
- 522. Schurmans S, Schiffmann SN, Gurden H, Lemaire M, Lipp HP, Schwam V, Pochet R, Imperato A, Böhme GA, Parmentier M. Impaired long-term potentiation induction in dentate gyrus of calretinin-deficient mice. Proc Natl Acad Sci USA 94: 10415–10420, 1997. doi: 10.1073/pnas.94.19.10415. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 523. Wöhr M, Orduz D, Gregory P, Moreno H, Khan U, Vörckel KJ, Wolfer DP, Welzl H, Gall D, Schiffmann SN, Schwaller B. Lack of parvalbumin in mice leads to behavioral deficits relevant to all human autism core symptoms and related neural morphofunctional abnormalities. Transl Psychiatry 5: e525, 2015. doi: 10.1038/tp.2015.19. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 524. Heizmann CW, Braun K. Changes in Ca2+-binding proteins in human neurodegenerative disorders. Trends Neurosci 15: 259–264, 1992. doi: 10.1016/0166-2236(92)90067-i. [DOI] [PubMed] [Google Scholar]
- 525. Marín O. Interneuron dysfunction in psychiatric disorders. Nat Rev Neurosci 13: 107–120, 2012. doi: 10.1038/nrn3155. [DOI] [PubMed] [Google Scholar]
- 526. McLachlan DR, Wong L, Bergeron C, Baimbridge KG. Calmodulin and calbindin D28K in Alzheimer disease. Alzheimer Dis Assoc Disord 1: 171–179, 1987. doi: 10.1097/00002093-198701030-00009. [DOI] [PubMed] [Google Scholar]
- 527. Ichimiya Y, Emson PC, Mountjoy CQ, Lawson DE, Heizmann CW. Loss of calbindin-28K immunoreactive neurones from the cortex in Alzheimer-type dementia. Brain Res 475: 156–159, 1988. doi: 10.1016/0006-8993(88)90210-7. [DOI] [PubMed] [Google Scholar]
- 528. Solodkin A, Veldhuizen SD, Van Hoesen GW. Contingent vulnerability of entorhinal parvalbumin-containing neurons in Alzheimer’s disease. J Neurosci 16: 3311–3321, 1996. doi: 10.1523/JNEUROSCI.16-10-03311.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 529. Sanchez-Mejias E, Nuñez-Diaz C, Sanchez-Varo R, Gomez-Arboledas A, Garcia-Leon JA, Fernandez-Valenzuela JJ, Mejias-Ortega M, Trujillo-Estrada L, Baglietto-Vargas D, Moreno-Gonzalez I, Davila JC, Vitorica J, Gutierrez A. Distinct disease-sensitive GABAergic neurons in the perirhinal cortex of Alzheimer’s mice and patients. Brain Pathol 30: 345–363, 2020. doi: 10.1111/bpa.12785. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 530. Fonseca M, Soriano E, Ferrer I, Martinez A, Tuñon T. Chandelier cell axons identified by parvalbumin-immunoreactivity in the normal human temporal cortex and in Alzheimer’s disease. Neuroscience 55: 1107–1116, 1993. doi: 10.1016/0306-4522(93)90324-9. [DOI] [PubMed] [Google Scholar]
- 531. Mikhaylova M, Bär J, van Bommel B, Schätzle P, YuanXiang P, Raman R, Hradsky J, Konietzny A, Loktionov EY, Reddy PP, Lopez-Rojas J, Spilker C, Kobler O, Raza SA, Stork O, Hoogenraad CC, Kreutz MR. Caldendrin directly couples postsynaptic calcium signals to actin remodeling in dendritic spines. Neuron 97: 1110–1125.e4, 2018. doi: 10.1016/j.neuron.2018.01.046. [DOI] [PubMed] [Google Scholar]
- 532. Bernstein HG, Sahin J, Smalla KH, Gundelfinger ED, Bogerts B, Kreutz MR. A reduced number of cortical neurons show increased Caldendrin protein levels in chronic schizophrenia. Schizophr Res 96: 246–256, 2007. doi: 10.1016/j.schres.2007.05.038. [DOI] [PubMed] [Google Scholar]
- 533. Hashimoto T, Volk DW, Eggan SM, Mirnics K, Pierri JN, Sun Z, Sampson AR, Lewis DA. Gene expression deficits in a subclass of GABA neurons in the prefrontal cortex of subjects with schizophrenia. J Neurosci 23: 6315–6326, 2003. doi: 10.1523/JNEUROSCI.23-15-06315.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 534. Hoenderop JG, Nilius B, Bindels RJ. Calcium absorption across epithelia. Physiol Rev 85: 373–422, 2005. doi: 10.1152/physrev.00003.2004. [DOI] [PubMed] [Google Scholar]
- 535. Moor MB, Bonny O. Ways of calcium reabsorption in the kidney. Am J Physiol Renal Physiol 310: F1337–F1350, 2016. doi: 10.1152/ajprenal.00273.2015. [DOI] [PubMed] [Google Scholar]
- 536. Taylor AN, Wasserman RH. Correlations between the vitamin D-induced calcium binding protein and intestinal absorption of calcium. Fed Proc 28: 1834–1838, 1969. [PubMed] [Google Scholar]
- 537. Pansu D, Bellaton C, Roche C, Bronner F. Duodenal and ileal calcium absorption in the rat and effects of vitamin D. Am J Physiol Gastrointest Liver Physiol 244: G695–G700, 1983. doi: 10.1152/ajpgi.1983.244.6.G695. [DOI] [PubMed] [Google Scholar]
- 538. Bronner F, Pansu D, Stein WD. An analysis of intestinal calcium transport across the rat intestine. Am J Physiol Gastrointest Liver Physiol 250: G561–G569, 1986. doi: 10.1152/ajpgi.1986.250.5.G561. [DOI] [PubMed] [Google Scholar]
- 539. Feher JJ, Wasserman RH. Calcium absorption and intestinal calcium-binding protein: quantitative relationship. Am J Physiol Endocrinol Metab 236: E556–E561, 1979. doi: 10.1152/ajpendo.1979.236.5.E556. [DOI] [PubMed] [Google Scholar]
- 540. Gross M, Kumar R. Physiology and biochemistry of vitamin D-dependent calcium binding proteins. Am J Physiol Renal Physiol 259: F195–F209, 1990. doi: 10.1152/ajprenal.1990.259.2.F195. [DOI] [PubMed] [Google Scholar]
- 541. Bronner F, Stein WD. CaBPr facilitates intracellular diffusion for Ca pumping in distal convoluted tubule. Am J Physiol Renal Physiol 255: F558–F562, 1988. doi: 10.1152/ajprenal.1988.255.3.F558. [DOI] [PubMed] [Google Scholar]
- 542. Adedeji AO, Gu YZ, Pourmohamad T, Kanerva J, Chen Y, Atabakhsh E, Tackett MR, Chen F, Bhatt B, Gury T, Dorchies O, Sonee M, Morgan M, Burkey J, Gautier JC, McDuffie JE. The utility of novel urinary biomarkers in mice for drug development studies. Int J Toxicol 40: 15–25, 2021. doi: 10.1177/1091581820970498. [DOI] [PubMed] [Google Scholar]
- 543. Bindels RJ, Timmermans JA, Hartog A, Coers W, van Os CH. Calbindin-D9k and parvalbumin are exclusively located along basolateral membranes in rat distal nephron. J Am Soc Nephrol 2: 1122–1129, 1991. doi: 10.1681/ASN.V261122. [DOI] [PubMed] [Google Scholar]
- 544. Belge H, Gailly P, Schwaller B, Loffing J, Debaix H, Riveira-Munoz E, Beauwens R, Devogelaer JP, Hoenderop JG, Bindels RJ, Devuyst O. Renal expression of parvalbumin is critical for NaCl handling and response to diuretics. Proc Natl Acad Sci USA 104: 14849–14854, 2007. doi: 10.1073/pnas.0702810104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 545. Loffing J, Loffing-Cueni D, Valderrabano V, Kläusli L, Hebert SC, Rossier BC, Hoenderop JG, Bindels RJ, Kaissling B. Distribution of transcellular calcium and sodium transport pathways along mouse distal nephron. Am J Physiol Renal Physiol 281: F1021–F1027, 2001. doi: 10.1152/ajprenal.0085.2001. [DOI] [PubMed] [Google Scholar]
- 546. Lambers TT, Mahieu F, Oancea E, Hoofd L, de Lange F, Mensenkamp AR, Voets T, Nilius B, Clapham DE, Hoenderop JG, Bindels RJ. Calbindin-D28K dynamically controls TRPV5-mediated Ca2+ transport. EMBO J 25: 2978–2988, 2006. doi: 10.1038/sj.emboj.7601186. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 547. Voets T, Nilius B, Hoefs S, van der Kemp AW, Droogmans G, Bindels RJ, Hoenderop JG. TRPM6 forms the Mg2+ influx channel involved in intestinal and renal Mg2+ absorption. J Biol Chem 279: 19–25, 2004. doi: 10.1074/jbc.M311201200. [DOI] [PubMed] [Google Scholar]
- 548. Olinger E, Schwaller B, Loffing J, Gailly P, Devuyst O. Parvalbumin: calcium and magnesium buffering in the distal nephron. Nephrol Dial Transplant 27: 3988–3994, 2012. doi: 10.1093/ndt/gfs457. [DOI] [PubMed] [Google Scholar]
- 549. Lee CT, Ng HY, Lee YT, Lai LW, Lien YH. The role of calbindin-D28k on renal calcium and magnesium handling during treatment with loop and thiazide diuretics. Am J Physiol Renal Physiol 310: F230–F236, 2016. doi: 10.1152/ajprenal.00057.2015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 550. von Tscharner V, Deranleau DA, Baggiolini M. Calcium fluxes and calcium buffering in human neutrophils. J Biol Chem 261: 10163–10168, 1986. doi: 10.1016/S0021-9258(18)67505-2. [DOI] [PubMed] [Google Scholar]
- 551. Hessian PA, Edgeworth J, Hogg N. MRP-8 and MRP-14, two abundant Ca2+-binding proteins of neutrophils and monocytes. J Leukoc Biol 53: 197–204, 1993. [PubMed] [Google Scholar]
- 552. Leukert N, Vogl T, Strupat K, Reichelt R, Sorg C, Roth J. Calcium-dependent tetramer formation of S100A8 and S100A9 is essential for biological activity. J Mol Biol 359: 961–972, 2006. doi: 10.1016/j.jmb.2006.04.009. [DOI] [PubMed] [Google Scholar]
- 553. Kehl-Fie TE, Chitayat S, Hood MI, Damo S, Restrepo N, Garcia C, Munro KA, Chazin WJ, Skaar EP. Nutrient metal sequestration by calprotectin inhibits bacterial superoxide defense, enhancing neutrophil killing of Staphylococcus aureus. Cell Host Microbe 10: 158–164, 2011. doi: 10.1016/j.chom.2011.07.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 554. Damo SM, Kehl-Fie TE, Sugitani N, Holt ME, Rathi S, Murphy WJ, Zhang Y, Betz C, Hench L, Fritz G, Skaar EP, Chazin WJ. Molecular basis for manganese sequestration by calprotectin and roles in the innate immune response to invading bacterial pathogens. Proc Natl Acad Sci USA 110: 3841–3846, 2013. doi: 10.1073/pnas.1220341110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 555. Kozlyuk N, Monteith AJ, Garcia V, Damo SM, Skaar EP, Chazin WJ. S100 proteins in the innate immune response to pathogens. Methods Mol Biol 1929: 275–290, 2019. doi: 10.1007/978-1-4939-9030-6_18. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 556. Kehl-Fie TE, Skaar EP. Nutritional immunity beyond iron: a role for manganese and zinc. Curr Opin Chem Biol 14: 218–224, 2010. doi: 10.1016/j.cbpa.2009.11.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 557. Brophy MB, Nolan EM. Manganese and microbial pathogenesis: sequestration by the Mammalian immune system and utilization by microorganisms. ACS Chem Biol 10: 641–651, 2015. doi: 10.1021/cb500792b. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 558. Hayden JA, Brophy MB, Cunden LS, Nolan EM. High-affinity manganese coordination by human calprotectin is calcium-dependent and requires the histidine-rich site formed at the dimer interface. J Am Chem Soc 135: 775–787, 2013. doi: 10.1021/ja3096416. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 559. Yin Y, Henzl MT, Lorber B, Nakazawa T, Thomas TT, Jiang F, Langer R, Benowitz LI. Oncomodulin is a macrophage-derived signal for axon regeneration in retinal ganglion cells. Nat Neurosci 9: 843–852, 2006. doi: 10.1038/nn1701. [DOI] [PubMed] [Google Scholar]
- 560. Kurimoto T, Yin Y, Habboub G, Gilbert HY, Li Y, Nakao S, Hafezi-Moghadam A, Benowitz LI. Neutrophils express oncomodulin and promote optic nerve regeneration. J Neurosci 33: 14816–14824, 2013. doi: 10.1523/JNEUROSCI.5511-12.2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 561. Scherbik SV, Brinton MA. Virus-induced Ca2+ influx extends survival of West Nile virus-infected cells. J Virol 84: 8721–8731, 2010. doi: 10.1128/JVI.00144-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 562. Siddharthan V, Wang H, Davies CJ, Hall JO, Morrey JD. Inhibition of West Nile virus by calbindin-D28k. PLoS One 9: e106535, 2014. doi: 10.1371/journal.pone.0106535. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 563. Pfannkuche H, Konrath A, Buchholz I, Richt JA, Seeger J, Müller H, Gäbel G. Infection of the enteric nervous system by Borna disease virus (BDV) upregulates expression of Calbindin D-28k. Vet Microbiol 127: 275–285, 2008. doi: 10.1016/j.vetmic.2007.09.005. [DOI] [PubMed] [Google Scholar]
- 564. Cannell MB. Effect of tetanus duration on the free calcium during the relaxation of frog skeletal muscle fibres. J Physiol 376: 203–218, 1986. doi: 10.1113/jphysiol.1986.sp016149. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 565. Henzl MT, Tanner JJ. Solution structure of Ca2+-free rat α-parvalbumin. Protein Sci 17: 431–438, 2008. doi: 10.1110/ps.073318308. [DOI] [PMC free article] [PubMed] [Google Scholar]