Abstract
Ventricular arrhythmia is the leading cause of sudden cardiac death (SCD). Deranged cardiac metabolism and abnormal redox state during cardiac diseases foment arrhythmogenic substrates through direct or indirect modulation of cardiac ion channel/transporter function. This review presents current evidence on the mechanisms linking metabolic derangement and excessive oxidative stress to ion channel/transporter dysfunction that predisposes to ventricular arrhythmias and SCD. As conventional anti-arrhythmic agents aiming at ion channels have proven challenging to use, targeting arrhythmogenic metabolic changes and redox imbalance may provide novel therapeutics to treat or prevent life-threatening arrhythmias and SCD.
Keywords: sudden cardiac death, arrhythmias, oxidative stress, metabolism
Introduction
Sudden cardiac death (SCD) refers to death following an unexpected sudden cardiac arrest in a patient with or without known structural heart disease. The incidence of SCD in the United States ranges from 300,000 to 460,000 events per year,1,2 depending on the criteria for SCD used for surveillance. SCD results from a complex interaction between pre-existing cardiac substrates, either structural or genetic, with superimposed physiological or environmental triggers. The most common underlying etiological disorders for SCD in adults age 35 and older are coronary heart disease (CHD, 65–80%)2, 3 and dilated cardiomyopathy (DCM, 10–15%).3 Various types of cardiomyopathy (e.g., hypertrophic cardiomyopathy, arrhythmogenic right ventricular cardiomyopathy, infiltrative, inflammatory, and valvular diseases), genetically determined rhythm disorders (e.g., long QT syndrome, Brugada syndrome, and catecholaminergic polymorphic ventricular tachycardia) or developmental disorders (anomalous origins of coronary arteries) account for most of the remaining SCDs.2, 4 The epidemiology and etiologies of SCD have been extensively reviewed previously4 and in the article on “Epidemiology of Sudden Cardiac Death” in this compendium series [Ref].
The physiological mechanisms reported to cause SCD vary with the patient population and the criteria used to define SCD.5, 6 In general, ventricular tachyarrhythmias, including ventricular fibrillation (VF) and pulseless ventricular tachycardia (VT), are the most common electrophysiological mechanisms leading to SCD.5, 7 Other physiological events that can result in SCD include pulseless electrical activity (PEA), bradyarrhythmias and asystole.8, 9 This review will focus only on SCDs attributed to ventricular tachyarrhythmias.
Ventricular tachyarrhythmias, like all cardiac arrhythmias, result from one of three primary mechanisms: reentry, abnormal automaticity, or triggered activity. Reentry is the most common mechanism for ventricular tachyarrhythmias, involving the presence of an anatomical or functional disturbance of cardiac electrical impulse propagation and of heterogeneous conduction. Abnormal automaticity results from the accelerated generation of an action potential by a region of ventricular cells. Triggered activity occurs when an action potential elicits subsequent depolarizations earlier (early afterdepolarization, EAD) or later (delayed afterdepolarization, DAD) in the repolarization phase.
All three mechanisms can result from abnormal functioning of myocardial ion channels and transporters, leading to disordered initiation or propagation of cardiac action potentials. Accumulating evidence suggests that altered cardiac ion channel/transporter function is closely linked to abnormal myocardial metabolic activity and imbalanced redox states in a wide range of cardiac pathology. This review presents the current evidence on the acute effects of abnormal myocardial metabolism and increased oxidative stress on myocardial ion channel/transporters that predispose to ventricular arrhythmias and SCD. It is important, however, to recognize that abnormal metabolism and oxidative stress in non-cardiac myocytes tissues can also contribute to the development of ventricular arrhythmias. For example, altered metabolism and redox state have been implicated in vascular tissue leading to atherosclerosis,10–13 which underlies the main substrate for SCD and CHD.2, 3 Another example is that abnormal metabolism and increased oxidative stress affect autonomic nervous system, thereby contributing to ventricular arrhythmias and SCD.14–17 Within the ventricle, chronic effects of such entities as diabetes and ischemia from atherosclerosis work through mechanisms involving metabolic and oxidative abnormalities to create the substrate of structural heart diseases that leads to SCD. Both the role of aberrant metabolism and oxidants in the chronic creation of such substrate will not be further considered.
Overview of cardiac ionic channels and membrane excitability
Normal functioning of the mammalian heart depends on proper electrical activity involving the initiation of the electrical impulse from pacemaker cells, the propagation of the electrical activity through specialized conduction system and myocardium, and the generation of action potentials in individual myocytes.18, 19 A normal cardiac cycle begins with the action potential originating from the cells in sinoatrial node, conducts through the atria, atrioventricular node, His bundle, Purkinje fibers and then spreads throughout the entire ventricular myocardium.20 The proper propagation of cardiac electrical impulse depends on low resistance pathways between cells via gap junctions, which are formed by docking of two connexin hemichannels on appositional sarcolemmal membranes.21
Cardiac action potentials are generated through the coordinated activity of various types of ion channels and transporters (Figure 1). Inward current conducting through the voltage-gated Na+ channel rapidly depolarizes the cell (Phase 0), which is followed by early repolarization (Phase 1) attributed to the transient outward K+ current (Ito). Depolarizing L-type Ca2+ currents (ICaL) and multiple repolarizing delayed rectifier K+ currents (IK) form the plateau phase (Phase 2), which is followed by rapid repolarization (Phase 3) with inactivated ICaL and increasing IK. The repolarization continues with the contribution of inwardly rectifying K+ current, IK1, until return to the resting membrane potential (Phase 4). The Ca2+ influx during the action potential triggers the opening of ryanodine receptor 2 (RyR2) on the sarcoplasmic reticulum (SR), triggering the release of Ca2+ from SR into the cytosol. Ca2+ then binds to troponin-C of the troponin-tropomyosin complex, resulting in sarcomere longitudinal shortening.
Figure 1. Action potential waveform and major ion channels/transporters in human ventricular cardiac myocytess.

The action potential (AP) waveform begins with the activation of Na+ channels (INa), leading to the rapid depolarization (phase 0), followed by inward repolarizing transient outward K+ currents (Ito) in phase 1, forming the notch of the AP waveform. The balance between the inward L-type Ca2+ current (ICaL) and outward delayed rectifier K+ currents (IK) contributes to the plateau (phase 2) of the AP waveform. IK continues to repolarize the membrane potential (phase 3) and the inwardly rectifying K+ currents (IK1) contributes to the later part of phase 3 repolarization and the maintenance of resting membrane potential (phase 4). During the AP, Ca2+ influx via ICaL triggers the release of Ca2+ ions from the sarcoplasmic reticulum (SR) via ryanodine receptor 2 (RyR2), coined “Ca2+-induced Ca2+-release.” During the diastolic phase, sarco/endoplasmic reticulum Ca2+-ATPase (SERCA) retrieves cytosolic Ca2+ to SR, and Na+/Ca2+ exchanger (NCX) extrudes Ca2+ from the cell, bringing cytosolic [Ca2+] back to baseline. Sarcolemmal KATP channels (sarcKATP), gated by intracellular ATP/ADP levels and acidosis, does not contribute to AP waveform under normal circumstances, but play an important role in regulating cellular metabolism and electrophysiological responses to metabolic and oxidative stresses. Na+/K+ ATPase consumes ATP to pump Na+ outside of the cell in exchange for K+, which is critical for the maintenance of Na+ and K+ gradients across the plasma membrane.
During the diastolic phase, Ca2+ is removed from the cytosol through sarco/endoplasmic reticulum Ca2+-ATPase (SERCA), which pumps cytosolic Ca2+ into SR, as well as Na+/Ca2+ exchanger (NCX), an antiporter membrane protein extruding one Ca2+ ion from the myocyte in exchange for three Na+ ions. Importantly, the maintenance of Na+ and K+ gradients across the plasma membrane is mediated by the Na+/K+ ATPase, one of the more energy-consuming cellular processes in cardiac myocytess. The energy stored in the resultant electrochemical Na+gradient and K+ gradient provides the driving force for many other transporter/exchangers. Conditions that impair the activity of the aforementioned ion channels/transporters can result in abnormal myocardial electrical functioning, leading to ventricular arrhythmias and SCD.
Myocardial metabolism and energetics
Under physiological conditions, more than 90% of ATP produced in the cardiac myocytes is supplied by mitochondria via oxidative phosphorylation (OXPHOS), and the remainder comes from glycolysis and GTP derived from the tricarboxylic acid (TCA) cycle.22 As the predominant energy generator in the heart, mitochondria occupy ~30% of the volume of cardiac myocytess.23 The mitochondria have double-membrane structure (inner and outer membranes) that forms separate compartments, the intermembrane space and mitochondrial matrix. Mitochondria metabolize predominantly fatty acids but can metabolize glucose to generate ATP via OXPHOS. Acetyl-CoA, the metabolic intermediate derived from sequential oxidation of glucose and fatty acid, is utilized to generate the reducing equivalents, nicotinamide adenine dinucleotide (NADH) and flavin adenine dinucleotide (FADH2), in the TCA cycle. NADH and FADH2 feed electrons to the electron-transport chain (ETC) residing on the mitochondrial inner membrane (Figure 2), where the electrons pass sequentially through complex I, II, III and IV and finally interact with molecular O2 to form H2O. Redox reactions occur at complex I, III and IV, driving proton (H+) from mitochondrial matrix into the intermembrane space. The resulting proton gradient and strongly negative mitochondrial membrane potential ΔΨm (~−150 to −180 mV)24 help to drive H+ flow back to the matrix through mitochondrial ATP synthase (complex V), allowing the conversion of ADP to ATP (Figure 2).
Figure 2. Oxidative phosphorylation and ROS production in mitochondria.

The diagrams of the mitochondrial inner membrane show key components of the electron transport chain (ETC) above, and channels and transporters (below). Reducing equivalents NADH and FADH2 produced from the TCA cycle feed electrons to the ETC along the mitochondrial inner membrane. The electrons flow through the ETC with the following sequence: complex I & II → coenzyme Q [Q] → complex III → cytochrome C [Cyt C] → complex IV → O2, during which coupled redox reactions drive H+ across the inner membrane, forming the proton gradient and the negative mitochondrial membrane potential, ΔΨm (~-150 to −180 mV). The free energy stored in the proton gradient and ΔΨm then drive H+ through the mitochondrial ATP synthase (complex V), converting ADP to ATP. An estimated 0.1–1% of the electrons leak prematurely to O2 at complexes I, II or III, resulting in the formation of superoxide (O2•−). Multiple important mitochondrial channels located on the inner membrane including mitochondrial KATP channels (mitoKATP), the mitochondrial Na+/Ca2+ exchanger (mitoNCX), the inner membrane anion channel (IMAC), the permeability transition pore (PTP) and the mitochondrial calcium uniporter (MCU) also contribute to the regulation of mitochondrial function, myocardial ROS and cellular cation homeostasis.
In healthy adult myocardium, 60–90% of acetyl-CoA comes from β-oxidation of fatty acids, whereas 10–40% comes from oxidation of pyruvate derived from glycolysis and lactate oxidation.25, 26 While the majority of cardiac ATP is consumed by the myofilaments, it is estimated that ~25% of cardiac ATP hydrolysis is used to fuel sarcolemmal and SR ion channels and transporters.27 The critical dependence of cardiac ion channels and transporters on energy supply from metabolized carbon fuels becomes evident under conditions such as myocardial ischemia, during which the mismatch between ATP supply and utilization can readily disrupt the cardiac rhythm through depleting energy supply to these channels and transporters.28, 29
Cardiac oxidants and oxidative stress
In biological systems, partially reduced forms of oxygen (O2) lead to the formation of “oxidants” or “reactive oxygen species (ROS)”, which oxidize other molecules when these molecules accept electrons. The main ROS in cardiac myocytess exist in radical forms such as superoxide (O2•−) and hydroxyl radicals (•OH) or in non-radical forms such as hydrogen peroxide (H2O2), hypochlorite (HOCl), nitric oxide (NO•) and peroxynitrite (ONOO−). It is important to note that although grouped under the acronym of “ROS”, different reactive species containing molecular oxygen vary greatly in diffusion coefficient, reactivity and oxidization potential.
Under physiological conditions, trace amounts of “signaling ROS”, which are tightly regulated by the balance of ROS production and ROS scavengers, form a network that coordinates metabolism with gene transcription and enzymatic activity.30, 31 Low levels of ROS produced during short periods of ischemia, for example, trigger signaling pathways that confer cardiac protection, a phenomenon coined “ischemic preconditioning (IPC)”, which is mediated, at least in part, through mitochondrial KATP channels (see below).32 In addition, low levels of NADPH oxidase 2 (NOX2)-mediated ROS production sensitizes nearby RyRs and triggers Ca2+ sparks in cardiac myocytess in response to physiological stretch with increased preload33 and afterload,34 a process termed X-ROS signaling. Excessive ROS leads to detrimental reactions with cellular lipids and proteins, eliciting maladaptive responses and other abnormalities that compromise cellular and tissue functions, ultimately leading to cell death.35–37 The deleterious effects caused by excessive ROS are defined as oxidative stress.
The main sources of ROS in cardiac myocytess include nicotinamide adenine dinucleotide phosphate (NADPH) oxidases, xanthine oxidase, nitric oxide synthase (NOS) and mitochondria.38, 39 Figure 3 illustrates the major ROS synthesis pathways in cardiac myocytess. NADPH oxidases are membrane-bound enzymes that use NADH or NADPH as the electron donor to generate O2•− from O2.40 Xanthine oxidase catalyzes the oxidation of hypoxanthine/xanthine and produces O2•−.41 NOS, on the other hand, produces a single nitrogen radical (NO•) through metabolizing L-arginine to L-citrulline. NOS becomes uncoupled under conditions such as depletion of L-arginine, oxidation of the NOS cofactor tetrahydrobiopterin (BH4), or increased S-glutathionylation of NOS. Under these conditions NOS produces O2•− instead of NO•.42, 43 NO• itself is a weak thiol oxidant, but it regulates various physiological functions by activating cGMP/protein kinase G (PKG) pathways or by covalent modification of protein cysteine residues via S-nitrosylation (SNO) or S-glutathiolation (GSS-);44, 45 NO• also reacts with O2•− and forms a more potent oxidative molecule peroxynitrite (ONOO−). Oxidative molecules derived from NO•, such as NO2•, N2O3 and ONOO−, are termed reactive nitrogen species (RNS).44, 45 As the metabolic hub in cardiac myocytess, mitochondria consume most of the O2 in cardiac myocytess for energy production, thereby serving as the major source of ROS in the heart.
Figure 3. Major ROS generation pathways in cardiac myocytess.

Mitochondria represent the predominant source of ROS in cardiac myocytess, producing superoxide (O2•−) by the premature leak of electrons from the respiratory chain at complexes I, II or III. Mitochondrial MnSOD can dismutate O2•− to hydrogen peroxide (H2O2), which can be broken down by transition metals such as Fe2+ and converted to hydroxyl radicals (•OH). Superoxide can also be generated by the reduction of O2 by free electrons released from (1) NAD(P)H oxidases, which oxidize NADH or NADPH to form NAD+ or NADP+, (2) xanthine oxidase, which catalyzes the conversion of hypoxanthine to xanthine or xanthine to uric acid, (3) endothelial nitric oxide synthase (eNOS), which switches from a coupled state (producing NO•) to an uncoupled oxide (producing O2•−) under conditions such as the reduction in L-arginine, the deficiency of cofactor tetrahydrobiopterin (BH4), or increased glutathionylation of eNOS. Superoxide can also interact with preformed NO• and generate peroxynitrite ONOO−.
In the mitochondria, ROS are generated as an inevitable byproduct of OXPHOS, resulting from the premature leak of electrons from the ETC to molecular oxygen. It has been estimated that 0.1–1% of the electrons passing through the ETC prematurely leak to O2 at complexes I, II or III, leading to O2•− formation (Figure 2).46 The ROS production rate in the mitochondria is accelerated by increased proton motive force, by the increased NADH/NAD+ ratio, by the increased ratio of reduced coenzyme Q10 (CoQH2) to coenzyme Q10 (CoQ), or when the O2 concentration is raised above physiological levels.47,48 Paradoxically, mitochondrial ROS production also increases under condition of very low local O2 concentration.47, 49 Mitochondrial O2•- is converted to hydrogen peroxide (H2O2) and O2 by manganese superoxide dismutase (MnSOD), the primary antioxidant enzyme in mitochondria.50 H2O2 is then converted to H2O and O2 by cytosolic and mitochondrial scavenging enzymes, glutathione peroxidases, peroxiredoxins and catalase. The emission of mitochondrial ROS is determined by the tight balance between the redox-dependent ROS production and elimination. The concept of “redox-optimized redox balance” implies that a reduced mitochondrial redox state (high NADH) not only favors mitochondrial ROS production, but at the same time optimizes ROS elimination through NADPH-dependent ROS scavenging system such as glutathione (GSH) and thioredoxin.51 Cardiac mitochondria are optimized to keep ROS emission at minimum by operating at an intermediate redox state, where ROS emission is highest under very oxidized or very reduced conditions.51, 52 During conditions with marked oxidation (such as heart failure), depletion of the antioxidative capacity leads to excessive ROS outflow in spite of reduced O2•− production from the ETC, whereas during highly reduced conditions (such as ischemia), excessive O2•− production from the ETC overwhelms the antioxidative capacity, in spite of the abundant NADPH levels.51–53
The mitochondrial ETC efficiency is impaired under pathological conditions such as myocardial ischemia54, 55 and heart failure,56, 57 resulting in increased electron leak and ROS production. Accumulating ROS levels trigger the opening of mitochondrial channels including mitochondrial permeability transition pore (PTP)58 and inner membrane anion channel (IMAC) (Figure 2),59 leading to depolarization of the ΔΨm and further acceleration of ROS production, a phenomenon termed “mitochondrial ROS-induced ROS release.” 60–62 The PTP is a non-selective channel residing on the inner mitochondrial membrane, and it opens with elevated matrix [Ca2+], increased ROS and phosphate levels, as well as by reduced adenine nucleotides (ADP or ATP),63 Brief and reversible PTP opening is considered physiological and serves as a release mechanism to prevent overt mitochondrial cation (especially Ca2+) and ROS accumulation.64 During myocardial ischemia, myocytes experience Ca2+ overload, high phosphate and low ATP concentrations. These conditions prime PTP for opening when reperfusion triggers ROS generation and mitochondrial Ca2+ accumulation, leading to prolonged PTP opening and burst ROS release from mitochondria.63, 64 Pharmacological inhibition of PTP (with cyclosporine A or sanglifehrin A) provides protection against ischemia-reperfusion injury in various experimental models.65, 66 Recent evidence suggests that dimers of mitochondrial ATP-synthase form the pore-forming component of PTP,67 and the c subunit of the ATP-synthase complex may be required for PTP-dependent mitochondrial fragmentation and cell death.68 IMAC, on the other hand, is responsible for anionic currents across the mitochondrial membrane, allowing the passage of O2•− generated by ETC from the matrix to the cytoplasmic side of the inner membrane, thereby serving as a mitochondrial O2•− efflux and ΔΨm depolarization mechanism.69 During ischemia and reperfusion, mitochondrial O2•− production is increased, which leads to increased permeability of IMAC, resulting in a burst in ROS release and ΔΨm depolarization. The resulting reversible Δψm collapse and membrane potential oscillation foster lethal ventricular arrhythmias, which can be abrogated by IMAC inhibitors.69 During mitochondrial ROS-induced ROS release, depolarized mitochondrial ΔΨm depletes NADPH (the level of which is coupled to [NADH] by nicotinamide nucleotide transhydrogenase in mitochondria), thereby leading to rapid dissipation of ROS-eliminating capacity, allowing acceleration of ROS emission from mitochondria. 51, 60–62
Elevated mitochondrial ROS also impairs ATP production through various mechanisms. Excessive ROS cause oxidative damage to components of the ETC such as complex I, II,70 IV71 and ATP synthase,70 impairing ATP generation and further aggravate electron leakage. In addition, O2•− activates mitochondrial uncoupling proteins, resulting in increased electron leak and uncoupled O2 consumption from ATP synthesis.72, 73 Taken together, excessive cardiac ROS leads to abnormal electrical function directly by ROS-mediated signaling and oxidative damage, as well as indirectly by decreasing ATP generation that is required for normal ion channel/transporters functioning.28, 29
Impaired cardiac metabolism and increased myocardial oxidative stress, depending on the etiology and clinical course, can be divided into two categories: acute and chronic. Acute cardiac oxidative and metabolic derangements typically occur with myocardial ischemia/reperfusion that results from coronary artery occlusion. Chronic oxidative and metabolic abnormalities, in contrast, occur with conditions such as cardiac hypertrophy, heart failure and diabetes mellitus. The impact of chronic oxidative and metabolic stresses on cardiac electrophysiology often involves changes in the transcript and protein expression of cardiac ion channels/transporters, which is commonly defined as “electrical remodeling.”74 Electrical remodeling during these chronic cardiac conditions have been extensively reviewed elsewhere.74, 75 Aside from a few exceptions, therefore, this review will focus on the effects of acute myocardial metabolic and oxidative stress on cardiac electrophysiology.
Mechanisms linking impaired cardiac metabolism to ventricular arrhythmias and SCD
A. Metabolic regulation of cardiac Ca2+ handling and homeostasis
Calcium ions are critical intracellular signaling molecules, responsible for the regulation of numerous cellular processes in cardiac myocytess including excitation-contraction coupling, gene transcription, enzyme activity and cell death.76 The intracellular [Ca2+] concentration fluctuates markedly between systole and diastole, yet the changes in cytosolic [Ca2+] are tightly regulated. Multiple signaling molecules, including calcium/calmodulin-dependent protein kinase II (CaMKII), protein kinase A (PKA) and protein kinase C (PKC), are involved in the regulation of these Ca2+ handling proteins (see review by Bers).76 Impaired Ca2+ homeostasis and handling have been implicated in the mechanical dysfunction and arrhythmogenesis observed in acute myocardial ischemia,77, 78 as well as in chronic cardiac conditions such as cardiac hypertrophy79, 80 and heart failure.81, 82
During myocardial ischemia or metabolic inhibition, various metabolic parameters are markedly altered in cardiac myocytess. For example, ATP levels decrease, cells become progressively acidic with elevated lactate levels,83 and the levels of phosphate84 and magnesium85 increase. Both reduced ATP and elevated phosphate levels can inhibit Na+/K+ ATPase activity,86, 87 leading to intracellular Na+ accumulation. Increased late Na+-currents (late INa) also contributes to increased intracellular [Na+] during myocardial ischemia88–90 and heart failure.91 Elevated intracellular [Na+] leads to increased cytosolic [Ca2+], at least in part, through decreased extrusion of Ca2+ or through actual Ca2+ entry with NCX activity in the reverse mode (Na+ out and Ca2+ in).92–94 The activity of SR Ca2+ uptake (through SERCA) and release (through RyR2) are both inhibited during myocardial ischemia;28 however, RyR2 inhibition appears to predominate over reduced SERCA activity during ischemia, reflected in lower frequency of spontaneous Ca2+ release through RyR228, 95 and increased SR Ca2+ load.28 Upon reperfusion or re-oxygenation, RyR2 is released from inhibition, producing spontaneous waves of Ca2+ release,96 which may contribute to calcium transient/action potential alternans and increase the risk of ventricular arrhythmias (Figure 4 and Table 1).97, 98
Figure 4. Effects of metabolic stresses on cardiac ion channel/transporter function.

Schematic illustration of the impact of acute (ischemia/hypoxia) and chronic (heart failure and diabetes) metabolic stresses on cardiac ion channel/transporter function and the electrophysiological consequences. Acidosis during acute ischemia leads to opening of sarcKATP channels, increased activity of NHE, and increased late INa. ATP depletion during ischemia contributes to the opening of sarcKATP channels, as well as impairs the function of Na/K ATPase. Other electrophysiological effects from ischemia include the reduction of peak INa, increased connexin conduction, and reduced RyR2 activity. Reduced RyR2 activity during acute ischemia is rapidly reversed upon reperfusion, contributing to the spontaneous waves of Ca2+ release and DADs. A few examples of electrophysiological impact from chronic metabolic stresses associated with heart failure and diabetes are also illustrated: heart failure reduces peak INa and multiple Kv (Ito, IK) currents. Diabetes is also associated with reduction in Kv currents. The increase in sarcKATP currents lead to APD shortening, whereas reduced Kv currents and increased late INa lead to prolonged APD and EADs. Finally, reduced Na/K ATPase activity, increased RyR2 activity and increased gap junction conduction during ischemia predispose to arrhythmogenic DADs.
Table 1.
Effects of metabolic changes on cardiac ion channel/transporter function and arrhythmogenicity
| Channel/Transporter effects | Metabolic changes | Effects on electrical/ionic homeostasis | Pro-arrhythmic mechanism |
|---|---|---|---|
| Na+/K+ ATPase ↓ | Ischemia/hypoxia | Na+ overload | Ca2+ overload and DAD |
| Cx hemichannel ↑ | Ischemia | Na+ overload | Ca2+ overload and DAD |
| Peak INa ↓ | Ischemia/heart failure | ↓ Na+ influx | Slow conduction |
| Late INa ↑ | Ischemia/hypoxia acidosis, ↑ LPC AMPK mutation | ↑ Na+ influx, prolonged APD | EAD |
| Kv ↓ | Diabetes, Heart failure | ↓ K+ influx, prolonged APD | EAD |
| Kv↑ | Insulin treatment in diabetic heart, PI3Kα activation, Exercise training | ↑ K+ channel expression | Protective |
| IKATP ↑ | Ischemia | ↑ K+ influx, shortened APD | Current sink, slow conduction |
| RyR2 | ↓ during ischemia; ↑ upon reperfusion | SR Ca2+ load ↑, spontaneous Ca2+ waves ↑ | Ca2+ transient/action potential alternans |
Cx: connexin; Peak INa: peak Na current ; Late INa: late Na current; Kv: voltage-gated K current; IKATP: ATP-sensitive K+ current; RyR2: ryanodine receptor 2; APD: action potential duration; SR: sarcoplasmic reticulum; EAD: early after-depolarization; DAD: delayed after-depolarization; LPC: lysophosphatidylcholine.
In addition to the ATP production machinery, mitochondria also harbor Ca2+ channels and transporters, allowing the transfer of Ca2+ ions between cytosol and mitochondria, thereby contributing to the dynamic regulation of sarcolemmal Ca2+ homeostasis. Mitochondria uptake Ca2+ predominantly through the mitochondrial Ca2+ uniporter (MCU) and extrude Ca2+ mainly through the mitochondrial Na+-Ca2+ exchanger (mitoNCX) (Figure 2).99,100 The capacity of Ca2+ uptake and release by mitochondria forms an additional buffer for cytosolic Ca2+ regulation, contributing to the spatiotemporal dynamics of Ca2+ signaling in cardiac cells.101, 102 Nevertheless, it remains debatable how much mitochondria contribute to cellular Ca2+ dynamics under physiological and pathological conditions.103 Recently Williams et al.102 provide quantitative data suggesting that mitochondria do not function as a significant buffer of cytosolic Ca2+ under physiological conditions; with prolonged elevation of cytosolic [Ca2+], however, mitochondrial Ca2+ uptake can increase up to 1000-fold and impact cellular Ca2+ dynamics significantly.102 Nevertheless, it has been shown that mitochondrial dysfunction results in cardiac myocytes Ca2+ transient alternans by impairing the mitochondrial capacity of Ca2+ handling, thereby predisposing to cardiac arrhythmias.104 In addition, pharmacological inhibition of mitochondrial uptake through MCU with Ru360 has been shown to reduce the incidence of ventricular arrhythmias induced by ischemia-reperfusion in the rodent heart.105 As the metabolic center in cardiac myocytess, therefore, mitochondria play an important role in transducing changes in cardiac metabolic states to the dynamic regulation of sarcolemmal Ca2+, membrane excitability and electrical functioning.
B. Metabolic regulation of Na+ homeostasis and Na+ channel function
Intracellular [Na+] plays a critical role in regulating the energetics, electrical functioning and contractility of cardiac myocytess, which can be attributed to the direct and indirect effects of cytosolic [Na+] on Ca2+ homeostasis,92 mitochondrial function,106 and cellular signaling.107 Multiple ion channels/transporters function to maintain the [Na+] homeostasis in cardiac myocytess: Na+/K+ ATPase, for example, consumes ATP to pump Na+ outside of the cell in exchange of K+, representing the main mechanism of Na+ efflux, whereas Na+ channels, NCX and the Na+/H+ exchanger (NHE) mediate Na+ influx in cardiac myocytess.
Cardiac voltage-gated Na+ channels form by the assembly of a pore-forming α subunit and auxiliary β subunits that modulate channel activities. Nav1.5 (SCN5A) is the predominant Nav α subunit expressed in the mammalian myocardium. Currents conducting through Na+ channels generate the rapid upstroke (phase 0) of the action potential. In addition, Na+ channels, together with cardiac gap junctions, govern the electrical conduction velocity in the myocardium. There are two effects of Na+ channel/transporter dysfunction on arrhythmogenesis: (1) through affecting Na+-Ca2+ homeostasis, thereby promoting delayed afterdepolarizations (DADs) and creating reentry substrates, and (2) through electrophysiological effects of the channel function. For example, increased late INa predisposes to APD prolongation and EADs, both of which are arrhythmogenic and predispose to SCD (Table 1 and Figure 4).
During the metabolic stress that ensues with myocardial ischemia/reperfusion, reduced Na+/K+ ATPase (because of reduced ATP and elevated ADP/phosphate levels), increased late INa (see below for detailed mechanisms)108 and enhanced NHE (owing to increased cytosolic acidosis) activity109, 110 result in increased cytosolic [Na+], thereby leading to Ca2+ overload and increased propensity for ventricular arrhythmias (see above). It has been shown that pharmacological inhibition or knockout of NHE renders cardiac function resistant to ischemia-reperfusion injury.111
In addition to altered Na+/K+ ATPase and NHE function, it has been demonstrated that ischemia/metabolic inhibition can trigger Na+ influx through connexin (Cx) hemichannels.112, 113 Under normal conditions Cx hemichannels located on adjacent intercellular sarcolemmal membranes combine to form gap junctions at the intercalated discs between cardiac myocytess. During ischemia/metabolic inhibition, Cx hemichannels localize to the nonjunctional membrane, turning into non-selective cation channels that are permeable to Na+, K+ and Ca2+.112, 114 With the high conductance of these Cx hemichannels, it is suggested that the opening of even a small number of Cx43 hemichannels can lead to doubling of Na+ influx in ventricular cardiac myocytess,114, 115 contributing to increased Na+ load and increased propensity for arrhythmias (Table 1 and Figure 4).
Elevated cytosolic [Na+] also impairs cardiac metabolism reciprocally through affecting mitochondrial function. Mitochondrial [Na+] is regulated by NHE-mediated Na+ uptake116 and mitoNCX-mediated Na+ efflux.117 Under physiological conditions, energized mitochondria extrude H+, and the resulting pH gradient drives the Na+ gradient between mitochondrial matrix (lower [Na+]) and cytosol (higher [Na+]).116, 118 Intracellular Na+ levels increase significantly in pathological conditions such as myocardial ischemia.109 The rise in cytosolic [Na+] during ischemia widens the Na+ gradient across mitochondria, leading to stronger driving force for Ca2+ extrusion from mitochondria via mitoNCX, thereby leading to reduced mitochondrial [Ca2+] and altered mitochondrial energetics,100, 106 including decreased activities of Ca2+-dependent dehydrogenases in the TCA cycle, the net oxidation of the matrix NADH/NADPH pool, and hence the deficiency of NADPH-dependent ROS scavengers, thereby increasing mitochondrial ROS emission.119, 120 Increased ROS levels further aggravate cytosolic Na+ accumulation by increasing late INa through direct Na+ channel modification121, 122 or by activating signaling molecules such as PKC123 or CaMKII (see below).124 Therefore, there may be a feed-forward mechanism between increased cytosolic [Na+], impaired mitochondrial metabolism, increased ROS emission, and ROS-induced elevation of cytosolic [Na+].125
In addition to cytosolic [Na+], cardiac metabolism also influences electrical functioning through modulating sarcolemmal Na+ channel activity.126 During myocardial ischemia, various metabolic derangements result in altered Na+ channel properties and function. The ischemic metabolite lysophosphatidylcholine (LPC), for example, has been shown to decrease peak Na+ current (peak INa) and to slow its inactivation.127 In addition, LPC also causes a marked increase in late INa, thereby increasing myocardial Na+ loading.88 Myocardial acidosis during ischemia is also known to affect the inactivation states of Na+ channel, leading to increased late INa (Table 1 and Figure 4).89, 90 Cardiac Na+ channels are reported to be modulated by adenosine monophosphate-activated protein kinase (AMPK). AMPK is a serine/threonine kinase that is activated by increased cytosolic AMP/ATP ratio to increase ATP production and decrease ATP consumption, thereby functioning as a master metabolic regulator in cardiac myocytess.128, 129 Mutations in PRKAG2 gene, which encodes the γ2 subunit of AMPK, have been reported to be associated with electrical instability and lethal ventricular arrhythmias.130–132 For example, PRKAG2 T172D, a constitutively active AMPK mutant, has been shown to slow Na+ channels inactivation, leading to increased late INa, prolonged APD and arrhythmogenic EADs130
C. Metabolic regulation of K+ channels and cardiac repolarization
Various types of voltage-gated K+ (Kv) channels and non-voltage-gated inwardly rectifying (Kir) channels contribute to myocardial action potential repolarization.19, 20 Under pathological conditions such as myocardial ischemia, diabetes and heart failure, repolarizing Kv currents in ventricular cardiac myocytess are reduced, leading to delayed repolarization and prolonged APD, which predispose to the development of ventricular arrhythmias and SCD (Figure 4).20
Among several types of Kir channels functioning in mammalian heart, IK1 contributes to the terminal phase of repolarization and the maintenance of resting membrane potentials in ventricular myocytes,20, 133, 134 whereas currents conducted via sarcolemmal KATP (sarcKATP) channels (IKATP), gated by intracellular ATP/ADP levels and acidosis,135 function as immediate sensors of cellular metabolism and play an essential role in electrophysiological responses to metabolic stresses such as myocardial ischemia.136, 137
Abnormal cardiac repolarization, including QT-interval prolongation and T wave abnormalities, are often observed in patients with diabetes, one of the leading chronic metabolic disorders associated with ventricular arrhythmias and sudden cardiac death.138, 139 Studies on cellular mechanisms of diabetes-induced repolarization abnormalities have consistently revealed downregulation of cardiac Kv currents in diabetic heart.140–142 Multiple mechanisms have been proposed to account for reduced Kv currents during diabetes. First, post-translational inhibition of Kv channel function occurs in diabetes. Plasma levels of free fatty acid metabolites, such as palmitoylcarnitine and palmitoyl-CoA, are increased with diabetes. They have been shown to inhibit cardiac Kv currents directly.143 Acute treatment with insulin has been demonstrated to reverse diabetes-induced Kv current reduction, suggesting the primary role of insulin-dependent signaling on Kv channel function.140 It has been suggested that improved glucose utilization by insulin may be responsible for restored Kv current expression, as compounds known to improve glucose utilization (such as dicholoroacetate or L-carnitine) have been shown to normalize cardiac Kv current density with diabetes.141 Increased ROS levels also contribute to reduced cardiac Kv currents during diabetes, and direct incubation with antioxidant GSH has been shown to restore Kv current expression in myocytes isolated from diabetic hearts.140 Second, transcriptional regulation of Kv channel expression may reduce Kv currents. Peroxisome proliferator-activated receptor α (PPARα), a critical regulator of glucose/fatty acid metabolism, is upregulated in diabetic heart and has been suggested to transcriptionally repress cardiac Kv channel expression.144 Cardiac-specific over-expression of PPARα downregulates the mRNA and protein expression levels of Kv4.2 and KChIP2, the channel subunits encoding Ito,144 whereas PPARα knock-out upregulates Ito.144 Interestingly, it has been recently demonstrated that the activation of phosphoinositide 3-kinase α (PI3Kα), a key component of insulin receptor signaling pathway that can be triggered by insulin treatment or exercise training, upregulates mRNA levels of various Kv channels145, 146 through an Akt-independent mechanism,147 which may provide the mechanistic explanation for insulin-mediated Kv current modulation, as well as the metabolic impact of exercise training on cardiac repolarization. It is important to note that the changes in Kv current observed with diabetes or exercise training reflect the chronic effects of metabolic stress on Kv channel remodeling, rather than acute channel effects.
The sarcKATP channels are present at high density in the sarcolemmal membrane. SarcKATP channels are inhibited by ATP and activated by ADP, Mg2+ or low pH, conditions that are associated with ischemia, insufficient fuel supply, and increased metabolic stress.148 Upon increased metabolic stress with ATP depletion and acidosis (e.g., during myocardial ischemia),149 sarcKATP channels are activated, allowing an inwardly rectifying repolarizing K+ current. SarcKATP channels open within seconds in response to acute ischemia, which is likely attributed to the rapid drop in pH (also within seconds of ischemia) in the ischemic tissue.150 ATP levels, by contrast, deplete at a slower rate and remain in the millimolar range (levels that are sufficient to keep sarcKATP channels closed) until 10–15 minutes after the onset of ischemia.151 Therefore, ATP depletion may not contribute to sarcKATP channel opening during the early phase of ischemia.
The cell surface expression density of sarcKATP channels is very high. It has been estimated that the opening of merely 1% of the sarcKATP channels is sufficient to significantly shorten myocardial APD.152 The opening of sarcKATP channels shortens APD and decreases inward Ca2+ currents, thereby reducing Ca2+-mediated cardiac energy consumption and preventing Ca2+ overload-induced cell death. With adequate numbers of sarcKATP channels in the open state, however, cardiac myocytess become hyperpolarized and rendered unexcitable.153 This creates a “current sink” that slows or blocks electrical propagation in the myocardium, predisposing to the development of ventricular arrhythmia (Table 1 and Figure 4).62, 154 Indeed, pharmacological inhibition of sarcKATP channels has been shown to reduce the incidence of ventricular arrhythmias in animal models155–157 and in humans.158–160 Reduced mitochondrial ATP production with impaired respiratory chain function leads to activation of sarcKATP channels. During ROS-induced ROS release, for example, dissipation of ΔΨm results in cessation of mitochondrial ATP production, leading to sarcKATP activation, altered myocyte APDs and even ventricular arrhythmias.59, 62, 161 Interestingly, it has been shown that ATP generated from glycolysis plays a bigger role in modulating sarcKATP activity than ATP generated from OXPHOS,162 which is in accordance with recent findings that the key glycolytic enzymes, including glyceraldehyde 3-phosphate dehydrogenase (GAPDH) and pyruvate kinase, serve as components of the sarcKATP channel macrocomplex and regulate its function.163, 164 SarcKATP channels have also been shown to exhibit reduced sensitivity to ATP-mediated inhibition in diabetic heart, resulting in higher sarcKATP conductance and greater APD shortening in response to hypoxia.165
D. Non-ion channel mechanisms linking metabolism to ventricular arrhythmias and SCD
As mentioned above, AMPK functions as a key metabolic sensor and regulator in cardiac myocytess. It has been reported that specific mutations in PRKAG2, the ϒ2 regulatory subunit of AMPK, lead to a glycogen storage cardiomyopathy characterized by ventricular pre-excitation and cardiac hypertrophy.166, 167 The structural abnormalities associated with PRKAG2 mutations, including the glycogen accumulation and ventricular hypertrophy, result in abnormal conduction that predisposes to ventricular arrhythmias.166, 167 In addition to AMPK, patients with inherited fatty acid oxidation disorders exhibit high incidence of ventricular arrhythmias, conduction defects and SCD,168 which can be attributed to the accumulation of arrhythmogenic intermediary metabolites of fatty acids.168 Fatty acid metabolites such as LPC accumulate during myocardial ischemia, contributing to the formation of arrhythmogenic substrates.88 Consistent with these findings, a transgenic mouse model of cardiac-specific overexpression of fatty acid transport protein 1 (α-MHC FATP1) has been shown to exhibit increased myocardial uptake and accumulation of free fatty acids, leading to arrhythmogenic electrical changes including QT prolongation.169 Taken together, these data suggest that abnormal cardiac metabolism can lead to ventricular arrhythmias and SCD through non-ion channel mechanisms.
Mechanisms linking increased cardiac oxidative stress to ventricular arrhythmias and SCD
A. Myocardial oxidative stress and cardiac Na+ channels
Abnormal cardiac Na+ channel function is often observed in cardiac diseases such as myocardial ischemia170, 171 and heart failure,91, 172 conditions that are associated with increased cardiac oxidative stress. Na+ channel abnormalities can be categorized as loss- or gain-of function. Loss-of-function Na+ channel abnormality leads to reduced peak INa, resulting in conduction block that potentiates re-entrant type ventricular arrhythmias.173 Gain-of-function Na+ channel abnormality, in contrast, results in the increase in late INa, which can lead to prolongation of APDs, EADs, reverse mode NCX transport and subsequent Ca2+ overload in sarcomere, all of which predispose to ventricular arrhythmias.122, 174
ROS have been demonstrated to affect Na+ channel function through multiple mechanisms. Transcriptionally, ROS have been shown to reduce Nav1.5 channel expression by reducing mRNA expression.175, 176 At the protein level, ROS have been shown to impair Nav1.5 channel inactivation through direct oxidation at the methionine residues of Nav1.5 channel protein.177 Elevated mitochondrial ROS are known to reduce peak INa by modifying Nav1.5 conductance through post-translational mechanisms without affecting cell surface expression of Nav1.5.57 The slowly inactivating component of Na+ current (late INa), on the other hand, is known to be augmented by ROS in cardiac myocytess.121, 122, 174 For example, increased H2O2 has been shown to increase the open probabilities of Na+ channel, leading to increased late INa,121, 122 and arrhythmogenic changes including prolongation of APDs, EADs, and cytosolic Ca2+ overload (Figure 5 and Table 2).122, 174 In addition to direct ROS-dependent effects on Na+ channels, ROS also modulate Na+ channel activities indirectly by altering membrane lipid environment178 or by activating signaling molecules such as PKC123 or CaMKII.124 Elevated mitochondrial ROS activate PKC, for example, which may reduce peak INa by affecting Nav1.5 phosphorylation. ROS-induced CaMKIIδ activation, on the other hand, has been shown to mediate H2O2-induced increases in late INa.124
Figure 5. Effects of increased oxidative stress on cardiac ion channel/transporter function.

Schematic figure to demonstrate the impact of increased oxidative stress on cardiac ion channel/transporter function. Increased ROS enhance the activity of RyR2 and sarcKATP channels and increase late INa. Excessive oxidative stress reduces peak INa and repolarizing K+ currents (Ito, IK and IK1) and leads to reduced gap junction protein Cx43 and reduced SERCA activity. Increased ROS inhibit normal operation of NCX (Na+ in, Ca2+ out) and promote the reverse mode of NCX activity (Na+ out, Ca2+ in). The increase in sarcKATP currents lead to APD shortening, whereas reduced Kv currents and increased late INa lead to prolonged APD and EADs. Increased RyR2 activity, reduced SERCA function, and reverse mode NCX activity lead to cytosolic Ca2+ overload and predispose to DADs.
Table 2.
Effects of increased oxidative stress on cardiac ion channel/transporter function and arrhythmogenicity
| Channel/Transporter | Effects of ROS | Effects on electrical/ionic homeostasis | Pro-arrhythmic mechanism |
|---|---|---|---|
| Peak INa | ↓ | ↓ Na+ influx | Slow conduction |
| Late INa | ↑ | ↑ Na+ influx, APD prolonged | EAD |
| Ito | ↓ | ↓ K+ influx, APD prolonged | EAD |
| IK | ↓ | ↓ K+ influx, APD prolonged | EAD |
| IK1 | ↓ | ↓ K+ influx, APD prolonged | EAD |
| IKATP | ↑ | ↑ K+ influx, APD shortened | Current sink, slow conduction |
| NCX | Reverse mode | Ca2+ overload | DAD |
| Cx43 | ↓ | Impaired gap junction function | Slow conduction |
| SERCA | ↓ | Ca2+ overload | DAD |
| RyR2 | ↑ | Ca2+ overload | DAD |
| mitoNCX | Reverse mode | Ca2+ overload | DAD |
| mitoKATP | ↑ | ↑Mitochondrial K+ influx | Protective, ischemic- preconditioning |
Peak INa: peak Na+ current; Late INa: late Na+ current; Ito: transient outward K+ current; IK: delayed rectifier K+ current; IK1: inwardly rectifying K+ current; IKATP: ATP-sensitive K+ current; NCX: Na+/Ca2+ exchanger; Cx43: connexin 43; SERCA: sarco/endoplasmic reticulum Ca2+-ATPase; RyR2: ryanodine receptor 2; mitoNCX: mitochondrial Na+/Ca2+ exchanger; mitoKATP: mitochondrial ATP-sensitive K+ current; APD: action potential duration; EAD: early after-depolarization; DAD: delayed after-depolarization.
Increased myocardial oxidative stress is often associated with impaired cardiac metabolism. The cardiac NADH level, for example, is elevated during cardiac ischemia and cardiomyopathy, which leads to increased mitochondrial ROS production and reduced peak INa. Using a mouse model of nonischemic cardiomyopathy, we have demonstrated recently that the cytosolic NADH and mitochondrial ROS levels are increased with cardiomyopathy, resulting in a cardiac peak INa reduction without altering membrane Na+ channel protein expression levels.57 The reduced peak INa can be restored by mitochondrial antioxidants NAD+ or Mito-TEMPO treatment.57 Consistent with these findings, NAD+ treatment has been shown to improve myocardial conduction abnormality that is associated with reduced peak INa in human failing heart.57
The link between mitochondrial ROS and INa regulation is also observed in the hereditary arrhythmia syndromes. One of the Brugada syndrome-causing mutations, A280V mutation in glycerol-3-phosphate dehydrogenase 1-like (GPD1-L) protein, has been demonstrated to reduce peak INa by increasing cytosolic NADH levels and mitochondrial ROS.123, 179 Peak INa reduction by A280V GPD1-L is abrogated by the treatment of NAD+ or Mito-TEMPO. 123 Taken together, these data suggest the important role of mitochondrial ROS in transducing the altered cardiac metabolic states (NADH/NAD+ levels) to the regulation of cardiac Na+ channels and membrane excitability.
During acute ischemia, myocardial NO• production is enhanced owing to increased NOS (especially eNOS) activity180 and conversion of tissue nitrate to NO•.181 Increased NO• production and elevated NOS activities play a critical role in the cardioprotective effects of IPC (see below Section C), which has been extensively reviewed previously.182, 183 Elevated NO• has been shown to increase late INa,184 which can be abolished by reducing agents GSH and DTT,184, 185 suggesting that NO• modulates Na+ channel activities by oxidizing Na+ channel or its regulatory proteins. Interestingly, others report that NO• decreases peak INa through reducing the open probability and the surface expression of functional Na+ channels, which are mediated indirectly through a PKG/PKA-dependent pathway.186 Peroxynitrite, formed by NO• and O2•− interaction, has also been shown to augment late INa. 187
B. Cardiac redox state and Ca2+ homeostasis
As mentioned earlier in this review, intracellular [Ca2+] is tightly regulated by various Ca2+ handling proteins, including voltage-gated Ca2+ channels, SERCA, RyR2, NCX and signaling molecules such as CaMKII, PKA and PKC. Many of these Ca2+ handling and regulatory proteins harbor methionines or thiol groups that are susceptible to the direct modification by ROS or reducing agents, thereby sensitive to the altered redox states during cardiac diseases.
In general, elevated ROS levels in cardiac myocytess are known to result in a net increase in intracellular [Ca2+].188 The direct application of H2O2 or increased mitochondrial ROS has been shown to increase ICaL and its sensitivity to isoproterenol stimulation in ventricular myocytes.189 In addition, oxidized LDL has been demonstrated to stimulate ICaL through LPC-induced mitochondrial ROS.190 Nevertheless, there are conflicting reports showing reduced cardiac ICaL in the presence of elevated oxidative stress.191, 192 The discrepant findings in the effects of ROS on cardiac ICaL may be explained by the differences in the experimental design, animal species and type of ROS involved.
The results on the impact of increased oxidative stress on other Ca2+ handling proteins, in contrast, are more consistent. Excessive ROS increases the open probability of RyR2, enhancing the release of Ca2+ from SR.193 Calcium sparks from RyR2 are increased in response to photoactivated or chemical-induced mitochondrial ROS.194 In contrast to RyR2, SERCA activity is reduced upon increased cardiac ROS,195–197 which can be, at least in part, attributed to reduced ATP supply for SERCA pump owing to mitochondrial dysfunction.38 Cardiac NCX, which normally extrudes Ca2+ from cytosol in exchange for Na+, has been shown to be activated in the reverse mode by increased oxidative stress, thereby increasing cytosolic [Ca2+] (Table 2, and Figure 5).198, 199
CaMKII is activated by ROS, and it is known to play a critical role in regulating calcium handling proteins in response to increased cardiac oxidative stress.200 For example, activated CaMKII augments ICaL by phosphorylating Cav1.2 channel subunit and increasing its open probability.201 Phosphorylation of RyR2 by CaMKII increases SR Ca2+ leak, promoting cytosolic Ca2+ overload and DADs.200 Taken together, the net impact of increased oxidative stress on Ca2+ handling proteins leads to cytosolic Ca2+ overload and depletion of SR Ca2+ store, resulting in detrimental changes such as arrhythmogenic DADs and contractile dysfunction.
As discussed earlier in this review, mitochondria also play an important role in the regulation of Ca2+ in cardiac myocytess. Increased oxidative stress is known to modulate mitochondrial [Ca2+] by altering mitochondrial ion channel activities, contributing to the perturbation of cytosolic Ca2+ homeostasis. For example, increased ROS during myocardial ischemia depolarize ΔΨm,202 forcing mitochondrial NCX into reverse mode and drives Ca2+ from cytosol into mitochondria.117, 203 Physiologically, mitochondrial Ca2+ is required for normal ETC function and serves as a positive effector of OXPHOS. Ca2+ overload, however, leads to increased mitochondrial ROS production through mechanisms including increased electron leakage,204, 205 enhanced NO production, which is known to inhibit complex IV and augment ROS production from complex III,206 and reduced cytochrome c-mediated respiration.207, 208 In addition, pathological mitochondrial Ca2+ overload has been shown to increase NO and ROS production in cardiac myocytess.209 These data suggest a positive feedback loop between ROS-induced Ca2+-overload and Ca2+-induced ROS production. Under pathological conditions such as myocardial ischemia, cellular and mitochondrial [Ca2+] increase, leading to increased ROS production. ROS over-production induced by elevated mitochondrial [Ca2+] results in further increase in mitochondrial Ca2+ and ROS levels. The Ca2+ load and ROS produced from this positive feed-forward loop can overwhelm the cellular capacity of ROS scavenging and Ca2+ clearance, resulting in cellular damage and electrical instability that predispose to ventricular arrhythmias and SCD.210 It is worth noting that during heart failure, mitochondrial Ca2+ uptake/accumulation is actually reduced (because accumulated cytosolic Na+ during heart failure inhibits mitochondrial Ca2+ uptake via activation of mitoNCX),119, 120 which also favors ROS production (by impairing mitochondrial energetics and oxidizing NAD(P)H ).119, 120
NO• and peroxynitrite also play important roles in modulating cardiac Ca2+ channel and Ca2+ handling proteins. Increased NO• inhibits L-type Ca2+ channel activity through increased S-nitrosylation211, 212 and cGMP-PKG-dependent phosphorylation213 of Ca2+ channel subunits, which may contribute to the cardioprotective effects of IPC by limiting Ca2+ influx during ischemia.212 The effects of NO• on RyR2 is concentration-dependent. Physiological NO• levels (~1 μM) do not affect RyR2 activity, whereas supraphysiological levels (≥100 μM) inhibit RyR2 activity.214 Peroxynitrite, on the other hand, activates RyR2 by oxidizing its cysteine residues.214 SERCA2a, by contrast, is inhibited by peroxynitrite-mediated tyrosine nitration. PKA pretreatment can prevent the inhibitory effects of peroxynitrite on SERCA2a by inducing dissociation of PLN from SERCA2a.215
C. Myocardial ROS and potassium channels
Increased oxidative stress is known to inhibit repolarizing Kv currents including Ito and several delayed rectifier IK (IKr, IKs and IKur) in mammalian ventricular myocytes.216–218 These effects are reversible by raising cellular antioxidant GSH levels.140, 219 ROS have been shown to regulate Kv current expression by reducing the transcript/protein expression of the Kv channel subunits218 and by modulating the phosphorylation of these channel subunits through protein kinases such as PKA, PKC or protein tyrosine phosphatases.220–222 Cardiac Kv channels are also regulated by NO•. It has been reported that NO• inhibits human cardiac Kv4.3 channel, thereby reducing the transient outward K+ current Ito.223 NO• also blocks Kv1.5 channel through S-nitrosylation and the activation of cGMP/PKG pathway.224
Myocardial ROS also regulates myocyte membrane excitability through sarcKATP channels. During myocardial ischemia, fuel substrate deprivation and increased oxidative stress depolarize the mitochondrial network, leading to fluctuated ΔΨm and ATP production levels, which result in oscillation of sarcKATP currents and APDs.69, 225 The ΔΨm depolarization of the mitochondrial network is potentiated by widespread ROS production induced by focal increases in mitochondrial ROS (ROS-induced ROS release).60, 62, 226 IMAC plays a critical role in ΔΨm depolarization upon increased ROS; pharmacologic inhibition of IMAC has been shown to prevent ΔΨm depolarization and the oscillation of sarcKATP currents and APD,59 as well as preventing ventricular arrhythmias in mammalian hearts227, 228 during myocardial ischemia.
Another group of KATP channels, residing on the mitochondrial inner membrane (mitoKATP channels), contribute importantly to the protective effects of IPC, an endogenous cellular protective mechanism involving brief periods of ischemia that confers protection against infarction produced by a subsequent prolonged ischemia.148, 229 Under physiological conditions, mitoKATP channel opening enhances mitochondrial ROS production, triggering downstream signaling pathways involved in gene transcription and cell growth.230 During IPC, the activation of mitoKATP channels allows mitochondrial K+ influx, resulting in partially depolarized ΔΨm, which leads to a compensatory increase in proton pumping and cellular respiration to maintain ΔΨm and ETC activity.229 Partially dissipated ΔΨm induced by mitoKATP channel opening blunts mitochondrial Ca2+ accumulation during ischemia and reduces PTP opening upon reperfusion,231, 232 both of which contribute to the protection against cell death by IPC. The opening of mitoKATP channel also triggers the production of “protective ROS”, which activates a PKC-dependent pathway that confers cardiac protection by reducing deleterious post-ischemic ROS production.233, 234 Interestingly, NO•, a physiologically important free radical, has been shown to potentiate mitoKATP opening235 and triggers both early and delayed IPC.236, 237 In addition, mitoKATP opening during ischemia provides additional K+ influx to maintain mitochondrial volume, which is critical to maintain the normal function of mitochondria and ETC.230 Pharmacologic inhibition of mitoKATP channels abrogates the anti-arrhythmic effects of IPC238 and several mitoKATP channel openers are shown to provide protective effects against ischemia-induced cardiac arrhythmias.157, 239, 240
D. Mitochondrial ROS and cardiac gap junction regulation
Cardiac gap junctions are formed by the assembly of a pair of juxtaposed intercellular hemichannels with each hemichannel consisting of six connexin (Cx) proteins. Gap junctions mediate the intercellular communication of small metabolites and ions and play an essential role in cardiac electrical conduction.241 Among the three principal connexin isoforms (Cx40, Cx43 and Cx45) expressed in the heart, Cx43 is the predominant isoform expressed in ventricular myocytes.242 Ventricular Cx43 expression has been shown to be downregulated with myocardial ischemia,243, 244 and heart failure,245, 246 which can lead to slowed conduction, increased electrical heterogeneity and abnormal anisotropic properties of the ventricles.247, 248 These changes can facilitate the initiation and maintenance of ventricular arrhythmias and SCDs.249, 250
Cardiac renin-angiotensin system (RAS) activity is increased upon acute ischemia, contributing to the acute and chronic myocardial remodeling in response to ischemia.251, 252 RAS activation is known to increase myocardial oxidative stress and downregulate ventricular gap junction protein Cx43.253–255 Transgenic mouse models with enhanced cardiac RAS activity255, 256 have high incidence of conduction abnormality, ventricular arrhythmias and sudden death owing to reduced cardiac Cx43 and abnormal gap junction function.255, 256 Using a transgenic mouse model of cardiac-restricted angiotensin converting enzyme overexpression (ACE8/8),255 we have demonstrated that increased cardiac RAS activity leads to increased activity of cSrc, a redox-sensitive tyrosine kinase, by increasing cSrc phosphorylation at Tyr416, in the ventricular myocardium. The activation of cSrc leads to Cx43 downregulation, impaired gap junction function, slowed cardiac conduction and increased incidence of ventricular arrhythmias and SCD.35, 257 The downregulation of Cx43 and increased risk for arrhythmias in ACE8/8 mice are alleviated by pharmacologic inhibition of RAS258 and cSrc.257 It has been shown that increased myocardial p-cSrc leads to Cx43 reduction through the competition between p-cSrc and Cx43 for the binding with zonula occludens-1, a scaffolding protein at the intercalated disk, resulting in Cx43 destabilization and degradation.259 Elevated p-cSrc levels also impair gap junction function by phosphorylating Cx43 at tyrosine residues.260 Using the same ACE8/8 mouse model, we have also shown that cardiac mitochondrial ROS were markedly increased with enhanced RAS activity.35, 257 Treatment with mitochondria-targeted antioxidant MitoTEMPO, but not with other types of antioxidants, decreases cSrc phosphorylation, preserves Cx43 expression, improves gap junction function, and abolishes ventricular arrhythmias and sudden cardiac death in ACE8/8 mice.35 Although the clinical evidence suggests that exogenous antioxidants do not prevent SCD,261 these data support a role for ROS, and suppression of endogenous, mitochondria-specific ROS may be an effective therapeutic approach.
Mechanistically, we have recently demonstrated that enhanced RAS signaling increases S-nitrosylation (SNO) of cardiac caveolin-1 (Cav1), an intrinsic inhibitor of cSrc, resulting in Cav1-cSrc dissociation and subsequent cSrc activation. Cav1 SNO upon enhanced RAS signaling is mediated by increased Cav1-eNOS binding that is dependent on elevated mitochondrial ROS.262 Consistent with these findings, Cav1 knock-out mice exhibit increased cSrc activation, reduced Cx43 expression, myocardial conduction defect and increased inducibility of ventricular arrhythmias.262 Taken together, these data suggest the critical roles of mitochondrial oxidative stress and Cav1 in AngII–induced gap junction remodeling and arrhythmia. As mitochondrial ROS are increased in myocardial ischemia251, 252 and heart failure,56, 57 both of which are associated with RAS activation, reduced ventricular Cx43 and increased risk for ventricular arrhythmias and SCD, it would be of great interest to test if the treatment with mitochondria-targeted antioxidant can normalize Cx43 expression and prevent ventricular arrhythmias and SCD in these pathological conditions.
Effects of chronic versus acute metabolic derangement and oxidative stress on arrhythmogenicity
In this review, we have primarily focused on the impact of acute myocardial metabolic and oxidative stress on cardiac electrophysiology, arrhythmogenicity and SCDs. It is important to note that chronic conditions associated with metabolic derangements and excessive oxidative stress such as aging,263 hypoxia/obstructive sleep apnea,264 chronic kidney disease265 and diabetes,138, 139 also result in arrhythmogenic changes that predispose to SCDs. Many of the arrhythmogenic changes related to these chronic conditions involve electrical remodeling74, 75 and the creation of arrhythmogenic substrates such as fibrosis,266,267 which may not be seen with acute metabolic and oxidative stress. Nevertheless, the effects of acute and chronic metabolic/oxidative stress on cardiac electrical function may not be mutually exclusive. The acute application of mitochondria-targeted antioxidant, for example, can restore the reduced peak INa observed with chronic heart failure.57 In addition, reduced Kv currents in diabetic heart can be reversed by acute insulin treatment.140 These observations suggest the contribution of the acute, modifiable metabolic/oxidative stress in these chronic conditions, which can be rapidly reversed by targeting the underlying metabolic/oxidative derangement. It is difficult to distinguish whether acute metabolic/oxidative changes exacerbate and amplify chronic changes or whether acute changes are unrelated to the chronic changes and simply occur on an arrhythmogenic background. Further studies are required to address these possibilities.
Conclusion
In summary, metabolic derangement and increased oxidative stress are prevalent in arrhythmogenic cardiac conditions, particularly during myocardial ischemia. Impaired cardiac metabolism and increased ROS production can lead to malfunction of various cellular mechanisms that are required to maintain normal electrical functioning and intracellular ionic homeostasis in cardiac myocytess. The impact of altered cardiac metabolism and increased oxidative stress on cardiac arrhythmogenicity is summarized in Table 1/Figure 4 and Table 2/Figure 5, respectively. As the conventional antiarrhythmic drugs targeting ion channels are often pro-arrhythmic, understanding the mechanisms linking abnormal metabolism and oxidative stress to cardiac arrhythmias may help to develop novel therapeutics to reduce the risk of life-threatening arrhythmias and SCD in patients with cardiac diseases. These observations suggest that therapeutics tailored to ameliorate metabolic derangement and oxidative stress may prove a more efficacious alternative to traditional ion channel blocking drugs to address arrhythmia in associated with cardiac diseases and a list of potential novel therapeutics are summarized in Table 3.
Table 3.
Potential novel therapeutics against ventricular arrhythmias and SCDs by targeting metabolic and oxidative stress
| Type of therapeutics | Mechanism | Effects on channel function/electrophysiology |
|---|---|---|
| NHE inhibitor111 | Inhibit NHE activation- induced Na+ and Ca2+ overload | Prevent ischemia/acidosis-induced Ca2+ overload |
| PI3Kα signaling pathway activator145, 146 | ↑ K+ channels transcriptionally | Restoring K+ currents with HF, cardiac hypertrophy and DM |
| MCU inhibitor105 | Inhibit mitochondrial Ca2+ uptake | Prevent mitochondrial Ca2+ overload and subsequent mitochondrial dysfunction |
| PTP inhibitor268 | Prevent ROS-induced Δψm depolarization | Prevent mitochondrial ROS-induced ROS release and subsequent Ca2+ overload |
| IMAC inhibitor62, 227, 228 | Prevent ROS-induced Δψm depolarization ↑Mitochondrial K+ | Prevent mitochondrial ROS-induced ROS release and subsequent Ca2+ overload |
| mitoKATP opener230,229 | influx, prevents Δψm depolarization | Stabilize Δψm and maintain mitochondrial function |
| Mitochondria-targeted anti-oxidants35, 57, 262 | ↓ mitochondrial ROS | Alleviate Cx43 and peak INa down- regulation with increased mitochondrial ROS |
| CaMKIIδ inhibitor124 | Prevent ROS-induced late INa | Prevent ROS-induced late INa and EAD |
NHE: Na+-H+ exchanger; PI3Kα: phosphoinositide 3-kinase α; MCU: mitochondrial calcium uniporter; PTP: mitochondrial permeability transition pore; IMAC: mitochondrial inner membrane anion channel; Δψm: mitochondrial membrane potential
Supplementary Material
Acknowledgments
Sources of Funding
This work was funded by National Institutes of Health (NIH) Grants RO1 HL104025 (SCD), HL106592 (SCD), a Veterans Affairs MERIT grants BX000859 (SCD), RO1 HL 122109 (JCM), RO1 HL 57414 (JCM), RO1 HL 071092 (JCM), a Taiwan Ministry of Science Technology Grant 103-2320-B-002-068-MY2 (KCY) and a Taiwan National Health Research Institute Career Development Grant NHRI-EX104-10418SC (KCY).
Nonstandard Abbreviations and Acronyms
- AMPK
adenosine monophosphate-activated protein kinase
- APD
action potential duration
- BH4
tetrahydrobiopterin
- CaMKII
calcium/calmodulin-dependent protein kinase II
- Cav1
caveolin-1
- CHD
coronary heart disease
- CoQ
coenzyme Q10
- CoQH2
reduced coenzyme Q10
- Cx
connexin
- DAD
delayed afterdepolarization
- DCM
dilated cardiomyopathy
- ΔΨm
mitochondrial membrane potential
- EAD
early afterdepolarization
- ETC
electron-transport chain
- FADH2
flavin adenine dinucleotide
- GSH
glutathione
- GAPDH
glyceraldehyde 3-phosphate dehydrogenase
- GPD1-L
glycerol-3-phosphate dehydrogenase 1-like
- GSS
S-glutathiolation
- H2O2
hydrogen peroxide
- HOCl
hypochlorite
- ICaL
L-type Ca2+ currents
- IK
delayed rectifier K+ currents
- IKATP
ATP-sensitive K+ current
- IMAC
inner membrane anion channel
- INa
Na+-current
- IPC
ischemic preconditioning
- Ito
transient outward K+ current
- Kir
inwardly rectifying K+ channels
- Kv
voltage-gated K+ channels
- LPC
lysophosphatidylcholine
- MnSOD
manganese superoxide dismutase
- MCU
mitochondrial Ca2+ uniporter
- mitoNCX
mitochondrial Na+-Ca2+ exchanger
- NADH
nicotinamide adenine dinucleotide
- NADPH
nicotinamide adenine dinucleotide phosphate
- NCX
Na+/Ca2+ exchanger
- NHE
Na+/H+ exchanger
- NO•
nitric oxide
- NOS
nitric oxide synthase
- NOX2
NADPH oxidase 2
- OXPHOS
oxidative phosphorylation
- O2•−
superoxide
- •OH
hydroxyl radicals
- ONOO−
peroxynitrite
- PEA
pulseless electrical activity
- PI3Kα
phosphoinositide 3-kinase α
- PKA
protein kinase A
- PKC
protein kinase C
- PKG
protein kinase G
- PPARα
peroxisome proliferator-activated receptor α
- PTP
mitochondrial permeability transition pore
- RAS
renin-angiotensin system
- RNS
reactive nitrogen species
- ROS
reactive oxygen species
- RyR2
ryanodine receptor 2
- sarcKATP
sarcolemmal KATP
- SCD
sudden cardiac death
- SERCA
sarco/endoplasmic reticulum Ca2+-ATPase
- SNO
S-nitrosylation
- SR
sarcoplasmic reticulum
- TCA cycle
tricarboxylic acid cycle
- VF
ventricular fibrillation
- VT
ventricular tachycardia
Footnotes
Disclosures
Dr. Dudley is an inventor of 13/551,790 A Method for Ameliorating or Preventing Arrhythmic Risk Associated with Cardiomyopathy by Improving Conduction Velocity, 13/507,319 A Method for Modulating or Controlling Connexin43 (Cx43) Level of a Cell and Reducing Arrhythmic Risk, PCT/US2008/011919 Modulation Of Sodium Current by Nicotinamide Adenine Dinucleotide, US 12/929,786 Modulating Mitochondrial Reactive Oxygen Species to Increase Cardiac Sodium Channel Current and Mitigate Sudden Death, and 13/551,790 Method for Ameliorating or Preventing Arrhythmic Risk Associated with Cardiomyopathy.
References
- 1.Roger VL, Go AS, Lloyd-Jones DM, Benjamin EJ, Berry JD, Borden WB, Bravata DM, Dai S, Ford ES, Fox CS, Fullerton HJ, Gillespie C, Hailpern SM, Heit JA, Howard VJ, Kissela BM, Kittner SJ, Lackland DT, Lichtman JH, Lisabeth LD, Makuc DM, Marcus GM, Marelli A, Matchar DB, Moy CS, Mozaffarian D, Mussolino ME, Nichol G, Paynter NP, Soliman EZ, Sorlie PD, Sotoodehnia N, Turan TN, Virani SS, Wong ND, Woo D, Turner MB. Heart disease and stroke statistics--2012 update: a report from the American Heart Association. Circulation. 2012;125(1):e2–e220. doi: 10.1161/CIR.0b013e31823ac046. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Zheng ZJ, Croft JB, Giles WH, Mensah GA. Sudden cardiac death in the United States, 1989 to 1998. Circulation. 2001;104(18):2158–2163. doi: 10.1161/hc4301.098254. [DOI] [PubMed] [Google Scholar]
- 3.Zipes DP, Wellens HJ. Sudden cardiac death. Circulation. 1998;98(21):2334–2351. doi: 10.1161/01.cir.98.21.2334. [DOI] [PubMed] [Google Scholar]
- 4.Deo R, Albert CM. Epidemiology and genetics of sudden cardiac death. Circulation. 2012;125(4):620–637. doi: 10.1161/CIRCULATIONAHA.111.023838. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Bayes de Luna A, Coumel P, Leclercq JF. Ambulatory sudden cardiac death: mechanisms of production of fatal arrhythmia on the basis of data from 157 cases. Am Heart J. 1989;117(1):151–159. doi: 10.1016/0002-8703(89)90670-4. [DOI] [PubMed] [Google Scholar]
- 6.Mitchell LB, Pineda EA, Titus JL, Bartosch PM, Benditt DG. Sudden death in patients with implantable cardioverter defibrillators: the importance of post-shock electromechanical dissociation. J Am Coll Cardiol. 2002;39(8):1323–1328. doi: 10.1016/s0735-1097(02)01784-9. [DOI] [PubMed] [Google Scholar]
- 7.Wood MA, Stambler BS, Damiano RJ, Greenway P, Ellenbogen KA. Lessons learned from data logging in a multicenter clinical trial using a late-generation implantable cardioverter-defibrillator. The Guardian ATP 4210 Multicenter Investigators Group. J Am Coll Cardiol. 1994;24(7):1692–1699. doi: 10.1016/0735-1097(94)90176-7. [DOI] [PubMed] [Google Scholar]
- 8.Nichol G, Thomas E, Callaway CW, Hedges J, Powell JL, Aufderheide TP, Rea T, Lowe R, Brown T, Dreyer J, Davis D, Idris A, Stiell I. Regional variation in out-of-hospital cardiac arrest incidence and outcome. JAMA. 2008;300(12):1423–1431. doi: 10.1001/jama.300.12.1423. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Teodorescu C, Reinier K, Dervan C, Uy-Evanado A, Samara M, Mariani R, Gunson K, Jui J, Chugh SS. Factors associated with pulseless electric activity versus ventricular fibrillation: the Oregon sudden unexpected death study. Circulation. 2010;122(21):2116–2122. doi: 10.1161/CIRCULATIONAHA.110.966333. [DOI] [PubMed] [Google Scholar]
- 10.Bornfeldt KE, Tabas I. Insulin resistance, hyperglycemia, and atherosclerosis. Cell Metab. 2011;14(5):575–585. doi: 10.1016/j.cmet.2011.07.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Pansuria M, Xi H, Li L, Yang XF, Wang H. Insulin resistance, metabolic stress, and atherosclerosis. Front Biosci (Schol Ed) 2012;4:916–931. doi: 10.2741/s308. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Stocker R, Keaney JF., Jr Role of oxidative modifications in atherosclerosis. Physiol Rev. 2004;84(4):1381–1478. doi: 10.1152/physrev.00047.2003. [DOI] [PubMed] [Google Scholar]
- 13.Madamanchi NR, Vendrov A, Runge MS. Oxidative stress and vascular disease. Arterioscler Thromb Vasc Biol. 2005;25(1):29–38. doi: 10.1161/01.ATV.0000150649.39934.13. [DOI] [PubMed] [Google Scholar]
- 14.Hirooka Y, Sagara Y, Kishi T, Sunagawa K. Oxidative stress and central cardiovascular regulation. - Pathogenesis of hypertension and therapeutic aspects. Circ J. 2010;74(5):827–835. doi: 10.1253/circj.cj-10-0153. [DOI] [PubMed] [Google Scholar]
- 15.Gerritsen J, Dekker JM, TenVoorde BJ, Kostense PJ, Heine RJ, Bouter LM, Heethaar RM, Stehouwer CD. Impaired autonomic function is associated with increased mortality, especially in subjects with diabetes, hypertension, or a history of cardiovascular disease: the Hoorn Study. Diabetes Care. 2001;24(10):1793–1798. doi: 10.2337/diacare.24.10.1793. [DOI] [PubMed] [Google Scholar]
- 16.Priori SG, Aliot E, Blomstrom-Lundqvist C, Bossaert L, Breithardt G, Brugada P, Camm JA, Cappato R, Cobbe SM, Di Mario C, Maron BJ, McKenna WJ, Pedersen AK, Ravens U, Schwartz PJ, Trusz-Gluza M, Vardas P, Wellens HJ, Zipes DP. Update of the guidelines on sudden cardiac death of the European Society of Cardiology. Eur Heart J. 2003;24(1):13–15. doi: 10.1016/s0195-0668x(02)00809-6. [DOI] [PubMed] [Google Scholar]
- 17.Dhalla NS, Adameova A, Kaur M. Role of catecholamine oxidation in sudden cardiac death. Fundam Clin Pharmacol. 2010;24(5):539–546. doi: 10.1111/j.1472-8206.2010.00836.x. [DOI] [PubMed] [Google Scholar]
- 18.Fozzard HA. Excitation-contraction coupling in the heart. Adv Exp Med Biol. 1991;308:135–142. doi: 10.1007/978-1-4684-6015-5_11. [DOI] [PubMed] [Google Scholar]
- 19.Roden DM, Balser JR, George AL, Jr, Anderson ME. Cardiac ion channels. Annu Rev Physiol. 2002;64:431–475. doi: 10.1146/annurev.physiol.64.083101.145105. [DOI] [PubMed] [Google Scholar]
- 20.Nerbonne JM, Kass RS. Molecular physiology of cardiac repolarization. Physiol Rev. 2005;85(4):1205–1253. doi: 10.1152/physrev.00002.2005. [DOI] [PubMed] [Google Scholar]
- 21.Kanno S, Saffitz JE. The role of myocardial gap junctions in electrical conduction and arrhythmogenesis. Cardiovasc Pathol. 2001;10(4):169–177. doi: 10.1016/s1054-8807(01)00078-3. [DOI] [PubMed] [Google Scholar]
- 22.Harris DA, Das AM. Control of mitochondrial ATP synthesis in the heart. Biochem J. 1991;280 (Pt 3):561–573. doi: 10.1042/bj2800561. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Schaper J, Meiser E, Stammler G. Ultrastructural morphometric analysis of myocardium from dogs, rats, hamsters, mice, and from human hearts. Circ Res. 1985;56(3):377–391. doi: 10.1161/01.res.56.3.377. [DOI] [PubMed] [Google Scholar]
- 24.Chen LB. Mitochondrial membrane potential in living cells. Annu Rev Cell Biol. 1988;4:155–181. doi: 10.1146/annurev.cb.04.110188.001103. [DOI] [PubMed] [Google Scholar]
- 25.Gertz EW, Wisneski JA, Stanley WC, Neese RA. Myocardial substrate utilization during exercise in humans. Dual carbon-labeled carbohydrate isotope experiments. J Clin Invest. 1988;82(6):2017–2025. doi: 10.1172/JCI113822. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Wisneski JA, Gertz EW, Neese RA, Gruenke LD, Morris DL, Craig JC. Metabolic fate of extracted glucose in normal human myocardium. J Clin Invest. 1985;76(5):1819–1827. doi: 10.1172/JCI112174. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Schramm M, Klieber HG, Daut J. The energy expenditure of actomyosin-ATPase, Ca2+-ATPase and Na+, K+-ATPase in guinea-pig cardiac ventricular muscle. J Physiol. 1994;481 (Pt 3):647–662. doi: 10.1113/jphysiol.1994.sp020471. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Overend CL, Eisner DA, O’Neill SC. Altered cardiac sarcoplasmic reticulum function of intact myocytes of rat ventricle during metabolic inhibition. Circ Res. 2001;88(2):181–187. doi: 10.1161/01.res.88.2.181. [DOI] [PubMed] [Google Scholar]
- 29.Silverman HS, Stern MD. Ionic basis of ischaemic cardiac injury: insights from cellular studies. Cardiovasc Res. 1994;28(5):581–597. doi: 10.1093/cvr/28.5.581. [DOI] [PubMed] [Google Scholar]
- 30.Pung YF, Sam WJ, Stevanov K, Enrick M, Chen CL, Kolz C, Thakker P, Hardwick JP, Chen YR, Dyck JR, Yin L, Chilian WM. Mitochondrial oxidative stress corrupts coronary collateral growth by activating adenosine monophosphate activated kinase-alpha signaling. Arterioscler Thromb Vasc Biol. 2013;33(8):1911–1919. doi: 10.1161/ATVBAHA.113.301591. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Sripathi SR, He W, Atkinson CL, Smith JJ, Liu Z, Elledge BM, Jahng WJ. Mitochondrial-nuclear communication by prohibitin shuttling under oxidative stress. Biochemistry. 2011;50(39):8342–8351. doi: 10.1021/bi2008933. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Sato T, Sasaki N, Seharaseyon J, O’Rourke B, Marban E. Selective pharmacological agents implicate mitochondrial but not sarcolemmal KATP channels in ischemic cardioprotection. Circulation. 2000;101(20):2418–2423. doi: 10.1161/01.cir.101.20.2418. [DOI] [PubMed] [Google Scholar]
- 33.Prosser BL, Ward CW, Lederer WJ. X-ROS signaling: rapid mechano-chemo transduction in heart. Science. 2011;333(6048):1440–1445. doi: 10.1126/science.1202768. [DOI] [PubMed] [Google Scholar]
- 34.Jian Z, Han H, Zhang T, Puglisi J, Izu LT, Shaw JA, Onofiok E, Erickson JR, Chen YJ, Horvath B, Shimkunas R, Xiao W, Li Y, Pan T, Chan J, Banyasz T, Tardiff JC, Chiamvimonvat N, Bers DM, Lam KS, Chen-Izu Y. Mechanochemotransduction during cardiac myocytes contraction is mediated by localized nitric oxide signaling. Sci Signal. 2014;7(317):ra27. doi: 10.1126/scisignal.2005046. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Sovari AA, Rutledge CA, Jeong EM, Dolmatova E, Arasu D, Liu H, Vahdani N, Gu L, Zandieh S, Xiao L, Bonini MG, Duffy HS, Dudley SC., Jr Mitochondria oxidative stress, connexin43 remodeling, and sudden arrhythmic death. Circ Arrhythm Electrophysiol. 2013;6(3):623–631. doi: 10.1161/CIRCEP.112.976787. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Sanders PN, Koval OM, Jaffer OA, Prasad AM, Businga TR, Scott JA, Hayden PJ, Luczak ED, Dickey DD, Allamargot C, Olivier AK, Meyerholz DK, Robison AJ, Winder DG, Blackwell TS, Dworski R, Sammut D, Wagner BA, Buettner GR, Pope RM, Miller FJ, Jr, Dibbern ME, Haitchi HM, Mohler PJ, Howarth PH, Zabner J, Kline JN, Grumbach IM, Anderson ME. CaMKII is essential for the proasthmatic effects of oxidation. Science translational medicine. 2013;5(195):195ra197. doi: 10.1126/scitranslmed.3006135. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Zafari AM, Harrison DG. Free radicals in heart failure: therapeutic targets for old and new drugs. Congest Heart Fail. 2002;8(3):129–130. doi: 10.1111/j.1527-5299.2002.00300.x. [DOI] [PubMed] [Google Scholar]
- 38.Aggarwal NT, Makielski JC. Redox control of cardiac excitability. Antioxid Redox Signal. 2013;18(4):432–468. doi: 10.1089/ars.2011.4234. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Yang KC, Bonini MG, Dudley SC., Jr Mitochondria and arrhythmias. Free Radic Biol Med. 2014;71:351–361. doi: 10.1016/j.freeradbiomed.2014.03.033. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Griendling KK, Sorescu D, Ushio-Fukai M. NAD(P)H oxidase: role in cardiovascular biology and disease. Circ Res. 2000;86(5):494–501. doi: 10.1161/01.res.86.5.494. [DOI] [PubMed] [Google Scholar]
- 41.Harrison R. Structure and function of xanthine oxidoreductase: where are we now? Free Radic Biol Med. 2002;33(6):774–797. doi: 10.1016/s0891-5849(02)00956-5. [DOI] [PubMed] [Google Scholar]
- 42.Landmesser U, Dikalov S, Price SR, McCann L, Fukai T, Holland SM, Mitch WE, Harrison DG. Oxidation of tetrahydrobiopterin leads to uncoupling of endothelial cell nitric oxide synthase in hypertension. J Clin Invest. 2003;111(8):1201–1209. doi: 10.1172/JCI14172. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Chen CA, Wang TY, Varadharaj S, Reyes LA, Hemann C, Talukder MA, Chen YR, Druhan LJ, Zweier JL. S-glutathionylation uncouples eNOS and regulates its cellular and vascular function. Nature. 2010;468(7327):1115–1118. doi: 10.1038/nature09599. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Klatt P, Lamas S. Regulation of protein function by S-glutathiolation in response to oxidative and nitrosative stress. Eur J Biochem. 2000;267(16):4928–4944. doi: 10.1046/j.1432-1327.2000.01601.x. [DOI] [PubMed] [Google Scholar]
- 45.Stamler JS, Lamas S, Fang FC. Nitrosylation. the prototypic redox-based signaling mechanism. Cell. 2001;106(6):675–683. doi: 10.1016/s0092-8674(01)00495-0. [DOI] [PubMed] [Google Scholar]
- 46.Turrens JF. Mitochondrial formation of reactive oxygen species. J Physiol. 2003;552(Pt 2):335–344. doi: 10.1113/jphysiol.2003.049478. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Murphy MP. How mitochondria produce reactive oxygen species. Biochem J. 2009;417(1):1–13. doi: 10.1042/BJ20081386. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Boveris A, Chance B. The mitochondrial generation of hydrogen peroxide. General properties and effect of hyperbaric oxygen. Biochem J. 1973;134(3):707–716. doi: 10.1042/bj1340707. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Schofield CJ, Ratcliffe PJ. Oxygen sensing by HIF hydroxylases. Nat Rev Mol Cell Biol. 2004;5(5):343–354. doi: 10.1038/nrm1366. [DOI] [PubMed] [Google Scholar]
- 50.Weisiger RA, Fridovich I. Superoxide dismutase. Organelle specificity. J Biol Chem. 1973;248(10):3582–3592. [PubMed] [Google Scholar]
- 51.Aon MA, Cortassa S, O’Rourke B. Redox-optimized ROS balance: a unifying hypothesis. Biochim Biophys Acta. 2010;1797(6–7):865–877. doi: 10.1016/j.bbabio.2010.02.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Cortassa S, O’Rourke B, Aon MA. Redox-optimized ROS balance and the relationship between mitochondrial respiration and ROS. Biochim Biophys Acta. 2014;1837(2):287–295. doi: 10.1016/j.bbabio.2013.11.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Nickel A, Kohlhaas M, Maack C. Mitochondrial reactive oxygen species production and elimination. J Mol Cell Cardiol. 2014;73:26–33. doi: 10.1016/j.yjmcc.2014.03.011. [DOI] [PubMed] [Google Scholar]
- 54.Novalija E, Kevin LG, Eells JT, Henry MM, Stowe DF. Anesthetic preconditioning improves adenosine triphosphate synthesis and reduces reactive oxygen species formation in mitochondria after ischemia by a redox dependent mechanism. Anesthesiology. 2003;98(5):1155–1163. doi: 10.1097/00000542-200305000-00018. [DOI] [PubMed] [Google Scholar]
- 55.Tompkins AJ, Burwell LS, Digerness SB, Zaragoza C, Holman WL, Brookes PS. Mitochondrial dysfunction in cardiac ischemia-reperfusion injury: ROS from complex I, without inhibition. Biochim Biophys Acta. 2006;1762(2):223–231. doi: 10.1016/j.bbadis.2005.10.001. [DOI] [PubMed] [Google Scholar]
- 56.Ide T, Tsutsui H, Kinugawa S, Utsumi H, Kang D, Hattori N, Uchida K, Arimura K, Egashira K, Takeshita A. Mitochondrial electron transport complex I is a potential source of oxygen free radicals in the failing myocardium. Circ Res. 1999;85(4):357–363. doi: 10.1161/01.res.85.4.357. [DOI] [PubMed] [Google Scholar]
- 57.Liu M, Gu L, Sulkin MS, Liu H, Jeong EM, Greener I, Xie A, Efimov IR, Dudley SC., Jr Mitochondrial dysfunction causing cardiac sodium channel downregulation in cardiomyopathy. J Mol Cell Cardiol. 2013;54:25–34. doi: 10.1016/j.yjmcc.2012.10.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Brenner C, Moulin M. Physiological roles of the permeability transition pore. Circ Res. 2012;111(9):1237–1247. doi: 10.1161/CIRCRESAHA.112.265942. [DOI] [PubMed] [Google Scholar]
- 59.Aon MA, Cortassa S, Marban E, O’Rourke B. Synchronized whole cell oscillations in mitochondrial metabolism triggered by a local release of reactive oxygen species in cardiac myocytes. J Biol Chem. 2003;278(45):44735–44744. doi: 10.1074/jbc.M302673200. [DOI] [PubMed] [Google Scholar]
- 60.Zorov DB, Juhaszova M, Sollott SJ. Mitochondrial ROS-induced ROS release: an update and review. Biochim Biophys Acta. 2006;1757(5–6):509–517. doi: 10.1016/j.bbabio.2006.04.029. [DOI] [PubMed] [Google Scholar]
- 61.Brady NR, Hamacher-Brady A, Westerhoff HV, Gottlieb RA. A wave of reactive oxygen species (ROS)-induced ROS release in a sea of excitable mitochondria. Antioxid Redox Signal. 2006;8(9–10):1651–1665. doi: 10.1089/ars.2006.8.1651. [DOI] [PubMed] [Google Scholar]
- 62.Akar FG, Aon MA, Tomaselli GF, O’Rourke B. The mitochondrial origin of postischemic arrhythmias. J Clin Invest. 2005;115(12):3527–3535. doi: 10.1172/JCI25371. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Halestrap AP, Richardson AP. The mitochondrial permeability transition: A current perspective on its identity and role in ischaemia/reperfusion injury. J Mol Cell Cardiol. 2014 doi: 10.1016/j.yjmcc.2014.08.018. [DOI] [PubMed] [Google Scholar]
- 64.Zorov DB, Juhaszova M, Sollott SJ. Mitochondrial reactive oxygen species (ROS) and ROS-induced ROS release. Physiol Rev. 2014;94(3):909–950. doi: 10.1152/physrev.00026.2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Argaud L, Gateau-Roesch O, Muntean D, Chalabreysse L, Loufouat J, Robert D, Ovize M. Specific inhibition of the mitochondrial permeability transition prevents lethal reperfusion injury. J Mol Cell Cardiol. 2005;38(2):367–374. doi: 10.1016/j.yjmcc.2004.12.001. [DOI] [PubMed] [Google Scholar]
- 66.Gomez L, Thibault H, Gharib A, Dumont JM, Vuagniaux G, Scalfaro P, Derumeaux G, Ovize M. Inhibition of mitochondrial permeability transition improves functional recovery and reduces mortality following acute myocardial infarction in mice. Am J Physiol Heart Circ Physiol. 2007;293(3):H1654–1661. doi: 10.1152/ajpheart.01378.2006. [DOI] [PubMed] [Google Scholar]
- 67.Giorgio V, von Stockum S, Antoniel M, Fabbro A, Fogolari F, Forte M, Glick GD, Petronilli V, Zoratti M, Szabo I, Lippe G, Bernardi P. Dimers of mitochondrial ATP synthase form the permeability transition pore. Proc Natl Acad Sci U S A. 2013;110(15):5887–5892. doi: 10.1073/pnas.1217823110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Bonora M, Bononi A, De Marchi E, Giorgi C, Lebiedzinska M, Marchi S, Patergnani S, Rimessi A, Suski JM, Wojtala A, Wieckowski MR, Kroemer G, Galluzzi L, Pinton P. Role of the c subunit of the FO ATP synthase in mitochondrial permeability transition. Cell Cycle. 2013;12(4):674–683. doi: 10.4161/cc.23599. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Aon MA, Cortassa S, O’Rourke B. Mitochondrial oscillations in physiology and pathophysiology. Adv Exp Med Biol. 2008;641:98–117. doi: 10.1007/978-0-387-09794-7_8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Murray J, Taylor SW, Zhang B, Ghosh SS, Capaldi RA. Oxidative damage to mitochondrial complex I due to peroxynitrite: identification of reactive tyrosines by mass spectrometry. J Biol Chem. 2003;278(39):37223–37230. doi: 10.1074/jbc.M305694200. [DOI] [PubMed] [Google Scholar]
- 71.Tatarkova Z, Kuka S, Racay P, Lehotsky J, Dobrota D, Mistuna D, Kaplan P. Effects of aging on activities of mitochondrial electron transport chain complexes and oxidative damage in rat heart. Physiol Res. 2011;60(2):281–289. doi: 10.33549/physiolres.932019. [DOI] [PubMed] [Google Scholar]
- 72.Echtay KS, Roussel D, St-Pierre J, Jekabsons MB, Cadenas S, Stuart JA, Harper JA, Roebuck SJ, Morrison A, Pickering S, Clapham JC, Brand MD. Superoxide activates mitochondrial uncoupling proteins. Nature. 2002;415(6867):96–99. doi: 10.1038/415096a. [DOI] [PubMed] [Google Scholar]
- 73.Echtay KS, Esteves TC, Pakay JL, Jekabsons MB, Lambert AJ, Portero-Otin M, Pamplona R, Vidal-Puig AJ, Wang S, Roebuck SJ, Brand MD. A signalling role for 4-hydroxy-2-nonenal in regulation of mitochondrial uncoupling. EMBO J. 2003;22(16):4103–4110. doi: 10.1093/emboj/cdg412. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Aiba T, Tomaselli GF. Electrical remodeling in the failing heart. Curr Opin Cardiol. 2010;25(1):29–36. doi: 10.1097/HCO.0b013e328333d3d6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Cutler MJ, Jeyaraj D, Rosenbaum DS. Cardiac electrical remodeling in health and disease. Trends Pharmacol Sci. 2011;32(3):174–180. doi: 10.1016/j.tips.2010.12.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Bers DM. Calcium cycling and signaling in cardiac myocytes. Annu Rev Physiol. 2008;70:23–49. doi: 10.1146/annurev.physiol.70.113006.100455. [DOI] [PubMed] [Google Scholar]
- 77.Varadarajan SG, An J, Novalija E, Smart SC, Stowe DF. Changes in [Na+](i), compartmental [Ca2+], and NADH with dysfunction after global ischemia in intact hearts. Am J Physiol Heart Circ Physiol. 2001;280(1):H280–293. doi: 10.1152/ajpheart.2001.280.1.H280. [DOI] [PubMed] [Google Scholar]
- 78.An J, Varadarajan SG, Novalija E, Stowe DF. Ischemic and anesthetic preconditioning reduces cytosolic [Ca2+] and improves Ca2+ responses in intact hearts. Am J Physiol Heart Circ Physiol. 2001;281(4):H1508–1523. doi: 10.1152/ajpheart.2001.281.4.H1508. [DOI] [PubMed] [Google Scholar]
- 79.Moore RL, Yelamarty RV, Misawa H, Scaduto RC, Jr, Pawlush DG, Elensky M, Cheung JY. Altered Ca2+ dynamics in single cardiac myocytes from renovascular hypertensive rats. Am J Physiol. 1991;260(2 Pt 1):C327–337. doi: 10.1152/ajpcell.1991.260.2.C327. [DOI] [PubMed] [Google Scholar]
- 80.Bentivegna LA, Ablin LW, Kihara Y, Morgan JP. Altered calcium handling in left ventricular pressure-overload hypertrophy as detected with aequorin in the isolated, perfused ferret heart. Circ Res. 1991;69(6):1538–1545. doi: 10.1161/01.res.69.6.1538. [DOI] [PubMed] [Google Scholar]
- 81.O’Rourke B, Kass DA, Tomaselli GF, Kaab S, Tunin R, Marban E. Mechanisms of altered excitation-contraction coupling in canine tachycardia-induced heart failure, I: experimental studies. Circ Res. 1999;84(5):562–570. doi: 10.1161/01.res.84.5.562. [DOI] [PubMed] [Google Scholar]
- 82.Pogwizd SM, Qi M, Yuan W, Samarel AM, Bers DM. Upregulation of Na+/Ca2+ exchanger expression and function in an arrhythmogenic rabbit model of heart failure. Circ Res. 1999;85(11):1009–1019. doi: 10.1161/01.res.85.11.1009. [DOI] [PubMed] [Google Scholar]
- 83.Koretsune Y, Marban E. Mechanism of ischemic contracture in ferret hearts: relative roles of [Ca2+]i elevation and ATP depletion. Am J Physiol. 1990;258(1 Pt 2):H9–16. doi: 10.1152/ajpheart.1990.258.1.H9. [DOI] [PubMed] [Google Scholar]
- 84.Elliott AC, Smith GL, Eisner DA, Allen DG. Metabolic changes during ischaemia and their role in contractile failure in isolated ferret hearts. J Physiol. 1992;454:467–490. doi: 10.1113/jphysiol.1992.sp019274. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85.Murphy E, Steenbergen C, Levy LA, Raju B, London RE. Cytosolic free magnesium levels in ischemic rat heart. J Biol Chem. 1989;264(10):5622–5627. [PubMed] [Google Scholar]
- 86.Eisner DA, Richards DE. Inhibition of the sodium pump by inorganic phosphate in resealed red cell ghosts. J Physiol. 1982;326:1–10. doi: 10.1113/jphysiol.1982.sp014172. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Apell HJ, Nelson MT, Marcus MM, Lauger P. Effects of the ATP, ADP and inorganic phosphate on the transport rate of the Na+, K+-pump. Biochim Biophys Acta. 1986;857(1):105–115. doi: 10.1016/0005-2736(86)90103-3. [DOI] [PubMed] [Google Scholar]
- 88.Undrovinas AI, Fleidervish IA, Makielski JC. Inward sodium current at resting potentials in single cardiac myocytes induced by the ischemic metabolite lysophosphatidylcholine. Circ Res. 1992;71(5):1231–1241. doi: 10.1161/01.res.71.5.1231. [DOI] [PubMed] [Google Scholar]
- 89.Jones DK, Peters CH, Tolhurst SA, Claydon TW, Ruben PC. Extracellular proton modulation of the cardiac voltage-gated sodium channel, Nav1.5. Biophys J. 2011;101(9):2147–2156. doi: 10.1016/j.bpj.2011.08.056. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90.Sokolov S, Peters CH, Rajamani S, Ruben PC. Proton-dependent inhibition of the cardiac sodium channel Nav1.5 by ranolazine. Front Pharmacol. 2013;4:78. doi: 10.3389/fphar.2013.00078. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91.Valdivia CR, Chu WW, Pu J, Foell JD, Haworth RA, Wolff MR, Kamp TJ, Makielski JC. Increased late sodium current in myocytes from a canine heart failure model and from failing human heart. J Mol Cell Cardiol. 2005;38(3):475–483. doi: 10.1016/j.yjmcc.2004.12.012. [DOI] [PubMed] [Google Scholar]
- 92.Mochizuki S, MacLeod KT. Effects of hypoxia and metabolic inhibition on increases in intracellular Ca2+ concentration induced by Na+/Ca2+ exchange in isolated guinea-pig cardiac myocytes. J Mol Cell Cardiol. 1997;29(11):2979–2987. doi: 10.1006/jmcc.1997.0542. [DOI] [PubMed] [Google Scholar]
- 93.Weber CR, Piacentino V, 3rd, Houser SR, Bers DM. Dynamic regulation of sodium/calcium exchange function in human heart failure. Circulation. 2003;108(18):2224–2229. doi: 10.1161/01.CIR.0000095274.72486.94. [DOI] [PubMed] [Google Scholar]
- 94.Armoundas AA, Hobai IA, Tomaselli GF, Winslow RL, O’Rourke B. Role of sodium-calcium exchanger in modulating the action potential of ventricular myocytes from normal and failing hearts. Circ Res. 2003;93(1):46–53. doi: 10.1161/01.RES.0000080932.98903.D8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95.Xu L, Mann G, Meissner G. Regulation of cardiac Ca2+ release channel (ryanodine receptor) by Ca2+, H+, Mg2+, and adenine nucleotides under normal and simulated ischemic conditions. Circ Res. 1996;79(6):1100–1109. doi: 10.1161/01.res.79.6.1100. [DOI] [PubMed] [Google Scholar]
- 96.Ferrier GR, Saunders JH, Mendez C. A cellular mechanism for the generation of ventricular arrhythmias by acetylstrophanthidin. Circ Res. 1973;32(5):600–609. doi: 10.1161/01.res.32.5.600. [DOI] [PubMed] [Google Scholar]
- 97.Clusin WT. Mechanisms of calcium transient and action potential alternans in cardiac cells and tissues. Am J Physiol Heart Circ Physiol. 2008;294(1):H1–H10. doi: 10.1152/ajpheart.00802.2007. [DOI] [PubMed] [Google Scholar]
- 98.Qian YW, Clusin WT, Lin SF, Han J, Sung RJ. Spatial heterogeneity of calcium transient alternans during the early phase of myocardial ischemia in the blood-perfused rabbit heart. Circulation. 2001;104(17):2082–2087. doi: 10.1161/hc4201.097136. [DOI] [PubMed] [Google Scholar]
- 99.Palty R, Silverman WF, Hershfinkel M, Caporale T, Sensi SL, Parnis J, Nolte C, Fishman D, Shoshan-Barmatz V, Herrmann S, Khananshvili D, Sekler I. NCLX is an essential component of mitochondrial Na+/Ca2+ exchange. Proc Natl Acad Sci U S A. 2010;107(1):436–441. doi: 10.1073/pnas.0908099107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100.Maack C, Cortassa S, Aon MA, Ganesan AN, Liu T, O’Rourke B. Elevated cytosolic Na+ decreases mitochondrial Ca2+ uptake during excitation-contraction coupling and impairs energetic adaptation in cardiac myocytes. Circ Res. 2006;99(2):172–182. doi: 10.1161/01.RES.0000232546.92777.05. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101.Drago I, De Stefani D, Rizzuto R, Pozzan T. Mitochondrial Ca2+ uptake contributes to buffering cytoplasmic Ca2+ peaks in cardiac myocytess. Proc Natl Acad Sci U S A. 2012;109(32):12986–12991. doi: 10.1073/pnas.1210718109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 102.Williams GS, Boyman L, Chikando AC, Khairallah RJ, Lederer WJ. Mitochondrial calcium uptake. Proc Natl Acad Sci U S A. 2013;110(26):10479–10486. doi: 10.1073/pnas.1300410110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 103.O’Rourke B, Blatter LA. Mitochondrial Ca2+ uptake: tortoise or hare? J Mol Cell Cardiol. 2009;46(6):767–774. doi: 10.1016/j.yjmcc.2008.12.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104.Florea SM, Blatter LA. The role of mitochondria for the regulation of cardiac alternans. Front Physiol. 2010;1:141. doi: 10.3389/fphys.2010.00141. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 105.de Garcia-Rivas GJ, Carvajal K, Correa F, Zazueta C. Ru360, a specific mitochondrial calcium uptake inhibitor, improves cardiac post-ischaemic functional recovery in rats in vivo. Br J Pharmacol. 2006;149(7):829–837. doi: 10.1038/sj.bjp.0706932. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 106.Liu T, O’Rourke B. Enhancing mitochondrial Ca2+ uptake in myocytes from failing hearts restores energy supply and demand matching. Circ Res. 2008;103(3):279–288. doi: 10.1161/CIRCRESAHA.108.175919. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 107.Haussinger D, Schliess F. Osmotic induction of signaling cascades: role in regulation of cell function. Biochem Biophys Res Commun. 1999;255(3):551–555. doi: 10.1006/bbrc.1998.9946. [DOI] [PubMed] [Google Scholar]
- 108.Haigney MC, Lakatta EG, Stern MD, Silverman HS. Sodium channel blockade reduces hypoxic sodium loading and sodium-dependent calcium loading. Circulation. 1994;90(1):391–399. doi: 10.1161/01.cir.90.1.391. [DOI] [PubMed] [Google Scholar]
- 109.Ingwall JS. How high does intracellular sodium rise during acute myocardial ischaemia? A view from NMR spectroscopy. Cardiovasc Res. 1995;29(2):279. [PubMed] [Google Scholar]
- 110.ten Hove M, van Emous JG, van Echteld CJ. Na+ overload during ischemia and reperfusion in rat hearts: comparison of the Na+/H+ exchange blockers EIPA, cariporide and eniporide. Mol Cell Biochem. 2003;250(1–2):47–54. doi: 10.1023/a:1024985931797. [DOI] [PubMed] [Google Scholar]
- 111.Wang Y, Meyer JW, Ashraf M, Shull GE. Mice with a null mutation in the NHE1 Na+-H+ exchanger are resistant to cardiac ischemia-reperfusion injury. Circ Res. 2003;93(8):776–782. doi: 10.1161/01.RES.0000094746.24774.DC. [DOI] [PubMed] [Google Scholar]
- 112.John SA, Kondo R, Wang SY, Goldhaber JI, Weiss JN. Connexin-43 hemichannels opened by metabolic inhibition. J Biol Chem. 1999;274(1):236–240. doi: 10.1074/jbc.274.1.236. [DOI] [PubMed] [Google Scholar]
- 113.Retamal MA, Schalper KA, Shoji KF, Bennett MV, Saez JC. Opening of connexin 43 hemichannels is increased by lowering intracellular redox potential. Proc Natl Acad Sci U S A. 2007;104(20):8322–8327. doi: 10.1073/pnas.0702456104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114.Kondo RP, Wang SY, John SA, Weiss JN, Goldhaber JI. Metabolic inhibition activates a non-selective current through connexin hemichannels in isolated ventricular myocytes. J Mol Cell Cardiol. 2000;32(10):1859–1872. doi: 10.1006/jmcc.2000.1220. [DOI] [PubMed] [Google Scholar]
- 115.Manjunath CK, Page E. Cell biology and protein composition of cardiac gap junctions. Am J Physiol. 1985;248(6 Pt 2):H783–791. doi: 10.1152/ajpheart.1985.248.6.H783. [DOI] [PubMed] [Google Scholar]
- 116.Jung DW, Apel LM, Brierley GP. Transmembrane gradients of free Na+ in isolated heart mitochondria estimated using a fluorescent probe. Am J Physiol. 1992;262(4 Pt 1):C1047–1055. doi: 10.1152/ajpcell.1992.262.4.C1047. [DOI] [PubMed] [Google Scholar]
- 117.Kim B, Matsuoka S. Cytoplasmic Na+-dependent modulation of mitochondrial Ca2+ via electrogenic mitochondrial Na+-Ca2+ exchange. J Physiol. 2008;586(6):1683–1697. doi: 10.1113/jphysiol.2007.148726. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 118.Donoso P, Mill JG, O’Neill SC, Eisner DA. Fluorescence measurements of cytoplasmic and mitochondrial sodium concentration in rat ventricular myocytes. J Physiol. 1992;448:493–509. doi: 10.1113/jphysiol.1992.sp019053. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 119.Liu T, Takimoto E, Dimaano VL, DeMazumder D, Kettlewell S, Smith G, Sidor A, Abraham TP, O’Rourke B. Inhibiting mitochondrial Na+/Ca2+ exchange prevents sudden death in a Guinea pig model of heart failure. Circ Res. 2014;115(1):44–54. doi: 10.1161/CIRCRESAHA.115.303062. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 120.Kohlhaas M, Liu T, Knopp A, Zeller T, Ong MF, Bohm M, O’Rourke B, Maack C. Elevated cytosolic Na+ increases mitochondrial formation of reactive oxygen species in failing cardiac myocytes. Circulation. 2010;121(14):1606–1613. doi: 10.1161/CIRCULATIONAHA.109.914911. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 121.Ward CA, Giles WR. Ionic mechanism of the effects of hydrogen peroxide in rat ventricular myocytes. J Physiol. 1997;500 (Pt 3):631–642. doi: 10.1113/jphysiol.1997.sp022048. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 122.Ma JH, Luo AT, Zhang PH. Effect of hydrogen peroxide on persistent sodium current in guinea pig ventricular myocytes. Acta Pharmacol Sin. 2005;26(7):828–834. doi: 10.1111/j.1745-7254.2005.00154.x. [DOI] [PubMed] [Google Scholar]
- 123.Liu M, Liu H, Dudley SC., Jr Reactive oxygen species originating from mitochondria regulate the cardiac sodium channel. Circ Res. 2010;107(8):967–974. doi: 10.1161/CIRCRESAHA.110.220673. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 124.Wagner S, Ruff HM, Weber SL, Bellmann S, Sowa T, Schulte T, Anderson ME, Grandi E, Bers DM, Backs J, Belardinelli L, Maier LS. Reactive oxygen species-activated Ca/calmodulin kinase IIdelta is required for late I(Na) augmentation leading to cellular Na and Ca overload. Circ Res. 2011;108(5):555–565. doi: 10.1161/CIRCRESAHA.110.221911. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 125.Bay J, Kohlhaas M, Maack C. Intracellular Na+ and cardiac metabolism. J Mol Cell Cardiol. 2013;61:20–27. doi: 10.1016/j.yjmcc.2013.05.010. [DOI] [PubMed] [Google Scholar]
- 126.Makielski JC, Farley AL. Na+ current in human ventricle: implications for sodium loading and homeostasis. J Cardiovasc Electrophysiol. 2006;17 (Suppl 1):S15–S20. doi: 10.1111/j.1540-8167.2006.00380.x. [DOI] [PubMed] [Google Scholar]
- 127.Shander GS, Undrovinas AI, Makielski JC. Rapid onset of lysophosphatidylcholine-induced modification of whole cell cardiac sodium current kinetics. J Mol Cell Cardiol. 1996;28(4):743–753. doi: 10.1006/jmcc.1996.0069. [DOI] [PubMed] [Google Scholar]
- 128.Hardie DG, Carling D. The AMP-activated protein kinase--fuel gauge of the mammalian cell? Eur J Biochem. 1997;246(2):259–273. doi: 10.1111/j.1432-1033.1997.00259.x. [DOI] [PubMed] [Google Scholar]
- 129.Hardie DG. AMP-activated protein kinase: the guardian of cardiac energy status. J Clin Invest. 2004;114(4):465–468. doi: 10.1172/JCI22683. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 130.Light PE, Wallace CH, Dyck JR. Constitutively active adenosine monophosphate-activated protein kinase regulates voltage-gated sodium channels in ventricular myocytes. Circulation. 2003;107(15):1962–1965. doi: 10.1161/01.CIR.0000069269.60167.02. [DOI] [PubMed] [Google Scholar]
- 131.Gollob MH, Green MS, Tang AS, Gollob T, Karibe A, Ali Hassan AS, Ahmad F, Lozado R, Shah G, Fananapazir L, Bachinski LL, Roberts R. Identification of a gene responsible for familial Wolff-Parkinson-White syndrome. N Engl J Med. 2001;344(24):1823–1831. doi: 10.1056/NEJM200106143442403. [DOI] [PubMed] [Google Scholar]
- 132.Gollob MH, Seger JJ, Gollob TN, Tapscott T, Gonzales O, Bachinski L, Roberts R. Novel PRKAG2 mutation responsible for the genetic syndrome of ventricular preexcitation and conduction system disease with childhood onset and absence of cardiac hypertrophy. Circulation. 2001;104(25):3030–3033. doi: 10.1161/hc5001.102111. [DOI] [PubMed] [Google Scholar]
- 133.Nichols CG, Lopatin AN. Inward rectifier potassium channels. Annu Rev Physiol. 1997;59:171–191. doi: 10.1146/annurev.physiol.59.1.171. [DOI] [PubMed] [Google Scholar]
- 134.Lopatin AN, Nichols CG. Inward rectifiers in the heart: an update on IK1. J Mol Cell Cardiol. 2001;33(4):625–638. doi: 10.1006/jmcc.2001.1344. [DOI] [PubMed] [Google Scholar]
- 135.Fan Z, Makielski JC. Intracellular H+ and Ca2+ modulation of trypsin-modified ATP-sensitive K+ channels in rabbit ventricular myocytes. Circ Res. 1993;72(3):715–722. doi: 10.1161/01.res.72.3.715. [DOI] [PubMed] [Google Scholar]
- 136.Zhuo ML, Huang Y, Liu DP, Liang CC. KATP channel: relation with cell metabolism and role in the cardiovascular system. Int J Biochem Cell Biol. 2005;37(4):751–764. doi: 10.1016/j.biocel.2004.10.008. [DOI] [PubMed] [Google Scholar]
- 137.Suzuki M, Saito T, Sato T, Tamagawa M, Miki T, Seino S, Nakaya H. Cardioprotective effect of diazoxide is mediated by activation of sarcolemmal but not mitochondrial ATP-sensitive potassium channels in mice. Circulation. 2003;107(5):682–685. doi: 10.1161/01.cir.0000055187.67365.81. [DOI] [PubMed] [Google Scholar]
- 138.Veglio M, Chinaglia A, Cavallo-Perin P. QT interval, cardiovascular risk factors and risk of death in diabetes. J Endocrinol Invest. 2004;27(2):175–181. doi: 10.1007/BF03346265. [DOI] [PubMed] [Google Scholar]
- 139.Aksnes TA, Schmieder RE, Kjeldsen SE, Ghani S, Hua TA, Julius S. Impact of new-onset diabetes mellitus on development of atrial fibrillation and heart failure in high-risk hypertension (from the VALUE Trial) Am J Cardiol. 2008;101(5):634–638. doi: 10.1016/j.amjcard.2007.10.025. [DOI] [PubMed] [Google Scholar]
- 140.Xu Z, Patel KP, Lou MF, Rozanski GJ. Up-regulation of K+ channels in diabetic rat ventricular myocytes by insulin and glutathione. Cardiovasc Res. 2002;53(1):80–88. doi: 10.1016/s0008-6363(01)00446-1. [DOI] [PubMed] [Google Scholar]
- 141.Xu Z, Patel KP, Rozanski GJ. Metabolic basis of decreased transient outward K+ current in ventricular myocytes from diabetic rats. Am J Physiol. 1996;271(5 Pt 2):H2190–2196. doi: 10.1152/ajpheart.1996.271.5.H2190. [DOI] [PubMed] [Google Scholar]
- 142.Shimoni Y, Firek L, Severson D, Giles W. Short-term diabetes alters K+ currents in rat ventricular myocytes. Circ Res. 1994;74(4):620–628. doi: 10.1161/01.res.74.4.620. [DOI] [PubMed] [Google Scholar]
- 143.Xu Z, Rozanski GJ. K+ current inhibition by amphiphilic fatty acid metabolites in rat ventricular myocytes. Am J Physiol. 1998;275(6 Pt 1):C1660–1667. doi: 10.1152/ajpcell.1998.275.6.C1660. [DOI] [PubMed] [Google Scholar]
- 144.Marionneau C, Aimond F, Brunet S, Niwa N, Finck B, Kelly DP, Nerbonne JM. PPARalpha-mediated remodeling of repolarizing voltage-gated K+ (Kv) channels in a mouse model of metabolic cardiomyopathy. J Mol Cell Cardiol. 2008;44(6):1002–1015. doi: 10.1016/j.yjmcc.2008.03.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 145.Yang KC, Foeger NC, Marionneau C, Jay PY, McMullen JR, Nerbonne JM. Homeostatic regulation of electrical excitability in physiological cardiac hypertrophy. J Physiol. 2010;588(Pt 24):5015–5032. doi: 10.1113/jphysiol.2010.197418. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 146.Yang KC, Jay PY, McMullen JR, Nerbonne JM. Enhanced cardiac PI3Kalpha signalling mitigates arrhythmogenic electrical remodelling in pathological hypertrophy and heart failure. Cardiovasc Res. 2012;93(2):252–262. doi: 10.1093/cvr/cvr283. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 147.Yang KC, Tseng YT, Nerbonne JM. Exercise training and PI3Kalpha-induced electrical remodeling is independent of cellular hypertrophy and Akt signaling. J Mol Cell Cardiol. 2012;53(4):532–541. doi: 10.1016/j.yjmcc.2012.07.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 148.O’Rourke B. Myocardial KATP channels in preconditioning. Circ Res. 2000;87(10):845–855. doi: 10.1161/01.res.87.10.845. [DOI] [PubMed] [Google Scholar]
- 149.Wu J, Cui N, Piao H, Wang Y, Xu H, Mao J, Jiang C. Allosteric modulation of the mouse Kir6.2 channel by intracellular H+ and ATP. J Physiol. 2002;543(Pt 2):495–504. doi: 10.1113/jphysiol.2002.025247. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 150.Gettes LS, Cascio WE, Johnson T, Fleet WF. Local myocardial biochemical and ionic alterations during myocardial ischaemia and reperfusion. Drugs. 1991;42 (Suppl 1):7–13. doi: 10.2165/00003495-199100421-00004. [DOI] [PubMed] [Google Scholar]
- 151.Steenbergen C, Murphy E, Watts JA, London RE. Correlation between cytosolic free calcium, contracture, ATP, and irreversible ischemic injury in perfused rat heart. Circ Res. 1990;66(1):135–146. doi: 10.1161/01.res.66.1.135. [DOI] [PubMed] [Google Scholar]
- 152.Faivre JF, Findlay I. Action potential duration and activation of ATP-sensitive potassium current in isolated guinea-pig ventricular myocytes. Biochim Biophys Acta. 1990;1029(1):167–172. doi: 10.1016/0005-2736(90)90450-3. [DOI] [PubMed] [Google Scholar]
- 153.Bhatnagar A. Contribution of ATP to oxidative stress-induced changes in action potential of isolated cardiac myocytes. Am J Physiol. 1997;272(4 Pt 2):H1598–1608. doi: 10.1152/ajpheart.1997.272.4.H1598. [DOI] [PubMed] [Google Scholar]
- 154.Aon MA, Cortassa S, Akar FG, Brown DA, Zhou L, O’Rourke B. From mitochondrial dynamics to arrhythmias. Int J Biochem Cell Biol. 2009;41(10):1940–1948. doi: 10.1016/j.biocel.2009.02.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 155.Vajda S, Baczko I, Lepran I. Selective cardiac plasma-membrane KATP channel inhibition is defibrillatory and improves survival during acute myocardial ischemia and reperfusion. Eur J Pharmacol. 2007;577(1–3):115–123. doi: 10.1016/j.ejphar.2007.08.016. [DOI] [PubMed] [Google Scholar]
- 156.Fischbach PS, White A, Barrett TD, Lucchesi BR. Risk of ventricular proarrhythmia with selective opening of the myocardial sarcolemmal versus mitochondrial ATP-gated potassium channel. J Pharmacol Exp Ther. 2004;309(2):554–559. doi: 10.1124/jpet.103.060780. [DOI] [PubMed] [Google Scholar]
- 157.Wirth KJ, Rosenstein B, Uhde J, Englert HC, Busch AE, Scholkens BA. ATP-sensitive potassium channel blocker HMR 1883 reduces mortality and ischemia-associated electrocardiographic changes in pigs with coronary occlusion. J Pharmacol Exp Ther. 1999;291(2):474–481. [PubMed] [Google Scholar]
- 158.Cacciapuoti F, Spiezia R, Bianchi U, Lama D, D’Avino M, Varricchio M. Effectiveness of glibenclamide on myocardial ischemic ventricular arrhythmias in non-insulin-dependent diabetes mellitus. Am J Cardiol. 1991;67(9):843–847. doi: 10.1016/0002-9149(91)90617-t. [DOI] [PubMed] [Google Scholar]
- 159.Aronson D, Mittleman MA, Burger AJ. Effects of sulfonylurea hypoglycemic agents and adenosine triphosphate dependent potassium channel antagonists on ventricular arrhythmias in patients with decompensated heart failure. Pacing Clin Electrophysiol. 2003;26(5):1254–1261. doi: 10.1046/j.1460-9592.2003.t01-1-00177.x. [DOI] [PubMed] [Google Scholar]
- 160.Lomuscio A, Vergani D, Marano L, Castagnone M, Fiorentini C. Effects of glibenclamide on ventricular fibrillation in non-insulin-dependent diabetics with acute myocardial infarction. Coron Artery Dis. 1994;5(9):767–771. [PubMed] [Google Scholar]
- 161.Sasaki N, Sato T, Marban E, O’Rourke B. ATP consumption by uncoupled mitochondria activates sarcolemmal KATP channels in cardiac myocytes. Am J Physiol Heart Circ Physiol. 2001;280(4):H1882–1888. doi: 10.1152/ajpheart.2001.280.4.H1882. [DOI] [PubMed] [Google Scholar]
- 162.Weiss JN, Lamp ST. Glycolysis preferentially inhibits ATP-sensitive K+ channels in isolated guinea pig cardiac myocytes. Science. 1987;238(4823):67–69. doi: 10.1126/science.2443972. [DOI] [PubMed] [Google Scholar]
- 163.Jovanovic S, Du Q, Crawford RM, Budas GR, Stagljar I, Jovanovic A. Glyceraldehyde 3-phosphate dehydrogenase serves as an accessory protein of the cardiac sarcolemmal K(ATP) channel. EMBO Rep. 2005;6(9):848–852. doi: 10.1038/sj.embor.7400489. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 164.Dhar-Chowdhury P, Harrell MD, Han SY, Jankowska D, Parachuru L, Morrissey A, Srivastava S, Liu W, Malester B, Yoshida H, Coetzee WA. The glycolytic enzymes, glyceraldehyde-3-phosphate dehydrogenase, triose-phosphate isomerase, and pyruvate kinase are components of the K(ATP) channel macromolecular complex and regulate its function. J Biol Chem. 2005;280(46):38464–38470. doi: 10.1074/jbc.M508744200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 165.Smith JM, Wahler GM. ATP-sensitive potassium channels are altered in ventricular myocytes from diabetic rats. Mol Cell Biochem. 1996;158(1):43–51. doi: 10.1007/BF00225881. [DOI] [PubMed] [Google Scholar]
- 166.Davies JK, Wells DJ, Liu K, Whitrow HR, Daniel TD, Grignani R, Lygate CA, Schneider JE, Noel G, Watkins H, Carling D. Characterization of the role of gamma2 R531G mutation in AMP-activated protein kinase in cardiac hypertrophy and Wolff-Parkinson-White syndrome. Am J Physiol Heart Circ Physiol. 2006;290(5):H1942–1951. doi: 10.1152/ajpheart.01020.2005. [DOI] [PubMed] [Google Scholar]
- 167.Sidhu JS, Rajawat YS, Rami TG, Gollob MH, Wang Z, Yuan R, Marian AJ, DeMayo FJ, Weilbacher D, Taffet GE, Davies JK, Carling D, Khoury DS, Roberts R. Transgenic mouse model of ventricular preexcitation and atrioventricular reentrant tachycardia induced by an AMP-activated protein kinase loss-of-function mutation responsible for Wolff-Parkinson-White syndrome. Circulation. 2005;111(1):21–29. doi: 10.1161/01.CIR.0000151291.32974.D5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 168.Bonnet D, Martin D, Pascale De L, Villain E, Jouvet P, Rabier D, Brivet M, Saudubray JM. Arrhythmias and conduction defects as presenting symptoms of fatty acid oxidation disorders in children. Circulation. 1999;100(22):2248–2253. doi: 10.1161/01.cir.100.22.2248. [DOI] [PubMed] [Google Scholar]
- 169.Chiu HC, Kovacs A, Blanton RM, Han X, Courtois M, Weinheimer CJ, Yamada KA, Brunet S, Xu H, Nerbonne JM, Welch MJ, Fettig NM, Sharp TL, Sambandam N, Olson KM, Ory DS, Schaffer JE. Transgenic expression of fatty acid transport protein 1 in the heart causes lipotoxic cardiomyopathy. Circ Res. 2005;96(2):225–233. doi: 10.1161/01.RES.0000154079.20681.B9. [DOI] [PubMed] [Google Scholar]
- 170.Janse MJ, Wit AL. Electrophysiological mechanisms of ventricular arrhythmias resulting from myocardial ischemia and infarction. Physiol Rev. 1989;69(4):1049–1169. doi: 10.1152/physrev.1989.69.4.1049. [DOI] [PubMed] [Google Scholar]
- 171.Fozzard HA, Makielski JC. The electrophysiology of acute myocardial ischemia. Annu Rev Med. 1985;36:275–284. doi: 10.1146/annurev.me.36.020185.001423. [DOI] [PubMed] [Google Scholar]
- 172.Janse MJ. Electrophysiological changes in heart failure and their relationship to arrhythmogenesis. Cardiovasc Res. 2004;61(2):208–217. doi: 10.1016/j.cardiores.2003.11.018. [DOI] [PubMed] [Google Scholar]
- 173.Liu M, Yang KC, Dudley SC., Jr Cardiac sodium channel mutations: why so many phenotypes? Nat Rev Cardiol. 2014 doi: 10.1038/nrcardio.2014.85. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 174.Song Y, Shryock JC, Wagner S, Maier LS, Belardinelli L. Blocking late sodium current reduces hydrogen peroxide-induced arrhythmogenic activity and contractile dysfunction. J Pharmacol Exp Ther. 2006;318(1):214–222. doi: 10.1124/jpet.106.101832. [DOI] [PubMed] [Google Scholar]
- 175.Shang LL, Sanyal S, Pfahnl AE, Jiao Z, Allen J, Liu H, Dudley SC., Jr NF-kappaB-dependent transcriptional regulation of the cardiac scn5a sodium channel by angiotensin II. Am J Physiol Cell Physiol. 2008;294(1):C372–379. doi: 10.1152/ajpcell.00186.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 176.Gao G, Xie A, Zhang J, Herman AM, Jeong EM, Gu L, Liu M, Yang KC, Kamp TJ, Dudley SC. Unfolded protein response regulates cardiac sodium current in systolic human heart failure. Circ Arrhythm Electrophysiol. 2013;6(5):1018–1024. doi: 10.1161/CIRCEP.113.000274. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 177.Kassmann M, Hansel A, Leipold E, Birkenbeil J, Lu SQ, Hoshi T, Heinemann SH. Oxidation of multiple methionine residues impairs rapid sodium channel inactivation. Pflugers Arch. 2008;456(6):1085–1095. doi: 10.1007/s00424-008-0477-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 178.Fukuda K, Davies SS, Nakajima T, Ong BH, Kupershmidt S, Fessel J, Amarnath V, Anderson ME, Boyden PA, Viswanathan PC, Roberts LJ, 2nd, Balser JR. Oxidative mediated lipid peroxidation recapitulates proarrhythmic effects on cardiac sodium channels. Circ Res. 2005;97(12):1262–1269. doi: 10.1161/01.RES.0000195844.31466.e9. [DOI] [PubMed] [Google Scholar]
- 179.Liu M, Sanyal S, Gao G, Gurung IS, Zhu X, Gaconnet G, Kerchner LJ, Shang LL, Huang CL, Grace A, London B, Dudley SC., Jr Cardiac Na+ current regulation by pyridine nucleotides. Circ Res. 2009;105(8):737–745. doi: 10.1161/CIRCRESAHA.109.197277. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 180.Depre C, Fierain L, Hue L. Activation of nitric oxide synthase by ischaemia in the perfused heart. Cardiovasc Res. 1997;33(1):82–87. doi: 10.1016/s0008-6363(96)00176-9. [DOI] [PubMed] [Google Scholar]
- 181.Wang P, Zweier JL. Measurement of nitric oxide and peroxynitrite generation in the postischemic heart. Evidence for peroxynitrite-mediated reperfusion injury. J Biol Chem. 1996;271(46):29223–29230. doi: 10.1074/jbc.271.46.29223. [DOI] [PubMed] [Google Scholar]
- 182.Simon JN, Duglan D, Casadei B, Carnicer R. Nitric oxide synthase regulation of cardiac excitation-contraction coupling in health and disease. J Mol Cell Cardiol. 2014;73:80–91. doi: 10.1016/j.yjmcc.2014.03.004. [DOI] [PubMed] [Google Scholar]
- 183.Weerateerangkul P, Chattipakorn S, Chattipakorn N. Roles of the nitric oxide signaling pathway in cardiac ischemic preconditioning against myocardial ischemia-reperfusion injury. Med Sci Monit. 2011;17(2):RA44–52. doi: 10.12659/MSM.881385. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 184.Ahern GP, Hsu SF, Klyachko VA, Jackson MB. Induction of persistent sodium current by exogenous and endogenous nitric oxide. J Biol Chem. 2000;275(37):28810–28815. doi: 10.1074/jbc.M003090200. [DOI] [PubMed] [Google Scholar]
- 185.Hammarstrom AK, Gage PW. Nitric oxide increases persistent sodium current in rat hippocampal neurons. J Physiol. 1999;520(Pt 2):451–461. doi: 10.1111/j.1469-7793.1999.t01-1-00451.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 186.Ahmmed GU, Xu Y, Hong Dong P, Zhang Z, Eiserich J, Chiamvimonvat N. Nitric oxide modulates cardiac Na+ channel via protein kinase A and protein kinase G. Circ Res. 2001;89(11):1005–1013. doi: 10.1161/hh2301.100801. [DOI] [PubMed] [Google Scholar]
- 187.Gautier M, Zhang H, Fearon IM. Peroxynitrite formation mediates LPC-induced augmentation of cardiac late sodium currents. J Mol Cell Cardiol. 2008;44(2):241–251. doi: 10.1016/j.yjmcc.2007.09.007. [DOI] [PubMed] [Google Scholar]
- 188.Goldhaber JI, Ji S, Lamp ST, Weiss JN. Effects of exogenous free radicals on electromechanical function and metabolism in isolated rabbit and guinea pig ventricle. Implications for ischemia and reperfusion injury. J Clin Invest. 1989;83(6):1800–1809. doi: 10.1172/JCI114085. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 189.Viola HM, Arthur PG, Hool LC. Transient exposure to hydrogen peroxide causes an increase in mitochondria-derived superoxide as a result of sustained alteration in L-type Ca2+ channel function in the absence of apoptosis in ventricular myocytes. Circ Res. 2007;100(7):1036–1044. doi: 10.1161/01.RES.0000263010.19273.48. [DOI] [PubMed] [Google Scholar]
- 190.Fearon IM. OxLDL enhances L-type Ca2+ currents via lysophosphatidylcholine-induced mitochondrial reactive oxygen species (ROS) production. Cardiovasc Res. 2006;69(4):855–864. doi: 10.1016/j.cardiores.2005.11.019. [DOI] [PubMed] [Google Scholar]
- 191.Coetzee WA, Opie LH. Effects of oxygen free radicals on isolated cardiac myocytes from guinea-pig ventricle: electrophysiological studies. J Mol Cell Cardiol. 1992;24(6):651–663. doi: 10.1016/0022-2828(92)91049-b. [DOI] [PubMed] [Google Scholar]
- 192.Hammerschmidt S, Wahn H. The effect of the oxidant hypochlorous acid on the L-type calcium current in isolated ventricular cardiac myocytess. J Mol Cell Cardiol. 1998;30(9):1855–1867. doi: 10.1006/jmcc.1998.0749. [DOI] [PubMed] [Google Scholar]
- 193.Kawakami M, Okabe E. Superoxide anion radical-triggered Ca2+ release from cardiac sarcoplasmic reticulum through ryanodine receptor Ca2+ channel. Mol Pharmacol. 1998;53(3):497–503. doi: 10.1124/mol.53.3.497. [DOI] [PubMed] [Google Scholar]
- 194.Yan Y, Liu J, Wei C, Li K, Xie W, Wang Y, Cheng H. Bidirectional regulation of Ca2+ sparks by mitochondria-derived reactive oxygen species in cardiac myocytes. Cardiovasc Res. 2008;77(2):432–441. doi: 10.1093/cvr/cvm047. [DOI] [PubMed] [Google Scholar]
- 195.Scherer NM, Deamer DW. Oxidative stress impairs the function of sarcoplasmic reticulum by oxidation of sulfhydryl groups in the Ca2+-ATPase. Arch Biochem Biophys. 1986;246(2):589–601. doi: 10.1016/0003-9861(86)90314-0. [DOI] [PubMed] [Google Scholar]
- 196.Kukreja RC, Kearns AA, Zweier JL, Kuppusamy P, Hess ML. Singlet oxygen interaction with Ca2+-ATPase of cardiac sarcoplasmic reticulum. Circ Res. 1991;69(4):1003–1014. doi: 10.1161/01.res.69.4.1003. [DOI] [PubMed] [Google Scholar]
- 197.Xu KY, Zweier JL, Becker LC. Hydroxyl radical inhibits sarcoplasmic reticulum Ca2+-ATPase function by direct attack on the ATP binding site. Circ Res. 1997;80(1):76–81. doi: 10.1161/01.res.80.1.76. [DOI] [PubMed] [Google Scholar]
- 198.Eigel BN, Gursahani H, Hadley RW. ROS are required for rapid reactivation of Na+/Ca2+ exchanger in hypoxic reoxygenated guinea pig ventricular myocytes. Am J Physiol Heart Circ Physiol. 2004;286(3):H955–963. doi: 10.1152/ajpheart.00721.2003. [DOI] [PubMed] [Google Scholar]
- 199.Goldhaber JI. Free radicals enhance Na+/Ca2+ exchange in ventricular myocytes. Am J Physiol. 1996;271(3 Pt 2):H823–833. doi: 10.1152/ajpheart.1996.271.3.H823. [DOI] [PubMed] [Google Scholar]
- 200.Rokita AG, Anderson ME. New therapeutic targets in cardiology: arrhythmias and Ca2+/calmodulin-dependent kinase II (CaMKII) Circulation. 2012;126(17):2125–2139. doi: 10.1161/CIRCULATIONAHA.112.124990. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 201.Wu Y, MacMillan LB, McNeill RB, Colbran RJ, Anderson ME. CaM kinase augments cardiac L-type Ca2+ current: a cellular mechanism for long Q-T arrhythmias. Am J Physiol. 1999;276(6 Pt 2):H2168–2178. doi: 10.1152/ajpheart.1999.276.6.H2168. [DOI] [PubMed] [Google Scholar]
- 202.Di Lisa F, Blank PS, Colonna R, Gambassi G, Silverman HS, Stern MD, Hansford RG. Mitochondrial membrane potential in single living adult rat cardiac myocytes exposed to anoxia or metabolic inhibition. J Physiol. 1995;486 (Pt 1):1–13. doi: 10.1113/jphysiol.1995.sp020786. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 203.Griffiths EJ, Ocampo CJ, Savage JS, Rutter GA, Hansford RG, Stern MD, Silverman HS. Mitochondrial calcium transporting pathways during hypoxia and reoxygenation in single rat cardiac myocytess. Cardiovasc Res. 1998;39(2):423–433. doi: 10.1016/s0008-6363(98)00104-7. [DOI] [PubMed] [Google Scholar]
- 204.Perez-Campo R, Lopez-Torres M, Cadenas S, Rojas C, Barja G. The rate of free radical production as a determinant of the rate of aging: evidence from the comparative approach. J Comp Physiol B. 1998;168(3):149–158. doi: 10.1007/s003600050131. [DOI] [PubMed] [Google Scholar]
- 205.Sohal RS, Allen RG. Relationship between metabolic rate, free radicals, differentiation and aging: a unified theory. Basic Life Sci. 1985;35:75–104. doi: 10.1007/978-1-4899-2218-2_4. [DOI] [PubMed] [Google Scholar]
- 206.Cleeter MW, Cooper JM, Darley-Usmar VM, Moncada S, Schapira AH. Reversible inhibition of cytochrome c oxidase, the terminal enzyme of the mitochondrial respiratory chain, by nitric oxide. Implications for neurodegenerative diseases. FEBS Lett. 1994;345(1):50–54. doi: 10.1016/0014-5793(94)00424-2. [DOI] [PubMed] [Google Scholar]
- 207.Ott M, Robertson JD, Gogvadze V, Zhivotovsky B, Orrenius S. Cytochrome c release from mitochondria proceeds by a two-step process. Proc Natl Acad Sci U S A. 2002;99(3):1259–1263. doi: 10.1073/pnas.241655498. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 208.Grijalba MT, Vercesi AE, Schreier S. Ca2+-induced increased lipid packing and domain formation in submitochondrial particles. A possible early step in the mechanism of Ca2+-stimulated generation of reactive oxygen species by the respiratory chain. Biochemistry. 1999;38(40):13279–13287. doi: 10.1021/bi9828674. [DOI] [PubMed] [Google Scholar]
- 209.Dedkova EN, Blatter LA. Characteristics and function of cardiac mitochondrial nitric oxide synthase. J Physiol. 2009;587(Pt 4):851–872. doi: 10.1113/jphysiol.2008.165423. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 210.Feissner RF, Skalska J, Gaum WE, Sheu SS. Crosstalk signaling between mitochondrial Ca2+ and ROS. Front Biosci (Landmark Ed) 2009;14:1197–1218. doi: 10.2741/3303. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 211.Hu H, Chiamvimonvat N, Yamagishi T, Marban E. Direct inhibition of expressed cardiac L-type Ca2+ channels by S-nitrosothiol nitric oxide donors. Circ Res. 1997;81(5):742–752. doi: 10.1161/01.res.81.5.742. [DOI] [PubMed] [Google Scholar]
- 212.Sun J, Picht E, Ginsburg KS, Bers DM, Steenbergen C, Murphy E. Hypercontractile female hearts exhibit increased S-nitrosylation of the L-type Ca2+ channel alpha1 subunit and reduced ischemia/reperfusion injury. Circ Res. 2006;98(3):403–411. doi: 10.1161/01.RES.0000202707.79018.0a. [DOI] [PubMed] [Google Scholar]
- 213.Wang H, Kohr MJ, Wheeler DG, Ziolo MT. Endothelial nitric oxide synthase decreases beta-adrenergic responsiveness via inhibition of the L-type Ca2+ current. Am J Physiol Heart Circ Physiol. 2008;294(3):H1473–1480. doi: 10.1152/ajpheart.01249.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 214.Sun J, Yamaguchi N, Xu L, Eu JP, Stamler JS, Meissner G. Regulation of the cardiac muscle ryanodine receptor by O2 tension and S-nitrosoglutathione. Biochemistry. 2008;47(52):13985–13990. doi: 10.1021/bi8012627. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 215.Lokuta AJ, Maertz NA, Meethal SV, Potter KT, Kamp TJ, Valdivia HH, Haworth RA. Increased nitration of sarcoplasmic reticulum Ca2+-ATPase in human heart failure. Circulation. 2005;111(8):988–995. doi: 10.1161/01.CIR.0000156461.81529.D7. [DOI] [PubMed] [Google Scholar]
- 216.Cerbai E, Ambrosio G, Porciatti F, Chiariello M, Giotti A, Mugelli A. Cellular electrophysiological basis for oxygen radical-induced arrhythmias. A patch-clamp study in guinea pig ventricular myocytes. Circulation. 1991;84(4):1773–1782. doi: 10.1161/01.cir.84.4.1773. [DOI] [PubMed] [Google Scholar]
- 217.Liang H, Li X, Li S, Zheng MQ, Rozanski GJ. Oxidoreductase regulation of Kv currents in rat ventricle. J Mol Cell Cardiol. 2008;44(6):1062–1071. doi: 10.1016/j.yjmcc.2008.03.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 218.Li X, Tang K, Xie B, Li S, Rozanski GJ. Regulation of Kv4 channel expression in failing rat heart by the thioredoxin system. Am J Physiol Heart Circ Physiol. 2008;295(1):H416–424. doi: 10.1152/ajpheart.91446.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 219.Li S, Li X, Li YL, Shao CH, Bidasee KR, Rozanski GJ. Insulin regulation of glutathione and contractile phenotype in diabetic rat ventricular myocytes. Am J Physiol Heart Circ Physiol. 2007;292(3):H1619–1629. doi: 10.1152/ajpheart.00140.2006. [DOI] [PubMed] [Google Scholar]
- 220.Chiarugi P. PTPs versus PTKs: the redox side of the coin. Free Radic Res. 2005;39(4):353–364. doi: 10.1080/10715760400027987. [DOI] [PubMed] [Google Scholar]
- 221.Nakamura TY, Coetzee WA, Vega-Saenz De Miera E, Artman M, Rudy B. Modulation of Kv4 channels, key components of rat ventricular transient outward K+ current, by PKC. Am J Physiol. 1997;273(4 Pt 2):H1775–1786. doi: 10.1152/ajpheart.1997.273.4.H1775. [DOI] [PubMed] [Google Scholar]
- 222.Li GR, Feng J, Wang Z, Fermini B, Nattel S. Adrenergic modulation of ultrarapid delayed rectifier K+ current in human atrial myocytes. Circ Res. 1996;78(5):903–915. doi: 10.1161/01.res.78.5.903. [DOI] [PubMed] [Google Scholar]
- 223.Gomez R, Nunez L, Vaquero M, Amoros I, Barana A, de Prada T, Macaya C, Maroto L, Rodriguez E, Caballero R, Lopez-Farre A, Tamargo J, Delpon E. Nitric oxide inhibits Kv4.3 and human cardiac transient outward potassium current (Ito1) Cardiovasc Res. 2008;80(3):375–384. doi: 10.1093/cvr/cvn205. [DOI] [PubMed] [Google Scholar]
- 224.Nunez L, Vaquero M, Gomez R, Caballero R, Mateos-Caceres P, Macaya C, Iriepa I, Galvez E, Lopez-Farre A, Tamargo J, Delpon E. Nitric oxide blocks hKv1.5 channels by S-nitrosylation and by a cyclic GMP-dependent mechanism. Cardiovasc Res. 2006;72(1):80–89. doi: 10.1016/j.cardiores.2006.06.021. [DOI] [PubMed] [Google Scholar]
- 225.Ryu SY, Lee SH, Ho WK. Generation of metabolic oscillations by mitoKATP and ATP synthase during simulated ischemia in ventricular myocytes. J Mol Cell Cardiol. 2005;39(6):874–881. doi: 10.1016/j.yjmcc.2005.08.011. [DOI] [PubMed] [Google Scholar]
- 226.Zorov DB, Filburn CR, Klotz LO, Zweier JL, Sollott SJ. Reactive oxygen species (ROS)-induced ROS release: a new phenomenon accompanying induction of the mitochondrial permeability transition in cardiac myocytes. J Exp Med. 2000;192(7):1001–1014. doi: 10.1084/jem.192.7.1001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 227.Brown DA, Aon MA, Akar FG, Liu T, Sorarrain N, O’Rourke B. Effects of 4′-chlorodiazepam on cellular excitation-contraction coupling and ischaemia-reperfusion injury in rabbit heart. Cardiovasc Res. 2008;79(1):141–149. doi: 10.1093/cvr/cvn053. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 228.Brown DA, Aon MA, Frasier CR, Sloan RC, Maloney AH, Anderson EJ, O’Rourke B. Cardiac arrhythmias induced by glutathione oxidation can be inhibited by preventing mitochondrial depolarization. J Mol Cell Cardiol. 2010;48(4):673–679. doi: 10.1016/j.yjmcc.2009.11.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 229.O’Rourke B. Evidence for mitochondrial K+ channels and their role in cardioprotection. Circ Res. 2004;94(4):420–432. doi: 10.1161/01.RES.0000117583.66950.43. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 230.Garlid KD, Dos Santos P, Xie ZJ, Costa AD, Paucek P. Mitochondrial potassium transport: the role of the mitochondrial ATP-sensitive K+ channel in cardiac function and cardioprotection. Biochim Biophys Acta. 2003;1606(1–3):1–21. doi: 10.1016/s0005-2728(03)00109-9. [DOI] [PubMed] [Google Scholar]
- 231.Murata M, Akao M, O’Rourke B, Marban E. Mitochondrial ATP-sensitive potassium channels attenuate matrix Ca2+ overload during simulated ischemia and reperfusion: possible mechanism of cardioprotection. Circ Res. 2001;89(10):891–898. doi: 10.1161/hh2201.100205. [DOI] [PubMed] [Google Scholar]
- 232.Wang L, Cherednichenko G, Hernandez L, Halow J, Camacho SA, Figueredo V, Schaefer S. Preconditioning limits mitochondrial Ca2+ during ischemia in rat hearts: role of K(ATP) channels. Am J Physiol Heart Circ Physiol. 2001;280(5):H2321–2328. doi: 10.1152/ajpheart.2001.280.5.H2321. [DOI] [PubMed] [Google Scholar]
- 233.Oldenburg O, Cohen MV, Downey JM. Mitochondrial KATP channels in preconditioning. J Mol Cell Cardiol. 2003;35(6):569–575. doi: 10.1016/s0022-2828(03)00115-9. [DOI] [PubMed] [Google Scholar]
- 234.Forbes RA, Steenbergen C, Murphy E. Diazoxide-induced cardioprotection requires signaling through a redox-sensitive mechanism. Circ Res. 2001;88(8):802–809. doi: 10.1161/hh0801.089342. [DOI] [PubMed] [Google Scholar]
- 235.Sasaki N, Sato T, Ohler A, O’Rourke B, Marban E. Activation of mitochondrial ATP-dependent potassium channels by nitric oxide. Circulation. 2000;101(4):439–445. doi: 10.1161/01.cir.101.4.439. [DOI] [PubMed] [Google Scholar]
- 236.Bolli R, Manchikalapudi S, Tang XL, Takano H, Qiu Y, Guo Y, Zhang Q, Jadoon AK. The protective effect of late preconditioning against myocardial stunning in conscious rabbits is mediated by nitric oxide synthase. Evidence that nitric oxide acts both as a trigger and as a mediator of the late phase of ischemic preconditioning. Circ Res. 1997;81(6):1094–1107. doi: 10.1161/01.res.81.6.1094. [DOI] [PubMed] [Google Scholar]
- 237.Bolli R, Bhatti ZA, Tang XL, Qiu Y, Zhang Q, Guo Y, Jadoon AK. Evidence that late preconditioning against myocardial stunning in conscious rabbits is triggered by the generation of nitric oxide. Circ Res. 1997;81(1):42–52. doi: 10.1161/01.res.81.1.42. [DOI] [PubMed] [Google Scholar]
- 238.Vegh A, Parratt JR. The role of mitochondrial KATP channels in antiarrhythmic effects of ischaemic preconditioning in dogs. Br J Pharmacol. 2002;137(7):1107–1115. doi: 10.1038/sj.bjp.0704966. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 239.Headrick JP, Willems L, Ashton KJ, Holmgren K, Peart J, Matherne GP. Ischaemic tolerance in aged mouse myocardium: the role of adenosine and effects of A1 adenosine receptor overexpression. J Physiol. 2003;549(Pt 3):823–833. doi: 10.1113/jphysiol.2003.041541. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 240.Das B, Sarkar C. Is the sarcolemmal or mitochondrial KATP channel activation important in the antiarrhythmic and cardioprotective effects during acute ischemia/reperfusion in the intact anesthetized rabbit model? Life Sci. 2005;77(11):1226–1248. doi: 10.1016/j.lfs.2004.12.042. [DOI] [PubMed] [Google Scholar]
- 241.Lo CW. Role of gap junctions in cardiac conduction and development: insights from the connexin knockout mice. Circ Res. 2000;87(5):346–348. doi: 10.1161/01.res.87.5.346. [DOI] [PubMed] [Google Scholar]
- 242.Litchenberg WH, Norman LW, Holwell AK, Martin KL, Hewett KW, Gourdie RG. The rate and anisotropy of impulse propagation in the postnatal terminal crest are correlated with remodeling of Cx43 gap junction pattern. Cardiovasc Res. 2000;45(2):379–387. doi: 10.1016/s0008-6363(99)00363-6. [DOI] [PubMed] [Google Scholar]
- 243.Kaprielian RR, Gunning M, Dupont E, Sheppard MN, Rothery SM, Underwood R, Pennell DJ, Fox K, Pepper J, Poole-Wilson PA, Severs NJ. Downregulation of immunodetectable connexin43 and decreased gap junction size in the pathogenesis of chronic hibernation in the human left ventricle. Circulation. 1998;97(7):651–660. doi: 10.1161/01.cir.97.7.651. [DOI] [PubMed] [Google Scholar]
- 244.Peters NS, Green CR, Poole-Wilson PA, Severs NJ. Reduced content of connexin43 gap junctions in ventricular myocardium from hypertrophied and ischemic human hearts. Circulation. 1993;88(3):864–875. doi: 10.1161/01.cir.88.3.864. [DOI] [PubMed] [Google Scholar]
- 245.Dupont E, Matsushita T, Kaba RA, Vozzi C, Coppen SR, Khan N, Kaprielian R, Yacoub MH, Severs NJ. Altered connexin expression in human congestive heart failure. J Mol Cell Cardiol. 2001;33(2):359–371. doi: 10.1006/jmcc.2000.1308. [DOI] [PubMed] [Google Scholar]
- 246.Kostin S, Rieger M, Dammer S, Hein S, Richter M, Klovekorn WP, Bauer EP, Schaper J. Gap junction remodeling and altered connexin43 expression in the failing human heart. Mol Cell Biochem. 2003;242(1–2):135–144. [PubMed] [Google Scholar]
- 247.Jongsma HJ, Wilders R. Gap junctions in cardiovascular disease. Circ Res. 2000;86(12):1193–1197. doi: 10.1161/01.res.86.12.1193. [DOI] [PubMed] [Google Scholar]
- 248.Gutstein DE, Morley GE, Tamaddon H, Vaidya D, Schneider MD, Chen J, Chien KR, Stuhlmann H, Fishman GI. Conduction slowing and sudden arrhythmic death in mice with cardiac-restricted inactivation of connexin43. Circ Res. 2001;88(3):333–339. doi: 10.1161/01.res.88.3.333. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 249.Efimov IR. Connections, connections, connexins: towards systems biology paradigm of cardiac arrhythmia. J Mol Cell Cardiol. 2006;41(6):949–951. doi: 10.1016/j.yjmcc.2006.09.003. [DOI] [PubMed] [Google Scholar]
- 250.van Rijen HV, van Veen TA, Gros D, Wilders R, de Bakker JM. Connexins and cardiac arrhythmias. Adv Cardiol. 2006;42:150–160. doi: 10.1159/000092567. [DOI] [PubMed] [Google Scholar]
- 251.Ikram H. The renin-angiotensin-aldosterone system and cardiac ischaemia. Heart. 1996;76(3 Suppl 3):60–67. doi: 10.1136/hrt.76.3_suppl_3.60. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 252.Santos RA, Brum JM, Brosnihan KB, Ferrario CM. The renin-angiotensin system during acute myocardial ischemia in dogs. Hypertension. 1990;15(2 Suppl):I121–127. doi: 10.1161/01.hyp.15.2_suppl.i121. [DOI] [PubMed] [Google Scholar]
- 253.De Mello WC. Renin-angiotensin system and cell communication in the failing heart. Hypertension. 1996;27(6):1267–1272. doi: 10.1161/01.hyp.27.6.1267. [DOI] [PubMed] [Google Scholar]
- 254.De Mello WC. Is an intracellular renin-angiotensin system involved in control of cell communication in heart? J Cardiovasc Pharmacol. 1994;23(4):640–646. doi: 10.1097/00005344-199404000-00018. [DOI] [PubMed] [Google Scholar]
- 255.Xiao HD, Fuchs S, Campbell DJ, Lewis W, Dudley SC, Jr, Kasi VS, Hoit BD, Keshelava G, Zhao H, Capecchi MR, Bernstein KE. Mice with cardiac-restricted angiotensin-converting enzyme (ACE) have atrial enlargement, cardiac arrhythmia, and sudden death. Am J Pathol. 2004;165(3):1019–1032. doi: 10.1016/S0002-9440(10)63363-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 256.Donoghue M, Wakimoto H, Maguire CT, Acton S, Hales P, Stagliano N, Fairchild-Huntress V, Xu J, Lorenz JN, Kadambi V, Berul CI, Breitbart RE. Heart block, ventricular tachycardia, and sudden death in ACE2 transgenic mice with downregulated connexins. J Mol Cell Cardiol. 2003;35(9):1043–1053. doi: 10.1016/s0022-2828(03)00177-9. [DOI] [PubMed] [Google Scholar]
- 257.Sovari AA, Iravanian S, Dolmatova E, Jiao Z, Liu H, Zandieh S, Kumar V, Wang K, Bernstein KE, Bonini MG, Duffy HS, Dudley SC. Inhibition of c-Src tyrosine kinase prevents angiotensin II-mediated connexin-43 remodeling and sudden cardiac death. J Am Coll Cardiol. 2011;58(22):2332–2339. doi: 10.1016/j.jacc.2011.07.048. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 258.Iravanian S, Sovari AA, Lardin HA, Liu H, Xiao HD, Dolmatova E, Jiao Z, Harris BS, Witham EA, Gourdie RG, Duffy HS, Bernstein KE, Dudley SC., Jr Inhibition of renin-angiotensin system (RAS) reduces ventricular tachycardia risk by altering connexin43. J Mol Med (Berl) 2011;89(7):677–687. doi: 10.1007/s00109-011-0761-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 259.Kieken F, Mutsaers N, Dolmatova E, Virgil K, Wit AL, Kellezi A, Hirst-Jensen BJ, Duffy HS, Sorgen PL. Structural and molecular mechanisms of gap junction remodeling in epicardial border zone myocytes following myocardial infarction. Circ Res. 2009;104(9):1103–1112. doi: 10.1161/CIRCRESAHA.108.190454. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 260.Toyofuku T, Yabuki M, Otsu K, Kuzuya T, Tada M, Hori M. Functional role of c-Src in gap junctions of the cardiomyopathic heart. Circ Res. 1999;85(8):672–681. doi: 10.1161/01.res.85.8.672. [DOI] [PubMed] [Google Scholar]
- 261.Myung SK, Ju W, Cho B, Oh SW, Park SM, Koo BK, Park BJ. Efficacy of vitamin and antioxidant supplements in prevention of cardiovascular disease: systematic review and meta-analysis of randomised controlled trials. BMJ. 2013;346:f10. doi: 10.1136/bmj.f10. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 262.Yang KC, Rutledge CA, Mao M, Bakhshi FR, Xie A, Liu H, Bonini MG, Patel HH, Minshall RD, Dudley SC., Jr Caveolin-1 Modulates Cardiac Gap Junction Homeostasis and Arrhythmogenecity by Regulating cSrc Tyrosine Kinase. Circ Arrhythm Electrophysiol. 2014;7(4):701–710. doi: 10.1161/CIRCEP.113.001394. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 263.Hatch F, Lancaster MK, Jones SA. Aging is a primary risk factor for cardiac arrhythmias: disruption of intracellular Ca2+ regulation as a key suspect. Expert Rev Cardiovasc Ther. 2011;9(8):1059–1067. doi: 10.1586/erc.11.112. [DOI] [PubMed] [Google Scholar]
- 264.Ludka O, Konecny T, Somers V. Sleep apnea, cardiac arrhythmias, and sudden death. Tex Heart Inst J. 2011;38(4):340–343. [PMC free article] [PubMed] [Google Scholar]
- 265.Roberts PR, Green D. Arrhythmias in chronic kidney disease. Heart. 2011;97(9):766–773. doi: 10.1136/hrt.2010.208587. [DOI] [PubMed] [Google Scholar]
- 266.Biernacka A, Frangogiannis NG. Aging and Cardiac Fibrosis. Aging Dis. 2011;2(2):158–173. [PMC free article] [PubMed] [Google Scholar]
- 267.Asbun J, Villarreal FJ. The pathogenesis of myocardial fibrosis in the setting of diabetic cardiomyopathy. J Am Coll Cardiol. 2006;47(4):693–700. doi: 10.1016/j.jacc.2005.09.050. [DOI] [PubMed] [Google Scholar]
- 268.Xie C, Kauffman J, Akar FG. Functional crosstalk between the mitochondrial PTP and KATP channels determine arrhythmic vulnerability to oxidative stress. Front Physiol. 2014;5:264. doi: 10.3389/fphys.2014.00264. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
