Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2014 Dec 1.
Published in final edited form as: Biochim Biophys Acta. 2013 Jun 14;1832(12):10.1016/j.bbadis.2013.06.008. doi: 10.1016/j.bbadis.2013.06.008

Alterations in Ryanodine Receptors and Related Proteins in Heart Failure

Sameer Ather 1,2,*, Jonathan L Respress 1,*, Na Li 1,*, Xander HT Wehrens 1,2
PMCID: PMC3800473  NIHMSID: NIHMS493689  PMID: 23770282

Abstract

Sarcoplasmic reticulum (SR) Ca2+ release plays an essential role in mediating cardiac myocyte contraction. Depolarization of the plasma membrane results in influx of Ca2+ through L-type Ca2+ channels (LTCCs) that in turn triggers efflux of Ca2+ from the SR through ryanodine receptor type-2 channels (RyR2). This process known as Ca2+-induced Ca2+release (CICR) occurs within the dyadic region, where the adjacent transverse (T)-tubules and SR membranes allow RyR2 clusters to release SR Ca2+ following Ca2+ influx through adjacent LTCCs. SR Ca2+ released during systole binds to troponin-C and initiates actin-myosin cross-bridging, leading to muscle contraction. During diastole, the cytosolic Ca2+ concentration is restored by the resequestration of Ca2+ into the SR by SR/ER Ca2+-ATPase (SERCA2a) and by the extrusion of Ca2+ via the Na+/Ca2+-exchanger (NCX1). This whole process, entitled excitation-contraction (EC) coupling, is highly coordinated and determines the force of contraction, providing a link between the electrical and mechanical activities of cardiac muscle. In response to heart failure (HF), the heart undergoes maladaptive changes that result in depressed intracellular Ca2+ cycling and decreased SR Ca2+ concentrations. As a result, the amplitude of CICR is reduced resulting in less force production during EC coupling. In this review, we discuss the specific proteins that alter the regulation of Ca2+ during HF. In particular, we will focus on defects in RyR2-mediated SR Ca2+ release.

Introduction

Excitation-contraction (EC) coupling represents a signaling cascade that allows the cardiac action potential to trigger myocyte contraction.[1] Depolarization of the cardiomyocyte membrane leads to activation of voltage-gated L-type Ca2+ channels (LTCC) located in plasma membrane invaginations known as transverse (T)-tubules. Ca2+ influx through LTCC triggers a much greater release of Ca2+ from the sarcoplasmic reticulum (SR) via the ryanodine receptor type-2 (RyR2), a process known as Ca2+-induced Ca2+ release (CICR).[2] This abrupt elevation in the cytosolic Ca2+ concentration allows Ca2+ to bind to troponin C, triggering the release of troponin I (TnI) from the myofilament, initiating contraction of cardiomyocytes. During muscle relaxation, cytosolic Ca2+ is actively pumped back into the SR via sarco/endoplasmic reticulum Ca2+−ATPase 2a (SERCA2a) and extruded from cardiomyocytes via the Na+/Ca2+ exchanger (NCX).[3, 4]

Heart failure (HF) is a major cause of morbidity and mortality in Western countries.[5] It is a physiological state in which cardiac output does not meet the demands of the body. Classically, systolic HF is caused by a reduction in ejection fraction and has higher mortality,, compared with HF with preserved ejection fraction.[6] In this review, we will focus on role of alterations in Ca2+ handling proteins in the pathogenesis of HF with reduced ejection fraction. In response to HF, the heart undergoes maladaptive changes that ultimately result in depressed intracellular Ca2+ cycling and decreased SR Ca2+ concentrations. Thus, subsequent action potentials lead to depressed CICR, releasing less Ca2+ and producing less force during EC coupling. Depressed CICR can result from a reduction in (1) trigger Ca2+ current through LTCC, (2) reuptake of Ca2+ into the SR, and/or (3) SR Ca2+ release through RyR2. Over time, these alterations contribute to reduced SR Ca2+ loading, and as such, interfere with the frequency-dependent enhancement of myocyte contractility. In this review, we discussed specific alterations in Ca2+ handling proteins associated with the pathogenesis of HF, with a focus on the complex pattern of alterations of RyR2 post-translational regulation during HF.

Regulation of ryanodine receptors during excitation-contraction coupling

RyR2 represents the primary channel mediating the intracellular Ca2+ release that triggers cardiomyocyte contraction during EC coupling.[7] This homotetrameric transmembrane protein with a molecular weight of 565-kDa per monomer is located on the SR membrane.[8] The amplitude of Ca2+ release via RyR2 is strongly modulated by second messengers (Ca2+, Mg2+, cAMP) and various intracellular proteins. One major regulatory subunit interacting with RyR2 is the 12.6-kDa cytosolic FK506-binding protein (FKBP12.6), also known as calstabin2.[9, 10] FKBP12.6, a peptidyl-prolyl cis-trans isomerase, tightly associates with RyR2, stabilizing its closed conformational state and facilitating channel closure.[11, 12] Other accessory proteins that bind to RyR2 and inhibit channel open probability include calmodulin and sorcin.[13, 14] Moreover, our studies have recently revealed that a structural protein junctophilin-2 (JPH2) also binds to RyR2 and inhibits its activity levels.[15]

Altered RyR2 regulation in HF

In addition to regulation by accessory and structural proteins, posttranslational modifications of RyR2, such as oxidation,[16, 17] S-nitrosylation,[1820] and phosphorylation[2123] have also been shown to regulate channel activity. The open probability of RyR2 is strongly regulated by protein kinases that phosphorylate distinct residues on the channel, and phosphatases that dephosphorylate RyR2 channel subunits. Several such enzymes are associated with the RyR2 channel macromolecular complex, including protein kinases A (PKA), Ca2+/CaM-dependent protein kinase II (CaMKII), and phosphatases PP1 and PP2A allowing for rapid and selective regulation of the channel activity.[21, 22, 24, 25] Several studies have been conducted to determine the relative importance of each of these regulatory proteins, especially PKA and CaMKII, in modulating RyR2 activity in the failing heart.[2631] Results of these studies are discordant regarding the relative importance of CaMKII versus PKA in modulating RyR2 activity. Our recent work suggests that this apparent discordance may be due to different pathophysiological mechanisms in ischemic versus non-ischemic HF.[ref 28] In the sections below, we have discussed in detail the current state of knowledge.

Regulation of RyR2 by PKA Phosphorylation in HF

One of the first reported changes in RyR2 regulation associated with HF pathogenesis was altered PKA phosphorylation.[21] This discovery by the Marks lab has led to many related studies on phosphorylation-dependent regulation of RyR2 in healthy and diseased hearts. Many, but not all, studies have demonstrated increased phosphorylation of the main PKA phosphorylation site on RyR2, serine 2808 (S2808) in HF.[21, 26, 28, 3236] Chronically increased S2808 phosphorylation on RyR2 may promote diastolic SR Ca2+ leak that, in turn, depletes SR Ca2+ stores and reduces EC coupling.[37, 38]

Recently, some new insights were obtained in genetically altered mice in which the S2808 site was either constitutively activated (S2808D) or genetically inactivated (S2808A). Shan et al. demonstrated that S2808D knock-in mice developed spontaneous HF and exhibited increased mortality after experimental myocardial infarction (MI), as a result of severe SR Ca2+ leak.[29] Conversely, it was shown that S2808A knock-in mice were relatively protected from the development of HF following experimental MI.[35] In contrast, another study showed that, although MI increased phosphorylation of RyR2 at the S2808 site, genetic inhibition of S2808 phosphorylation in S2808A mice did not protect mice from developing ischemic HF.[26] The reasons for these discrepant findings remain unknown at this time, although differences in the time-points chosen for the assessment of cardiac function after MI and differences in the size of the infarction may have contributed.

In contrast to the above findings, a recent study by Benkusky et al.[36] demonstrated that S2808A knock-in mice were not protected from non-ischemic HF due to pressure overload. Consistent with this finding, Respress et al.[28] showed that there was only a non-significant increase in S2808 phosphorylation on RyR2 in WT mice subjected to non-ischemic HF. In this particular model of non-ischemic HF caused by transverse aortic constriction, no significant changes were observed in the phosphorylation status of S2808 on RyR2, as well as the PKA phosphorylation site S16 on phospholamban (PLN). Consistent with these findings in animals, patients with non-ischemic dilated cardiomyopathy (DCM) also did not show increases in the phosphorylation status of S2808 on RyR2.[28] Thus, it seems that the potential involvement of S2808 phosphorylation on RyR2 in HF development may depend on the disease etiology, with current evidence suggesting that enhanced PKA-dependent phosphorylation of S2808 on RyR2 may only play a role in the development of ischemic HF.

Regulation of RyR2 by CaMKII in HF

In addition to PKA, RyR2 channels are also phosphorylated by CaMKII, which primarily phosphorylates a nearby site (serine 2814) on RyR2.[22, 27] Hoch et al. were among the first group of scientists to demonstrate increased expression of the CaMKII-δC isoform in cardiac muscle of patients with HF.[39] Following this report, Zhang et al. showed that cardiac-specific overexpression of CaMKII-δC in mice causes abnormal Ca2+ handling and development of HF.[40] Conversely, Zhang et al. showed that overexpression of AC3-I, a peptide inhibitor of CaMKII, delayed the onset of HF in mice following MI.[41] Various studies have demonstrated that increased RyR2 phosphorylation by CaMKII upregulation contributes to enhanced SR Ca2+ leak in HF.[37, 38, 42] The role of CaMKII phosphorylation of RyR2 in HF is further supported by the findings that pressure overload increases CaMKII phosphorylation of RyR2, enhances diastolic SR Ca2+ release events, and promotes HF in mice, and the fact that this phenotype can be reversed by genetic inhibition of CaMKII activity in CaMKII-δ deficient mice.[43, 44] These studies suggest a dominant role of abnormal CaMKII regulation of RyR2 in HF.

Our recent studies provided additional insights into the importance of the S2814 phosphorylation site on RyR2 as a downstream target of CaMKII in HF. Respress et al.[28] demonstrated that patients with non-ischemic DCM, but not ischemic cardiomyopathy, exhibited increased levels of CaMKII phosphorylation of RyR2. Genetic inhibition of CaMKII phosphorylation of S2814 on RyR2 (S2814A) in mice protected against pressure overload induced non-ischemic HF but not against MI-induced ischemic HF.[28] Taken together, these studies identified RyR2 as a downstream target of CaMKII in HF and suggest that CaMKII phosphorylation of RyR2 is important in the pathogenesis of non-ischemic HF.[28]

Regulation of RyR2 by PP1 and PP2A in HF

The fact that RyR2 phosphorylation can be chronically increased in HF despite downregulation of β-adrenergic receptors raised the idea that RyR2 hyperphosphorylation might be the result of altered local signaling, rather than defective global cAMP and Ca/CaM-dependent signaling. Seminal by Marx et al. revealed that RyR2 is in fact a macromolecular channel complex that contains protein kinases (PKA, CaMKII) and protein phosphatases (PP1, PP2A) targeted to the pore-forming channel subunits.[21] Although PP1 and PP2A have similar catalytic mechanisms, substrate recognition mechanisms on RyR2 for PP1 and PP2A may be distinct. For example, Huke et al. showed that whereas PP1 can dephosphorylate both S2808 (the major PKA site) and S2814 (the major CaMKII site), PP2A can only dephosphorylate the S2814 site on RyR2.[45]

There is also emerging evidence that reduced PP activity may contribute to altered RyR2 phosphorylation levels in HF. For example, Ai et al. demonstrated reduced amounts of PP1 and PP2A within the RyR2 complex.[37] In addition, pharmacological inhibition of PP2A activity in cardiomyocytes results in hyperphosphorylation of RyR2 at site S2814, promoting diastolic Ca2+ release events.[46, 47] Together, these studies suggest that reduced PP1 and PP2A activity in the RyR2 macromolecular complex can underlie the abnormal Ca2+ handling seen in the failing heart.[46, 47] However, similar to the controversy regarding the role of PKA and CaMKII in the specific etiologies of HF, recent studies by Belevych et al. showed that levels of PP1A catalytic subunits did not change in a tachycardia-induced HF dog model, whereas both regulatory and catalytic subunits of PP2A were downregulated in this model.[46] This is in contrast to other studies which showed reduced protein levels and PKA phosphorylation of inhibitor-1 (I-1, an inhibitor of phosphatase-1) in human failing hearts, implicating the role of altered PP1 activity.[48] Overexpression of a truncated, constitutively active I-1 form reversed contractile dysfunction in previously failing myocytes.[49, 50] Moreover, genetic overexpression of I-1 in mice enhanced contractility, leading to an increase in RyR2 phosphorylation at S2814, but not S2808, and increased diastolic SR Ca2+ release events and arrhythmogenicity.[51]

In addition to PPs, cyclic nucleotide phosphodiesterases (PDEs) have also been identified in the RyR2 complex, where they regulate intracellular levels of cAMP and cGMP.[21] Within the cardiac PDE family, multiple isoforms are expressed varying in mechanistic action as well as subcellular location.[52] Of the six major classes, PDEs 2/3/4 regulate the activity of PKA through the hydrolyzing of cAMP.[53] Specifically, PDE4B mediates cAMP at the level of the plasma membrane,[54] while PDE3A and PDE4D are thought to directly regulate cAMP levels on the SR.[55] Therefore, these isoforms aid in the control of β-adrenergic stimulation, PKA-mediated phosphorylation of RyR2, and cardiomyocyte contraction.[33] In contrast, PDE4 inhibition increases the frequency of SR Ca2+ release,[33] and the incidence of atrial fibrillation upon β-adrenergic stimulation, suggesting that PDE is another critical regulator of RyR2 phosphorylation through cAMP/ PKA activation. Based on current literature, lack of PDE-mediated regulation of PKA through dysregulation of β-adrenergic receptor/cAMP compartmentalization also contributes to RyR2 hyperphosphorylation, cardiac remodeling, and HF. Specifically, it has been shown that reduced levels of PDE4D3 in patients with HF may contribute to RyR2 hyperphopshorylation and channel defects.[33] In summary, the exact roles of PP1, PP2A, and PDE4 in RyR2 phosphorylation and HF pathogenesis remains to be elucidated, although emerging data suggest that the balance between kinase, phosphatase, and phosphodiesterase activity is altered in HF.

RyR2 and oxidation in HF

Redox signaling serves as another important posttranslational modulator of RyR2 open probability.[16, 17, 56] Each monomer of RyR2 contains roughly 90 cysteine residues, 20 of which are in a reduced state under normal conditions. Many of these cysteines are potential targets for various redox modifications, including S-nitrosylation, S-glutathionylation, and disulfide crosslinking.[18] Endothelial nitric oxide synthase (eNOS), which is localized to caveolae where it compartmentalizes with beta-adrenergic receptors and LTCC, allows NO to inhibit β-adrenergic-induced inotropy.[5760] In contrast, neuronal NOS (nNOS) has an opposite, facilitative effect on contractility, wherein, it stimulates SR Ca2+ release via RyR2.[18, 19] This opposite effect is further evidenced by the observation that nNOS deficiency reduces inotropic response, whereas eNOS deficiency enhances contractility due to corresponding changes in SR Ca2+ release.[20] In addition to redox modulation by nitrosylation, cardiac SR is reported to contain NADH oxidase as well as NOX2 oxidase (that utilizes NADPH instead of NADH) both of which regulate the RyR2 complex under physiological conditions.[6163]

In HF, a shift in the cellular redox potential creates a more oxidized state with increases in reactive oxygen species (ROS) and NO.[6466] This shift in balance has a significant impact on SR Ca2+ handling proteins affecting EC coupling, in particular RyR2.[6466] In the initial phases of HF, increases in NO levels may be an adaptive response to myocardial dysfunction, elevated cytokines, and increases in ROS.[64] This is evidenced by the observation that S-nitrosylation (>3 sites per monomer) leads to reversible RyR2 activation.[18] However, as HF becomes more advanced, oxidation of proteins supersedes nitrosylation. Whereas, nitrosylation is a reversible process, oxidation leads to irreversible activation (beyond a certain threshold).[18] Moreover, oxidation of reactive thiols induces cross-linking between RyR2 subunits resulting in channel activation that can be reversed by S-nitrosylation.[56] In addition, nNOS deficiency leads to an increase in ROS that may lead to oxidation of reactive thiols on RyR2.[67] Thus, whereas S-nitrosylation seems to control the basal redox state of the channel and is an adaptive process in initial stages of HF, oxidation is a pathological process found in advanced HF.[16] Finally, in addition to the direct effects of ROS on RyR2, oxidation can also activate RyR2 by means of CaMKII oxidation and activation.[68] Recent evidence suggests that CaMKII activity can be enhanced by methionine oxidation of M281/M282.[68] These findings highlight the complex and interdependent nature of various posttranslational modifications of RyR2 that can alter channel activity during HF. For more information regarding the role of oxidative stress in HF, please see the review article by Tsutsui et al.[69]

RyR2 and structural proteins in HF

Structural remodeling also plays an important role in the pathogenesis of HF.[70] One of the important aspects of structural remodeling at the cellular level is the remodeling of T-tubules, which can indirectly impact Ca2+ handling.[71] HF leads to a reduction in the overall density as well as irregularities in T-tubule patterning.[70, 72, 73] Specifically, the changes include an increase in longitudinal component and a reduction in the transverse component.[74] As indicated above, proper CICR is critical for contraction in cardiomyocytes. T-tubules ensure that LTCC and RyR2 are in close proximity to each other for optimal CICR. Disruptions in the dyadic distance leads to desynchronized CICR and a reduction in the force of contractility.[15, 70, 71, 74] An important role of T-tubule disruption in HF pathogenesis is also indicated by the fact that mechanical unloading reverses T-tubule remodeling and normalizes CICR in left ventricle in a rat model of ischemic HF as well as in right ventricle in a rat model of pulmonary artery hypertension.[70, 75] In addition to its effect on CICR, T-tubule disruption in HF also alters propagation of action potential further hampering EC coupling.[76] Recently, our lab and Wei et al.[15, 73] found that junctophilin-2 (JPH2), a structural protein, may be a critical mediator of T-tubule remodeling in HF. Our data revealed that an acute reduction in JPH2 levels lead to variability in dyadic distance, loss of EC coupling gain, resulting in aberrant SR Ca2+ leak through RyR2 and eventual HF development.[15] Complementary to our results, Wei et al showed that in early stages of heart failure in a rat model of HF, the T tubule disruption was accompanied by reduction in JPH2 levels.[73] Overall, there is strong evidence that alterations in T-tubule structure alter SR Ca2+ regulation and play an important role in the pathogenesis of HF. These structural changes are intertwined with electrical remodeling, and together, have become critical therapeutic targets for the treatment of HF. Current research is aimed at determining whether downregulation of JPH2 is in response to or an actual cause of HF.

L-type Ca2+ channels in HF

EC coupling in failing hearts can also be depressed due to altered regulation of the LTCC (CaV1.2) on the plasma membrane. The LTCC is composed of a pore-forming α1C subunit and the auxiliary subunits β, α, and γ, which are involved in channel trafficking to the sarcolemma and modulation of the voltage-dependence of channel gating.[77, 78] While some investigators have reported an increase in ICa,L in left ventricular hypertrophy and HF, others have demonstrated an unaltered or downregulated LTCC in HF.[78] This contradiction is evidenced by the fact that although the single-channel activity of LTCC is enhanced in failing heart, whole-cell current is unaltered.[79] A recent study showed that failing hearts have impaired Cav1.2 trafficking.[80, 81] These investigators demonstrated that `Bridging integrator 1' (BIN1), a membrane scaffolding protein that causes Cav1.2 to traffic to T tubules, is reduced in human failing cardiomyocytes. Hong et al. reported that although the total Cav1.2 was not changed in the failing heart, lack of BIN1 leads to reduced abundance of Cav1.2 channels in T-tubules that in turn reduces ICa,L.[81] In a more recent study, Goonasekera et al. showed that reduced LTCC led to a compensatory elevation in neuroendocrine stimulation and RyR2-mediated SR Ca2+ leak, which actually increased the gain of ECC.[82] This enhanced ECC gain resulted in pathologic cardiac hypertrophy and failure through calcineurin/nuclear factor of activated T cells (calcineurin/NFAT) signaling. Although these findings require further investigation, they suggest complex reciprocal signaling between the LTCC and intracellular pro-hypertrophic signaling pathways.

SERCA2a in HF

The cardiac isoform of SERCA (SERCA2a) controls the rate of cytosolic Ca2+ removal and the refilling of SR Ca2+ load in cardiomyocytes. Decreased expression and activity of SERCA2a leads to impaired SR Ca2+ uptake and reduced SR Ca2+ load in cardiomyocytes, which, in turn, compromises SR Ca2+ release and impairs cardiomyocyte contractility.[83] Reduced SERCA2a function is a hallmark of failing hearts and has been found in many experimental models of HF and end-stage HF patients.[83] Normalization of SERCA2a function has been shown to increase contractility and improve hemodynamics along with survival in rodent and large animal models of HF.[8488] These animal studies have led to SERCA2a gene therapy that is currently being evaluated in a clinical trial.[83]

The activity of SERCA2a is directly regulated by accessory subunits, including phospholamban (PLN) and sarcolipin (SLN). PLN, a small protein comprised of 52 amino acid residues, has proven to be the major inhibitory regulator of SERCA2a. Similar to the regulation of RyR2, PKA and CaMKII also play an important role in the regulation of PLN, as they can phosphorylate PLN at sites Ser16 and Thr17, respectively.[89, 90] In its dephosphorylated state, PLN exists predominantly as monomers that bind to and inhibit SERCA2a activity. On the other hand, phosphorylated PLN forms pentameric complexes, which have a lower affinity for SERCA2a, allowing for enhanced activity. The role of PLN in the pathogenesis of HF has been confirmed in animal models where inhibition of PLN restored SR Ca2+ load, delaying the progression of cardiac dysfunction and pathological remodeling.[90, 91] However, loss of PLN did not rescue other genetic mouse models of HF, questioning a causal association between the two.[92]

In addition to PLN, SLN, a 31-amino acid protein, also regulates SERCA2a activity.[93] SLN can bind to SERCA2a through PLN and cause a “superinhibition” of SERCA2a.[94, 95] Transgenic mice with cardiac-specific overexpression of SLN exhibit an impairment of cardiac contractility and ventricular hypertrophy, associated with a reduced amplitude and increased decay time of Ca2+ transients in isolated papillary muscle.[96] Moreover, mice lacking both PLN and SLN display increased SERCA2a pump activity, resulting in elevations in the amount of SR Ca2+ load and systolic Ca2+ transient.[97] However, over time, constitutive activation of SERCA2a in this mutant model becomes detrimental to cardiac function.[97] These findings suggest the importance of the dynamic regulation of SERCA2a by PLN and/or SLN in the maintenance of cardiac contractility under normal conditions and during pathological states.

Furthermore, the small ubiquitin-like modifier type 1 (SUMO1) has also been shown to regulate the activity of SERCA2a through a novel posttranslational modification, termed SUMOylation, which is thought to increase the intrinsic activity of SERCA2a ATPase.[98, 99] Kho et al. found that the level of SERCA2a and SUMO1 were reduced in animal models of HF and in failing cardiomyocytes isolated from HF patients.[99] In the same study, using a transverse aortic constriction-induced HF mouse model, the authors demonstrated that SERCA2a levels could be restored by increasing SUMO1 expression using gene transfer, thereby improving cardiac performance and decreasing mortality.

NCX1 in HF

Another membrane protein essential in Ca2+ homeostasis in cardiomyocytes is the Na+/Ca2+-exchanger type 1 (NCX1). During diastole, NCX1 removes 1 Ca2+ ion from the cytosol in exchange for influx of 3 Na+ ions. As such, this represents another major mechanism, along with SERCA2a, of restoring cytosolic Ca2+ to pre-systolic levels. HF is characterized by elevated cytosolic Ca2+ levels due to reduction in CICR and reduced SERCA2a activity as detailed above, which in turn leads to a compensatory increase in NCX1 expression and activity.[100103] However, the upregulation of NCX1 may become maladaptive over long periods of time due to several reasons. First, in contrast to SERCA2a, NCX1 does not replete SR Ca2+ stores and, thus, does not help in restoring the diminished CICR. Second, upregulation and/or increased activity of NCX1 can be potentially pro-arrhythmogenic.[102105] Overall, the role of NCX is unclear in the pathogenesis of contractile dysfunction in HF but its detrimental role in causing lethal arrhythmias in the setting of HF has been established by several studies.

Conclusion

Several studies have confirmed that diastolic Ca2+ leak from the SR is increased in myocytes isolated from animals as well as patients with HF. In addition, there is abundant evidence that remodeling of the RyR2 macromolecular channel complex and changes in posttranslational modifications contribute to channel dysfunctions that underlie this SR Ca2+ leak. This enhanced SR Ca2+ leak observed in HF can impair cardiac performance, although defects in SERCA2a and NCX1 regulation may further exacerbate the excitation-contraction coupling defects in HF. The cellular consequences of persistent SR Ca2+ leak probably include reduced SR Ca2+ loading, which suppresses the next systolic SR Ca2+ release and contractile event. Additional studies are required to determine how different post-translational modifications conspire to generate SR Ca2+ leak in failing hearts, and how one can intervene therapeutically to restore SR Ca2+ handling and cardiac function. A better understanding of the molecular mechanisms underlying defects in RyR2 may spur the development of novel therapeutic approaches to specifically inhibit SR Ca2+ leak pathways in HF.

Figure 1.

Figure 1

Schematic representation of changes in Type 2 Ryanodine receptors (RyR2) and related proteins in heart failure (HF). Left, represents Ca2+ handling in a cell from normal heart, and right, represents Ca2+ handling in a cell from a failing heart. 1) shows increased RyR2 activity represented by a wide open RyR2 due to post-translational modification by increased protein kinase A (PKA) and Ca2+/CaM-dependent protein kinase II (CaMKII) activity, reduced protein phosphatases (PP) 1 and PP 2A activity, reduced junctophilin 2 (JPH2) leading to abnormalities in T-tubule structure and increased oxidation by reactive oxygen species (ROS), 2) represents complex changes in L-type Ca2+ channel (LTCC) as discussed in the text, 3) represents reduced SR/ER Ca2+-ATPase (SERCA2a) activity and corresponding increased phospholamban (PLN) and Sarcolipin (SLN) activity, and 4) represents increased Na+/Ca2+ exchanger (NCX) activity, in heart failure. ↑ represents increase and ↓ represents decrease.

Acknowledgements

This work was supported in part by NHLBI grants R01-HL089598 and R01-HL091947 (to X.H.T.W.), and T32-HL007676 (to J.L.R.). This work was also supported by American Heart Association grants 10PRE4160124 (to S.A.), 12BGIA12050207 (to N.L.), and 13EIA14560061 (to X.H.T.W.), the Muscular Dystrophy Association, and the Fondation Leducq (`Alliance for CaMKII Signaling in Heart' to X.H.T.W.). S.A. was also supported by an Alkek Foundation fellowship from Baylor College of Medicine.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  • [1].Nabauer M, Callewart G, Cleeman L, Morad M. Regulation of calcium release is gated by calcium current, not gating charge, in cardiac, myocytes. Science. 1989;244:800–803. doi: 10.1126/science.2543067. [DOI] [PubMed] [Google Scholar]
  • [2].Fabiato A. Time and calcium dependence of activation and inactivation of calcium-induced release of calcium from the sarcoplasmic reticulum of a skinned canine cardiac purkinje cell. J Gen Physiol. 1985;85:247–289. doi: 10.1085/jgp.85.2.247. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [3].Bers DM, Guo T. Calcium signaling in cardiac ventricular myocytes. Ann N Y Acad Sci. 2005;1047:86–98. doi: 10.1196/annals.1341.008. [DOI] [PubMed] [Google Scholar]
  • [4].Wehrens XHT, Lehnart SE, Marks AR. Intracellular calcium release channels and cardiac disease. Annu Rev Physiol. 2005;67:69–98. doi: 10.1146/annurev.physiol.67.040403.114521. [DOI] [PubMed] [Google Scholar]
  • [5].Bui AL, Horwich TB, Fonarow GC. Epidemiology and risk profile of heart failure. Nat Rev Cardiol. 2011;8:30–41. doi: 10.1038/nrcardio.2010.165. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [6].Ather S, Chan W, Bozkurt B, Aguilar D, Ramasubbu K, Zachariah AA, Wehrens XH, Deswal A. Impact of noncardiac comorbidities on morbidity and mortality in a predominantly male population with heart failure and preserved versus reduced ejection fraction. J Am Coll Cardiol. 2012;59:998–1005. doi: 10.1016/j.jacc.2011.11.040. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [7].Tunwell RE, Wickenden C, Bertrand BM, Shevchenko VI, Walsh MB, Allen PD, Lai FA. The human cardiac muscle ryanodine receptor-calcium release channel: identification, primary structure and topological analysis. Biochem J. 1996;318:477–487. doi: 10.1042/bj3180477. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [8].Otsu K, Willard HF, Khanna VK, Zorzato F, Green NM, MacLennan DH. Molecular cloning of cDNA encoding the Ca2+ release channel (ryanodine receptor) of rabbit cardiac muscle sarcoplasmic reticulum. J Biol Chem. 1990;265:13472–13483. [PubMed] [Google Scholar]
  • [9].Marks AR, Tempst P, Hwang KS, Taubman MB, Inui M, Chadwick C, Fleischer S, Nadal-Ginard B. Molecular cloning and characterization of the ryanodine receptor/junctional channel complex cDNA from skeletal muscle sarcoplasmic reticulum. Proc Natl Acad Sci U S A. 1989;86:8683–8687. doi: 10.1073/pnas.86.22.8683. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [10].Lam E, Martin MM, Timerman AP, Sabers C, Fleischer S, Lukas T, Abraham RT, O'Keefe SJ, O'Neill EA, Wiederrecht GJ. A novel FK506 binding protein can mediate the immunosuppressive effects of FK506 and is associated with the cardiac ryanodine receptor. J Biol Chem. 1995;270:26511–26522. doi: 10.1074/jbc.270.44.26511. [DOI] [PubMed] [Google Scholar]
  • [11].Brillantes AB, Ondrias K, Scott A, Kobrinsky E, Ondriasova E, Moschella MC, Jayaraman T, Landers M, Ehrlich BE, Marks AR. Stabilization of calcium release channel (ryanodine receptor) function by FK506-binding protein. Cell. 1994;77:513–523. doi: 10.1016/0092-8674(94)90214-3. [DOI] [PubMed] [Google Scholar]
  • [12].Wehrens XH, Lehnart SE, Huang F, Vest JA, Reiken SR, Mohler PJ, Sun J, Guatimosim S, Song LS, Rosemblit N, D'Armiento JM, Napolitano C, Memmi M, Priori SG, Lederer WJ, Marks AR. FKBP12.6 deficiency and defective calcium release channel (ryanodine receptor) function linked to exercise-induced sudden cardiac death. Cell. 2003;113:829–840. doi: 10.1016/s0092-8674(03)00434-3. [DOI] [PubMed] [Google Scholar]
  • [13].Chen SR, MacLennan DH. Identification of calmodulin-, Ca(2+)-, and ruthenium red-binding domains in the Ca2+ release channel (ryanodine receptor) of rabbit skeletal muscle sarcoplasmic reticulum. J Biol Chem. 1994;269:22698–22704. [PubMed] [Google Scholar]
  • [14].Meyers MB, Pickel VM, Sheu SS, Sharma VK, Scotto KW, Fishman GI. Association of sorcin with the cardiac ryanodine receptor. J Biol Chem. 1995;270:26411–26418. doi: 10.1074/jbc.270.44.26411. [DOI] [PubMed] [Google Scholar]
  • [15].van Oort RJ, Garbino A, Wang W, Dixit SS, Landstrom AP, Gaur N, De Almeida AC, Skapura DG, Rudy Y, Burns AR, Ackerman MJ, Wehrens XH. Disrupted junctional membrane complexes and hyperactive ryanodine receptors after acute junctophilin knockdown in mice. Circulation. 2011;123:979–988. doi: 10.1161/CIRCULATIONAHA.110.006437. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [16].Yano M, Okuda S, Oda T, Tokuhisa T, Tateishi H, Mochizuki M, Noma T, Doi M, Kobayashi S, Yamamoto T, Ikeda Y, Ohkusa T, Ikemoto N, Matsuzaki M. Correction of defective interdomain interaction within ryanodine receptor by antioxidant is a new therapeutic strategy against heart failure. Circulation. 2005;112:3633–3643. doi: 10.1161/CIRCULATIONAHA.105.555623. [DOI] [PubMed] [Google Scholar]
  • [17].Santos CX, Anilkumar N, Zhang M, Brewer AC, Shah AM. Redox signaling in cardiac myocytes. Free Radic Biol Med. 2011;50:777–793. doi: 10.1016/j.freeradbiomed.2011.01.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [18].Xu L, Eu JP, Meissner G, Stamler JS. Activation of the cardiac calcium release channel (ryanodine receptor) by poly-S-nitrosylation. Science. 1998;279:234–237. doi: 10.1126/science.279.5348.234. [DOI] [PubMed] [Google Scholar]
  • [19].Eu JP, Sun J, Xu L, Stamler JS, Meissner G. The skeletal muscle calcium release channel: coupled O2 sensor and NO signaling functions. Cell. 2000;102:499–509. doi: 10.1016/s0092-8674(00)00054-4. [DOI] [PubMed] [Google Scholar]
  • [20].Barouch LA, Harrison RW, Skaf MW, Rosas GO, Cappola TP, Kobeissi ZA, Hobai IA, Lemmon CA, Burnett AL, O'Rourke B, Rodriguez ER, Huang PL, Lima JA, Berkowitz DE, Hare JM. Nitric oxide regulates the heart by spatial confinement of nitric oxide synthase isoforms. Nature. 2002;416:337–339. doi: 10.1038/416337a. [DOI] [PubMed] [Google Scholar]
  • [21].Marx SO, Reiken S, Hisamatsu Y, Jayaraman T, Burkhoff D, Rosemblit N, Marks AR. PKA phosphorylation dissociates FKBP12.6 from the calcium release channel (ryanodine receptor): Defective regulation in failing hearts. Cell. 2000;101:365–376. doi: 10.1016/s0092-8674(00)80847-8. [DOI] [PubMed] [Google Scholar]
  • [22].Wehrens XH, Lehnart SE, Reiken SR, Marks AR. Ca2+/calmodulin-dependent protein kinase II phosphorylation regulates the cardiac ryanodine receptor. Circ Res. 2004;94:e61–70. doi: 10.1161/01.RES.0000125626.33738.E2. [DOI] [PubMed] [Google Scholar]
  • [23].Braz JC, Gregory K, Pathak A, Zhao W, Sahin B, Klevitsky R, Kimball TF, Lorenz JN, Nairn AC, Liggett SB, Bodi I, Wang S, Schwartz A, Lakatta EG, DePaoli-Roach AA, Robbins J, Hewett TE, Bibb JA, Westfall MV, Kranias EG, Molkentin JD. PKC-alpha regulates cardiac contractility and propensity toward heart failure. Nat Med. 2004;10:248–254. doi: 10.1038/nm1000. [DOI] [PubMed] [Google Scholar]
  • [24].Marx SO, Reiken S, Hisamatsu Y, Gaburjakova M, Gaburjakova J, Yang YM, Rosemblit N, Marks AR. Phosphorylation-dependent regulation of ryanodine receptors. A novel role for leucine/isoleucine zippers. J Cell Biol. 2001;153:699–708. doi: 10.1083/jcb.153.4.699. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [25].Currie S, Loughrey CM, Craig MA, Smith GL. Calcium/calmodulin-dependent protein kinase IIdelta associates with the ryanodine receptor complex and regulates channel function in rabbit heart. Biochem J. 2004;377:357–366. doi: 10.1042/BJ20031043. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [26].Zhang H, Makarewich CA, Kubo H, Wang W, Duran JM, Li Y, Berretta RM, Koch WJ, Chen X, Gao E, Valdivia HH, Houser SR. Hyperphosphorylation of the cardiac ryanodine receptor at serine 2808 is not involved in cardiac dysfunction after myocardial infarction. Circ Res. 2012;110:831–840. doi: 10.1161/CIRCRESAHA.111.255158. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [27].van Oort RJ, McCauley MD, Dixit SS, Pereira L, Yang Y, Respress JL, Wang Q, De Almeida AC, Skapura DG, Anderson ME, Bers DM, Wehrens XH. Ryanodine receptor phosphorylation by calcium/calmodulin-dependent protein kinase II promotes life-threatening ventricular arrhythmias in mice with heart failure. Circulation. 2010;122:2669–2679. doi: 10.1161/CIRCULATIONAHA.110.982298. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [28].Respress JL, van Oort RJ, Li N, Rolim N, Dixit SS, Dealmeida A, Voigt N, Lawrence WS, Skapura DG, Skardal K, Wisloff U, Wieland T, Ai X, Pogwizd SM, Dobrev D, Wehrens XH. Role of RyR2 Phosphorylation at S2814 During Heart Failure Progression. Circ Res. 2012;110:1474–1483. doi: 10.1161/CIRCRESAHA.112.268094. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [29].Shan J, Betzenhauser MJ, Kushnir A, Reiken S, Meli AC, Wronska A, Dura M, Chen BX, Marks AR. Role of chronic ryanodine receptor phosphorylation in heart failure and beta-adrenergic receptor blockade in mice. J Clin Invest. 2010;120:4375–4387. doi: 10.1172/JCI37649. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [30].Bers DM. Ryanodine receptor S2808 phosphorylation in heart failure: smoking gun or red herring. Circ Res. 2012;110:796–799. doi: 10.1161/CIRCRESAHA.112.265579. [DOI] [PubMed] [Google Scholar]
  • [31].McCauley MD, Wehrens XH. Ryanodine receptor phosphorylation, calcium/calmodulin-dependent protein kinase II, and life-threatening ventricular arrhythmias. Trends Cardiovasc Med. 2011;21:48–51. doi: 10.1016/j.tcm.2012.02.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [32].Reiken S, Wehrens XH, Vest JA, Barbone A, Klotz S, Mancini D, Burkhoff D, Marks AR. Beta-blockers restore calcium release channel function and improve cardiac muscle performance in human heart failure. Circulation. 2003;107:2459–2466. doi: 10.1161/01.CIR.0000068316.53218.49. [DOI] [PubMed] [Google Scholar]
  • [33].Lehnart SE, Wehrens XH, Reiken S, Warrier S, Belevych AE, Harvey RD, Richter W, Jin SL, Conti M, Marks AR. Phosphodiesterase 4D deficiency in the ryanodine-receptor complex promotes heart failure and arrhythmias. Cell. 2005;123:25–35. doi: 10.1016/j.cell.2005.07.030. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [34].Wehrens XH, Lehnart SE, Reiken S, van der Nagel R, Morales R, Sun J, Cheng Z, Deng SX, de Windt LJ, Landry DW, Marks AR. Enhancing calstabin binding to ryanodine receptors improves cardiac and skeletal muscle function in heart failure. Proc Natl Acad Sci U S A. 2005;102:9607–9612. doi: 10.1073/pnas.0500353102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [35].Wehrens XH, Lehnart SE, Reiken S, Vest JA, Wronska A, Marks AR. Ryanodine receptor/calcium release channel PKA phosphorylation: a critical mediator of heart failure progression. Proc Natl Acad Sci U S A. 2006;103:511–518. doi: 10.1073/pnas.0510113103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [36].Benkusky NA, Weber CS, Scherman JA, Farrell EF, Hacker TA, John MC, Powers PA, Valdivia HH. Intact beta-adrenergic response and unmodified progression toward heart failure in mice with genetic ablation of a major protein kinase A phosphorylation site in the cardiac ryanodine receptor. Circ Res. 2007;101:819–829. doi: 10.1161/CIRCRESAHA.107.153007. [DOI] [PubMed] [Google Scholar]
  • [37].Ai X, Curran JW, Shannon TR, Bers DM, Pogwizd SM. Ca2+/calmodulin-dependent protein kinase modulates cardiac ryanodine receptor phosphorylation and sarcoplasmic reticulum Ca2+ leak in heart failure. Circ Res. 2005;97:1314–1322. doi: 10.1161/01.RES.0000194329.41863.89. [DOI] [PubMed] [Google Scholar]
  • [38].Shannon TR, Pogwizd SM, Bers DM. Elevated sarcoplasmic reticulum Ca2+ leak in intact ventricular myocytes from rabbits in heart failure. Circ Res. 2003;93:592–594. doi: 10.1161/01.RES.0000093399.11734.B3. [DOI] [PubMed] [Google Scholar]
  • [39].Hoch B, Meyer R, Hetzer R, Krause EG, Karczewski P. Identification and expression of delta-isoforms of the multifunctional Ca2+/calmodulin-dependent protein kinase in failing and nonfailing human myocardium. Circ Res. 1999;84:713–721. doi: 10.1161/01.res.84.6.713. [DOI] [PubMed] [Google Scholar]
  • [40].Zhang T, Maier LS, Dalton ND, Miyamoto S, Ross J, Jr., Bers DM, Brown JH. The deltaC isoform of CaMKII is activated in cardiac hypertrophy and induces dilated cardiomyopathy and heart failure. Circ Res. 2003;92:912–919. doi: 10.1161/01.RES.0000069686.31472.C5. [DOI] [PubMed] [Google Scholar]
  • [41].Zhang R, Khoo MS, Wu Y, Yang Y, Grueter CE, Ni G, Price EE, Jr., Thiel W, Guatimosim S, Song LS, Madu EC, Shah AN, Vishnivetskaya TA, Atkinson JB, Gurevich VV, Salama G, Lederer WJ, Colbran RJ, Anderson ME. Calmodulin kinase II inhibition protects against structural heart disease. Nat Med. 2005;11:409–417. doi: 10.1038/nm1215. [DOI] [PubMed] [Google Scholar]
  • [42].Sag CM, Wadsack DP, Khabbazzadeh S, Abesser M, Grefe C, Neumann K, Opiela MK, Backs J, Olson EN, Brown JH, Neef S, Maier SK, Maier LS. Calcium/calmodulin-dependent protein kinase II contributes to cardiac arrhythmogenesis in heart failure. Circ Heart Fail. 2009;2:664–675. doi: 10.1161/CIRCHEARTFAILURE.109.865279. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [43].Ling H, Zhang T, Pereira L, Means CK, Cheng H, Gu Y, Dalton ND, Peterson KL, Chen J, Bers D, Brown JH. Requirement for Ca2+/calmodulin-dependent kinase II in the transition from pressure overload-induced cardiac hypertrophy to heart failure in mice. J Clin Invest. 2009;119:1230–1240. doi: 10.1172/JCI38022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [44].Backs J, Backs T, Neef S, Kreusser MM, Lehmann LH, Patrick DM, Grueter CE, Qi X, Richardson JA, Hill JA, Katus HA, Bassel-Duby R, Maier LS, Olson EN. The delta isoform of CaM kinase II is required for pathological cardiac hypertrophy and remodeling after pressure overload. Proc Natl Acad Sci U S A. 2009;106:2342–2347. doi: 10.1073/pnas.0813013106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [45].Huke S, Bers DM. Ryanodine receptor phosphorylation at Serine 2030, 2808 and 2814 in rat cardiomyocytes. Biochem Biophys Res Commun. 2008;376:80–85. doi: 10.1016/j.bbrc.2008.08.084. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [46].Belevych AE, Terentyev D, Terentyeva R, Nishijima Y, Sridhar A, Hamlin RL, Carnes CA, Gyorke S. The relationship between arrhythmogenesis and impaired contractility in heart failure: role of altered ryanodine receptor function. Cardiovasc Res. 2011;90:493–502. doi: 10.1093/cvr/cvr025. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [47].Terentyev D, Belevych AE, Terentyeva R, Martin MM, Malana GE, Kuhn DE, Abdellatif M, Feldman DS, Elton TS, Gyorke S. miR-1 overexpression enhances Ca(2+) release and promotes cardiac arrhythmogenesis by targeting PP2A regulatory subunit B56alpha and causing CaMKII-dependent hyperphosphorylation of RyR2. Circ Res. 2009;104:514–521. doi: 10.1161/CIRCRESAHA.108.181651. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [48].El-Armouche A, Pamminger T, Ditz D, Zolk O, Eschenhagen T. Decreased protein and phosphorylation level of the protein phosphatase inhibitor-1 in failing human hearts. Cardiovasc Res. 2004;61:87–93. doi: 10.1016/j.cardiores.2003.11.005. [DOI] [PubMed] [Google Scholar]
  • [49].Bauce B, Rampazzo A, Basso C, Bagattin A, Daliento L, Tiso N, Turrini P, Thiene G, Danieli GA, Nava A. Screening for ryanodine receptor type 2 mutations in families with effort-induced polymorphic ventricular arrhythmias and sudden death: early diagnosis of asymptomatic carriers. J Am Coll Cardiol. 2002;40:341–349. doi: 10.1016/s0735-1097(02)01946-0. [DOI] [PubMed] [Google Scholar]
  • [50].Pathak A, del Monte F, Zhao W, Schultz JE, Lorenz JN, Bodi I, Weiser D, Hahn H, Carr AN, Syed F, Mavila N, Jha L, Qian J, Marreez Y, Chen G, McGraw DW, Heist EK, Guerrero JL, DePaoli-Roach AA, Hajjar RJ, Kranias EG. Enhancement of cardiac function and suppression of heart failure progression by inhibition of protein phosphatase 1. Circ Res. 2005;96:756–766. doi: 10.1161/01.RES.0000161256.85833.fa. [DOI] [PubMed] [Google Scholar]
  • [51].Wittkopper K, Fabritz L, Neef S, Ort KR, Grefe C, Unsold B, Kirchhof P, Maier LS, Hasenfuss G, Dobrev D, Eschenhagen T, El-Armouche A. Constitutively active phosphatase inhibitor-1 improves cardiac contractility in young mice but is deleterious after catecholaminergic stress and with aging. J Clin Invest. 2010;120:617–626. doi: 10.1172/JCI40545. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [52].Lugnier C. Cyclic nucleotide phosphodiesterase (PDE) superfamily: a new target for the development of specific therapeutic agents. Pharmacol Ther. 2006;109:366–398. doi: 10.1016/j.pharmthera.2005.07.003. [DOI] [PubMed] [Google Scholar]
  • [53].Fischmeister R, Castro LR, Abi-Gerges A, Rochais F, Jurevicius J, Leroy J, Vandecasteele G. Compartmentation of cyclic nucleotide signaling in the heart: the role of cyclic nucleotide phosphodiesterases. Circ Res. 2006;99:816–828. doi: 10.1161/01.RES.0000246118.98832.04. [DOI] [PubMed] [Google Scholar]
  • [54].Leroy J, Richter W, Mika D, Castro LR, Abi-Gerges A, Xie M, Scheitrum C, Lefebvre F, Schittl J, Mateo P, Westenbroek R, Catterall WA, Charpentier F, Conti M, Fischmeister R, Vandecasteele G. Phosphodiesterase 4B in the cardiac L-type Ca(2)(+) channel complex regulates Ca(2)(+) current and protects against ventricular arrhythmias in mice. J Clin Invest. 2011;121:2651–2661. doi: 10.1172/JCI44747. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [55].Conti M, Richter W, Mehats C, Livera G, Park JY, Jin C. Cyclic AMP-specific PDE4 phosphodiesterases as critical components of cyclic AMP signaling. J Biol Chem. 2003;278:5493–5496. doi: 10.1074/jbc.R200029200. [DOI] [PubMed] [Google Scholar]
  • [56].Aghdasi B, Reid MB, Hamilton SL. Nitric oxide protects the skeletal muscle Ca2+ release channel from oxidation induced activation. J Biol Chem. 1997;272:25462–25467. doi: 10.1074/jbc.272.41.25462. [DOI] [PubMed] [Google Scholar]
  • [57].Feron O, Saldana F, Michel JB, Michel T. The endothelial nitric-oxide synthase-caveolin regulatory cycle. J Biol Chem. 1998;273:3125–3128. doi: 10.1074/jbc.273.6.3125. [DOI] [PubMed] [Google Scholar]
  • [58].Schwencke C, Yamamoto M, Okumura S, Toya Y, Kim SJ, Ishikawa Y. Compartmentation of cyclic adenosine 3',5'-monophosphate signaling in caveolae. Molecular endocrinology. 1999;13:1061–1070. doi: 10.1210/mend.13.7.0304. [DOI] [PubMed] [Google Scholar]
  • [59].Hare JM, Lofthouse RA, Juang GJ, Colman L, Ricker KM, Kim B, Senzaki H, Cao S, Tunin RS, Kass DA. Contribution of caveolin protein abundance to augmented nitric oxide signaling in conscious dogs with pacing-induced heart failure. Circ Res. 2000;86:1085–1092. doi: 10.1161/01.res.86.10.1085. [DOI] [PubMed] [Google Scholar]
  • [60].Hare JM, Givertz MM, Creager MA, Colucci WS. Increased sensitivity to nitric oxide synthase inhibition in patients with heart failure: potentiation of beta-adrenergic inotropic responsiveness. Circulation. 1998;97:161–166. doi: 10.1161/01.cir.97.2.161. [DOI] [PubMed] [Google Scholar]
  • [61].Cherednichenko G, Zima AV, Feng W, Schaefer S, Blatter LA, Pessah IN. NADH oxidase activity of rat cardiac sarcoplasmic reticulum regulates calcium-induced calcium release. Circ Res. 2004;94:478–486. doi: 10.1161/01.RES.0000115554.65513.7C. [DOI] [PubMed] [Google Scholar]
  • [62].Sanchez G, Pedrozo Z, Domenech RJ, Hidalgo C, Donoso P. Tachycardia increases NADPH oxidase activity and RyR2 S-glutathionylation in ventricular muscle. J Mol Cell Cardiol. 2005;39:982–991. doi: 10.1016/j.yjmcc.2005.08.010. [DOI] [PubMed] [Google Scholar]
  • [63].Hidalgo C, Sanchez G, Barrientos G, Aracena-Parks P. A transverse tubule NADPH oxidase activity stimulates calcium release from isolated triads via ryanodine receptor type 1 S -glutathionylation. J Biol Chem. 2006;281:26473–26482. doi: 10.1074/jbc.M600451200. [DOI] [PubMed] [Google Scholar]
  • [64].Choudhary G, Dudley SC., Jr. Heart failure, oxidative stress, and ion channel modulation. Congest Heart Fail. 2002;8:148–155. doi: 10.1111/j.1527-5299.2002.00716.x. [DOI] [PubMed] [Google Scholar]
  • [65].Gonzalez DR, Treuer AV, Castellanos J, Dulce RA, Hare JM. Impaired S-nitrosylation of the ryanodine receptor caused by xanthine oxidase activity contributes to calcium leak in heart failure. J Biol Chem. 2010;285:28938–28945. doi: 10.1074/jbc.M110.154948. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [66].Terentyev D, Gyorke I, Belevych AE, Terentyeva R, Sridhar A, Nishijima Y, de Blanco EC, Khanna S, Sen CK, Cardounel AJ, Carnes CA, Gyorke S. Redox modification of ryanodine receptors contributes to sarcoplasmic reticulum Ca2+ leak in chronic heart failure. Circ Res. 2008;103:1466–1472. doi: 10.1161/CIRCRESAHA.108.184457. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [67].Khan SA, Lee K, Minhas KM, Gonzalez DR, Raju SV, Tejani AD, Li D, Berkowitz DE, Hare JM. Neuronal nitric oxide synthase negatively regulates xanthine oxidoreductase inhibition of cardiac excitation-contraction coupling. Proc Natl Acad Sci U S A. 2004;101:15944–15948. doi: 10.1073/pnas.0404136101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [68].Erickson JR, Joiner ML, Guan X, Kutschke W, Yang J, Oddis CV, Bartlett RK, Lowe JS, O'Donnell SE, Aykin-Burns N, Zimmerman MC, Zimmerman K, Ham AJ, Weiss RM, Spitz DR, Shea MA, Colbran RJ, Mohler PJ, Anderson ME. A dynamic pathway for calcium-independent activation of CaMKII by methionine oxidation. Cell. 2008;133:462–474. doi: 10.1016/j.cell.2008.02.048. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [69].Tsutsui H, Kinugawa S, Matsushima S. Oxidative stress and heart failure. Am J Physiol Heart Circ Physiol. 301:H2181–2190. doi: 10.1152/ajpheart.00554.2011. [DOI] [PubMed] [Google Scholar]
  • [70].Ibrahim M, Navaratnarajah M, Siedlecka U, Rao C, Dias P, Moshkov AV, Gorelik J, Yacoub MH, Terracciano CM. Mechanical unloading reverses transverse tubule remodelling and normalizes local Ca2+-induced Ca2+release in a rodent model of heart failure. Eur J Heart Fail. 2012;14:571–580. doi: 10.1093/eurjhf/hfs038. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [71].Gomez AM, Valdivia HH, Cheng H, Lederer MR, Santana LF, Cannell MB, McCune SA, Altschuld RA, Lederer WJ. Defective excitation-contraction coupling in experimental cardiac hypertrophy and heart failure. Science. 1997;276:800–806. doi: 10.1126/science.276.5313.800. [DOI] [PubMed] [Google Scholar]
  • [72].Lyon AR, MacLeod KT, Zhang Y, Garcia E, Kanda GK, Lab MJ, Korchev YE, Harding SE, Gorelik J. Loss of T-tubules and other changes to surface topography in ventricular myocytes from failing human and rat heart. Proc Natl Acad Sci U S A. 2009;106:6854–6859. doi: 10.1073/pnas.0809777106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [73].Wei S, Guo A, Chen B, Kutschke W, Xie YP, Zimmerman K, Weiss RM, Anderson ME, Cheng H, Song LS. T-tubule remodeling during transition from hypertrophy to heart failure. Circ Res. 2010;107:520–531. doi: 10.1161/CIRCRESAHA.109.212324. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [74].Song LS, Sobie EA, McCulle S, Lederer WJ, Balke CW, Cheng H. Orphaned ryanodine receptors in the failing heart. Proceedings of the National Academy of Sciences of the United States of America. 2006;103:4305–4310. doi: 10.1073/pnas.0509324103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [75].Xie YP, Chen B, Sanders P, Guo A, Li Y, Zimmerman K, Wang LC, Weiss RM, Grumbach IM, Anderson ME, Song LS. Sildenafil prevents and reverses transverse-tubule remodeling and Ca(2+) handling dysfunction in right ventricle failure induced by pulmonary artery hypertension. Hypertension. 2012;59:355–362. doi: 10.1161/HYPERTENSIONAHA.111.180968. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [76].Sacconi L, Ferrantini C, Lotti J, Coppini R, Yan P, Loew LM, Tesi C, Cerbai E, Poggesi C, Pavone FS. Action potential propagation in transverse-axial tubular system is impaired in heart failure. Proc Natl Acad Sci U S A. 2012;109:5815–5819. doi: 10.1073/pnas.1120188109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [77].Anderson ME. Calmodulin kinase and L-type calcium channels; a recipe for arrhythmias? Trends Cardiovasc Med. 2004;14:152–161. doi: 10.1016/j.tcm.2004.02.005. [DOI] [PubMed] [Google Scholar]
  • [78].Bodi I, Mikala G, Koch SE, Akhter SA, Schwartz A. The L-type calcium channel in the heart: the beat goes on. J Clin Invest. 2005;115:3306–3317. doi: 10.1172/JCI27167. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [79].Schroder F, Handrock R, Beuckelmann DJ, Hirt S, Hullin R, Priebe L, Schwinger RH, Weil J, Herzig S. Increased availability and open probability of single L-type calcium channels from failing compared with nonfailing human ventricle. Circulation. 1998;98:969–976. doi: 10.1161/01.cir.98.10.969. [DOI] [PubMed] [Google Scholar]
  • [80].Hong TT, Cogswell R, James CA, Kang G, Pullinger CR, Malloy MJ, Kane JP, Wojciak J, Calkins H, Scheinman MM, Tseng ZH, Ganz P, De Marco T, Judge DP, Shaw RM. Plasma BIN1 correlates with heart failure and predicts arrhythmia in patients with arrhythmogenic right ventricular cardiomyopathy. Heart Rhythm. 2012;9:961–967. doi: 10.1016/j.hrthm.2012.01.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [81].Hong TT, Smyth JW, Chu KY, Vogan JM, Fong TS, Jensen BC, Fang K, Halushka MK, Russell SD, Colecraft H, Hoopes CW, Ocorr K, Chi NC, Shaw RM. BIN1 is reduced and Cav1.2 trafficking is impaired in human failing cardiomyocytes. Heart Rhythm. 2012;9:812–820. doi: 10.1016/j.hrthm.2011.11.055. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [82].Goonasekera SA, Hammer K, Auger-Messier M, Bodi I, Chen X, Zhang H, Reiken S, Elrod JW, Correll RN, York AJ, Sargent MA, Hofmann F, Moosmang S, Marks AR, Houser SR, Bers DM, Molkentin JD. Decreased cardiac L-type Ca(2)(+) channel activity induces hypertrophy and heart failure in mice. J Clin Invest. 2012;122:280–290. doi: 10.1172/JCI58227. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [83].Kranias EG, Hajjar RJ. Modulation of Cardiac Contractility by the Phopholamban/SERCA2a Regulatome. Circ Res. 2012;110:1646–1660. doi: 10.1161/CIRCRESAHA.111.259754. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [84].Kumarswamy R, Lyon AR, Volkmann I, Mills AM, Bretthauer J, Pahuja A, Geers-Knorr C, Kraft T, Hajjar RJ, Macleod KT, Harding SE, Thum T. SERCA2a gene therapy restores microRNA-1 expression in heart failure via an Akt/FoxO3A-dependent pathway. Eur Heart J. 2012;33:1067–1075. doi: 10.1093/eurheartj/ehs043. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [85].Lyon AR, Bannister ML, Collins T, Pearce E, Sepehripour AH, Dubb SS, Garcia E, O'Gara P, Liang L, Kohlbrenner E, Hajjar RJ, Peters NS, Poole-Wilson PA, Macleod KT, Harding SE. SERCA2a gene transfer decreases sarcoplasmic reticulum calcium leak and reduces ventricular arrhythmias in a model of chronic heart failure. Circ Arrhythm Electrophysiol. 2011;4:362–372. doi: 10.1161/CIRCEP.110.961615. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [86].Sakata S, Lebeche D, Sakata N, Sakata Y, Chemaly ER, Liang LF, Tsuji T, Takewa Y, del Monte F, Peluso R, Zsebo K, Jeong D, Park WJ, Kawase Y, Hajjar RJ. Restoration of mechanical and energetic function in failing aortic-banded rat hearts by gene transfer of calcium cycling proteins. J Mol Cell Cardiol. 2007;42:852–861. doi: 10.1016/j.yjmcc.2007.01.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [87].Xin W, Lu X, Li X, Niu K, Cai J. Attenuation of endoplasmic reticulum stress-related myocardial apoptosis by SERCA2a gene delivery in ischemic heart disease. Mol Med. 2011;17:201–210. doi: 10.2119/molmed.2010.00197. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [88].Hirsch JC, Borton AR, Albayya FP, Russell MW, Ohye RG, Metzger JM. Comparative analysis of parvalbumin and SERCA2a cardiac myocyte gene transfer in a large animal model of diastolic dysfunction. Am J Physiol Heart Circ Physiol. 2004;286:H2314–2321. doi: 10.1152/ajpheart.01137.2003. [DOI] [PubMed] [Google Scholar]
  • [89].Mattiazzi A, Mundina-Weilenmann C, Guoxiang C, Vittone L, Kranias E. Role of phospholamban phosphorylation on Thr17 in cardiac physiological and pathological conditions. Cardiovasc Res. 2005;68:366–375. doi: 10.1016/j.cardiores.2005.08.010. [DOI] [PubMed] [Google Scholar]
  • [90].Zhang YH, Zhang MH, Sears CE, Emanuel K, Redwood C, El-Armouche A, Kranias EG, Casadei B. Reduced phospholamban phosphorylation is associated with impaired relaxation in left ventricular myocytes from neuronal NO synthase-deficient mice. Circ Res. 2008;102:242–249. doi: 10.1161/CIRCRESAHA.107.164798. [DOI] [PubMed] [Google Scholar]
  • [91].Zhang T, Guo T, Mishra S, Dalton ND, Kranias EG, Peterson KL, Bers DM, Brown JH. Phospholamban ablation rescues sarcoplasmic reticulum Ca(2+) handling but exacerbates cardiac dysfunction in CaMKIIdelta(C) transgenic mice. Circ Res. 2010;106:354–362. doi: 10.1161/CIRCRESAHA.109.207423. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [92].Song Q, Schmidt AG, Hahn HS, Carr AN, Frank B, Pater L, Gerst M, Young K, Hoit BD, McConnell BK, Haghighi K, Seidman CE, Seidman JG, Dorn GW, 2nd, Kranias EG. Rescue of cardiomyocyte dysfunction by phospholamban ablation does not prevent ventricular failure in genetic hypertrophy. J Clin Invest. 2003;111:859–867. doi: 10.1172/JCI16738. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [93].Autry JM, Rubin JE, Pietrini SD, Winters DL, Robia SL, Thomas DD. Oligomeric interactions of sarcolipin and the Ca-ATPase. J Biol Chem. 2011;286:31697–31706. doi: 10.1074/jbc.M111.246843. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [94].Asahi M, Nakayama H, Tada M, Otsu K. Regulation of sarco(endo)plasmic reticulum Ca2+ adenosine triphosphatase by phospholamban and sarcolipin: implication for cardiac hypertrophy and failure. Trends Cardiovasc Med. 2003;13:152–157. doi: 10.1016/s1050-1738(03)00037-9. [DOI] [PubMed] [Google Scholar]
  • [95].Babu GJ, Bhupathy P, Carnes CA, Billman GE, Periasamy M. Differential expression of sarcolipin protein during muscle development and cardiac pathophysiology. J Mol Cell Cardiol. 2007;43:215–222. doi: 10.1016/j.yjmcc.2007.05.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [96].Asahi M, Otsu K, Nakayama H, Hikoso S, Takeda T, Gramolini AO, Trivieri MG, Oudit GY, Morita T, Kusakari Y, Hirano S, Hongo K, Hirotani S, Yamaguchi O, Peterson A, Backx PH, Kurihara S, Hori M, MacLennan DH. Cardiac-specific overexpression of sarcolipin inhibits sarco(endo)plasmic reticulum Ca2+ ATPase (SERCA2a) activity and impairs cardiac function in mice. Proc Natl Acad Sci U S A. 2004;101:9199–9204. doi: 10.1073/pnas.0402596101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [97].Shanmugam M, Gao S, Hong C, Fefelova N, Nowycky MC, Xie LH, Periasamy M, Babu GJ. Ablation of phospholamban and sarcolipin results in cardiac hypertrophy and decreased cardiac contractility. Cardiovasc Res. 2011;89:353–361. doi: 10.1093/cvr/cvq294. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [98].Schwartz RJ, Yeh ET. Weighing in on heart failure: the role of SERCA2a SUMOylation. Circ Res. 2012;110:198–199. doi: 10.1161/RES.0b013e318246f187. [DOI] [PubMed] [Google Scholar]
  • [99].Kho C, Lee A, Jeong D, Oh JG, Chaanine AH, Kizana E, Park WJ, Hajjar RJ. SUMO1-dependent modulation of SERCA2a in heart failure. Nature. 2011;477:601–605. doi: 10.1038/nature10407. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [100].Mork HK, Sjaastad I, Sande JB, Periasamy M, Sejersted OM, Louch WE. Increased cardiomyocyte function and Ca2+ transients in mice during early congestive heart failure. J Mol Cell Cardiol. 2007;43:177–186. doi: 10.1016/j.yjmcc.2007.05.004. [DOI] [PubMed] [Google Scholar]
  • [101].Hasenfuss G, Schillinger W, Lehnart SE, Preuss M, Pieske B, Maier LS, Prestle J, Minami K, Just H. Relationship between Na+-Ca2+-exchanger protein levels and diastolic function of failing human myocardium. Circulation. 1999;99:641–648. doi: 10.1161/01.cir.99.5.641. [DOI] [PubMed] [Google Scholar]
  • [102].Pogwizd SM, Qi M, Yuan W, Samarel AM, Bers DM. Upregulation of Na(+)/Ca(2+) exchanger expression and function in an arrhythmogenic rabbit model of heart failure. Circ Res. 1999;85:1009–1019. doi: 10.1161/01.res.85.11.1009. [DOI] [PubMed] [Google Scholar]
  • [103].Sipido KR, Volders PG, de Groot SH, Verdonck F, Van de Werf F, Wellens HJ, Vos MA. Enhanced Ca(2+) release and Na/Ca exchange activity in hypertrophied canine ventricular myocytes: potential link between contractile adaptation and arrhythmogenesis. Circulation. 2000;102:2137–2144. doi: 10.1161/01.cir.102.17.2137. [DOI] [PubMed] [Google Scholar]
  • [104].de Groot SH, Schoenmakers M, Molenschot MM, Leunissen JD, Wellens HJ, Vos MA. Contractile adaptations preserving cardiac output predispose the hypertrophied canine heart to delayed afterdepolarization-dependent ventricular arrhythmias. Circulation. 2000;102:2145–2151. doi: 10.1161/01.cir.102.17.2145. [DOI] [PubMed] [Google Scholar]
  • [105].Pogwizd SM, Schlotthauer K, Li L, Yuan W, Bers DM. Arrhythmogenesis and contractile dysfunction in heart failure: Roles of sodium-calcium exchange, inward rectifier potassium current, and residual beta-adrenergic responsiveness. Circ Res. 2001;88:1159–1167. doi: 10.1161/hh1101.091193. [DOI] [PubMed] [Google Scholar]

RESOURCES