In healthy muscle, the rapid release of calcium ions (Ca2+) with excitation–contraction (E-C) coupling, results in elevations in Ca2+ concentrations which can exceed 10-fold that of resting values. The sizable transient changes in Ca2+ concentrations are necessary for the activation of signaling pathways, which rely on Ca2+ as a second messenger, including those involved with force generation, fiber type distribution and hypertrophy. However, prolonged elevations in intracellular Ca2+ can result in the unwanted activation of Ca2+ signaling pathways that cause muscle damage, dysfunction, and disease. Muscle employs several calcium handling and calcium transport proteins that function to rapidly return Ca2+ concentrations back to resting levels following contraction. This review will detail our current understanding of calcium handling during the decay phase of intracellular calcium transients in healthy skeletal and cardiac muscle. We will also discuss how impairments in Ca2+ transport can occur and how mishandling of Ca2+ can lead to the pathogenesis and/or progression of skeletal muscle myopathies and cardiomyopathies.

Calcium (Ca2+) is a divalent cation which is indispensably involved with molecular signaling. Innately, Ca2+ has flexibility in its bonding angles and lengths, allowing for an array of potential ligation patterns [1]. The diversity in bond arrangements allows molecules to present binding sites with numerous variations. These slight differences among binding sites allow the kinetics of bond formation to vary among Ca2+-binding molecules [1,2]. Thus, the complexity of Ca2+ signaling becomes apparent as it can be regulated by the location and expression of the Ca2+-binding proteins and the kinetics of the Ca2+-binding site.

Within myofibers, cytosolic Ca2+ concentrations ([Ca2+]cyt) can fluctuate from resting concentrations of 100 nM to values above 1000 nM during tetanus [3]. When [Ca2+]cyt is elevated above resting concentrations, Ca2+ interacts with two categories of Ca2+ binding molecules: buffers and sensors [4]. Ca2+ sensors elicit a downstream signal when binding occurs. The temporal range of Ca2+ signaling can be as brief as seconds but may also exist on the timescale of days [5]. Ca2+ signaling is effectively diminished when [Ca2+]cyt reverts back to resting concentrations. One strategy for increasing the rate of [Ca2+]cyt decay is through Ca2+ buffering. Buffers act to sequester Ca2+ without being directly incorporated into molecular signaling. The presence of Ca2+ buffers allows for the regulation of Ca2+ diffusion, which can indirectly affect molecular signaling pathways [6–8].

Although Ca2+ buffering does have a role in altering intracellular Ca2+ transients (ICT), the decay of [Ca2+]cyt is mediated primarily through the movement of Ca2+ across phospholipid membranes [1]. Within striated muscle, the main strategy for lowering [Ca2+]cyt is through Ca2+ sequestration in the sarcoplasmic reticulum (SR), a membrane bound organelle which surrounds the contractile myofilaments [3]. Ca2+ is transported against a concentration gradient into the SR in an ATP-dependent manner. This occurs through the sarco(endo)plasmic reticulum Ca2+-ATPase (SERCA), which, under ideal conditions, will transport two Ca2+ for every ATP hydrolyzed [9,10]. SERCA not only has a critical role in the regulation of [Ca2+]cyt but also in the regulation of the energy expenditure of the cell [11]. Within cardiac tissue, a smaller, yet notable proportion of the cytosolic Ca2+ is also extruded across the plasma membrane to effectively reduce [Ca2+]cyt during the cardiac cycle [12]. The role of Ca2+ transport is essential for the homeostatic function of cells, especially within excitable, contractile tissue. During dysregulation, inadequate control of [Ca2+]cyt can result in the unwanted activation of proteolytic and apoptotic pathways, leading to muscle damage, dysfunction, and even disease. In this review, we discuss the role of Ca2+ transport in the maintenance of healthy muscle as well as the role it can have in the genesis and exacerbation of pathological states.

As voluntary tissue, force generation by skeletal muscle follows the excitation of the sarcolemmal membrane by an associated motor neuron [13]. Neural signaling results in depolarization of the myofiber plasma membrane [13], which is detected by the voltage sensitive dihydropyridine receptors (DHPR) located within the transverse tubule (T-tubule) membrane [14]. In skeletal muscle, the DHPR is physically linked with the ryanodine receptor (RyR) and when changes in voltage are detected, the DHPR acts to increase the open probability of the RyR leading to increased Ca2+ release [15]. In cardiac muscle, SR Ca2+ release is slightly different such that when Ca2+ enters the cytosol via the DHPR channels, it binds to and opens RyR channels through a mechanism referred to as Ca2+-induced Ca2+ release (CICR) [12].

The RyR is a Ca2+ channel embedded in the terminal cisternae of the SR that functions as the major Ca2+ release channel inside muscle cells and, when activated, increases the [Ca2+]cyt [15]. Within healthy muscle tissue, the increase in [Ca2+]cyt results in the immediate activation of SERCA, which acts to pump Ca2+ back into the SR. However, with the continuance of high frequency neural signaling the rate of Ca2+ outflow is greater than the ability for SERCA to re-sequester Ca2+. Consequently, [Ca2+]cyt rises within the myofiber and binds to troponin C resulting in the movement of tropomyosin and thus uncovering the myosin binding site on the thin filaments [16,17]. With the myosin binding site exposed, crossbridge formation between the thick and thin filaments results in the generation of force [18,19]. Muscle relaxation will not occur until the termination of high frequency neural signaling and inactivation of RYR. Upon the cessation of Ca2+ release, the decay phase of [Ca2+]cyt begins.

In a single-twitch stimulus, a rapid rise in [Ca2+]cyt to a peak is quickly followed by a decay, all within milliseconds [20–22]. The decay phase of [Ca2+]cyt has a negative exponential relationship in which the initial rate of decay is rapid but as [Ca2+]cyt approaches pre-stimulation resting concentrations the rate of calcium uptake is attenuated. Within skeletal muscle tissue, the characteristics of a Ca2+ transient varies across different types of muscle fibers. Baylor and Hollingworth (2003) compared the ICTs of slow fibers from soleus tissue and fast fibers from EDL muscle [23]. They found identical times to peak [Ca2+]cyt during Ca2+ release in the slow and fast fibre types; however, the peak [Ca2+]cyt amplitude was two times greater and the ICT half duration was ∼1.6 times shorter in fast EDL fibers compared with the slow soleus fibers [23]. The rate of Ca2+ sequestration in the final half of the ICT decay phase was also three times greater in fast fibers compared with slow fibers [23]. The fiber type differences in the rate of ICT decay are believed to contribute to different force summation responses at submaximal stimulation frequencies. With evoked contractions at 67 Hz, the ICT amplitudes and force grew only slightly with continued stimulation in fast fibers whereas ICT amplitude and force increased with each subsequent stimulation in slow fibers [23].

The myocardium of the heart contracts in response to neural excitation to pump blood out of the heart and relaxes upon the cessation of neural stimulation to allow the heart to refill with blood. Cardiomyocytes within the myocardium express similar but different isoforms of Ca2+ handling proteins than found in skeletal muscle as discussed below; however, their regulation and contribution to excitation–contraction (E-C) coupling vary slightly (see Figure 1). Like skeletal muscle, SERCA is the dominant Ca2+ transport protein in cardiomyocytes that contributes to the decay of Ca2+ transients during E-C coupling [12,24]. Unlike skeletal muscle, the NCX and the slow Ca2+ sequestering systems, which include mitochondrial Ca2+ uniport (MCU) and sarcolemmal Ca2+ ATPase (PMCA), also contribute to Ca2+ decay in cardiac muscle [12,24]. During cardiomyocyte relaxation in rat ventricle, SERCA accounts for as much as 92% of ICT decay, while that of NCX, and the slow system contribute 7% and 1%, respectively [25]. In contrast, within rabbit ventricle myocytes incubated at 25°C, the contribution from SERCA, NCX, and the slow systems has been reported as 70%, 27%, and 3%, respectively [26]; however, under stimulated conditions (in the presence of isoproterenol), the contribution by SERCA has been found to be increased to 83% in rabbits [27].

Movement of Ca 2+ in skeletal myofibers and cardiomyocytes during E-C coupling

Figure 1
Movement of Ca 2+ in skeletal myofibers and cardiomyocytes during E-C coupling

Movement of Ca2+ in skeletal myofibers (A) and cardiomyocytes (B) during E-C coupling. The relative contribution of Ca2+ transport proteins to Ca2+ ion removal during the decay phase of Ca2+ transients is indicated by different arrows with thicker solid arrows indicating a major contribution, thinner dashed arrows indicating a minor contribution, and no contribution of expressed proteins where arrows are not shown. Refer to the text for further details.

Figure 1
Movement of Ca 2+ in skeletal myofibers and cardiomyocytes during E-C coupling

Movement of Ca2+ in skeletal myofibers (A) and cardiomyocytes (B) during E-C coupling. The relative contribution of Ca2+ transport proteins to Ca2+ ion removal during the decay phase of Ca2+ transients is indicated by different arrows with thicker solid arrows indicating a major contribution, thinner dashed arrows indicating a minor contribution, and no contribution of expressed proteins where arrows are not shown. Refer to the text for further details.

Close modal

Sarco(endo)plasmic reticulum Ca2+-ATPase

SERCA is a 110 kDa P-type ATPase, which is embedded in the membrane of the longitudinal SR. Structurally, SERCA is composed of three cytoplasmic domains: phosphorylation (P-domain), actuator (A-domain) and nucleotide-binding (N-domain); and 10 transmembrane helices (M1–M10) [28]. During the catalytic cycle, SERCA undergoes drastic conformational changes with the binding of cations and ATP by alternating between an E1 and E2 state [29]. Previous genomic and proteomic analyses have shown that there are multiple SERCA isoforms derived from three distinct genes: ATP2A1, ATP2A2, and ATP2A3. From these genes, 13 different mRNA splice variants and 10 protein isoforms have been identified [30–34]. SERCA isoforms are between 75 and 84% homologous [35]. The SERCA isoform expression changes with development and aging and varies between tissues. Within fetal and neonatal rat skeletal muscle, SERCA1b and SERCA2a are expressed in fast fibers while SERCA2a is expressed in slow fibers [35]. In rat heart, SERCA2a and SERCA3 are initially expressed during early developmental stages but as development continues only SERCA2a mRNA remains expressed [35]. In adult rats, SERCA1a becomes most commonly found within fast twitch muscle fibers, and SERCA2a is found most commonly in slow twitch fibers and myocardial tissue [35]. Smooth muscle expresses both SERCA2a and SERCA2b [35,36]. Nonmuscle tissue is known to express SERCA2b and SERCA3a–c [35].

Despite considerable homology among SERCA isoforms, there are drastic differences in the lusitropic measures between these tissue types. Differences do exist in the kinetics of SERCA isoforms [37]; however, the large contrast in relaxation appears to be more so associated with differences in the quantity of SERCA molecules expressed, with fast twitch fibers expressing ∼5-fold greater levels of SERCA compared with slow twitch fibers [11,38]. Another factor affecting ICT decay differences among fiber types is the expression of the SR luminal protein calsequestrin, which is greater in type II muscle fibers [38]. Greater calsequestrin expression results in more Ca2+ buffering capacity within the SR resulting in less back-inhibition while SERCA pumps Ca2+ [38]. Thus, with greater Ca2+ buffering within the SR lumen, SERCA can maintain a higher rate of Ca2+ uptake [38].

The major isoform of SERCA that is expressed in cardiac tissue is SERCA2a with greater expression in the atria compared with ventricles [39], which is associated with faster contractile kinetics of atria compared with ventricles [40]. The ventricles of the heart are also thicker and produce stronger contractions than atria [39] and considering the importance of SR Ca2+ load to contractile force, SERCA2a expression and function is vital in these chambers. Studies in transgenic animal models provide good evidence for this. In animal models with reduced SERCA2a activity, the contractility of the left ventricle (LV) is significantly impaired compared with the wild-type controls [41]. Conversely, in models with increased SERCA2a activity, contractility of the LV is significantly enhanced [42].

Phospholamban

Several proteins have been identified which act to regulate SERCA function either positively or negatively [43–45]. Among these SERCA regulators, phospholamban (PLN) is one of the most studied due to its vital role in cardiac function and disease [46–48]. PLN is expressed in cardiac tissue and in type I fibers of skeletal muscle [49,50], where it predominantly associates with SERCA2a [44]. However, it should be noted that work from our group has identified that in human muscles PLN can be found in type II fibres as well [44]. PLN is a 52-amino acid protein located in the SR membrane where it can interact with SERCA (Figure 2) [51,52]. PLN is composed of a small luminal domain, a transmembranous domain and a cytosolic domain [46,53,54]. The transmembranous domain consists of a single helix, which can directly bind to the Ca2+-binding sites formed by the M2, M4, M6 and M9 helices of SERCA [53,55]. In binding to SERCA, PLN elicits an inhibitory effect on the Ca2+ pump by reducing the apparent affinity of Ca2+ binding [56,57]; however, at maximal [Ca2+]cyt, PLN dissociates from SERCA and does not affect Vmax [58–60], but the mechanism behind this remains to be elucidated.

Regulation of SERCA by PLN and SLN

Figure 2
Regulation of SERCA by PLN and SLN

(A) A cartoon model showing a SERCA molecule with no protein inhibitors bound (left), PLN bound (middle), and SLN bound (right) illustrates the physical and functional interaction between SERCA and its endogenous protein regulators. (B) A graphical summary shows the independent effects of PLN and SLN on SERCA activity. Refer to the text for further details.

Figure 2
Regulation of SERCA by PLN and SLN

(A) A cartoon model showing a SERCA molecule with no protein inhibitors bound (left), PLN bound (middle), and SLN bound (right) illustrates the physical and functional interaction between SERCA and its endogenous protein regulators. (B) A graphical summary shows the independent effects of PLN and SLN on SERCA activity. Refer to the text for further details.

Close modal

The inhibitory action of PLN on SERCA can be disrupted via phosphorylation of two sites within the cytosolic domain of PLN. These phosphorylation sites are targeted by two different kinases: Ser16 by cAMP dependent kinase A, and Thr17 by Ca2+ calmodulin dependent kinase II (CAMKII) [46,61]. Upon phosphorylation of either of these sites, PLN will dissociate from SERCA while remaining in the SR membrane either in its monomeric or homopentameric form [61]. PLN phosphorylation acts as a mechanism for the cell to quickly alter the rate of Ca2+ sequestration. During times where [Ca2+]cyt is elevated, CAMKII will detect these elevations and respond by phosphorylating PLN, thus relieving its inhibitory effect on SERCA and allowing for faster calcium clearance [62]. PLN inhibition is also alleviated by β-adrenergic stimulation through increasing the kinase activity of PKA [46,62–64]. This is especially important in cardiac tissue when the work demand increases such as during exercise. Exercise elicits an increased oxygen demand of the recruited skeletal muscles. In order to ensure adequate oxygen delivery to meet the demands of the recruited muscles, both heart rate and cardiac contractility must increase to elevate cardiac output. With increases in the rate of cardiac cycling, the rate of Ca2+ removal from the cytosol must also increase to ensure adequate relaxation of the myocardium during diastole [27]. In return, this would contribute to increased SR Ca2+ load and thus greater Ca2+ release and systolic force in subsequent beats [46,63]. The ability to modulate the inhibitory effect of PLN on SERCA is an important mechanism to quickly alter the rate of calcium sequestration, which may be necessary depending on the situational needs of the contractile tissue.

Sarcolipin

Another SERCA regulatory protein, which has garnered a lot of recent interest is sarcolipin (SLN). SLN is also a transmembrane protein found with the SR membrane (Figure 2). Structurally, it contains an 8 amino acid cytoplasmic domain, a 19 amino acid transmembrane alpha helix, and a 4 amino acid luminal tail [57,65–67]. SLN shares considerable homology with PLN within the transmembrane helix and thus associates with SERCA in a similar manner [46,68]. In humans, SLN is most abundantly expressed in fast-twitch type IIA muscle fibers where it most commonly associates with SERCA1a, although a small percentage of slow-twitch type I fibers also co-express SLN and SERCA1a [44]. Seemingly contradictory, muscles with the highest reported SLN expression in mice are the slower oxidative muscles (i.e soleus, red gastrocnemius, and diaphragm) [49,69,70]. This is likely explained by the fact that these slow oxidative mouse muscles contain a fiber type distribution in which 40–50% of fibers are type IIA [71]. Thus, like in human skeletal muscle, it appears SLN may be expressed within mouse type IIA fibers where it can regulate SERCA1a; however, single fiber data from mice are required to confirm this. SLN is also highly expressed in atrial cardiomyocytes where it regulates SERCA2a function, at least in mice [69,72].

Like PLN, SLN is an inhibitory regulator of SERCA, reducing the apparent affinity of cytosolic Ca2+ with SERCA. Unlike PLN, SLN binding can reduce the maximal catalytic activity of SERCA [44,57]. SLN has also been identified for its thermogenic properties [73]. SLN remains bound to SERCA during its catalytic cycle and interacts with one of the SERCA-Ca2+ binding sites [74]. Consequently, SLN but not PLN can reduce the efficiency of Ca2+ transport [74–76]. When this uncoupling of Ca2+ pumping occurs, more ATP must be consumed to pump the same quantity of Ca2+. Research has shown that this innate uncoupling mechanism is involved in the maintenance of core body temperature during cold exposure and energy balance during caloric surplus [73,75,77,78]. However, it should be noted that thermogenesis is also regulated through Ca2+ cycling mechanisms in resting mammalian muscles that don’t express SLN [79]. Additionally, the uncoupling mechanism elicited by SLN may also influence Ca2+ signaling as previous work from our laboratory has shown that SLN expression can affect calcineurin activity in models of muscle overload, disease and disuse [71,80–82]. Further evidence for the role of SLN in Ca2+ signaling has been shown by Maurya and colleagues (2017), who found that changes in SLN expression can affect the activity of the transcriptional coactivator, peroxisome proliferator-activated receptor-gamma coactivator-1alpha (PGC1α), CAMKII, and mitochondrial biogenesis pathways [83].

Other emerging SERCA regulators

Although PLN and SLN are the two most researched SERCA regulatory proteins, recent bioinformatics studies have identified other proteins with a similar helical pattern, which could potentially also regulate SERCA. Of the proteins identified, dwarf opening reading frame (DWORF) and myoregulin (MLN) appear to have gained the most interest [84,85]. Unlike negative regulators such as PLN and SLN, DWORF is believed to be a positive regulator of SERCA. By binding to SERCA, DWORF prevents the binding of inhibitory SERCA regulators such as PLN and SLN [45]. In doing this, DWORF maintains the sensitivity of SERCA at low [ Ca2+]cyt. MLN, on the other hand, is believed to be another inhibitory regulator of SERCA [43]. Interestingly, MLN also appears to have multiple sites, which could be phosphorylated like that of PLN [43]. More studies are needed to fully understand the function of these proteins in vivo.

Redox regulation of SERCA

Redox signaling plays an important role in the regulation of several of the major physiological systems of muscle including the SR Ca2+ regulatory system [86]. SERCAs are redox-sensitive proteins that may be activated by low levels of reactive oxygen (ROS) and nitrogen (RNS) species [87–89]. Physiologically, activation of SERCA2b in vascular smooth muscle occurs through redox signaling where nitric oxide and superoxide anion, through the formation of peroxynitrite, activate SERCA2b by reversible S-glutathiolation on Cys674 resulting in arterial relaxation [87]. The same molecular mechanism is also involved in the activation of SERCA2a in cardiac myocytes by the nitric oxide derivative nitroxyl, which may require an interaction with oxidized PLN [89]. Our work showing that glutathione depletion and cellular oxidation increased SERCA2a content, and maximal Ca2+-ATPase activity in rat diaphragm [88] supports the view that SERCA pump activity in skeletal muscle is also regulated through redox signaling.

Na+/Ca2+ exchanger

As mentioned, the NCX plays an important role in myocardial Ca2+ transport. Lying on the sarcolemma of cardiomyocytes, NCX is responsible for controlling the exchange of 1 Ca2+ ion for every 3 Na+ ions across the membrane [12,90–92]. The direction of this exchange can either be forward (Ca2+ out and Na+ in) or reverse (Ca2+ in and Na+ out) depending on the electrochemical gradient across the membrane [12,90–92]. When the membrane is depolarized, initially the reverse mode is favored because the membrane potential is greater than the combined (Na+ and Ca2+ ions) equilibrium potential of NCX, but the amount of Ca2+ entry is <1 μM [12,92]. As the [Ca2+]cyt rises due to SR Ca2+ release, the membrane potential effect becomes negligible, and the direction of the exchanger now changes to the forward mode with Ca2+ being extruded from the cytosol [12,92]. Although small changes in the subsarcolemmal [Na+] and [Ca2+] can impact the direction of the exchanger, under physiological conditions, the direction is largely driven by the [Ca2+]cyt [92]. After the release of Ca2+ into the cytosol during a contraction, the flow of ions observed in the NCX will be inward, which will result in the entry of Na+ and the extrusion of Ca2+ from the cell [12,91]. The opposite will occur with a positive membrane potential or increased cytosolic Na+ levels [12,92]. Under physiological conditions, NCX will be largely in its inward state [12]. Ion transport by NCX works in a similar stepwise fashion to SERCA, with ion binding sites facing the cytoplasmic side in the E1 state and facing the extracellular side in the E2 state [93–96].

NCX is regulated by Na+, Ca2+, and H+. Regulation by Na+ is referred to as Na+- dependent inactivation and this occurs in the presence of elevated intracellular Na+ levels [91,93,97]. High cytoplasmic Na+, such as during depolarization of the cell, leads to Na+ ions binding to their site on NCX in the E1 state, which then switches to an inactivated E1 state instead of cycling to an E2 state [93,95] This method of regulation is modulated by a part of the exchanger within the large cytoplasmic loop domain known as the XIP region as certain mutations within this region have been shown to abolish Na+-dependent regulation [97–99]. Ca2+ ions, on the other hand, up-regulate NCX activity by interacting with two other sites on the large cytoplasmic loop domain referred to as the Ca2+ binding domains 1 and 2 (CBD1 and CBD2) [91,100,101]. Although both CBD1 and CBD2 are involved in Ca2+-dependent activation, CBD1 is the primary Ca2+ sensor due to its 7-fold higher affinity for Ca2+ compared with CBD2, binding up to 4 Ca2+ ions at a range of 200 nM to 1 μM [101]. Upon a rise in cytosolic Ca2+ such as during E-C coupling, CBD1 and CBD2 change conformation and undergo an electrostatic shift bringing them closer together [101]. This conformational change allows the signal to be relayed via a small domain, called the α-catenin-like domain (CLD), to the transmembrane domain to increase NCX activity [100,101]. CBD2 can bind up to two Ca2+ ions during times of very high [Ca2+] to potentially help to overcome Na+-dependent inactivation by increasing the electrostatic potential [101]. Finally, intracellular pH modulates activity of NCX by increasing or decreasing its activity if the pH is increased or decreased, respectively [91,102–105]. When intracellular pH decreases, NCX inhibition occurs in two steps: (1) primary or fast blockade that occurs with rising H+ levels and works independent of the [Na+]cyt levels and (2) secondary or slow blockade that is dependent on [Na+]cyt levels, which enhance the affinity of NCX for H+ [102,103,105]. The mechanism for pH modulation of NCX was previously thought as H+ competing with Ca2+ for sites at CBD1 [91]. However, mutagenesis studies have identified histidine residues 124 and 165 as two important players which modulate NCX function allosterically through H+ binding, which is distinct from regulation by Na+ or Ca2+ [104].

Within skeletal muscle, two isoforms of NCX are expressed, NCX1 and NCX3 [106], with the latter being the predominant isoform [107]. Although NCX is expressed in skeletal muscle, it doesn’t appear to be directly involved in E-C coupling [108]. The rate of Ca2+ transport by the skeletal muscle NCX has been reported to be ∼30 times slower compared with cardiac muscle [109]. It has been postulated this transporter is more involved in regulation of [Ca2+]cyt during repeated muscle contractions when [Ca2+]cyt is very high [109,110]. Furthermore, NCX3 knockout mice display low levels of necrosis within the muscle fibers, suggesting that the NCX is important for maintenance of Ca2+ homeostasis within skeletal myofibers [107].

Calcium buffering proteins

The presence of cytosolic Ca2+ buffering proteins can also influence the decay phase of Ca2+ transients in muscle. Within the cytosol of myofibers, the most well documented Ca2+ buffering protein is parvalbumin (PV), which has a role in lowering [Ca2+]cyt [23,111,112]. Within small mammalian animal models such as mice and rats, PV has been found in neural tissue and type II muscle fibers [113]. In contrast, little or no PV is expressed in cardiac and type I muscle fibers [111–113]. Furthermore, PV has not been detected in human skeletal muscle [114]. PV is a 12 kDa protein that binds Ca2+ at an optimal ratio of 2 mol Ca2+:1 mol PV [115,116]. The two high affinity binding sites are occupied by Mg2+ at resting [Ca2+]cyt (<100 nM) [117]. The rate at which PV can act to sequester Ca2+ is dependent on the rate of dissociation of Mg2+ from the binding sites [112,118]. Consequently, PV buffering does not occur immediately following rises in [Ca2+]cyt, which is why it is often considered a slow Ca2+ buffer [117]. The binding kinetics of PV explain why its increased expression has been shown to reduce 1⁄2 relaxation time and increase the rate of relaxation (−df/dt) without altering the single twitch contractile properties of skeletal muscle [6,7,112,119,120]. However, during low frequency stimulation (30 and 50 Hz) of type I fibres, PV overexpression attenuates force production [7]. Taken together, this research provides evidence that PV is a Ca2+ buffering protein, which can increase rates of Ca2+ sequestration during E-C coupling.

Although the mitochondria are not traditionally known for Ca2+ buffering, there is evidence showing that Ca2+ uptake through the mitochondrial Ca2+ uniporter is important for the maintenance of myocyte homeostasis [121]. Although it’s fairly well established that mitochondrial Ca2+ uptake in cardiomyocytes does not alter [Ca2+]cyt transients or cardiac contractility, there is evidence that mitochondria can modulate [Ca2+]cyt in fast-twitch skeletal muscle fibers under certain conditions [121], a view that is supported by recent evidence indicating that mitochondria in fast-twitch mouse fibers have a high Ca2+-buffering capacity [122]. A recent paper by Butera and colleagues (2021) highlighted the dynamic relationship between the mitochondria and PV regarding Ca2+ buffering [123]. Butera and colleagues reported that when PV was overexpressed, mitochondrial volume and number decrease but when PV was down-regulated, the mitochondria increased in size and quantity [123]. This alone may not indicate that mitochondria are being up-regulated to buffer Ca2+ but rather mitochondrial biogenesis results from the activation of Ca2+-dependent signaling pathways in the absence of PV. However, an interesting observation made by Butera and collogues was that when PV was down-regulated there was a significantly greater proportion of mitochondria located near Ca2+ release units suggesting mitochondria may have a larger role in Ca2+ buffering than originally thought. Another interesting consideration is that there is a grouping of mitochondria localized around the longitudinal SR, where SERCA is located [124]. If mitochondria do have a role in buffering Ca2+, then the location of the mitochondria may have a role in altering Ca2+ concentrations in areas around SERCA [123,125].

Calcium related myopathies without muscle wasting

Among skeletal muscle myopathies, there is a grouping that involves prolonged contractions due to a reduced ability to lower ICTs during the relaxation phase of E-C coupling. Most notably among this grouping of myopathies is Brody’s disease and Brody’s syndrome, both of which involve reductions in SERCA1a activity [126]. Although similar in name and presentation, Brody’s disease but not Brody’s syndrome involves a mutation in the ATP2A1 gene encoding SERCA1a [127]. Even within the myopathies classified as Brody’s disease, there is a degree of heterogeneity in the ATP2A1 mutations as they have been documented in the A-, P-, and N-domains as well as several others at various points along the transmembrane helices [126]. Among the genetically unresolved cases of Brody’s syndrome, one possibility is that mutations of a SERCA inhibitor (e.g., SLN, PLN, or MLN) could be causing reductions in SERCA function [128], but none have been documented to date.

Although there are differences in the mutations leading to myopathy pathogenesis of Brody’s disease and Brody’s syndrome, the symptoms of these diseases are similar in presentation. Research has shown that reductions in SERCA activity appear to be specific to the SERCA1a isoform in type II muscle fibers [129]. As a result, prolonged contractions with delays in relaxation become apparent during faster contractions, which require the activation of type II fibres. Sustained contractions in this disease have been associated with increased plasma creatine kinase, cases of rhabdomyolysis and malignant hyperthermia-like episodes [126,130,131]. Despite these acute harmful effects brought on by the reduced rate of skeletal muscle relaxation, there does not appear to be a muscle wasting aspect to these group of diseases [132]. Although some reports have suggested atrophy, which is likely a result of the underutilization of the type II fibres [126], there does not appear to be progressive wasting brought on by this disease.

Immunohistochemical analysis of muscle samples from Brody’s disease patients revealed a severe reduction in the SERCA content of type II fibres while the SERCA content in type I fibres was normal [126]. Furthermore, immunoblot quantification from whole muscle homogenates revealed an absolute reduction in SERCA1a content [126]. Within human tissue it has been shown that type II fibres do express small amounts of the SERCA2a isoform [44,133]. This may explain why Ca2+ concentrations are able to return to resting levels in type II fibres even when the SERCA1a protein is either partially functioning or completely nonfunctional. However, since the endogenous expression of SERCA2a in type II fibres is relatively low, in Brody’s disease a substantially smaller quantity of SERCA pumps contributes to Ca2+ sequestration, thus resulting in a slower decay of the ICT.

Pan and colleagues attempted to create a murine model of Brody’s disease by knocking out the gene encoding SERCA1 [134]. In creating this model, it was found that neonatal mice were born healthy but became cyanotic and died shortly after birth [134]. It was hypothesized that the SERCA1 knockout mice pups were unable to properly ventilate due to the relatively high proportion of type II fibers in mouse diaphragm resulting in a greater reliance on SERCA1a for diaphragm relaxation compared with humans.

Myopathies involving elevations in resting calcium concentrations

Another grouping of muscle diseases is characterized by muscle wasting. These diseases share a commonality in that elevations in resting Ca2+ concentrations appear to contribute to the pathogenesis, and in some cases, progression of the disease phenotype. These diseases include Duchenne’s muscular dystrophy (DMD) and centronuclear myopathy (CNM).

Muscular dystrophies are a heterogenous group of inherited disorders, which vary genetically and in clinical presentation [135–137]. These disorders involve increased muscle turnover resulting in progressive atrophy of the skeletal muscles [137–139]. Among these diseases, DMD is considered the most common and the most severe [139]. DMD is an X-linked recessive disease which results in atrophy of the limb, axial, and facial muscles. The life expectancy of individuals with this disease is reduced by about 75% at which point death is usually caused by respiratory or cardiac complications [139].

From a molecular standpoint, DMD arises from mutations in the DMD gene resulting in a deficiency in the dystrophin protein. Dystrophin is a rod-shaped cytoskeletal protein which forms a critical link between the submembrane cytoskeleton and proteins of the extracellular matrix. Importantly, dystrophin associates with proteins within the sarcolemma, which are integrated in pathways associated with Ca2+ influx and ROS signaling. Muscular dystrophy affects fibres specialized for faster contractions, as the instability from the absence of dystrophin results in greater myofiber damage [140]. The myofiber damage results in a continuous cycle of muscle turnover through the up-regulation of degeneration and regeneration pathways [141,142]. The increased muscle turnover subsequently results in inflammation and the infiltration of collogen into the muscle making it more fibrotic [141,143,144].

There is still some ambiguity in how the deficiency of dystrophin results in the progression of DMD. Currently, there are two leading theories as to how DMD progresses, both of which involve elevations in [Ca2+]cyt. The first theory posits the lack of dystrophin results in transient tearing of the sarcolemma membrane during contractions. The membrane damage could allow for an influx of extracellular Ca2+ into the myofiber [145–147]. A second theory lies in dystrophin’s role as a scaffolding protein and its role in positioning membrane ion channels. In the absence of dystrophin, these ion channels may function improperly and increase Ca2+ influx into the myofiber [145–147]. Prolonged elevations in [Ca2+]cyt can increase proteolytic activity and the production of ROS [82,148–150]. Furthermore, increased ROS can inactivate SERCA function which can further augment Ca2+ dysregulation [88,151–155]. Previous work by Goonasekera and colleagues (2011) has shown that elevations in [Ca2+]cyt alone are enough to promote a shift toward a DMD like phenotype [156]. This was shown through preventing increases in [Ca2+]cyt by overexpressing SERCA in dystrophic tissue, which ameliorated the pathology [156]. This work has provided strong evidence for the role of Ca2+ in promoting the progression of DMD. Recently, the use of CDN1163, a pharmaceutical SERCA activator, has been used as a therapeutic intervention for DMD through its ability to promote Ca2+ sequestration in the SR. Specifically, Nogami et al. (2021) administered CDN1163 to mdx mice, which reduced [Ca2+]cyt and decreased muscular degeneration and fibrosis [157]. Using a different approach to restore SERCA function in DMD models, Gehrig and colleagues (2012) demonstrated that both transgenic and pharmacological overexpression of the chaperone protein, heat shock protein 72 (HSP72), could attenuate the progression of DMD [158]. HSP72 associates with SERCA and acts to protect it against oxidative damage and inactivation [152,158,159].

CNM is a heterogeneous group of inherited neuromuscular diseases characterized by increased localization of centralized nuclei. Although variance exists among disease phenotypes, other common histological indicators include an increased proportion of type I fibres, centralized aggregations of oxidative activity, muscle atrophy and weakness [160]. The most severe cases of CNM arise in a X-linked inheritance pattern from a mutation in the MTM1 gene. MTM1 encodes for myotubularin, which functions to regulate PI(3)P, endocytosis and endolysosomal function [161–163]. Milder cases of CNM have been reported in individuals with mutations in DNM2, which encodes dynamin 2, and BIN1, which encodes amphiphysin-2 [163]. MTM1, DNM2, and BIN1 are all involved with membrane trafficking suggesting they all work through the same pathogenic pathway [161–166]. Interestingly, Ca2+ dysregulation is also a common feature within each of these genetic variations of CNM [163,167,168]. Recently, mutations in RYR1 and TTN, the genes encoding skeletal muscle RyR and the cytoskeletal protein titin, respectively, have also been implicated in the development of a CNM phenotype providing greater evidence for the involvement of Ca2+ in CNM pathogenesis [169,170]. RyR1-related CNM differs from other forms of CNM as there appears to be no direct links to defective membrane trafficking [163]. Mutations to TTN which cause CNM are relatively heterogenous but most appear to involve C-terminus truncations that are associated with a reduction in calpain-3 and nebulin-2, two proteins which interact with the C-terminal region, which may cause irregularities in Ca2+ release [163,169,171].

While trying to characterize the role of PLN in SERCA regulation in skeletal muscle, Song and colleagues (2004) discovered that PLN overexpression specifically in type I muscle fibres results in muscle disease [172], which was later recognized as a CNM-like phenotype [173]. Fajardo et al. (2015) determined that mice with targeted overexpression of PLN in type I fibres displayed an increased centralization of nuclei and oxidative activity as well as type I fibre hypotrophy at 1 month of age [173]. Furthermore, a fibre type shift toward type I fibres was evident by 4–6 months [173]. Interestingly, because the Pln transgene is attached to the β-MHC promoter in this model, the shift to type I fibres leads to greater expression of PLN causing greater disease severity. In this model, there is also atrophy of type I fibres while there appears to be a compensatory hypertrophy of type II fibres [173]. Through histological analysis, it was suggested by Fajardo and colleagues that similarities in oxidative staining and fibrosis make PLN-related CNM resemble a phenotype which appears more closely related to TTN- and RyR-CNM [173]. Furthermore, Fajardo and colleagues reported a significant 53% reduction in SERCA activity and increases in both total and monomeric PLN content in muscle samples from three human CNM patients compared with five healthy subjects. Overall, this work has shown that overexpressing a SERCA inhibitor and severely impairing SR Ca2+ transport can cause CNM. Therefore, future studies should assess whether targeting SERCA function is a viable therapeutic strategy for this disease.

Abnormal Ca2+ loading in heart failure

Heart failure (HF) can present with either a reduced (HFrEF) or preserved (HFpEF) ejection fraction [174,175]. HFrEF involves the inability of the LV to generate sufficient force to pump out enough blood due to cardiomyocyte loss and dysfunction (e.g., following a myocardial infarction) [174,176]. This leads to eccentric remodeling whereby the existing myocytes stretch and become thinner, alongside an increase in fibrosis in the extracellular space [174]. HFpEF involves the inability of the LV to sufficiently relax and allow for filling due to the enlargement of cardiomyocytes that is caused by an increase in afterload (e.g., hypertension) [174,176]. This leads to concentric remodeling where the walls of the LV thicken and elevate filling pressures [174]. Overall, in both cases, the LV walls enlarge causing dysfunction in the chamber [174–176].

It is widely observed from electrophysiological studies that myocytes from failing hearts have abnormal Ca2+ transients that are smaller in amplitude and longer in duration which result in abnormal force production [177–181]. It is well established that cardiomyocytes from nonfailing hearts display a positive force–frequency relationship with increasing frequencies of stimulation causing increases in systolic force [182–184]. However, in myocytes from failing hearts, this is only true at lower frequencies as force decreases at higher frequencies of stimulation (i.e., negative force–frequency relationship) and accordingly, the peak [Ca2+]cyt is decreased in these cells [181,183–185]. In contrast, diastolic force increases with increasing frequency at all stimulation frequencies in the cells from failing hearts and this is associated with an increase in diastolic [Ca2+]cyt [181,186].

While decreases in Ca2+ release with HF could be due to defective coupling between the RyR and DHPR, reduced SR Ca2+ content could also contribute. This was assessed in one study using a canine model of HF by assessing the CICR gain, which is the amount of Ca2+ released from the SR for a given level of current density from the DHPR channel and ε, which is the effectiveness of coupling between DHPR activation and SR Ca2+ release [178]. The CICR gain was lower in the failing cardiomyocytes along with lower SR Ca2+ content compared with the healthy cardiomyocytes, but no difference was observed for ε between the two groups. This indicates that in this model of HF it was the reduced SR Ca2+ content that led to reduced Ca2+ release in the failing cardiomyocytes and not reduced ε. Furthermore, after bathing the failing cardiomyocytes in a high Ca2+ solution (5 mmol/L) to bring the SR Ca2+ content back to levels observed in their healthy counterparts, CICR gain was recovered to levels seen in the healthy cells. This is further supported by another study which showed that despite smaller and longer Ca2+ transients, RyR content and activity was intact in failing canine and human hearts [179]. Nonetheless, it should be noted that studies using other models of HF have found ineffective coupling to contribute to the pathology [187–189].

As previously mentioned, SR Ca2+ uptake by SERCA is the primary pathway involved in lowering [Ca2+]cyt and inducing relaxation of the myocardium during diastole. Several studies have shown that SERCA Ca2+ uptake is reduced in failing hearts from both human and animal models [177,179,181,185,190,191]. In theory, reduced SERCA activity could be due to either decreased SERCA content or increased PLN content leading to greater inhibition of SERCA. Interestingly, reductions in both SERCA and PLN content are consistently observed in HF [179,185,190,192,193], which translates to either unchanged or decreased SERCA:PLN ratios (i.e., greater inhibition of SERCA by PLN). Another reason behind the lower Ca2+ uptake in HF could be decreased phosphorylation of PLN leading to increased inhibition of SERCA. Although results have varied across models of HF, there is good evidence to suggest that PLN phosphorylation is probably reduced [194–196]. This may be surprising given that CAMKII expression and activity have been found to be increased in HF [197]; however, given that PLN phosphorylation status is dependent on a balance of kinase and phosphatase activity, this can be explained by increased protein phosphatase 1 (PP1) activity [198,199]. With regards to other studies which show contrasting results, the stage in the progression of HF at which phosphorylation of PLN was assessed can impact the results [194,195]. Regardless of changes in the levels of the aforementioned proteins observed across different models, Ca2+ uptake into the SR is reduced.

NCX, being the second major contributor in the myocardium to the transport of Ca2+ out of the cytosol, plays a compensatory role in HF. In the face of decreased SERCA content, NCX content, and accordingly its activity, is found to be increased in HF [192,200,201]. A study in failing canine cardiomyocytes found that while SERCA levels decreased by approximately 28%, NCX levels increased by 104% [192]. In order to assess the contribution of SERCA and NCX to Ca2+ removal during relaxation, the authors first inhibited SERCA and found the time it takes to reach 50% of maximal force during relaxation was significantly greater (i.e., slower relaxation) in the nonfailing hearts compared with the failing hearts indicating that SERCA has a greater contribution to relaxation in these cells. They then inhibited NCX by placing the cells in a Na+-free solution, which increased the time to 50% maximal force during relaxation in both sets of cardiomyocytes but only up to 30% in normal cells and up to 97% in failing cells, indicating that NCX contributes more to relaxation in the failing cells. More recently, a study using a guinea pig HF model found that while the contribution to Ca2+ removal by SERCA decreased by 28% that of NCX increased by 63% [202]. Although this compensation of increased NCX content helps to remove [Ca2+]cyt, it can actually increase the probability of arrhythmogenic events. In failing hearts, with increasing frequency of stimulation, the intracellular Na+ ([Na+]cyt) levels rise [186,203], due to increased activity of the Na+/H+ exchanger on the sarcolemma [204] and reduced Na+/K+ ATPase activity [202] causing increased reverse-mode NCX activity [186,205]. Under β-adrenergic stimulation, the rise in NCX-reverse current would increase [Ca2+]cyt and SERCA uptake of Ca2+ into the SR would also increase leading to SR Ca2+ overload [186,205]. Therefore, the SR Ca2+ overload, alongside an increased membrane potential due to reduced inward rectifier K+ channel current, would increase the susceptibility to arrhythmias [186,205].

Abnormalities in Ca2+ handling in HF are not isolated to cardiac muscle but have similarly been found in skeletal muscle. Several animal studies have shown abnormalities in Ca2+ handling in skeletal muscle with HF, including reduced SERCA1 and 2 content [206], lower Ca2+ uptake [207], and lower Ca2+ release [208], which impairs muscle force generation capacity and increases fatiguability [208–210]. Middlekauff and colleagues (2012) sought to examine if such skeletal muscle Ca2+ handling abnormalities also occurred in humans with HF and whether oxidative stress may be involved [211]. Analyses of vastus lateralis biopsy samples revealed lower DHPR and SERCA2a content alongside lower phosphorylated PLN in advanced HF patients compared with healthy controls; however, there were no differences in markers of oxidative stress [211]. In cardiac muscle, chronic augmentation of sympathetic nerve activity leads to increased CAMKII expression and activity resulting in a phenomenon known as excitation transcription coupling whereby CAMKII acts as a long-term regulator of hypertrophic genes and Ca2+ handling proteins [212,213]. Thus, Middlekauf and colleagues postulated, because skeletal muscle is also subject to elevated sympathetic activity, that excitation-transcription coupling alongside increased PP1 activity may also be at play in this tissue causing the Ca2+ handling abnormalities observed. Therefore, the Ca2+ dysregulation seen in both tissues are mirroring each other likely due to the same mechanism.

Altered redox regulation of SERCA in disease

As mentioned earlier, SERCAs are redox-sensitive proteins that are activated by low levels of ROS and RNS, but they are also highly susceptible to oxidative damage and inactivation by high levels of ROS and RNS (i.e., oxidative stress) [153,154,214–216]. Cohen and colleagues have shown that in disease, irreversible oxidation of key thiols, specifically cysteine-674, prevents activation of SERCAs through redox signaling mechanisms [87,217,218]. Several studies have also noted increased oxidation and nitration of SERCA in aged skeletal muscle [155,214,219], which we have shown results in the loss of redox control of SERCA activity and expression [220]. However, our work showing that several SERCA-binding proteins, including heat shock protein 70, PLN, and SLN, can protect SERCA structure and function during cellular stress [152,159,221], suggests that these SERCA-binding proteins may play a crucial role in protecting cellular Ca2+ homeostasis and preserving cardiac and skeletal muscle function under conditions of chronic oxidative stress and disease.

Maintenance of Ca2+ homeostasis is essential for healthy muscle. Within skeletal myofibers and cardiomyocytes, transient increases in [Ca2+]cyt during E-C coupling must be rapidly lowered to re-establish resting concentrations and induce muscle relaxation, which is primarily accomplished by SERCA-mediated Ca2+ transport into the SR. During short contractions of fast-twitch rodent and amphibian myofibers, the cytosolic Ca2+ buffer PV also contributes to muscle relaxation by rapidly lowering [Ca2+]cyt. In the heart, Ca2+ extrusion from cardiomyocytes through the NCX contributes significantly to ventricular relaxation, and the PMCA and MCU also contribute to a very minor extent. Impairments in Ca2+ transport and dysfunctional Ca2+ homeostasis has consistently been shown to be involved with the pathogenesis and/or the progression of multiple myopathies within skeletal muscle and in HF. A better understanding of the function and regulation of muscle Ca2+ transport proteins can allow for greater insights into skeletal and cardiac muscle physiology and disease. Furthermore, this body of work can lead to novel approaches in the treatment of diseases involving Ca2+ dysregulation.

The authors declare that there are no competing interests associated with the manuscript.

The original work from the authors cited in this review was supported by Natural Sciences and Engineering Research Council of Canada (NSERC) [grant number #RGPIN-2020-05632 (to A.R.T.)].

Mark A. Valentim: Conceptualization, Writing—original draft, Writing—review and editing. Aditya N. Brahmbhatt: Conceptualization, Writing—original draft, Writing—review and editing. A. Russell Tupling: Conceptualization, supervision, funding acquisition, Writing—original draft, Writing—review and editing.

α-CLD

α-catenin-like domain

[Ca2+]cyt

cytosolic calcium concentration

[Na+]cyt

cytosolic sodium concentration

Ca2+

calcium

CAMKII

calcium/calmodulin protein kinase II

CBD1

calcium binding domain 1

CBD2

calcium binding domain 2

CICR

calcium-induced calcium release

CNM

centronuclear myopathy

DHPR

dihydropyridine receptor

DMD

Duchenne’s Muscular Dystrophy

DWORF

Dwarf Open Reading Frame

E-C Coupling

excitation–contraction coupling

HF

heart failure

HFpEF

heart failure with preserved ejection fraction

HFrEF

heart failure with reduced ejection fraction

HSP72

heat shock protein 72

ICT

intracellular calcium transient

LV

left ventricle

MCU

mitochondrial calcium uniporter

MLN

myoregulin

Na+

sodium

NCX

sodium (Na+)/calcium (Ca2+) exchanger

PI(3)P

phosphatidyl inositol (3) phosphate

PLN

phospholamban

PV

parvalbumin

RNS

reactive nitrogen species

ROS

reactive oxygen species

RyR

ryanodine receptor

SERCA

sarco(endo)plasmic reticulum Ca2+-ATPase

SLN

sarcolipin

SR

sarcoplasmic reticulum

Vmax

maximal rate

1.
Carafoli
E.
and
Krebs
J.
(
2016
)
Why calcium? How calcium became the best communicator
J. Biol. Chem.
291
,
20849
20857
[PubMed]
2.
Drake
S.K.
and
Falke
J.J.
(
1996
)
Kinetic tuning of the EF-hand calcium binding motif: The gateway residue independently adjusts (i) barrier height and (ii) equilibrium
.
Biochemistry
35
,
1753
1760
[PubMed]
3.
Schiaffino
S.
and
Reggiani
C.
(
2011
)
Fiber types in mammalian skeletal muscles
.
Physiol. Rev.
91
,
1447
1531
[PubMed]
4.
Gifford
J.L.
,
Walsh
M.P.
and
Vogel
H.J.
(
2007
)
Structures and metal-ion-binding properties of the Ca2+-binding helix-loop-helix EF-hand motifs
.
Biochem. J.
405
,
199
221
[PubMed]
5.
Iino
M.
(
2010
)
Spatiotemporal dynamics of Ca2+ signaling and its physiological roles
.
Proc. Japan Acad. Ser. B.
86
,
244
256
6.
Chen
G.
,
Carroll
S.
,
Racay
P.
,
Dick
J.
,
Pette
D.
,
Traub
I.
et al.
(
2001
)
Deficiency in parvalbumin increases fatigue resistance in fast-twitch muscle and upregulates mitochondria
.
Am. J. Physiol. Cell Physiol.
281
,
114
122
7.
Chin
E.R.
,
Grange
R.W.
,
Viau
F.
,
Simard
A.R.
,
Humphries
C.
,
Shelton
J.
et al.
(
2003
)
Alterations in slow-twitch muscle phenotype in transgenic mice overexpressing the Ca2+ buffering protein parvalbumin
.
J. Physiol.
547
,
649
663
[PubMed]
8.
Racay
P.
,
Gregory
P.
and
Schwaller
B.
(
2006
)
Parvalbumin deficiency in fast-twitch muscles leads to increased “slow-twitch type” mitochondria, but does not affect the expression of fiber specific proteins
.
FEBS J.
273
,
96
108
[PubMed]
9.
Maclennan
D.H.
,
Rice
W.J.
and
Green
N.M.
(
1997
)
The mechanism of Ca2+ transport by sarco (endo) plasmic reticulum Ca2+-ATPase
.
J. Biol. Chem.
272
,
28815
28818
[PubMed]
10.
Toyoshima
C.
and
Inesi
G.
(
2004
)
Structural basis of ion pumping by Ca2+-ATPase of the sarcoplasmic reticulum
.
Annu. Rev. Biochem.
73
,
269
292
[PubMed]
11.
Smith
I.
,
Bombardier
E.
,
Vigna
C.
and
Tupling
A.R.
(
2013
)
ATP consumption by sarcoplasmic reticulum Ca2+ pumps accounts for 40-50% of resting metabolic rate in mouse fast and slow twitch skeletal muscle
.
PLoS ONE
8
,
1
11
12.
Bers
D.M.
(
2002
)
Cardiac excitation-contraction coupling
.
Nature
415
,
198
205
[PubMed]
13.
Calderón
J.C.
,
Bolaños
P.
and
Caputo
C.
(
2014
)
The excitation-contraction coupling mechanism in skeletal muscle
.
Biophys. Rev.
6
,
133
160
[PubMed]
14.
Rios
E.
and
Brum
G.
(
1987
)
Involvement of dihydropyridine receptors in excitation-contraction coupling in skeletal muscle
.
Nature
325
,
717
720
[PubMed]
15.
Rebbeck
R.T.
,
Karunasekara
Y.
,
Board
P.G.
,
Beard
N.A.
,
Casarotto
M.G.
and
Dulhunty
A.F.
(
2014
)
Skeletal muscle excitation-contraction coupling: Who are the dancing partners?
Int. J. Biochem. Cell Biol.
48
,
28
38
[PubMed]
16.
Toyoshima
C.
and
Wakabayashi
T.
(
1985
)
Three-dimensional image analysis of the complex of thin filaments and myosin molecules from skeletal muscle. IV. Reconstitution from minimal- and high-dose images of the actin-tropomyosin-myosin subfragment-1 complex
.
J. Biochem.
97
,
245
263
[PubMed]
17.
Gorecka
A.
,
Aksoy
M.
and
Hartsthorne
J.
(
2016
)
The effect of phosphorylation of glizzard myosin on actin activation
.
Biochem. Biophys. Res. Commun.
71
,
325
331
18.
Huxley
H.E.
(
1969
)
The mechanism of muscular contraction
.
Am. Heart J.
164
,
1356
1366
19.
Huxley
H.E.
(
1968
)
Structural difference between resting and rigor muscle. Evidence from intensity changes in the low-angle equatorial X-ray diagram
.
J. Mol. Biol.
37
,
507
520
[PubMed]
20.
Baylor
S.M.
and
Hollingworth
S.
(
2003
)
Sarcoplasmic reticulum calcium release compared in slow-twitch and fast-twitch fibres of mouse muscle
.
J. Physiol.
551
,
125
138
[PubMed]
21.
Smith
I.C.
,
Gittings
W.
,
Huang
J.
,
McMillan
E.M.
,
Quadrilatero
J.
,
Tupling
R.R.
et al.
(
2013
)
Potentiation in mouse lumbrical muscle without myosin light chain phosphorylation: Is resting calcium responsible?
J. Gen. Physiol.
141
,
297
308
[PubMed]
22.
Smith
I.C.
,
Vandenboom
R.
and
Tupling
A.R.
(
2014
)
Juxtaposition of the changes in intracellular calcium and force during staircase potentiation at 30 and 37°C
.
J. Gen. Physiol.
144
,
561
570
[PubMed]
23.
Baylor
S.M.
and
Hollingworth
S.
(
2003
)
Sarcoplasmic reticulum calcium release compared in slow-twitch and fast-twitch fibres of mouse muscle
.
J. Physiol.
551
,
125
138
[PubMed]
24.
Biesiadecki
B.J.
,
Davis
J.P.
,
Ziolo
M.T.
and
Janssen
P.M.L.
(
2014
)
Tri-modal regulation of cardiac muscle relaxation; intracellular calcium decline, thin filament deactivation, and cross-bridge cycling kinetics
.
Biophys. Rev.
6
,
273
289
[PubMed]
25.
Bassani
J.W.
,
Bassani
R.A.
and
Bers
D.M.
(
1994
)
Relaxation in rabbit and rat cardiac cells: species‐dependent differences in cellular mechanisms
.
J. Physiol.
476
,
279
293
[PubMed]
26.
Puglisi
J.L.
,
Bassani
R.A.
,
Bassani
J.W.M.
,
Amin
J.N.
and
Bers
D.M.
(
1996
)
Temperature and relative contributions of Ca transport systems in cardiac myocyte relaxation
.
Am. J. Physiol. Heart Circ. Physiol.
270
,
H1772
H1778
27.
Waggoner
J.R.
,
Ginsburg
K.S.
,
Mitton
B.
,
Haghighi
K.
,
Robbins
J.
,
Bers
D.M.
et al.
(
2009
)
Phospholamban overexpression in rabbit ventricular myocytes does not alter sarcoplasmic reticulum Ca transport
.
Am. J. Physiol. Heart Circ. Physiol.
296
,
698
703
28.
Toyoshima
C.
,
Nakasako
M.
,
Nomura
H.
and
Ogawa
H.
(
2000
)
Crystal structure of the calcium pump of sarcoplasmic reticulum at 2.6 Å resolution
.
Nature
405
,
647
655
[PubMed]
29.
Toyoshima
C.
,
Iwasawa
S.
,
Ogawa
H.
,
Hirata
A.
,
Tsueda
J.
and
Inesi
G.
(
2013
)
Crystal structures of the calcium pump and sarcolipin in the Mg2+-bound E1 state
.
Nature
495
,
260
264
[PubMed]
30.
Brandl
C.J.
,
Green
N.M.
,
Korczak
B.
and
MacLennan
D.H.
(
1986
)
Two Ca2+ ATPase genes: Homologies and mechanistic implications of deduced amino acid sequences
.
Cell
44
,
597
607
[PubMed]
31.
Brandl
C.J.
,
de Leon
S.
,
Martin
D.R.
and
MacLennan
D.H.
(
1987
)
Adult forms of the Ca2+ATPase of sarcoplasmic reticulum. Expression in developing skeletal muscle
.
J. Biol. Chem.
262
,
3768
3774
[PubMed]
32.
Lytton
J.
and
MacLennan
D.H.
(
1988
)
Molecular cloning of cDNAs from human kidney coding for two alternatively spliced products of the cardiac Ca2+-ATPase gene
.
J. Biol. Chem.
263
,
15024
15031
[PubMed]
33.
MacLennan
D.H.
,
Brandl
C.J.
,
Korczak
B.
and
Green
N.M.
(
1985
)
Amino-acid sequence of a Ca2+-Mg2+-dependent ATPase from rabbit muscle sarcoplasmic reticulum, deduced from its complementary DNA sequence
.
Nature
316
,
696
700
[PubMed]
34.
De la Bastie
D.
,
Levitsky
D.
,
Rappaport
L.
,
Mercadier
J.J.
,
Marotte
F.
,
Wisnewsky
C.
et al.
(
1990
)
Function of the sarcoplasmic reticulum and expression of its Ca2+-ATPase gene in pressure overload-induced cardiac hypertrophy in the rat
.
Circ. Res.
66
,
554
564
[PubMed]
35.
Periasamy
M.
and
Kalyanasundaram
A.
(
2007
)
SERCA pump isoforms: their role in calcium transport and disease
.
Muscle Nerve
35
,
430
442
[PubMed]
36.
Wu
K.D.
,
Bungard
D.
and
Lytton
J.
(
2001
)
Regulation of SERCA Ca2+ pump expression by cytoplasmic [Ca2+] in vascular smooth muscle cells
.
Am. J. Physiol. - Cell Physiol.
280
,
843
851
37.
Lytton
J.
,
Westlin
M.
,
Burk
S.E.
,
Shull
G.E.
and
MacLennan
D.H.
(
1992
)
Functional comparisons between isoforms of the sarcoplasmic or endoplasmic reticulum family of calcium pumps
.
J. Biol. Chem.
267
,
14483
14489
[PubMed]
38.
Murphy
R.M.
,
Larkins
N.T.
,
Mollica
J.P.
,
Beard
N.A.
and
Lamb
G.D.
(
2009
)
Calsequestrin content and SERCA determine normal and maximal Ca2+ storage levels in sarcoplasmic reticulum of fast- and slow-twitch fibres of rat
.
J. Physiol.
587
,
443
460
[PubMed]
39.
Lüss
I.
,
Boknik
P.
,
Jones
L.R.
,
Kirchhefer
U.
,
Knapp
J.
,
Linck
B.
et al.
(
1999
)
Expression of cardiac calcium regulatory proteins in atrium v ventricle in different species
.
J. Mol. Cell Cardiol.
31
,
1299
1314
[PubMed]
40.
Minajeva
A.
,
Kaasik
A.
,
Paju
K.
,
Seppet
E.
,
Lompré
A.M.
,
Veksler
V.
et al.
(
1997
)
Sarcoplasmic reticulum function in determining atrioventricular contractile differences in rat heart
.
Am. J. Physiol. Heart Circ. Physiol.
273
,
2498
2507
41.
Periasamy
M.
,
Reed
T.D.
,
Liu
L.H.
,
Ji
Y.
,
Loukianov
E.
,
Paul
R.J.
et al.
(
1999
)
Impaired cardiac performance in heterozygous mice with a null mutation in the sarco(endo)plasmic reticulum Ca2+-ATPase isoform 2 (SERCA2) gene
.
J. Biol. Chem.
274
,
2556
2562
[PubMed]
42.
He
H.
,
Giordano
F.J.
,
Hilal-Dandan
R.
,
Choi
D.J.
,
Rockman
H.A.
,
McDonough
P.M.
et al.
(
1997
)
Overexpression of the rat sarcoplasmic reticulum Ca2+ ATPase gene in the heart of transgenic mice accelerates calcium transients and cardiac relaxation
.
J. Clin. Invest.
100
,
380
389
[PubMed]
43.
Anderson
D.M.
,
Anderson
K.M.
,
Chang
C.
,
Makarewich
C.A.
,
Nelson
B.R.
,
Mcanally
J.R.
et al.
(
2015
)
A micropeptide encoded by a putative long non-coding RNA regulates muscle performance
.
Cell
160
,
595
606
[PubMed]
44.
Fajardo
V.A.
,
Bombardier
E.
,
Vigna
C.
,
Devji
T.
,
Bloemberg
D.
,
Gamu
D.
et al.
(
2013
)
Co-Expression of SERCA isoforms, phospholamban and sarcolipin in human skeletal muscle fibers
.
PLoS ONE
8
,
1
13
45.
Makarewich
C.A.
,
Munir
A.Z.
,
Schiattarella
G.G.
,
Bezprozvannaya
S.
,
Raguimova
O.N.
,
Cho
E.E.
et al.
(
2018
)
The DWORF micropeptide enhances contractility and prevents heart failure in a mouse model of dilated cardiomyopathy
.
Elife
7
,
1
23
46.
MacLennan
D.H.
and
Kranias
E.G.
(
2003
)
Phospholamban: A crucial regulator of cardiac contractility
.
Nat. Rev. Mol. Cell Biol.
4
,
566
577
[PubMed]
47.
Hughes
G.
,
Starling
A.P.
,
Sharma
R.P.
,
East
J.M.
and
Lee
A.G.
(
1996
)
An investigation of the mechanism of inhibition of the Ca2+-ATPase by phospholamban
.
Biochem. J.
318
,
973
979
[PubMed]
48.
Cantilina
T.
,
Sagara
Y.
,
Inesi
G.
and
Jones
L.R.
(
1993
)
Comparative studies of cardiac and skeletal sarcoplasmic reticulum ATPases. Effect of a phospholamban antibody on enzyme activation by Ca2+
.
J. Biol. Chem.
268
,
17018
17025
[PubMed]
49.
Vangheluwe
P.
,
Schuermans
M.
,
Zádor
E.
,
Waelkens
E.
,
Raeymaekers
L.
and
Wuytack
F.
(
2005
)
Sarcolipin and phospholamban mRNA and protein expression in cardiac and skeletal muscle of different species
.
Biochem. J.
389
,
151
159
[PubMed]
50.
Bhupathy
P.
,
Babu
G.J.
and
Periasamy
M.
(
2007
)
Sarcolipin and phospholamban as regulators of cardiac sarcoplasmic reticulum Ca2+ ATPase
.
J. Mol. Cell Cardiol.
42
,
903
911
[PubMed]
51.
Fujii
J.
,
Ueno
A.
,
Kitano
K.
,
Tanaka
S.
,
Kadoma
M.
and
Tada
M.
(
1987
)
Complete complementary DNA-derived amino acid sequence of canine cardiac phospholamban
.
J. Clin. Invest.
79
,
301
304
[PubMed]
52.
Tada
M.
,
Kirchberger
M.A.
and
Katz
A.M.
(
1975
)
Phosphorylation of a 22,000 dalton component of the cardiac sarcoplasmic reticulum by adenosine 3’,5’ monophosphate dependent protein kinase
.
J. Biol. Chem.
250
,
2640
2647
[PubMed]
53.
Toyoshima
C.
,
Asahi
M.
,
Sugita
Y.
,
Khanna
R.
,
Tsuda
T.
and
MacLennan
D.H.
(
2003
)
Modeling of the inhibitory interaction of phospholamban with the Ca2+ ATPase
.
Proc. Natl. Acad. Sci. U.S.A.
100
,
467
472
[PubMed]
54.
Lamberth
S.
,
Holger
S.
,
Muenchbach
M.
,
Vorherr
T.
,
Krebs
J.
,
Carafoli
E.
et al.
(
2000
)
NMR solution structure of Phospholamban
.
Helv. Chim. Acta
83
,
2141
2152
55.
Hutter
M.C.
,
Krebs
J.
,
Meiler
J.
,
Griesinger
C.
,
Carafoli
E.
and
Helms
V.
(
2002
)
A structural model of the complex formed by phospholamban and the calcium pump of sarcoplasmic reticulum obtained by molecular mechanics
.
Chem. Biol. Chem.
3
,
1200
1208
56.
Asahi
M.
,
Sugita
Y.
,
Kurzydlowski
K.
,
De Leon
S.
,
Tada
M.
,
Toyoshima
C.
et al.
(
2003
)
Sarcolipin regulates sarco(endo)plasmic reticulum Ca2+-ATPase (SERCA) by binding to transmembrane helices alone or in association with phospholamban
.
Proc. Natl. Acad. Sci. U.S.A.
100
,
5040
5045
[PubMed]
57.
Gorski
P.A.
,
Glaves
J.P.
,
Vangheluwe
P.
and
Young
H.S.
(
2013
)
Sarco(endo)plasmic reticulum calcium ATPase (SERCA) inhibition by sarcolipin is encoded in its luminal tail
.
J. Biol. Chem.
288
,
8456
8467
[PubMed]
58.
Mundiña Weilenmann
C.
,
Ferrero
P.
,
Said
M.
,
Vittone
L.
,
Kranias
E.G.
and
Mattiazzi
A.
(
2005
)
Role of phosphorylation of Thr17 residue of phospholamban in mechanical recovery during hypercapnic acidosis
.
Cardiovasc. Res.
66
,
114
122
[PubMed]
59.
Asahi
M.
,
McKenna
E.
,
Kurzydlowski
K.
,
Tada
M.
and
MacLennan
D.H.
(
2000
)
Physical interactions between phospholamban and sarco(endo)plasmic reticulum Ca2+-ATPases are dissociated by elevated Ca2+, but not by phospholamban phosphorylation, vanadate, or thapsigargin, and are enhanced by ATP
.
J. Biol. Chem.
275
,
15034
15038
[PubMed]
60.
Kranias
E.G.
and
Bers
D.M.
(
2007
)
Calcium and cardiomyopathies
.
Subcell. Biochem.
45
,
523
537
[PubMed]
61.
Wegener
A.D.
and
Jones
L.R.
(
1984
)
Phosphorylation-induced mobility shift in phospholamban in sodium dodecyl sulfate-polyacrylamide gels. Evidence for a protein structure consisting of multiple identical phosphorylatable subunits
.
J. Biol. Chem.
259
,
1834
1841
[PubMed]
62.
Wegener
A.D.
,
Simmerman
H.K.B.
,
Lindemann
J.P.
and
Jones
L.R.
(
1989
)
Phospholamban phosphorylation in intact ventricles. Phosphorylation of serine 16 and threonine 17 in response to β-adrenergic stimulation
.
J. Biol. Chem.
264
,
11468
11474
[PubMed]
63.
Brittsan
A.G.
and
Kranias
E.G.
(
2000
)
Phospholamban and cardiac contractile function
.
J. Mol. Cell Cardiol.
32
,
2131
2139
[PubMed]
64.
Kirchberger
M.
(
1972
)
Cyclic adenosine 3′,5′-monophosphate-dependent protein kinase stimulation of calcium uptake by canine cardiac microsomes
.
J. Mol. Cell Cardiol.
4
,
673
680
[PubMed]
65.
Wang
S.
,
Gopinath
T.
,
Larsen
E.K.
,
Weber
D.K.
,
Walker
C.
,
Uddigiri
V.R.
et al.
(
2021
)
Structural basis for sarcolipin’s regulation of muscle thermogenesis by the sarcoplasmic reticulum Ca2+-ATPase
.
Sci. Adv.
7
,
1
11
66.
Mascioni
A.
,
Karim
C.
,
Barany
G.
,
Thomas
D.D.
and
Veglia
G.
(
2002
)
Structure and orientation of sarcolipin in lipid environments
.
Biochemistry
41
,
475
482
[PubMed]
67.
Wawrzynow
A.
,
Theibert
J.L.
,
Murphy
C.
,
Jona
I.
,
Martonosi
A.
and
Collins
J.H.
(
1992
)
Sarcolipin, the “proteolipid” of skeletal muscle sarcoplasmic reticulum, is a unique, amphipathic, 31-residue peptide
.
Arch. Biochem. Biophys.
298
,
620
623
[PubMed]
68.
Winther
A.M.L.
,
Bublitz
M.
,
Karlsen
J.L.
,
Møller
J.V.
,
Hansen
J.B.
,
Nissen
P.
et al.
(
2013
)
The sarcolipin-bound calcium pump stabilizes calcium sites exposed to the cytoplasm
.
Nature
495
,
265
269
[PubMed]
69.
Babu
G.J.
,
Bhupathy
P.
,
Carnes
C.A.
,
Billman
G.E.
and
Periasamy
M.
(
2007
)
Differential expression of sarcolipin protein during muscle development and cardiac pathophysiology
.
J. Mol. Cell Cardiol.
43
,
215
222
[PubMed]
70.
Tupling
A.R.
,
Bombardier
E.
,
Gupta
S.C.
,
Hussain
D.
,
Vigna
C.
,
Bloemberg
D.
et al.
(
2011
)
Enhanced ca 2 transport and muscle relaxation in skeletal muscle from sarcolipin-null mice
.
Am. J. Physiol. Cell Physiol.
301
,
C841
C849
[PubMed]
71.
Chambers
P.J.
,
Juracic
E.S.
,
Fajardo
V.A.
and
Tupling
A.R.
(
2022
)
Role of SERCA and sarcolipin in adaptive muscle remodeling
.
Am. J. Physiol. Cell Physiol.
322
,
C382
C394
72.
Minamisawa
S.
,
Wang
Y.
,
Chen
J.
,
Ishikawa
Y.
,
Chien
K.R.
and
Matsuoka
R.
(
2003
)
Atrial chamber-specific expression of sarcolipin is regulated during development and hypertrophic remodeling
.
J. Biol. Chem.
278
,
9570
9575
[PubMed]
73.
Bal
N.C.
,
Maurya
S.K.
,
Sopariwala
D.H.
,
Sahoo
S.K.
,
Gupta
S.C.
,
Shaikh
S.A.
et al.
(
2012
)
Sarcolipin is a newly identified regulator of muscle-based thermogenesis in mammals
.
Nat. Med.
18
,
1575
1579
[PubMed]
74.
Sahoo
S.K.
,
Shaikh
S.A.
,
Sopariwala
D.H.
,
Bal
N.C.
and
Periasamy
M.
(
2013
)
Sarcolipin protein interaction with sarco(endo)plasmic reticulum Ca2+ ATPase (SERCA) is distinct from phospholamban protein, and only sarcolipin can promote uncoupling of the serca pump
.
J. Biol. Chem.
288
,
6881
6889
[PubMed]
75.
Bombardier
E.
,
Smith
I.C.
,
Vigna
C.
,
Fajardo
V.A.
and
Tupling
A.R.
(
2013
)
Ablation of sarcolipin decreases the energy requirements for Ca2+ transport by sarco(endo)plasmic reticulum Ca2+-ATPases in resting skeletal muscle
.
FEBS Lett.
587
,
1687
1692
[PubMed]
76.
Gamu
D.
,
Juracic
E.S.
,
Fajardo
V.A.
,
Rietze
B.A.
,
Tran
K.
,
Bombardier
E.
et al.
(
2019
)
Phospholamban deficiency does not alter skeletal muscle SERCA pumping efficiency or predispose mice to diet-induced obesity
.
Am. J. Physiol. Endocrinol. Metab.
316
,
E432
E442
[PubMed]
77.
Rowland
L.A.
,
Bal
N.C.
,
Kozak
L.P.
and
Periasamy
M.
(
2015
)
Uncoupling protein 1 and sarcolipin are required to maintain optimal thermogenesis, and loss of both systems compromises survival of mice under cold stress
.
J. Biol. Chem.
290
,
12282
12289
[PubMed]
78.
Rowland
L.A.
,
Maurya
S.K.
,
Bal
N.C.
,
Kozak
L.P.
and
Periasamy
M.
(
2016
)
Sarcolipin and uncoupling protein 1 play distinct roles in diet induced thermogenesis and do not compensate for one another
.
Obes. (Silver Spring)
24
,
1430
1433
79.
Meizoso-Huesca
A.
,
Pearce
L.
,
Barclay
C.J.
and
Launikonis
B.S.
(
2022
)
Ca2+ leak through ryanodine receptor 1 regulates thermogenesis in resting skeletal muscle
.
Proc. Natl. Acad. Sci.
119
,
1
8
80.
Fajardo
V.A.
,
Chambers
P.J.
,
Juracic
E.S.
,
Rietze
B.A.
,
Gamu
D.
,
Bellissimo
C.
et al.
(
2018
)
Sarcolipin deletion in mdx mice impairs calcineurin signalling and worsens dystrophic pathology
.
Hum. Mol. Genet.
27
,
4094
4102
[PubMed]
81.
Fajardo
V.A.
,
Rietze
B.A.
,
Chambers
P.J.
,
Bellissimo
C.
,
Bombardier
E.
,
Quadrilatero
J.
et al.
(
2017
)
Effects of sarcolipin deletion on skeletal muscle adaptive responses to functional overload and unload
.
Am. J. Physiol. Cell Physiol.
313
,
C154
C161
[PubMed]
82.
Fajardo
V.A.
,
Gamu
D.
,
Mitchell
A.
,
Bloemberg
D.
,
Bombardier
E.
,
Chambers
P.J.
et al.
(
2017
)
Sarcolipin deletion exacerbates soleus muscle atrophy and weakness in phospholamban overexpressing mice
.
PLoS ONE
12
,
1
17
83.
Maurya
S.K.
,
Herrera
J.L.
,
Sahoo
S.K.
,
Reis
F.C.
,
Vega
R.B.
,
Kelly
D.P.
et al.
(
2017
)
Sarcolipin signaling promotes mitochondrial biogenesis and oxidative metabolism in skeletal muscle
.
Cell Rep.
24
,
2919
2931
84.
Anderson
D.M.
,
Makarewich
C.A.
,
Anderson
K.M.
,
Shelton
J.M.
,
Bezprozvannaya
S.
,
Bassel-Duby
R.
et al.
(
2017
)
Widespread control of calcium signaling by a family of SERCA-inhibiting micropeptides
.
Sci Signal.
9
,
ra119
85.
Rathod
N.
,
Bak
J.J.
,
Primeau
J.O.
,
Fisher
M.E.
,
Espinoza-fonseca
L.M.
,
Lemieux
M.J.
et al.
(
2021
)
Nothing regular about the regulins: Distinct functional properties of SERCA transmembrane peptide regulatory subunits
.
Int. J. Mol. Sci.
22
,
8891
[PubMed]
86.
Lecarpentier
Y.
(
2007
)
Physiological role of free radicals in skeletal muscles
.
J. Appl. Physiol.
103
,
1917
1918
87.
Adachi
T.
,
Weisbrod
R.M.
,
Pimentel
D.R.
,
Ying
J.
,
Sharov
V.S.
,
Schöneich
C.
et al.
(
2004
)
S-glutathiolation by peroxynitrite activates SERCA during arterial relaxation by nitric oxide
.
Nat. Med.
10
,
1200
1207
[PubMed]
88.
Tupling
A.R.
,
Vigna
C.
,
Ford
R.J.
,
Tsuchiya
S.C.
,
Graham
D.A.
,
Denniss
S.G.
et al.
(
2007
)
Effects of buthionine sulfoximine treatment on diaphragm contractility and SR Ca2+ pump function in rats
.
J. Appl. Physiol.
103
,
1921
1928
89.
Lancel
S.
,
Zhang
J.
,
Evangelista
A.
,
Trucillo
M.P.
,
Tong
X.
,
Siwik
D.A.
et al.
(
2009
)
Nitroxyl activates SERCA in cardiac myocytes via glutathiolation of cysteine 674
.
Circ. Res.
104
,
720
723
[PubMed]
90.
Ottolia
M.
,
Torres
N.
,
Bridge
J.H.B.
,
Philipson
K.D.
and
Goldhaber
J.I.
(
2013
)
Na/Ca exchange and contraction of the heart
.
J. Mol. Cell Cardiol.
61
,
28
33
[PubMed]
91.
Scranton
K.
,
John
S.
,
Escobar
A.
,
Goldhaber
J.I.
and
Ottolia
M.
(
2020
)
Modulation of the cardiac Na+-Ca2+ exchanger by cytoplasmic protons: molecular mechanisms and physiological implications
.
Cell Calcium
87
,
1
21
92.
Bers
D.M.
(
2002
)
Cardiac Na/Ca exchange function in rabbit, mouse and man: what’s the difference?
J. Mol. Cell Cardiol.
34
,
369
373
[PubMed]
93.
Hilgemann
D.W.
,
Matsuoka
S.
,
Nagel
G.A.
and
Collins
A.
(
1992
)
Steady-state and dynamic properties of cardiac sodium-calcium exchange: sodium-dependent inactivation
.
J. Gen. Physiol.
100
,
905
932
[PubMed]
94.
Hilgemann
D.W.
,
Nicoll
D.A.
and
Philipson
K.D.
(
1991
)
Charge movement during Na+ translocation by native and cloned cardiac Na+/Ca2+ exchanger
.
Nature
352
,
715
718
[PubMed]
95.
Matsuoka
S.
and
Hilgemann
D.W.
(
1994
)
Inactivation of outward Na+‐Ca2+ exchange current in guinea‐pig ventricular myocytes
.
J. Physiol.
476
,
443
458
[PubMed]
96.
Shigekawa
M.
and
Iwamoto
T.
(
2001
)
Cardiac Na+-Ca2+ exchange
.
Circ. Res.
88
,
864
876
[PubMed]
97.
Matsuoka
S.
,
Nicoll
D.A.
,
He
Z.
and
Philipson
K.D.
(
1997
)
Regulation of the cardiac Na+-Ca2+ exchanger by the endogenous XIP region
.
J. Gen. Physiol.
109
,
273
286
[PubMed]
98.
Ottolia
M.
and
Philipson
K.D.
(
2013
)
NCX1: mechanism of transport
.
Adv. Exp. Med. Biol.
961
,
49
54
[PubMed]
99.
Scranton
K.
,
John
S.
,
Escobar
A.
,
Goldhaber
J.I.
and
Ottolia
M.
(
2020
)
Modulation of the cardiac Na+-Ca2+ exchanger by cytoplasmic protons: molecular mechanisms and physiological implications
.
Cell Calcium
87
,
102140
[PubMed]
100.
Hilge
M.
,
Aelen
J.
and
Vuister
G.W.
(
2006
)
Ca2+ regulation in the Na+/Ca2+ exchanger involves two markedly different Ca2+ sensors
.
Mol. Cell.
22
,
15
25
[PubMed]
101.
Hilge
M.
,
Aelen
J.
,
Foarce
A.
,
Perrakis
A.
and
Vuister
G.W.
(
2009
)
Ca2+ regulation in the Na+/Ca2+ exchanger features a dual electrostatic switch mechanism
.
Proc. Natl. Acad. Sci. U.S.A.
106
,
14333
14338
[PubMed]
102.
Doering
A.E.
and
Lederer
W.J.
(
1993
)
The mechanism by which cytoplasmic protons inhibit the sodium-calcium exchanger in guinea-pig heart cells
.
J. Physiol.
466
,
481
499
[PubMed]
103.
Doering
A.E.
and
Lederer
W.J.
(
1994
)
The action of Na+ as a cofactor in the inhibition by cytoplasmic protons of the cardiac Na(+)‐Ca2+ exchanger in the guinea‐pig
.
J. Physiol.
480
,
9
20
[PubMed]
104.
John
S.
,
Kim
B.
,
Olcese
R.
,
Goldhaber
J.I.
and
Ottolia
M.
(
2018
)
Molecular determinants of pH regulation in the cardiac Na+-Ca2+ exchanger
.
J. Gen. Physiol.
150
,
245
257
[PubMed]
105.
Philipson
P.D.
,
Bersohn
M.M.
and
Nishimoto
A.Y.
(
1982
)
Effects of pH on Na+-Ca2+ exchange in canine cardiac sarcolemmal vesicles
.
Circ. Res.
50
,
287
293
[PubMed]
106.
Brini
M.
and
Carafoli
E.
(
2011
)
The plasma membrane Ca2+ ATPase and the NCX cooperate in the regulation of cell calcium
.
Cold Spring Harb. Perspect. Biol.
3
,
1
15
[PubMed]
107.
Sokolow
S.
,
Manto
M.
,
Gailly
P.
,
Molgó
J.
,
Vandebrouck
C.
,
Vanderwinden
J.M.
et al.
(
2004
)
Impaired neuromuscular transmission and skeletal muscle fiber necrosis in mice lacking Na/Ca exchanger 3
.
J. Clin. Invest.
113
,
265
273
[PubMed]
108.
Garcia
M.C.
,
Diaz
A.F.
,
Godinez
R.
and
Sanchez
J.A.
(
1992
)
Effect of sodium deprivation on contraction and charge movement in frog skeletal muscle fibres
.
J. Muscle Res. Cell Motil.
13
,
354
365
[PubMed]
109.
Donoso
P.
and
Hidalgo
C.
(
1989
)
Sodium-calcium exchange in transverse tubules isolated from frog skeletal muscle
.
BBA - Biomembr.
978
,
8
16
110.
Balnave
C.D.
and
Allen
D.G.
(
1998
)
Evidence for Na+/Ca2+ exchange in intact single skeletal muscle fibers from the mouse
.
Am. J. Physiol. Physiol.
274
,
C940
C946
111.
Heizmann
C.W.
(
1984
)
Parvalbumin, and intracellular calcium-binding protein; distribution, properties and possible roles in mammalian cells
.
Experientia
40
,
910
921
[PubMed]
112.
Schwaller
B.
,
Dick
J.
,
Dhoot
G.
,
Carroll
S.
,
Vrbova
G.
,
Nicotera
P.
et al.
(
1999
)
Prolonged contraction-relaxation cycle of fast-twitch muscles in parvalbumin knockout mice
.
Am. J. Physiol. Cell Physiol.
276
,
395
403
113.
Ecob-Prince
M.S.
and
Leberer
E.
(
1989
)
Parvalbumin in mouse muscle in vivo and in vitro
.
Differentiation
40
,
10
16
[PubMed]
114.
Fohr
U.G.
,
Weber
B.R.
,
Muntener
M.
,
Staudenmann
W.
,
Hughes
G.J.
,
Frutiger
S.
et al.
(
1993
)
Human alpha and beta parvalbumins. Structure and tissue-specific expression
.
Eur. J. Biochem.
215
,
719
727
[PubMed]
115.
Kretsinger
R.H.
and
Nockolds
C.E.
(
1973
)
Carp muscle calcium-binding protein. II. Structure determination and general description
.
J. Biol. Chem.
248
,
3313
3326
[PubMed]
116.
Fuchtbauer
E.M.
,
Rowlerson
A.M.
,
Gotz
K.
,
Friedrich
G.
,
Mabuchi
K.
,
Gergely
J.
et al.
(
1991
)
Direct correlation of parvalbumin levels with myosin isoforms and succinate dehydrogenase activity on frozen sections of rodent muscle
.
J. Histochem. Cytochem.
39
,
355
361
[PubMed]
117.
Haiech
J.
,
Derancourt
J.
,
Pechère
J.F.
and
Demaille
J.G.
(
1979
)
Magnesium and calcium binding to parvalbumins: Evidence for differences between parvalbumins and an explanation of their relaxing function
.
Biochemistry
18
,
2752
2758
[PubMed]
118.
Hou
T.
,
Johnson
D.
and
Rall
J.A.
(
1991
)
Effect of temperature on relaxation rate and Ca2+, Mg2+ dissociation rates from parvalbumin of frog muscle fibres
.
J. Physiol.
499
,
399
410
119.
Müntener
M.
,
Käser
L.
,
Weber
J.
and
Berchtold
M.W.
(
1995
)
Increase of skeletal muscle relaxation speed by direct injection of parvalbumin cDNA
.
Proc. Natl. Acad. Sci. U.S.A.
92
,
6504
6508
[PubMed]
120.
Schmidt
U.
,
Zhu
X.
,
Lebeche
D.
,
Huq
F.
,
Guerrero
J.L.
and
Hajjar
R.J.
(
2005
)
In vivo gene transfer of parvalbumin improves diastolic function in aged rat hearts
.
Cardiovasc. Res.
66
,
318
323
[PubMed]
121.
Reggiani
C.
and
Marcucci
L.
(
2022
)
A controversial issue: can mitochondria modulate cytosolic calcium and contraction of skeletal muscle fibers?
J. Gen. Physiol.
154
,
1
15
122.
Lamboley
C.R.
,
Pearce
L.
,
Seng
C.
,
Meizoso-Huesca
A.
,
Singh
D.P.
,
Frankish
B.P.
et al.
(
2021
)
Ryanodine receptor leak triggers fiber Ca2+ redistribution to preserve force and elevate basal metabolism in skeletal muscle
.
Sci. Adv.
7
,
eabi7166
[PubMed]
123.
Butera
G.
,
Vecellio Reane
D.
,
Canato
M.
,
Pietrangelo
L.
,
Boncompagni
S.
,
Protasi
F.
et al.
(
2021
)
Parvalbumin affects skeletal muscle trophism through modulation of mitochondrial calcium uptake
.
Cell Rep.
35
,
109087
[PubMed]
124.
Rizzuto
R.
,
Pinton
P.
,
Carrington
W.
,
Fay
F.S.
,
Fogarty
K.E.
,
Lifshitz
L.M.
et al.
(
1998
)
Close contacts with the endoplasmic reticulum as determinants of mitochondrial Ca2+ responses
.
Science
280
,
1763
1766
[PubMed]
125.
Rossini
M.
and
Filadi
R.
(
2020
)
Sarcoplasmic reticulum-mitochondria kissing in cardiomyocytes: Ca2+, ATP, and undisclosed secrets
.
Front. Cell Dev. Biol.
8
,
532
[PubMed]
126.
Molenaar
J.P.
,
Verhoeven
J.I.
,
Rodenburg
R.J.
,
Kamsteeg
E.J.
,
Erasmus
C.E.
,
Vicart
S.
et al.
(
2020
)
Clinical, morphological and genetic characterization of Brody disease: An international study of 40 patients
.
Brain
143
,
452
466
[PubMed]
127.
Voermans
N.C.
,
Laan
A.E.
,
Oosterhof
A.
,
van Kuppevelt
T.H.
,
Drost
G.
,
Lammens
M.
et al.
(
2012
)
Brody syndrome: a clinically heterogeneous entity distinct from Brody disease. A review of literature and a cross-sectional clinical study in 17 patients
.
Neuromuscul. Disord.
22
,
944
954
[PubMed]
128.
Odermatt
A.
,
Taschner
P.E.M.
,
Scherer
S.W.
,
Beatty
B.
,
Khanna
V.K.
,
Cornblath
D.R.
et al.
(
1997
)
Characterization of the gene encoding human sarcolipin (SLN), a proteolipid associated with SERCA1: Absence of structural mutations in five patients with brody disease
.
Genomics
45
,
541
553
[PubMed]
129.
Odermatt
A.
,
Taschner
P.E.M.
,
Khanna
V.K.
,
Busch
H.F.M.
,
Karpati
G.
,
Jablecki
C.K.
et al.
(
1996
)
Mutations in the gene-encoding SERCA1, the fast-twitch skeletal muscle sarcoplasmic reticulum Ca2+ ATPase, are associated with Brody disease
.
Nat. Genet.
14
,
191
194
[PubMed]
130.
Sambuughin
N.
,
Zvaritch
E.
,
Kraeva
N.
,
Sizova
O.
,
Sivak
E.
,
Dickson
K.
et al.
(
2014
)
Exome analysis identifies brody myopathy in a family diagnosed with malignant hyperthermia susceptibility
.
Mol. Genet. Genomic Med.
2
,
472
483
[PubMed]
131.
Bergstrom
C.
,
Remz
M.
,
Khan
S.
and
McNutt
M.
(
2021
)
Brody myopathy presenting as recurrent Rhabdomyolysis
.
Am. J. Med.
134
,
e429
e430
[PubMed]
132.
Maclennan
D.H.
(
2000
)
Ca2+ signalling and muscle disease
.
Eur. J. Biochem.
267
,
5291
5297
[PubMed]
133.
Lee
E.H.
(
2010
)
Ca2+ channels and skeletal muscle diseases
.
Prog. Biophys. Mol. Biol.
103
,
35
43
[PubMed]
134.
Pan
Y.
,
Zvaritch
E.
,
Tupling
A.R.
,
Rice
W.J.
,
De Leon
S.
,
Rudnicki
M.
et al.
(
2003
)
Targeted disruption of the ATP2A1 gene encoding the sarco(endo)plasmic reticulum Ca2+ ATPase isoform 1 (SERCA1) impairs diaphragm function and is lethal in neonatal mice
.
J. Biol. Chem.
278
,
13367
13375
[PubMed]
135.
Aartsma-Rus
A.
,
Van Deutekom
J.C.T.
,
Fokkema
I.F.
,
Van Ommen
G.J.B.
and
Den Dunnen
J.T.
(
2006
)
Entries in the Leiden Duchenne muscular dystrophy mutation database: An overview of mutation types and paradoxical cases that confirm the reading-frame rule
.
Muscle Nerve
34
,
135
144
[PubMed]
136.
Allen
D.G.
,
Gervasio
O.L.
,
Yeung
E.W.
and
Whitehead
N.P.
(
2010
)
Calcium and the damage pathways in muscular dystrophy
.
Can. J. Physiol. Pharmacol.
88
,
83
91
[PubMed]
137.
Bulfield
G.
,
Siller
W.G.
,
Wight
P.A.L.
and
Moore
K.J.
(
1984
)
X chromosome-linked muscular dystrophy (mdx) in the mouse
.
Proc. Natl. Acad. Sci. U.S.A.
81
,
1189
1192
[PubMed]
138.
Cooper
S.T.
and
Head
S.I.
(
2015
)
Membrane injury and repair in the muscular dystrophies
.
Neuroscientist
21
,
653
668
[PubMed]
139.
Mercuri
E.
and
Muntoni
F.
(
2013
)
Muscular dystrophies
.
Lancet
381
,
845
860
[PubMed]
140.
Webster
C.
,
Silberstein
L.
,
Hays
A.P.
and
Blau
H.M.
(
1988
)
Fast muscle fibers are preferentially affected in Duchenne muscular dystrophy
.
Cell
52
,
503
513
[PubMed]
141.
McDonald
A.A.
,
Hebert
S.L.
,
Kunz
M.D.
,
Ralles
S.J.
and
McLoon
L.K.
(
2015
)
Disease course in mdx:Utrophin+/− mice: comparison of three mouse models of duchenne muscular dystrophy
.
Physiol. Rep.
3
,
e12391
[PubMed]
142.
DiMario
J.X.
,
Uzman
A.
and
Strohman
R.C.
(
1991
)
Fiber regeneration is not persistent in dystrophic (mdx) mouse skeletal muscle
.
Dev. Biol.
148
,
314
321
[PubMed]
143.
Carnwath
J.W.
and
Shotton
D.M.
(
1987
)
Muscular dystrophy in the mdx mouse: histopathology of the soleus and extensor digitorum longus muscles
.
J. Neurol. Sci.
80
,
39
54
[PubMed]
144.
Stedman
H.H.
,
Sweeney
H.L.
,
Shrager
J.B.
,
Maguire
H.C.
,
Panettieri
R.A.
,
Petrof
B.
et al.
(
1991
)
The mdx mouse diaphragm reproduces the degenerative changes of Duchenne muscular dystrophy
.
Nature
352
,
536
539
[PubMed]
145.
Whitehead
N.P.
,
Yeung
E.W.
and
Allen
D.G.
(
2006
)
Muscle damage in mdx (dystrophic) mice: Role of calcium and reactive oxygen species
.
Clin. Exp. Pharmacol. Physiol.
33
,
657
662
[PubMed]
146.
Hopf
F.W.
,
Turner
P.R.
and
Steinhardt
R.A.
(
2007
)
Calcium misregulation and the pathogenesis of muscular dystrophy
.
Subcell. Biochem.
45
,
429
464
[PubMed]
147.
Allen
D.G.
,
Gervasio
O.L.
,
Yeung
E.W.
and
Whitehead
N.P.
(
2010
)
Calcium and the damage pathways in muscular dystrophy
.
Can. J. Physiol. Pharmacol.
88
,
83
91
[PubMed]
148.
Murphy
R.M.
,
Verburg
E.
and
Lamb
G.D.
(
2006
)
Ca2+ activation of diffusible and bound pools of μ-calpain in rat skeletal muscle
.
J. Physiol.
576
,
595
612
[PubMed]
149.
Murphy
R.M.
,
Dutka
T.L.
,
Horvath
D.
,
Bell
J.R.
,
Delbridge
L.M.
and
Lamb
G.D.
(
2013
)
Ca2+-dependent proteolysis of junctophilin-1 and junctophilin-2 in skeletal and cardiac muscle
.
J. Physiol.
591
,
719
729
[PubMed]
150.
Gailly
P.
,
De Backer
F.
,
Van Schoor
M.
and
Gillis
J.M.
(
2007
)
In situ measurements of calpain activity in isolated muscle fibres from normal and dystrophin-lacking mdx mice
.
J. Physiol.
582
,
1261
1275
[PubMed]
151.
Grover
A.K.
,
Kwan
C.Y.
and
Samson
S.E.
(
2003
)
Effects of peroxynitrite on sarco/endoplasmic reticulum Ca2+ pump isoforms SERCA2b and SERCA3a
.
Am. J. Physiol. Cell Physiol.
285
,
1537
1543
152.
Fu
M.H.
and
Tupling
A.R.
(
2009
)
Protective effects of Hsp70 on the structure and function of SERCA2a expressed in HEK-293 cells during heat stress
.
Am. J. Physiol. Heart Circ. Physiol.
296
,
H1175
H1183
153.
Xu
K.Y.
,
Zweier
J.L.
and
Becker
L.C.
(
1997
)
Hydroxyl radical inhibits sarcoplasmic reticulum Ca2+-ATPase function by direct attack on the ATP binding site
.
Circ. Res.
80
,
76
81
[PubMed]
154.
Morris
T.E.
and
Sulakhe
P.V.
(
1997
)
Sarcoplasmic reticulum Ca2+-pump dysfunction in rat cardiomyocytes briefly exposed to hydoxyl radicals
.
Free Radic. Biol. Med.
22
,
37
47
[PubMed]
155.
Viner
R.I.
,
Krainev
A.G.
,
Williams
T.D.
,
Schöneich
C.
and
Bigelow
D.J.
(
1997
)
Identification of oxidation-sensitive peptides within the cytoplasmic domain of the sarcoplasmic reticulum Ca2+-ATPase
.
Biochemistry
36
,
7706
7716
[PubMed]
156.
Goonasekera
S.A.
,
Lam
C.K.
,
Millay
D.P.
,
Sargent
M.A.
,
Hajjar
R.J.
,
Kranias
E.G.
et al.
(
2011
)
Mitigation of muscular dystrophy in mice by SERCA overexpression in skeletal muscle
.
J. Clin. Invest.
121
,
1044
1052
[PubMed]
157.
Nogami
K.
,
Maruyama
Y.
,
Sakai-Takemura
F.
,
Motohashi
N.
,
Elhussieny
A.
,
Imamura
M.
et al.
(
2021
)
Pharmacological activation of SERCA ameliorates dystrophic phenotypes in dystrophin-deficient mdx mice
.
Hum. Mol. Genet.
30
,
1006
1019
[PubMed]
158.
Gehrig
S.M.
,
Van Der Poel
C.
,
Sayer
T.A.
,
Schertzer
J.D.
,
Henstridge
D.C.
,
Church
J.E.
et al.
(
2012
)
Hsp72 preserves muscle function and slows progression of severe muscular dystrophy
.
Nature
484
,
394
398
[PubMed]
159.
Tupling
A.R.
,
Gramolini
A.O.
,
Duhamel
T.A.
,
Kondo
H.
,
Asahi
M.
,
Tsuchiya
S.C.
et al.
(
2004
)
HSP70 binds to the fast-twitch skeletal muscle sarco(endo)plasmic reticulum Ca2+-ATPase (SERCA1a) and prevents thermal inactivation
.
J. Biol. Chem.
279
,
52382
52389
[PubMed]
160.
Jungbluth
H.
,
Wallgren-Pettersson
C.
and
Laporte
J.
(
2008
)
Centronuclear (myotubular) myopathy
.
Orphanet. J. Rare Dis.
3
,
1
13
[PubMed]
161.
Blondeau
F.
,
Laporte
J.
,
Bodin
S.
,
Superti-Furga
G.
,
Payrastre
B.
and
Mandel
J.L.
(
2000
)
Myotubularin, a phosphatase deficient in myotubular myopathy, acts on phosphatidylinositol 3-kinase and phosphatidylinositol 3-phosphate pathway
.
Hum. Mol. Genet.
9
,
2223
2229
[PubMed]
162.
Cowling
B.S.
,
Toussaint
A.
,
Muller
J.
and
Laporte
J.
(
2012
)
Defective membrane remodeling in neuromuscular diseases: Insights from animal models
.
PLos Genet.
8
,
e1002595
[PubMed]
163.
Jungbluth
H.
and
Gautel
M.
(
2014
)
Pathogenic mechanisms in centronuclear myopathies
.
Front. Aging Neurosci.
6
,
1
11
[PubMed]
164.
Cowling
B.S.
,
Chevremont
T.
,
Prokic
I.
,
Kretz
C.
,
Ferry
A.
,
Coirault
C.
et al.
(
2014
)
Reducing dynamin 2 expression rescues X-linked centronuclear myopathy
.
J. Clin. Invest.
124
,
1350
1363
[PubMed]
165.
Cowling
B.S.
,
Prokic
I.
,
Tasfaout
H.
,
Rabai
A.
,
Humbert
F.
,
Rinaldi
B.
et al.
(
2017
)
Amphiphysin (BIN1) negatively regulates dynamin 2 for normal muscle maturation
.
J. Clin. Invest.
127
,
4477
4487
[PubMed]
166.
Praefcke
G.J.K.
and
McMahon
H.T.
(
2004
)
The dynamin superfamily: universal membrane tubulation and fission molecules?
Nat. Rev. Mol. Cell Biol.
5
,
133
147
[PubMed]
167.
Shen
J.
,
Yu
W.
,
Brotto
M.
,
Scherman
J.A.
,
Guo
C.
,
Stoddard
C.
et al.
(
2009
)
Deficiency of MIP phosphatase induces a muscle disorder by disrupting Ca2+ homeostasis
.
Nat. Cell Biol.
11
,
769
776
[PubMed]
168.
Fraysse
B.
,
Guicheneu
P.
and
Bitoun
M.
(
2016
)
Calcium homeostasis alterations in a mouse model of the dynamin 2-related centronuclear myopathy
.
Biol. Open.
5
,
1691
1696
[PubMed]
169.
Ceyhan-Birsoy
O.
,
Agrawal
P.B.
,
Hidalgo
C.
,
Schmitz-Abe
K.
,
Dechene
E.T.
,
Swanson
L.C.
et al.
(
2013
)
Recessive truncating titin gene, TTN, mutations presenting as centronuclear myopathy
.
Neurology
81
,
1205
1214
[PubMed]
170.
Wilmshurst
J.M.
,
Lillis
S.
,
Zhou
H.
,
Pillay
K.
,
Henderson
H.
,
Kress
W.
et al.
(
2010
)
RYR1 mutations are a common cause of congenital myopathies with central nuclei
.
Ann. Neurol.
68
,
717
726
[PubMed]
171.
Kramerova
I.
,
Kudryashova
E.
,
Wu
B.
,
Ottenheijm
C.
,
Granzier
H.
and
Spencer
M.J.
(
2008
)
Novel role of calpain-3 in the triad-associated protein complex regulating calcium release in skeletal muscle
.
Hum. Mol. Genet.
17
,
3271
3280
[PubMed]
172.
Song
Q.
,
Young
K.B.
,
Chu
G.
,
Gulick
J.
,
Gerst
M.
,
Grupp
I.L.
et al.
(
2004
)
Overexpression of phospholamban in slow‐twitch skeletal muscle is associated with depressed contractile function and muscle remodeling
.
FASEB J.
18
,
1
19
[PubMed]
173.
Fajardo
V.A.
,
Bombardier
E.
,
McMillan
E.
,
Tran
K.
,
Wadsworth
B.J.
,
Gamu
D.
et al.
(
2015
)
Phospholamban overexpression in mice causes a centronuclear myopathy-like phenotype
.
DMM Dis. Model Mech.
8
,
999
1009
174.
Schwinger
R.H.G.
(
2021
)
Pathophysiology of heart failure
.
Cardiovasc. Diagn. Ther.
11
,
263
276
[PubMed]
175.
Mosterd
A.
and
Hoes
A.W.
(
2007
)
Clinical epidemiology of heart failure
.
Heart
93
,
1137
1146
[PubMed]
176.
McMurray JJ
V.
and
Pfeffer
M.A.
(
2005
)
Heart failure
.
Lancet
365
,
1877
1889
[PubMed]
177.
Piacentino
V.
,
Weber
C.R.
,
Chen
X.
,
Weisser-Thomas
J.
,
Margulies
K.B.
,
Bers
D.M.
et al.
(
2003
)
Cellular basis of abnormal calcium transients of failing human ventricular myocytes
.
Circ. Res.
92
,
651
658
[PubMed]
178.
Hobai
I.A.
and
O'Rourke
B.
(
2001
)
Decreased sarcoplasmic reticulum calcium content is responsible for defective excitation-contraction coupling in canine heart failure
.
Circulation
103
,
1577
1584
[PubMed]
179.
Jiang
M.T.
,
Lokuta
A.J.
,
Farrell
E.F.
,
Wolff
M.R.
,
Haworth
R.A.
and
Valdivia
H.H.
(
2002
)
Abnormal Ca2+ release, but normal ryanodine receptors, in canine and human heart failure
.
Circ. Res.
91
,
1015
1022
[PubMed]
180.
Lindner
M.
,
Erdmann
E.
and
Beuckelmann
D.J.
(
1998
)
Calcium content of the sarcoplasmic reticulum in isolated ventricular myocytes from patients with terminal heart failure
.
J. Mol. Cell Cardiol.
30
,
743
749
[PubMed]
181.
Schmidt
U.
,
Hajjar
R.J.
,
Helm
P.A.
,
Kim
C.S.
,
Doye
A.A.
and
Gwathmey
J.K.
(
1998
)
Contribution of abnormal sarcoplasmic reticulum ATPase activity to systolic and diastolic dysfunction in human heart failure
.
J. Mol. Cell Cardiol.
30
,
1929
1937
[PubMed]
182.
Janssen
P.M.L.
and
Periasamy
M.
(
2007
)
Determinants of frequency-dependent contraction and relaxation of mammalian myocardium
.
J. Mol. Cell Cardiol.
43
,
523
531
[PubMed]
183.
Endoh
M.
(
2004
)
Force-frequency relationship in intact mammalian ventricular myocardium: physiological and pathophysiological relevance
.
Eur. J. Pharmacol.
500
,
73
86
[PubMed]
184.
Palomeque
J.
,
Vila Petroff
M.G.
and
Mattiazzi
A.
(
2004
)
Pacing staircase phenomenon in the heart: From Bodwitch to the XXI century
.
Hear Lung. Circ.
13
,
410
420
185.
Hasenfuss
G.
,
Reinecke
H.
,
Studer
R.
,
Meyer
M.
,
Pieske
B.
,
Holtz
J.
et al.
(
1994
)
Relation between myocardial function and expression of sarcoplasmic reticulum Ca2+-ATPase in failing and nonfailing human myocardium
.
Circ. Res.
75
,
434
442
[PubMed]
186.
Pieske
B.
,
Maier
L.S.
,
Piacentino
V.
,
Weisser
J.
,
Hasenfuss
G.
and
Houser
S.
(
2002
)
Rate dependence of [Na+]i and contractility in nonfailing and failing human myocardium
.
Circulation
106
,
447
453
[PubMed]
187.
Beuckelmann
D.J.
,
Näbauer
M.
and
Erdmann
E.
(
1992
)
Intracellular calcium handling in isolated ventricular myocytes from patients with terminal heart failure
.
Circulation
85
,
1046
1055
[PubMed]
188.
Dries
E.
,
Santiago
D.J.
,
Gilbert
G.
,
Lenaerts
I.
,
Vandenberk
B.
,
Nagaraju
C.K.
et al.
(
2018
)
Hyperactive ryanodine receptors in human heart failure and ischaemic cardiomyopathy reside outside of couplons
.
Cardiovasc. Res.
114
,
1512
1524
[PubMed]
189.
Gómez
A.M.
,
Guatimosim
S.
,
Dilly
K.W.
,
Vassort
G.
and
Lederer WJ
B.
(
2001
)
Heart failure after myocardial infarction
.
Circulation
104
,
688
693
[PubMed]
190.
Kiss
E.
,
Ball
N.A.
,
Kranias
E.G.
and
Walsh
R.A.
(
1995
)
Differential changes in cardiac phospholamban and sarcoplasmic reticular Ca2+ -ATPase protein levels
.
Circ. Res.
77
,
759
764
[PubMed]
191.
Schwinger
R.H.G.
,
Münch
G.
,
Bölck
B.
,
Karczewski
P.
,
Krause
E.G.
and
Erdmann
E.
(
1999
)
Reduced Ca2+-sensitivity of SERCA 2a in failing human myocardium due to reduced serin-16 phospholamban phoshorylation
.
J. Mol. Cell Cardiol.
31
,
479
491
[PubMed]
192.
O'Rourke
B.
,
Kass
D.A.
,
Tomaselli
G.F.
,
Kääb
S.
,
Tunin
R.
and
Marbán
E.
(
1999
)
Mechanisms of altered excitation-contraction coupling in canine tachycardia-induced heart failure, I: experimental studies
.
Circ. Res.
84
,
562
570
[PubMed]
193.
Currie
S.
and
Smith
G.L.
(
1999
)
Enhanced phosphorylation of phospholamban and downregulation of sarco/endoplasmic reticulum Ca2+ ATPase type 2 (SERCA 2) in cardiac sarcoplasmic reticulum from rabbits with heart failure
.
Cardiovasc. Res.
41
,
135
146
[PubMed]
194.
Vittone
L.
,
Mundina-Weilenmann
C.
and
Mattiazzi
A.
(
2008
)
Phospholamban phosphorylation by CaMKII under pathophysiological conditions
.
Front. Biosci.
13
,
5988
6005
[PubMed]
195.
Mattiazzi
A.
,
Mundiña-Weilenmann
C.
,
Guoxiang
C.
,
Vittone
L.
and
Kranias
E.
(
2005
)
Role of phospholamban phosphorylation on Thr17 in cardiac physiological and pathological conditions
.
Cardiovasc. Res.
68
,
366
375
[PubMed]
196.
Mattiazzi
A.
and
Kranias
E.G.
(
2014
)
The role of CaMKII regulation of phospholamban activity in heart disease
.
Front. Pharmacol.
5
,
1
11
[PubMed]
197.
Maier
L.S.
(
2005
)
CaMKIIδ overexpression in hypertrophy and heart failure: Cellular consequences for excitation-contraction coupling
.
Brazilian J. Med. Biol. Res.
38
,
1293
1302
198.
Gupta
R.C.
,
Mishra
S.
,
Rastogi
S.
,
Imai
M.
,
Habib
O.
and
Sabbah
H.N.
(
2003
)
Cardiac SR-coupled PP1 activity and expression are increased and inhibitor I protein expression is decreased in failing hearts
.
Am. J. Physiol. Heart Circ. Physiol.
285
,
2373
2381
199.
Nicolaou
P.
,
Hajjar
R.J.
and
Kranias
E.G.
(
2009
)
Role of protein phosphatase-1 inhibitor-1 in cardiac physiology and pathophysiology
.
J. Mol. Cell Cardiol.
47
,
365
371
[PubMed]
200.
Xu
L.
,
Chen
J.
,
Li
X.Y.
,
Ren
S.
,
Huang
C.X.
,
Wu
G.
et al.
(
2012
)
Analysis of Na+/Ca2+ exchanger (NCX) function and current in murine cardiac myocytes during heart failure
.
Mol. Biol. Rep.
39
,
3847
3852
[PubMed]
201.
Pogwizd
S.M.
,
Qi
M.
,
Yuan
W.
,
Samarel
A.M.
and
Bers
D.M.
(
1999
)
Upregulation of Na+/Ca2+ exchanger expression and-function in an arrhythmogenic rabbit model of heart failure
.
Circ. Res.
85
,
1009
1019
[PubMed]
202.
Ke
H.Y.
,
Yang
H.Y.
,
Francis
A.J.
,
Collins
T.P.
,
Surendran
H.
,
Alvarez-Laviada
A.
et al.
(
2020
)
Changes in cellular Ca2+ and Na+ regulation during the progression towards heart failure in the guinea pig
.
J. Physiol.
598
,
1339
1359
[PubMed]
203.
Despa
S.
,
Islam
M.A.
,
Weber
C.R.
,
Pogwizd
S.M.
and
Bers
D.M.
(
2002
)
Intracellular Na+ concentration is elevated in heart failure but Na/K pump function is unchanged
.
Circulation
105
,
2543
2548
[PubMed]
204.
Baartscheer
A.
,
Schumacher
C.A.
,
Van Borren
M.M.G.J.
,
Belterman
C.N.W.
,
Coronel
R.
and
Fiolet
J.W.T.
(
2003
)
Increased Na+/H+-exchange activity is the cause of increased [Na+]i and underlies disturbed calcium handling in the rabbit pressure and volume overload heart failure model
.
Cardiovasc. Res.
57
,
1015
1024
[PubMed]
205.
Pogwizd
S.M.
,
Schlotthauer
K.
,
Li
L.
,
Yuan
W.
and
Bers
D.M.
(
2001
)
Arrhythmogenesis and contractile dysfunction in heart failure
.
Circ. Res.
88
,
1159
1167
[PubMed]
206.
Bueno
C.R.
,
Ferreira
J.C.B.
,
Pereira
M.G.
,
Bacurau
A.V.N.
and
Brum
P.C.
(
2010
)
Aerobic exercise training improves skeletal muscle function and Ca 2+ handling-related protein expression in sympathetic hyperactivity-induced heart failure
.
J. Appl. Physiol.
109
,
702
709
207.
Shah
K.R.
,
Ganguly
P.K.
,
Netticadan
T.
,
Arneja
A.S.
and
Dhalla
N.S.
(
2004
)
Changes in skeletal muscle SR Ca2+ pump in congestive heart failure due to myocardial infarction are prevented by angiotensin II blockade
.
Can. J. Physiol. Pharmacol.
82
,
438
447
[PubMed]
208.
Perreault
C.L.
,
Gonzalez-Serratos
H.
,
Litwin
S.E.
,
Sun
X.
,
Franzini-Armstrong
C.
and
Morgan
J.P.
(
1993
)
Alterations in contractility and intracellular Ca2+ transients in isolated bundles of skeletal muscle fibers from rats with chronic heart failure
.
Circ. Res.
73
,
405
412
[PubMed]
209.
Lunde
P.K.
,
Dahlstedt
A.J.
,
Bruton
J.D.
,
Lännergren
J.
,
Thorén
P.
,
Sejersted
O.M.
et al.
(
2001
)
Contraction and intracellular Ca2+ handling in isolated skeletal muscle of rats with congestive heart failure
.
Circ. Res.
88
,
1299
1305
[PubMed]
210.
Harrington
D.
,
Anker
S.D.
,
Chua
T.P.
,
Webb-Peploe
K.M.
,
Ponikowski
P.P.
,
Poole-Wilson
P.A.
et al.
(
1997
)
Skeletal muscle function and its relation to exercise tolerance in chronic heart failure
.
J. Am. Coll. Cardiol.
30
,
1758
1764
[PubMed]
211.
Middlekauff
H.R.
,
Vigna
C.
,
Verity
M.A.
,
Fonarow
G.C.
,
Horwich
T.B.
,
Hamilton
M.A.
et al.
(
2012
)
Abnormalities of calcium handling proteins in skeletal muscle mirror those of the heart in humans with heart failure : a shared mechanism?
J. Card. Fail.
18
,
724
733
[PubMed]
212.
Maier
L.S.
,
Zhang
T.
,
Chen
L.
,
DeSantiago
J.
,
Brown
J.H.
and
Bers
D.M.
(
2003
)
Transgenic CaMKIIδc overexpression uniquely alters cardiac myocyte Ca2+ handling: Reduced SR Ca2+ load and activated SR Ca2+ release
.
Circ. Res.
92
,
904
911
[PubMed]
213.
Maier
L.S.
(
2009
)
Role of CaMKII for signaling and regulation in the heart
.
Front. Biosci.
14
,
486
496
214.
Viner
R.I.
,
Williams
T.D.
and
Schöneich
C.
(
1999
)
Peroxynitrite modification of protein thiols: oxidation, nitrosylation, and S-glutathiolation of functionally important cysteine residue(s) in the sarcoplasmic reticulum Ca-ATPase
.
Biochemistry
38
,
12408
12415
[PubMed]
215.
Gutiérrez-Martín
Y.
,
Martín-Romero
F.J.
,
Iñesta-Vaquera
F.A.
,
Gutiérrez-Merino
C.
and
Henao
F.
(
2004
)
Modulation of sarcoplasmic reticulum Ca2+-ATPase by chronic and acute exposure to peroxynitrite
.
Eur. J. Biochem.
271
,
2647
2657
[PubMed]
216.
Favero
T.G.
,
Colter
D.
,
Hooper
P.F.
and
Abramson
J.J.
(
1998
)
Hypochlorous acid inhibits Ca2+ -ATPase from skeletal muscle sarcoplasmic reticulum
.
J. Appl. Physiol.
84
,
425
430
217.
Tong
X.
,
Ying
J.
,
Pimentel
D.R.
,
Trucillo
M.
,
Adachi
T.
and
Cohen
R.A.
(
2008
)
High glucose oxidizes SERCA cysteine-674 and prevents inhibition by nitric oxide of smooth muscle cell migration
.
J. Mol. Cell Cardio.
44
,
361
369
,
218.
Ying
J.
,
Sharov
V.
,
Xu
S.
,
Jiang
B.
,
Gerrity
R.
,
Schöneich
C.
et al.
(
2008
)
Cysteine-674 oxidation and degradation of sarcoplasmic reticulum Ca2+ ATPase in diabetic pig aorta
.
Free Radic. Biol. Med.
45
,
756
762
,
[PubMed]
219.
Fugere
N.A.
,
Ferrington
D.A.
and
Thompson
L.V.
(
2006
)
Protein Nitration With Aging in the Rat Semimembranosus and Soleus Muscles
.
J. Gerontol. Ser. A Biol. Sci. Med. Sci.
61
,
806
812
220.
Smith
I.C.
,
Vigna
C.
,
Levy
A.S.
,
Denniss
S.G.
,
Rush
J.W.E.
and
Tupling
A.R.
(
2015
)
The effects of buthionine sulfoximine treatment on diaphragm contractility and SERCA pump function in adult and middle aged rats
.
Physiol. Rep.
3
,
e12547
[PubMed]
221.
Fu
M.
,
Bombardier
E.
,
Gamu
D.
and
Tupling
A.R.
(
2020
)
Phospholamban and sarcolipin prevent thermal inactivation of sarco(endo)plasmic reticulum Ca2+-ATPases
.
Biochem. J.
477
,
4281
4294
,
[PubMed]
This is an open access article published by Portland Press Limited on behalf of the Biochemical Society and distributed under the Creative Commons Attribution License 4.0 (CC BY).