Chapter 15
Cardiomyocyte Metabolism: All Is in Flux Heinrich Taegtmeyer Department of Medicine/Cardiology, The University of Texas School of Medicine at Houston, Houston, TX
The intermediary metabolism of energy-providing substrates supports the rhythmic contraction of the heart (Figure 15.1). However, metabolism also encompasses the universal principle of nature that all is in flux. From the Krebs cycle to the cell cycle, from protein synthesis to degradation, from phosphorylation to dephosphorylation, from histone acteylation to deacetylation, all is in flux. First enunciated by Heraclitus in 513 BCE, in modern times no one has seen this principle more clearly than Hans Krebs in the discovery of the citric acid cycle (1) and Rudolph Schoenheimer in “The Dynamic State of Body Constituents” (2). This chapter highlights principles of cardiac energetics, and includes multiple newly discovered roles of intermediary metabolites that regulate cardiac cell function. The chapter also highlights emerging concepts in metabolic regulation at the transcriptional, translational, and post-translational level, and discusses the putative role of metabolic signals in the regulation of myocardial protein turnover. In the cold light of twenty-first century molecular biology metabolic pathways seem a distant memory. Yet the importance of fuel metabolism in the heart is newly appreciated in heart disease, cancer, and diabetes. A brief review of the basics is in order.
FUEL METABOLISM IN PERSPECTIVE As in any other organ of the body, it is impossible to separate function from metabolism in the heart. The bulk of the heart’s energy for contraction, ion movements, and intracellular protein turnover is provided by the metabolism of oxidizable substrates. The final product of oxidative metabolism in the mitochondria is ATP. In the simple scheme (Figure 15.1), a decrease in the flux of metabolic energy results in a decrease in ATP production, a decrease in oxygen consumption and a decrease in contractile function. Vice versa, an increase in energy demand results in increased ATP turnover, an increase in oxygen consumption and an increase in contractile function. The heart’s requirement for calcium, nutrients, and oxygen has been recognized for well over a century (3,4)
Muscle. DOI: http://dx.doi.org/10.1016/B978-0-12-381510-1.00015-6 © 2012 Elsevier Inc. All rights reserved.
and the field of cardiac metabolism has a rich history of integrative and molecular physiology (5). By measuring arteriovenous differences in the heart lung preparation (6) of the dog or in the human heart through cannulation of the coronary sinus (7), early investigators have laid the foundation for current concepts of myocardial energy substrate metabolism. The more recent reviews survey the knowledge gained through isolated heart preparations, isolated cell preparations, and isolated cell organelles (5,8 17). As will be discussed later in the chapter, the toolbox for the discovery of metabolic regulation in genetic animal models and in cardiomyocytes has grown even more dramatically than the clinically useful instrumentation.
MITOCHONDRIA AND THE DYNAMICS OF METABOLISM IN THE HEART Cardiac metabolism maintains a dynamic state of equilibrium for efficient energy transfer at the site of ATP production in the mitochondria and at the site of ATP usage in the cross-bridges. This has three consequences. Hence, the rate of energy turnover, and not the tissue content of ATP, is the main driving force of energy metabolism (10,18 20). The greater the work output, the higher the rate of ATP turnover, the higher the rate of oxygen consumption, and the higher the rate of substrate input and utilization (Figure 15.1). Secondly, energy transfer in the heart obeys the first and second laws of thermodynamics. Within a closed system, energy can only be converted from one form into another, and a process occurs spontaneously only if it is associated with an increase in randomness (or entropy) of the system. Everything else requires catalysts or enzymes. Third, when the heart’s ability to convert chemical into mechanical energy is impaired for any reason (21) the consequences manifest themselves as functional and metabolic derangements in the rest of the body, i.e. heart failure from a disease of impaired energy transfer in the heart.
187
188
PART | II
A close correlation exists among mitochondrial volume fraction, heart rate, and total body oxygen consumption, with mitochondrial volume fractions ranging from 25% humans to 38% in mice (22). Heart muscle mitochondria are not only abundant, they also contain a far greater number of cristae (the location of respiratory chain enzymes) than do mitochondria of other organs, such as liver, brain, or skeletal muscle (23). Lastly, while the term is immensely important, it is no longer sufficient to call mitochondria only the “power stations” of the cell (24). Mitochondria are dynamic cell organelles which divide and fuse, which are subject to organized destruction (mitophagy), which are the sources of cell death signals (cytochrome c) and which are a major target of metabolic dysregulation in diabetes and heart failure. Input: Substrates + O2
Metabolism
ADP + Pi
ADP
PCr
Contraction
Output: Pump action, Heat FIGURE 15.1 The heart converts substrates and oxygen (input) to pump action and heat (output). Metabolism of energy providing substrates fuels ATP production for the contractile process.
THE HIGHWAYS OF ENERGY TRANSFER In the cardiomyocytes, complex chemical reactions proceed rapidly at a relatively low temperature and at low substrate concentrations through enzyme-catalyzed pathways, an essential part of which are cycles (e.g. the Krebs cycle, the ATP cycle), because cycles have evolved as the most efficient form of energy transfer (Figure 15.2) (5,25). The function of both extra- and intra-mitochondrial cycles requires conservation of their constituent moieties. The replenishment of intermediate moieties in a cycle requires anaplerotic pathways (26). Anaplerotic pathways of the Krebs cycle are critical for energy transfer in the heart (27). In addition to the concept of cycles, a few definitions may be useful. A metabolic pathway is defined as a series of enzyme-catalyzed reactions beginning with a flux-generating step (usually a reaction catalyzed by a non-equilibrium reaction or transport of the metabolite across a membrane) and ending with the removal of a product (28,29). Characteristic of most metabolic pathways is that, once flux has been initiated, there is a rapid and concerted response of the entire pathway. In this system of flux, metabolite levels control enzyme activities and, in turn, enzyme activities are controlled by metabolite levels. Control is the power to change metabolic flux in response to an external signal, whereas regulation is the inherent capacity of enzymes geared toward maintaining a constant internal state (30). In such a system, large changes in the flux through metabolic pathways correspond to only very small changes in myocardial metabolite concentration (31). Regulatory sites of metabolism, or pacemaker enzymes (32), have become targets for the manipulation of metabolism with drugs (33,34).
FIGURE 15.2 From the circulation of the blood to the cross-bridges of the sarcomeres energy transfer makes use of a series of moiety conserved cycles. The metabolic cycles are located in the mitochondria (red). Ca21 is considered the main regulator of both contraction and Krebs cycle flux.
Circulation Ca2+ Perfusion Krebs cycle Ca2+
NADH H+ FADH H+ H+ gradient
Ca2+
ATP PCr Mitochondria
Cardiac Muscle
Crossbridges Ca2+
Chapter | 15
Cardiomyocyte Metabolism: All Is in Flux
189
ATP production is tightly coupled to ATP utilization. Hence, substrate oxidation is also tightly coupled to cardiac work (35 37). An acute increase in work load of the isolated working rat heart results in an acute increase in myocardial oxygen consumption (measured) and CO2 production (Figure 15.4a) (predicted); and the heart uses carbohydrates (glycogen, glucose, and lactate) as fuel for respiration to meet the increased demand for chemical energy (Figure 15.4b). In other words, with an increase in
The breakdown of substrates can be divided into four stages (Figure 15.3). The first stage consists of substrate delivery and uptake by the cell. The second stage consists of intracellular pathways leading to acetyl coenzyme A (acetyl-CoA). The third stage consists of the oxidation of acetyl-CoA in the Krebs cycle; and the fourth stage consists of the reaction of reducing equivalents with molecular oxygen in the respiratory chain, where electron transfer is coupled to rephosphorylation of ADP to ATP.
Lactate Substrate uptake
GLUT G-6-P
FIGURE 15.3 The metabolism of energy-providing substrates is organized into three stages, all of which are regulated: substrate uptake, intermediary metabolism, and the oxidation of smaller carbon units in the Krebs cycle providing reducing equivalents for the respiratory chain. (With kind permission from Springer Science 1 Business Media: Cardiovascular Medicine, 3rd edition, 2007, p. 1159, eds Willerson, Cohn, Wellens, Holmes, figure 50.3.)
Free fatty acids
Glucose
Triglycerides
FFA
Glycogen ACS
Lactate
Acyl-CoA
Pyruvate
Metabolism
CPT Acyl-CoA
β Oxidation
Pyruvate PDC Acetyl-CoA
Oxidation CO2
Krebs cycle
NADH H+ ½O2 H2O NADH H+ FADH2
Respiratory chain
ATP
ATP ADP
α-MHC
ATP ADP
ADP
ATP
Ca2+
SERCA
Sarcoplasmic reticulum
7 0 6 0 5 0 4 0 5 0
5 5
measured predicted from oxidation of Exogenous Lipid + Endogenous Lipid + Exogenous Glucose + Endogenous Glucose + Exogenous Lactate + Pyruvate Release 6 6 0 5 Perfusion Time (min)
7 0
7 5
Lactate
4.0
Epinephrine
8 0
Oxidation Rate (umol/min/g dry wt.)
4.5
Epinephrine
Oxygen Consumption, measured predicted (umol.min/g dry wt.)
ATP
β-MHC
9 0
0
ATP + Pi ATP + Pi
Ca2+
Energy utilization
NAD+ FAD
3.5 3.0 2.5
Glucose
2.0
Oleate
1.5 1.0 0.5
Glycogen
0.0 −0.5
5 0
5 5
6 6 0 5 Perfusion Time (min)
7 0
7 5
FIGURE 15.4 In the isolated working heart an acute increase in work load results in an acute increase of oxygen consumption (a) and oxidation of carbohydrates (b). Note the instant, but transient, increase in glycogen oxidation, and the sustained increase in glucose and lactate oxidation. All three substrates support the increase in cardiac work. See text for further detail. (From Goodwin et al., 1998 (35).)
190
energy demand the heart readily adds a second substrate for energy provision. Reduced flavin adenine dinucleotide (FADH2) and NADH are generated, and the reducing equivalents enter the electron transport chain producing an electrochemical gradient across the inner mitochondrial membrane that drives ATP synthesis in the presence of molecular oxygen, ADP, and inorganic phosphate. The exact mechanism by which respiration is coupled to energy expenditure in vivo is, however, still not known (38 40). It can be stated with certainly that for a given physiologic environment the heart oxidizes the most efficient substrate.
TRACING METABOLIC PATHWAYS EX VIVO AND IN VIVO The tracing of metabolic pathways can be accomplished qualitatively owing to the development of new, nondestructive imaging techniques such as nuclear magnetic resonance spectroscopy or positron emission tomography (PET). Both techniques also permit the assessment of regional metabolic processes in the beating heart both ex vivo and in vivo (41 48). Analysis of energy-rich phosphates in the beating heart by nuclear magnetic resonance spectroscopy of 31P lends further support to the view that over a relatively wide range of work output, tissue content of ATP is not correlated with rates of ATP turnover, which can be assessed by oxygen consumption or contractile performance of the heart (49). The recent adaptation of isotopomer analysis of 13C natural abundance or 13C-labeled compounds allows the analysis of flux through specific pathways, especially the citric acid cycle and glycogen turnover, to be studied quantitatively as serial spectra are obtained (48,50 52). More recently, the use of hyperpolarization of 13C biomolecules (53) has made it possible to assess Krebs cycle metabolism with magnetic resonance spectroscopy (MRS) (54). Collectively, MRS signals from 31P, 13C, and other stable isotopes hold the promise to provide further insight into the dynamics of heart metabolism (55). Short-lived positron-emitting tracers have been so successful in their clinical application because they permit a visual assessment of regional differences in metabolic activity of the heart (47,56). Two types can be distinguished. The first approach entails uptake and retention of a tracer analogue, such as [18F-]-2-deoxy, 2-fluoroglucose (FDG). The second approach entails the uptake and clearance of 11C tracers (fatty acids, glucose, or lactate) where the rapid phase of clearance from the tissue represents their rate of oxidation or, in the case of [11C]acetate, oxidation in the citric acid cycle alone (56). Uptake and retention of the tracer analogue FDG increase linearly
PART | II
Cardiac Muscle
with time (57), whereas the clearance of a labeled fatty acid after an initial peak in tissue activity is biexponential (56,58,59). Both types of approaches have been used clinically to assess substrate metabolism in normal and ischemic myocardium. Enhanced glucose uptake (assessed with FDG) is used to identify reversible ischemia in “hibernating” myocardium where heart muscle remodels metabolically and the mismatch of perfusion and metabolism is an indicator of myocardial viability (60). Assessment of myocardial metabolism with [11C] tracers, which are oxidized by the heart, suggests that patients with idiopathic dilated cardiomyopathy exhibit changes in substrate preference characterized by decreased fatty acid metabolism and increased glucose metabolism (61) in keeping with the constitutive expression of the fetal metabolic gene program in the failing human heart (62).
GENETIC MODELS FOR THE ELUCIDATION OF CARDIAC METABOLISM Redirecting energy metabolism is largely orchestrated by proteins that are involved in regulating flux through entire pathways. Accordingly, much has been learned from genetic models in which specific proteins are either mutated, overexpressed, or deleted. Examples are given in the sections on substrate metabolism below (63 66).
MAJOR ENERGY-PROVIDING SUBSTRATES A legitimate question is: Why is it important to know metabolic pathways in detail? Metabolic pathways can be likened to a power grid, and the heart can be likened to a light bulb. Fuels are the various sources of energy that are all converted to electricity. A simple explanation for this redundancy is that the grid uses those fuels that are most readily available, but all fuels are converted to the same form of energy. The same principle applies to the heart: For a given environment the heart uses the most efficient source of fuel. However, the fuel distribution system is far more complex than an electrical power grid.
Glucose, Lactate, and Glycogen Although long chain fatty acids are the predominant fuel for energy provision to the post-natal heart, carbohydrates are the fuel for the fetal heart (67) and also for the stressed adult heart in the state of exercise or pressure overload (68,69). The reasons for such a “hierarchy” of fuels can be deduced from a number of observations:
Chapter | 15
Cardiomyocyte Metabolism: All Is in Flux
1. In the normal, i.e. the non-diabetic, mammalian organism glucose levels in the blood are tightly regulated at around 5 mM or (90 mg%). 2. When we exercise and blood lactate levels rise, lactate replaces all other substrates as fuel for respiration of the heart (68). 3. When the normal heart is stressed, it oxidizes first glycogen, then glucose and lactate (35,70). 4. Glucose is an anaplerotic substrate for the citrate cycle (37,71) and glucose is essential for the initiation of fatty acid oxidation in heart (72). Isotopic tracer studies have demonstrated that the normal human heart produces lactate at the same time it oxidizes lactate (35,73). It appears that a large portion of exogenous glucose, when taken up by the cardiomyocyte, is shunted into first glycogen before it is oxidized (74,75).
Regulation of Glucose Metabolism There are three energy-yielding stages of glucose metabolism: the glycolytic pathway leading to pyruvate, oxaloacetate, and lactate; the Krebs cycle; and the respiratory chain. Each state is regulated by its own set of checkpoints, so that overall flux through the pathways (which may be assessed externally on a second-by-second timescale with the glucose tracer analogue FDG) (57) proceeds at a rate just sufficient to satisfy the heart’s beat-to-beat needs for ATP. We discuss the regulatory sites of metabolism in more detail because dysregulation of one of these steps can affect the rate and efficacy of energy transfer and, hence, contractile function of the heart.
Glucose Entry into the Myocyte Glucose uptake by the cardiomyocyte follows classical Michaelis Menten kinetics, and the transport of glucose occurs along a steep concentration gradient and is regulated by specific transporters (76). The stereospecificity of the transporter for sugars of the carbon configuration is not matched by the same degree of selectivity, and various tracer analogues, including 2-deoxyglucose and FDG, are transported in the same way as glucose. Glucose transport, the rate-limiting step in myocardial glucose utilization (77), requires facilitative glucose transporters. The family of glucose transporters (GLUT) is conserved over a wide range of organisms, suggesting a common evolutionary origin (78 81). GLUT-1 and GLUT-4 are the major glucose transporter isoforms expressed in the heart (62,82). GLUT-4, the insulinsensitive transporter, is also expressed by skeletal muscle and in adipose tissue (78). Recruitment of GLUT-4 from a microsomal cytosolic pool to the sarcolemma by insulin (or ischemia or adrenergic stimulation) (83 86) increases
191
the maximal velocity of glucose transport. Alphaadrenergic stimulation uses the same signaling pathway as insulin to promote glucose uptake (87), whereas the effects of ischemia, beta-receptor stimulation, and insulin on glucose uptake are additive (88). Cardiomyocytes also express the GLUT-1 transporter isoform, which is presumably independent of insulin regulation and predominates in fetal, hypertrophied, atrophied and failing myocardium (62,79,89). GLUT-1 is the first gene whose transcription is dually stimulated in response to hypoxia and inhibition of oxidative phosphorylation (90), and overexpression of GLUT-1 prevents the functional decline of hypertrophied heart (91). Both transporters have a Km for glucose (i.e. the concentration at which the rate of glucose transport is half maximal) that is in the range of plasma glucose concentrations under fasting conditions (92). The normal heart also expresses a low amount of GLUT-3, which has a Km below the normal plasma glucose concentration (93). Over the years a number of novel GLUTs have been identified in heart muscle including GLUT-8, GLUT-11, and GLUT-12 (94). A role of these proteins in myocardial glucose metabolism has not yet been defined. Phosphorylation of glucose by hexokinase becomes rate-limiting for glycolysis at high rates of glucose transport. Rates of glucose phosphorylation measured in vitro are more than twice as high as the maximal measured rates of glucose utilization by the heart at a physiologic workload and with glucose as the only substrate (95). However, intracellular glucose concentrations rise during starvation, in diabetes, and with the concomitant oxidation of fatty acids, ketone bodies, or lactate, which indicates inhibition of the phosphorylation step. This is most likely to be due to accumulation of glucose-6-phosphate, which is an allosteric inhibitor of hexokinase II, the cardiac isoform of hexokinase. Reduction of hexokinase II levels results in decreased cardiac function and altered remodeling after ischemia and reperfusion (96). Glucose-6-phosphate is at the branch point of four distinct pathways: 1. degradation via the Embden Meyerhof pathway (also termed the glycolytic pathway when it entails metabolism of glucose to lactate only); 2. conversion to glycogen via the glycogen synthase reaction; 3. metabolism (oxidation) via the pentose-phosphate pathway, which yields ribose and the reduced form of nicotinamide adenine dinucleotide phosphate (NADPH) (97); 4. entry into the hexosamine biosynthetic pathway (98). While the latter two pathways are of quantitatively lesser importance in heart muscle than are the former two, they provide glycogen.
192
Glycogen Glycogen metabolism was the first paradigm for the molecular basis of hormone action (99). In this pathway the control of enzyme activity by an allosteric regulator (activation of phosphorylase by adenosine monophosphate [AMP]) was first described (100), enzyme regulation by covalent modification was discovered (101), and the molecular basis of hormone action by signal transduction was elucidated by the discovery of cyclic adenosine monophosphate (cAMP) (102). Although the enzyme glycogen synthase kinase 3β (GSK3β), which phosphorylates and inhibits glycogen synthesis, is a well-known regulator of cardiac growth, its role in intermediary metabolism must not be overlooked. Hypertrophic stimuli inhibit GSK3β to regulate changes in metabolism, gene expression, and cytoskeletal integrity needed to promote cell growth (103). The role of glycogen metabolism in the heart is still not completely understood (104). Here two points should be made. Firstly, the vast amount of glycogen in fetal cardiac muscle probably allows the heart to maintain its contractile activity in the face of severe hypoxia (105) during birth. Secondly, glycogen and glycogen phosphorylase are closely associated with the sarcoplasmic reticulum (106) and in skeletal muscle a decreased glycogen content results in a reversible reduction in force, Ca21 release from the sarcoplasmic reticulum, and contractile protein function (107).
Glycolysis Glucose is special among the energy-providing substrates for the heart in its ability to provide a small, but significant, amount of ATP through substrate-level phosphorylation in the glycolytic pathway. This occurs especially in the setting of hypoxia and ischemia when flux through the glycolytic pathway is enhanced resulting in the formation of lactate as well as alanine (108). Other stimulants of flux through the glycolytic pathway are increases in cardiac work, either acutely with exercise (35,109) or chronically with pressure overload without (110) or with hypertrophy (69,111). Both hypertrophy and atrophy are associated with increased glucose oxidation rates in the face of decreased insulin responsiveness of the heart (112). In ischemia the accumulation of glycolytic intermediates may worsen contractile function (113) and acute hyperglycemia may abolish ischemic preconditioning in vivo (114) while provision of glucose together with insulin and potassium improves contractile function in the acutely ischemic, reperfused myocardium (115 118). The first step committing glucose to the glycolytic pathway is 6-phospho-fructo-1-kinase (PFK-1), which catalyzes the phosphorylation of fructose-6-phosphate to fructose 1,6-bisphosphate. Because of the complex
PART | II
Cardiac Muscle
allosteric regulation of PFK, this is a rate-limiting step (pacemaker enzyme) for glycolysis (119). ATP, citrate, and protons are negative allosteric effectors, whereas AMP and fructose 1,6-bisphosphate are positive effectors (120 122). Fructose 2,6-bisphosphate is the main activator of PFK-1 in normoxic heart (123). Further down in the glycolytic pathway, the oxidation of the triose-phosphate glyceraldehydes 3-phosphate to 1,3-diphsophoglycerate couples in the energy-conserving step in the glycolytic pathway that leads to the nonoxidative formation of ATP with high cardiac work (124) or ischemia (125) when PFK becomes strongly activated, glycolysis is controlled further downstream, at the triosephosphate dehydrogenase step.
Pyruvate Metabolism: An Branch Point The last glycolytic intermediate, pyruvate, is substrate for another branchpoint in metabolism. Pyruvate can be reduced to lactate (which completes the glycolytic pathway), transaminated to alanine (108), carboxylated to oxaloacetate or malate (126,127), or, most importantly, oxidized to acetyl-CoA. In well-oxygenated, working heart muscle, however, the bulk of pyruvate enters the mitochondrion through a transporter which can be inhibited by 4-hydroxy-alpha-cyanocinnamate (128). Once inside the mitochondrial matrix, pyruvate is either decarboxylated to acetyl-CoA or carboxylated to oxaloacetate (127). The capture of metabolically produced carbon dioxide from the pyruvate dehydrogenase reaction to form oxaloacetate is an example for the efficient use of one substrate supplying two precursors for citrate synthesis and for the efficient recycling of carbon dioxide. Oxidative decarboxylation of pyruvate is highly regulated by activation and inactivation of the pyruvate dehydrogenase complex (PDC) (129). The conversion of pyruvate to acetyl-CoA requires the sequential action of three different enzymes: pyruvate dehydrogenase, dihydrolipoyl transacetylase, and dihydrolipoyl dehydrogenase. The reaction also requires five different coenzymes or prosthetic groups: thiamine pyrophosphate, lipoic acid, uncombined coenzyme A (CoA SH), FAD1, and NAD1. These enzymes and coenzymes are organized into a multienzyme cluster. Much work was done in the 1970s on the regulation of PDC by covalent modification through a phosphorylation and dephosphorylation cycle (8). Multi-site phosphorylation of the pyruvate dehydrogenase component of the complex provides an indirect means by which the entire complex is regulated by the relative activities of the PDC kinase and phosphatase reactions. Of the metabolite pairs ATP ADP, acetyl-CoA CoA SH, NADH NAD1, and lactate pyruvate, the first member either activates the kinase or serves as substrate, and the
Chapter | 15
Cardiomyocyte Metabolism: All Is in Flux
second member inhibits the enzyme. Ca21 and Mg21 both inhibit the kinase and activate the phosphatase reaction (i.e. lead to PDC-activation). The effects of fatty acids or ketone body oxidation are likely to be mediated by the increase in acetyl-CoA because the primary effect or the inhibition-inactivation of PDC by fatty acids or ketone bodies in the acetyl-CoA COA SH ratio (130). Conversely, an increase in cardiac work may inhibit the PDC kinase owing to a decrease in NADH, acetyl-CoA, and ATP, leading to activation of PDC (131).
Fatty Acids Fatty acids, esterified as triglycerides, are the heart’s predominant fuel for respiration (7,132,133). In spite of their preeminence as a source for energy, fatty acids are also the only fuel capable of uncoupling oxidative phosphorylation (134,135), which lowers the efficiency of fatty acids as energy substrates. The predominant forms of fatty acids in the blood stream are the monounsaturated long chain fatty acid oleate (C18:1) and the saturated fatty acid palmitate (C16:0). The pathway of long chain fatty acid oxidation starts with the liberation of fatty acids from triglycerides and ends with the entry of acetyl-CoA into the citric acid cycle (Figure 15.3). Fatty acids cross the plasma membrane with the help of carrier proteins (136), are bound by a heart-specific fatty acid-binding protein (h-FABP) (137 139) and ushered to mitochondria (140) via activation to fatty-acyl coenzyme A (acyl-CoA). Deletion of fatty acyl-CoA synthetase in the heart causes hypertrophy, most likely by rerouting glucose metabolism (141). AcylCoA is transesterified with carnitine for transport across the inner mitochondrial membrane in exchange for carnitine, re-esterification with CoA SH, beta-oxidation, and finally, oxidation of acetyl-CoA. A variable amount of FFA taken up by heart muscle is also esterified with glycerol in the cytosol to form triglycerides. Increased net triglyceride synthesis by the heart has been observed with starvation, diabetes, ischemia, and in heart failure (142). At present it is not known whether increased triglyceride levels are the result of increased rates of esterification or a decreased rate of lipolysis (143).
Carnitine Palmitoyl Transferase and a Critical Role for Malonyl-CoA Transfer of long chain fatty acyl-CoA into mitochondria is rate-limiting for β-oxidation of long chain fatty acids and requires a three-step membrane transport process. The first step in this sequence is the transfer of the acyl group from CoA to carnitine, catalyzed by the enzyme carnitine palmitoyl transferase I (CPT I).
193
The CPT I system is inhibited by physiologic concentrations of malonyl-CoA, the product of acetyl-CoA carboxylation, and was first characterized in the liver, where malonyl-CoA serves as a signal regulating the relative rates of fatty acid oxidation, ketogenesis, and triglyceride synthesis (144) and subsequently found in the heart (145), where the inhibition of cardiac CPT I by malonyl-CoA was also demonstrated (146). A tissue-specific acetylCoA carboxylase has been identified (147) and so has malonyl-CoA decarboxylase (148 150). Because of the importance of malonyl-CoA for the regulation of fatty acid oxidation, the regulation of ACC and MCD is intensely investigated. Specifically, ACC is inhibited by phosphorylation through AMP kinase. This inhibition results in lower malonyl-CoA levels and increased rates of fatty acid oxidations. Not surprisingly, AMP kinase has earned its name as “fuel gauge” of the cell (151). The carnitine-acyl unit traverses the inner mitochondrial membrane and is transferred to CoA SH inside the mitochondrial matrix by a second carnitine acyl-CoA transferase (CPT II) located in the inner surface of the inner mitochondrial membrane. Once inside the mitochondria, acyl-CoA is committed to oxidation by the system of β-oxidation (Figure 15.3). The β-oxidation pathway is controlled by fluctuations in the concentrations of substrates (acyl-CoA, NAD1, and FAD1) (152) or, expressed in more physiologic terms, by workload and oxygen supply of the heart. It is not clear how the rate of oxidation of FFA in the intact heart is related to citric acid cycle activity and the rate of oxidative phosphorylation, but it is clear that the accumulation of long chain acyl-CoA esters in ischemia inhibits the adenine nucleotide translocase in mitochondria (153). This inhibition provides a possible explanation for contractile dysfunction. To explore the gene regulatory mechanisms involved in the metabolic control of long chain fatty acid oxidation in the heart, the expression of muscle-type CPT-1 (mCPT-1) has been characterized in primary cardiac myocytes after incubation with oleate, which regulates mCPT-1 expression via the peroxisome proliferators-activated receptor alpha (PPAR-α) (154). PPAR-α belongs to a nuclear receptor family with multiple functions (155). Malonyl-CoA decarboxylase (MCD) is transcriptionally regulated by PPAR-α (150) and MCD inhibition protects the ischemic heart most likely by inhibiting fatty acid oxidation (156) and promoting glucose oxidation (157).
Transcriptional Regulation of Myocardial Fatty Acid Metabolism One way by which the heart responds to increased plasma fatty acid levels is by increasing the expression of various proteins involved in fatty acid utilization (154,158 160)
194
including uncoupling protein 3 (135). The mechanism by which fatty acids activate the transcription of these genes is through activation of the nuclear receptor PPAR-α which heterodimerizes with 9-cis retinoic acid receptor (RXR) and activates the transcription of genes whose promoter contains the PPAR response element (PPRE). Of the co-factors that bind to the PPAR-α/RXR heterodimer, PCG-1α is highly expressed in the heart (161,162) and plays a key role in mitochondrial biogenesis (163), as discussed in another chapter.
Defective Fatty Acid Metabolism as Cause for Heart Failure Inborn errors in myocardial fatty acid metabolism are causes of cardiomyopathy and sudden death in children (164 166). The advent of tandem mass spectrometry has greatly advanced the screening for such metabolic cardiomyopathies (166). Perhaps best identified are the defects of the acyl-CoA dehydrogenase enzyme family, which catalyzes the initial reactions in mitochondrial β-oxidation of fatty acids (167 169). Enzymatic defects involving long and very long chain fatty acids (C14 or greater) or defects in cellular carnitine import of enzymes involved in the carnitine shuttle cause more severe heart failure than defects involving the metabolism of shorter-chain fatty acids (170). Impaired β-oxidation and accumulation of triglycerides in cardiomyocytes are hallmarks of contractile dysfunction of the heart from obese Zucker rats (171). In Zucker diabetic fatty rats lipotoxicity and contractile dysfunction are reversed by the administration of agents that lower plasma triglyceride and fatty acid levels (172,173). Reactivation of PPAR-α in adapted hypertrophied or hibernating myocardium results in severe contractile dysfunction (174,175).
Ketone Bodies Even in physiologic situations, such as short-term starvation or exercise, ketone body concentrations in the plasma can rise by up to 50-fold (176) and a key feature of myocardial ketone body metabolism is their concentrationdependent uptake in vivo (177). The ketone bodies, acetoacetate and beta-hydoxybutyrate, have ready access to the enzymes of the citric acid cycle in the heart (178). Ketone bodies do not sustain the full work output of the heart in vivo (179) and their rate of oxidation in the citric acid cycle is insufficient to meet the energy needs of the heart ex vivo (37). The relative inhibition of the citric acid cycle at the level of 2-oxoglutarate dehydrogenase provides a rare example of a defect in metabolism that causes reversible contractile dysfunction in the heart (37,180), which is relieved by anaplerosis (i.e. the provision of oxaloacetate or other citric acid cycle
PART | II
Cardiac Muscle
intermediates in addition to acetyl-CoA (127)). Pyruvate carboxylation in the in vivo heart accounts for at least 3 6% of citric acid cycle flux despite considerable variations in flux through pyruvate decarboxylase (181), supporting the principle that the heart functions best when it oxidizes several substrates simultaneously.
Amino Acids Although they are the building blocks of all proteins, amino acids are also an integral part of myocardial energy metabolism. Under physiologic circumstances one of the main functions of transaminases in heart muscle is the provision of carbon skeletons for the citric acid cycle (182). Furthermore, the amino acids aspartate and glutamate play an important role in the transfer of reducing equivalents across the mitochondrial membrane for oxidation of cytosolic NADH by the mitochondrial electron transport chain (183). In the malate aspartate cycle, cytosolic NADH, which cannot cross the inner mitochondrial membrane, is oxidized by the reduction of oxaloacetate to malate. Malate enters the mitochondrion and is oxidized to oxaloacetate, which is transaminated with glutamate to form 2-oxoglutarate and aspartate. The aspartate and 2-oxoglutarate leave the mitochondrion. Transmission of these metabolites regenerates oxaloacetate and glutamate in the cytosol, the net effect being a transfer of hydrogen ions across the mitochondrial membrane (184 186). A second postulated function for the malate aspartate shuttle is the delivery of intermediates (malate) to the citric acid cycle, serving as a modulator of cycle activity (187). Myocardial amino acid metabolism during hypoxia, ischemia, and reperfusion has been studied in the intact and isolated heart muscle (188 190). The amino acid alanine is like lactate, an end product of anaerobic glucose breakdown, arising through transamination of glutamate in the reaction: glutamate 1 pyruvate-2-oxoglutarate 1 alanine Where alanine leaves the cell, 2-oxoglutarate is further metabolized and decarboxylated to succinate via substrate level phosphorylation in the citric acid cycle (191): 2-oxoglutarate 1 NAD1 1 GDP 1 Pi -succinate 1 CO2 1 NADH2 1 GTP: GTP is readily transphosphorylated to ATP; this reaction is a source of anaerobic energy independent of lactate formation. Enhanced glutamate uptake in vivo has been shown by 13N accumulation from labeled glutamate in ischemic heart muscle (192) and glutamate enrichment has become a successful strategy to lessen the effects of ischemia with blood cardioplegia (193).
Chapter | 15
Cardiomyocyte Metabolism: All Is in Flux
195
METABOLIC REMODELING
intracellular long chain fatty acids are also highly regulated. Consequently, there is a balance between transport of fatty acids from the interstitial space into the cell, transport from the cytosol into mitochondria and subsequent oxidation. This balance is regulated mainly by malonyl-CoA, which inhibits transport of fatty acids into mitochondria. Prolonged inhibition of CPT I promotes intramyocellular lipid accumulation and insulin resistance in rats (196). Conversely, leptin inhibits malonyl-CoA synthesis in skeletal muscle leading to greater rates of fatty acid oxidation (197). Metabolic dysregulation, especially in the situation where the rate of fatty acid uptake exceeds the rate of fatty acid oxidation (198,199) leads to a wide range of disturbances that have been attributed to mitochondrial dysfunction (200,201) and which include triglyceride accumulation (143,202). Like a sentinel, lipid accumulation also is a feature of the failing human heart (Figure 15.5) (142) and is associated with a broad range of cellular and metabolic derangements collectively called lipotoxicity (143,203). Possible mechanisms causing a decline in contractile function of the “lipotoxic heart” include the inhibition of the adenine nucleotide translocator by fatty-acylCoA, the upregulation of TNF-α, increased ceramide levels, ROS accumulation, iNOS induction, chronic activation
Changes in metabolic fluxes are the first responders to changes in the physiologic environment of the heart. It follows that metabolic remodeling triggers and sustains functional and structural remodeling (194). Many of the acute changes in metabolic flux are brought about by the same signal transduction cascades believed to be involved in the adaptation to the heart’s environment, e.g. phosphotidylinositol 3-kinase (87), diacylglycerol, Ca21, and protein kinase C (195). Thus, metabolic signals are an integral part of cardiac adaptation to its environment (194).
Dysregulated Fatty Acid Metabolism A case in point are long chain fatty acids. In addition to their role as energy-providing substrates, fatty acids serve as mediators of signal transduction and as ligands for the nuclear receptor PPAR-α. Changes in substrate supply to the heart are reflected in changes of flux through fatty acid metabolizing pathways, which are both transcriptionally and post-transcriptionally regulated. In other words, redirecting energy metabolism is largely orchestrated by proteins that are, in turn, transcriptionally or post-transcriptionally regulated by metabolites. Levels of
Oil Red O Staining in Failing Human Heart
Low
Moderate
High
30%
44% 26%
Low
N = 27
Arbitrary units
High
Moderate
Oil Red O
2 1.8 1.6 1.4 1.2 1 0.8 0.6 0.4 0.2 0
*#
NF
HF
HF+O
HF+DM
FIGURE 15.5 Triglyceride accumulation in the failing heart. Note the high prevalence of intracellular fat droplets (stained with oil red O), especially in patients showing signs of metabolic dysregulation (obesity and type 2 diabetes). See text for further detail. (From Sharma et al., 2004 (142).)
196
of PKCs, myosin isoform switches (204), downregulation of MEF2C regulated genes (205), cytochrome C release (206), and apoptosis (172). Interestingly, the heart’s ability to secrete lipoproteins is inversely related to pathologic lipid accumulation (207). In addition, the ability to secrete lipoproteins protects the heart against the development of cardiomyopathy in a diabetes model (208). Lastly, in rodent models of lipotoxicity the administration of leptin decreases non-adipose tissue lipid accumulation and ameliorates lipotoxicity (209). Thus, a wide array of potentially reversible disturbances of fatty acid and lipid metabolism are involved in the phenomenon of lipotoxicity.
Dysregulated Glucose Metabolism Glucose and its metabolites also have multiple functions in the cardiac myocyte. Failure to adequately control levels of intracellular glucose metabolites has been implicated in the development of insulin resistance and in the generation of reactive oxygen species (ROS). Compared to the liver (210), relatively little is known about the effects of glucose metabolites on gene expression in the heart. Through investigations of the glucose/ carbohydrate response elements (GIRE/ChoRE) in the promoter regions of various glucose-regulated genes, a number of candidate transcription factors have been identified that are believed to be involved in glucosemediated gene expression. In the liver and in fat cells upstream stimulatory factor (USF), stimulatory protein 1 (Sp1), and sterol regulatory element binding protein 1 (SREBP1) play a role in glucose sensing (211), and it is reasonable to assume that the same transcription factors play similar roles in the heart (194). Excessive glucose metabolic accumulation is also associated with various cardiac pathologies. Although studied mainly in the vasculature, the four main mechanisms proposed for hyperglycemia induced diabetic complications (209) are likely to be operative in the cardiomyocyte as well. The four hypotheses are increased polyol pathway flux (via aldose reductase), increased intracellular formation of advanced glycation end products (AGE) and AGEinduced ROS generation activation of protein kinase C mostly through activation by the lipid second messenger diacylglycerol (DAG), and increased flux through the hexosamine biosynthetic pathway (209). As mentioned, many of these hypotheses have thus far been tested only in vascular tissue and await further confirmation in cardiomyocytes. However, it is already known that O-linked β-N-acetylglucosamine (O-GlcNAc) modifies many different nuclear and cytoplasmic proteins and that O-GlcNAc plays an important role in signal transduction (212). The basis is a “spillover” of intermediates of glucose metabolism into the hexosamine
PART | II
Cardiac Muscle
biosynthetic pathway and into the pentose phosphate pathway (194).
CONCLUSIONS Myocardial metabolism is an integral part of the function of the heart as consumer and provider of energy and has pleiotropic roles. The bulk of the energy for contraction of the heart comes from oxidative phosphorylation of ADP. A vast network of highly regulated metabolic pathways matches demand and supply with precision. The complexity of intermediary metabolism is also a rich source of speculation on impaired energy transfer as either cause or consequence of impaired cardiac function. Not surprisingly, restoration of efficient energy substrate metabolism as target for the treatment of the failing heart has thus far been an elusive goal for most investigators. All is in flux.
ACKNOWLEDGMENTS I thank Roxy Ann Tate for help with the preparation of this manuscript as well as past and present members of my laboratory for many discussions. Work in my lab is supported by the National Heart, Lung and Blood Institute of the US Public Health Service.
REFERENCES 1. Krebs HA, Johnson WA. The role of citric acid in intermediate metabolism in animal tissues. Enzymologia 1937;4:148 56. 2. Schoenheimer R. The dynamic state of body constituents. Cambridge, MA: Harvard University Press;1942. 3. Langendorff O. Untersuchungen am uberlebenden Saugethierherzen. Arch Ges Physiol Menschen Tiere 1895; 61:291 332. 4. Winterstein H. Ueber die Sauerstoffatmung des isolierten Saeugetierherzens. Z Allg Physiol 1904;4:339 59. 5. Taegtmeyer H. Energy metabolism of the heart: from basic concepts to clinical applications. Curr Prob Cardiol 1994;19:57 116. 6. Evans CL. The metabolism of cardiac muscle. In: Newton W, editor. Evans’ recent advances in physiology. Philadelphia, PA: Blakinston’s Sons and Co;1939. pp. 157 82. 7. Bing RJ. The metabolism of the heart. Harvey Lect 1955;50:27 70. 8. Randle PJ, Tubbs PK. Carbohydrate and fatty acid metabolism. In: Berne R, Sperelakis V, Geiger S, editors. Handbook of physiology: the cardiovascular system. Bethesda, MD: American Physiological Society;1979. pp. 805 44. 9. Barger PM, Kelly DP. PPAR signaling in the control of cardiac energy metabolism. Trends Cardiovasc Med 2000;10:238 45. 10. Balaban RS. Cardiac energy metabolism homeostasis: role of cytosolic calcium. J Mol Cell Cardiol 2002;34:1259 71. 11. Stanley WC, Recchia FA, Lopaschuk GD. Myocardial substrate metabolism in the normal and failing heart. Physiol Rev 2005;85:1093 129.
Chapter | 15
Cardiomyocyte Metabolism: All Is in Flux
12. Bugger H, Abel ED. Molecular mechanisms for myocardial mitochondrial dysfunction in the metabolic syndrome. Clin Sci (Lond) 2008;114:195 210. 13. Ingwall JS. Energy metabolism in heart failure and remodelling. Cardiovasc Res 2009;81:412 9. 14. O’Rourke B. Be still my beating heart. Never! Circ Res 2010;106:238 9. 15. Rosca MG, Hoppel CL. Mitochondria in heart failure. Cardiovasc Res 2010;88:40 50. 16. Abel ED, Doenst T. Mitochondrial adaptations to physiological vs. pathological cardiac hypertrophy. Cardiovasc Res 2011;90:234 42. 17. Baskin K, Taegtmeyer H. Taking pressure off the heart: the ins and outs of atrophic remodeling. Cardiovasc Res 2011;90:243 50. 18. Taegtmeyer H, Roberts AF, Raine AE. Energy metabolism in reperfused heart muscle: metabolic correlates to return of function. J Am Coll Cardiol 1985;6:864 70. 19. Balaban RS, Kontor HL, Katz LA, Briggs RW. Relation between work and phosphate metabolite in the in vivo paced mammalian heart. Science 1986;232:1121 3. 20. Kupriyanov VV, Lakomkin VL, Kapelko VI, Steinschneider AY, Ruuge EK, Saks VA. Dissociation of adenosine triphosphate levels and contractile function in isovolumic hearts perfused with 2-deoxyglucose. J Mol Cell Cardiol 1987;19:729 40. 21. Katz A. Is the failing heart energy depleted? Cardiol Clin 1998;16:633 44. 22. Barth E, Sta¨mmler G, Speiser B, Schaper J. Ultrastructural quantitation of mitochondria and myofilaments in cardiac muscle from 10 different animal species including man. J Mol Cell Cardiol 1992;24:669 81. 23. McNutt N, Fawcett D. Myocardial ultrastructure. In: Lauger G, Brady A, editors. The mammalian myocardium. New York: John Wiley & Sons;1973. pp. 1 50. 24. Lane N. Power, sex, suicide: Mitochondria and the meaning of life. New York: Oxford University Press;2005. 25. Baldwin JE, Krebs HA. The evolution of metabolic cycles. Nature 1981;291:381 2. 26. Russell RR, Taegtmeyer H. Anaplerosis. Encyclopedia of biological chemistry. New York: Elsevier;2010. 27. Russell RR, Taegtmeyer H. Changes in citric acid cycle flux and anaplerosis antedate the functional decline in isolated rat hearts utilizing acetoacetate. J Clin Invest 1991;87:384 90. 28. Newsholme EA, Start C. Regulation in metabolism. Chichester: J. Wiley & Sons;1973. 29. Newsholme EA, Leech AR. Biochemistry for the medical sciences. Chichester: J. Wiley;1983. 30. Fell D. Understanding the control of metabolism. Miami, FL: Portland Press;1997. 31. Brown GC. Control of respiration and ATP synthesis in mammalian mitochondria and cells. Biochem J 1992;284:1 13. 32. Krebs HA. Control of metabolic processes. Endeavour 1957;16:125 32. 33. Taegtmeyer H, King LM, Jones BE. Energy substrate metabolism, myocardial ischemia, and targets for pharmacotherapy. Am J Cardiol 1998;82:54K 60K. 34. Taegtmeyer H. Cardiac metabolism as a target for the treatment of heart failure. Circulation 2004;110:894 6.
197
35. Goodwin GW, Taylor CS, Taegtmeyer H. Regulation of energy metabolism of the heart during acute increase in heart work. J Biol Chem 1998;273:29530 9. 36. Neely JR, Liebermeister H, Battersby EJ, Morgan HE. Effect of pressure development on oxygen consumption by isolated rat heart. Am J Physiol 1967;212:804 14. 37. Taegtmeyer H, Hems R, Krebs HA. Utilization of energyproviding substrates in the isolated working rat heart. Biochem J 1980;186:701 11. 38. Hochachka P, Somero G. Strategies of biochemical adaptations. Philadelphia, PA: W.B. Saunders;1973. 39. Balaban RS. Regulation of oxidative phosphorylation in the mammalian cell. Am J Physiol 1990;258:C377 89. 40. Brown G, Cooper C. Control analysis applied to single enzymes: can an isolated enzyme have a unique rate-limiting step? Biochem J 1993;294:87 94. 41. Gadian D, Hoult D, Radda G, Seeley P, Chance B, Barlow C. Phosphorous nuclear magnetic resonance studies in normoxic and ischemic cardiac tissue. Proc Natl Acad Sci USA 1976;73:291 332. 42. Jacobus W, Taylor G, Hollis D, Nunnally R. Phosphorous nuclear magnetic resonance of perfused working rat hearts. Nature 1977;265:756 8. 43. Bottomley P. Noninvasive study of high energy phosphate metabolism in human heart by depth-resolved 31P NMR spectroscopy. Science 1985;229:769 72. 44. McMillin-Wood J. Biochemical approaches in metabolism: application to positron emission tomography. Circulation 1985;72: IV145 50. 45. Schelbert HR. Assessment of myocardial metabolism by PET: a sophisticated dream or clinical reality? Eur J Nucl Med 1986;12:570 5. 46. Taegtmeyer H, Mossberg KA, Nguyen VT. Positron labeled tracers. A window for the assessment of energy metabolism in heart and skeletal muscle. Acta Radiol Suppl 1991;376:40 4. 47. Schwaiger M, Hicks R. The clinical role of metabolic imaging of the heart by positron emission tomography. J Nucl Med 1991;32:565 78. 48. Lewandowski ED. Nuclear magnetic resonance evaluation of metabolic and respiratory support of work load in intact rabbit hearts. Circ Res 1992;70:576 82. 49. Russell III RR, Cline GW, Guthrie PH, Goodwin GW, Shulman GI, Taegtmeyer H. Regulation of exogenous and endogenous glucose metabolism by insulin and acetoacetate in the isolated working rat heart. A three tracer study of glycolysis, glycogen metabolism, and glucose oxidation. J Clin Invest 1997;100:2892 9. 50. Malloy C, Sherry A, Jeffrey F. Carbon flux through citric acid cycle pathways in perfused heart by 13C NMR spectroscopy. FEBS Lett 1987;212:58 62. 51. Weiss RG, Gloth ST, Kalil-Filho R, Chacko VP, Stern MD, Gerstenblith G. Indexing tricarboxylic acid cycle flux in intact hearts by carbon-13 nuclear magnetic resonance. Circ Res 1992;70:392 408. 52. Laughlin MR, Fleming Taylor J, Chesnik AS, Balaban RS. Regulation of glycogen metabolism in canine myocardium: Effects of insulin and epinephrine in vivo. Am J Physiol 1992;262:E875 83.
198
53. Hovener JB, Chekmenev EY, Harris KC, Perman WH, Robertson LW, Ross BD, et al. PASADENA hyperpolarization of 13C biomolecules: equipment design and installation. Magma 2009;22:111 21. 54. Schroeder MA, Atherton HJ, Ball DR, Cole MA, Heather LC, Griffin JL, et al. Real-time assessment of Krebs cycle metabolism using hyperpolarized 13C magnetic resonance spectroscopy. Faseb J 2009;23:2529 38. 55. Holloway C, ten Hove M, Clarke K, Neubauer S. MR spectroscopy in heart failure. Front Biosci (Schol Ed) 2011;3:331 40. 56. Dence CS, Herrero P, Schwarz SW, Mach RH, Gropler RJ, Welch MJ. Imaging myocardium enzymatic pathways with carbon-11 radiotracers. Methods Enzymol 2004;385:286 315. 57. Nguyen VT, Mossberg KA, Tewson TJ, Wong WH, Rowe RW, Coleman GM, et al. Temporal analysis of myocardial glucose metabolism by 2-[18F]fluoro-2-deoxy-D-glucose. Am J Physiol 1990;259:H1022 31. 58. Schelbert H, Henze E, Sochor H. Effects of substrate availability on myocardial 11C palmitate kinetics by positron emission tomography in normal subjects and patients with ventricular dysfunction. Am Heart J 1986;111:1055 65. 59. Brown MA, Marshall DR, Sobel BE, Bergmann SR. Delineation of myocardial oxygen utilization with carbon-11 labelled acetate. Circulation 1987;76:687 96. 60. Tillisch J, Brunken R, Marshall R, Schwaiger M, Mandelkern M, Phelps M, et al. Reversibility of cardiac wall-motion abnormalities predicted by positron tomography. N Engl J Med 1986;314:884 8. 61. Davila-Roman VG, Vedala G, Herrero P, de las Fuentes L, Rogers JG, Kelly DP, et al. Altered myocardial fatty acid and glucose metabolism in idiopathic dilated cardiomyopathy. J Am Coll Cardiol 2002;40:271 7. 62. Razeghi P, Young ME, Alcorn JL, Moravec CS, Frazier OH, Taegtmeyer H. Metabolic gene expression in fetal and failing human heart. Circulation 2001;104:2923 31. 63. Ingwall JS. Transgenesis and cardiac energetics: new insights into cardiac metabolism. J Mol Cell Cardiol 2004;37:613 23. 64. Russell LK, Finck BN, Kelly DP. Mouse models of mitochondrial dysfunction and heart failure. J Mol Cell Cardiol 2005;38:81 91. 65. Hsueh W, Abel ED, Breslow JL, Maeda N, Davis RC, Fisher EA, et al. Recipes for creating animal models of diabetic cardiovascular disease. Circ Res 2007;100:1415 27. 66. Brookheart RT, Michel CI, Schaffer JE. As a matter of fat. Cell Metab 2009;10:9 12. 67. Fisher DJ, Heymann MA, Rudolph AM. Myocardial oxygen and carbohydrate consumption in fetal lambs in utero and in adult sheep. Am J Physiol Heart Circ Physiol 1980;238:H399 405. 68. Goodwin GW, Taegtmeyer H. Improved energy homeostasis of the heart in the metabolic state of exercise. Am J Physiol Heart Circ Physiol 2000;279:H1490 501. 69. Bishop S, Altschuld R. Increased glycolytic metabolism in cardiac hypertrophy and congestive heart failure. Am J Physiol 1970;218:153 9. 70. Goodwin GW, Ahmad F, Doenst T, Taegtmeyer H. Energy provision from glycogen, glucose, and fatty acids on adrenergic stimulation of isolated working rat hearts. Am J Physiol 1998;274: H1239 47. 71. Russell RR, Taegtmeyer H. Coenzyme A sequestration in rat hearts oxidizing ketone bodies. J Clin Invest 1992;89:968 73.
PART | II
Cardiac Muscle
72. Tirosh R, Mishor T, Pinson A. Glucose is essential for the initiation of fatty acid oxidation in ATP-depleted cultured myocytes. Mol Cell Biochem 1996;162:159 63. 73. Gertz EW, Wisneski JA, Neese RA, Bristow JD, Searle GL, Hanlon JT. Myocardial lactate metabolism: evidence of lactate release during net chemical extraction in man. Circulation 1981;63:1273 9. 74. 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;76:1819 27. 75. Wisneski JA, Stanley WC, Neese RA, Gertz EW. Effects of acute hyperglycemia on myocardial glycolytic activity in humans. J Clin Invest 1990;85:1648 56. 76. Gould GW. Facilitative glucose transporters. Georgetown, TX: Landes Comp., Chapman & Hall;1997. 77. Manchester J, Kong X, Nerbonne J, Lowry O, Lawrence Jr J. Glucose transport and phosphorylation in single cardiac myocytes: rate limiting steps in glucose metabolism. Am J Physiol 1994;266:E326 33. 78. Pessin JE, Bell GI. Mammalian facilitative glucose transporter family: structure and molecular regulation. Annu Rev Physiol 1992;54:911 30. 79. Gould GW, Holman GD. The glucose transporter family: structure, function and tissue-specific expression. Biochem J 1993;295: 329 41. 80. Mueckler M. Facilitative glucose transporters. Eur J Biochem 1994;219:713 25. 81. Shepherd PR, Kahn BB. Glucose transporters and insulin action implications for insulin resistance and diabetes mellitus. N Engl J Med 1999;341:248 57. 82. Olson AL, Pessin JE. Structure, function, and regulation of the mammalian facilitative glucose transporter gene family. Annu Rev Nutr 1996;16:235 56. 83. Cushman S, Wardzala L. Potential mechanism of insulin action on glucose transport in the isolated rat adipose cell. J Biol Chem 1980;255:4755 62. 84. Wheeler TJ. Translocation of glucose transporters in response to anoxia in heart. J Biol Chem 1988;263:19447 54. 85. Sun D, Nguyen N, Delgrado TR, Schwaiger M, Brosius FCI. Ischemia induces translocation of the insulin-responsive glucose transporter GLUT 4 to the plasma membrane of cardiac myocytes. Circulation 1994;89:793 8. 86. Russell RR, Yin R, Caplan MJ, Hu X, Ren J, Young LH. Ischemia on myocardial GLUT1 and GLUT4 translocation in vivo. Circulation 1998;98:2180 6. 87. Doenst T, Taegtmeyer H. Alpha-adrenergic stimulation mediates glucose uptake through phosphatidylinositol 3-kinase in rat heart. Circ Res 1999;84:467 74. 88. Doenst T, Taegtmeyer H. Ischemia-stimulated glucose uptake does not require catecholamines in rat heart. J Mol Cell Cardiol 1999;31:435 43. 89. Depre C, Shipley GL, Chen W, Han Q, Doenst T, Moore ML, et al. Unloaded heart in vivo replicates fetal gene expression of cardiac hypertrophy. Nat Med 1998;4:1269 75. 90. Behrooz A, Ismail-Beigi F. Stimulation of glucose transport by hypoxia: signals and mechanisms. News Physiol Sci 1999;14:105 10. 91. Liao R, Mohit J, Lei C, D’Agostino J, Aiello F, Luptak I, et al. Cardiac-specific overexpression of GLUT1 prevents the
Chapter | 15
92.
93.
94.
95.
96.
97.
98.
99. 100.
101.
102. 103.
104. 105.
106.
107.
108.
Cardiomyocyte Metabolism: All Is in Flux
development of heart failure due to pressure-overload in mice. Circulation 2002;106:2125 31. Nishimura H, Pallardo FV, Seidner GA, Vannucci S, Simpson IA, Birnbaum MF. Kinetics of GLUT1 and GLUT4 glucose transporters expressed in Xenopus oocytes. J Biol Chem 1993;268:8514 20. Shepherd PR, Gould GW, Colville CA, McCoid SC, Gibbs EM, Kahn BB. Distribution of GLUT3 glucose transporter protein in human tissues. Biochem Biophys Res Commun 1992;188: 149 54. Joost HG, Bell GI, Best JD, Birnbaum MJ, Charron MJ, Chen YT, et al. Nomenclature of the GLUT/SLC2A family of sugar/ polyol transport facilitators. Am J Physiol Endocrinol Metab 2002;282:E974 6. Newsholme EA, Crabtree B. Theoretical principles in the approaches to the control of metabolic pathways and their application to glycolysis in muscle. J Molec Cell Cardiol 1979;11:839 56. Wu R, Smeele KM, Wyatt E, Ichikawa Y, Eerbeek O, Sun L, et al. Reduction in hexokinase II levels results in decreased cardiac function and altered remodeling after ischemia/reperfusion injury. Circ Res 2011;108:60 9. Goodwin GW, Cohen DM, Taegtmeyer H. [5-3H]glucose overestimates glycolytic flux in isolated working rat heart: role of the pentose phosphate pathway. Am J Physiol Endocrinol Metab 2001;280:E502 8. Taegtmeyer H, McNulty P, Young ME. Adaptation and maladaptation of the heart in diabetes: Part I: general concepts. Circulation 2002;105:1727 33. Cohen P. Control of enzyme activity. London: Chapman & Hall;1976. Cori G, Colowick S, Cori C. The action of nucleotides on the disruptive phosphorylation of glycogen. J Biol Chem 1938;123:381 9. Krebs EG, Fischer EH. The phosphorylase b to a converting enzyme of rabbit skeletal muscle. Biochim Biophys Acta 1956;20:150 7. Robison G, Butcher R, Sutherland E. Cyclic AMP. London: Academic Press;1971. Antos CL, McKinsey TA, Frey N, Kutschke W, McAnally J, Shelton JM, et al. Activated glycogen synthase-3 beta suppresses cardiac hypertrophy in vivo. Proc Natl Acad Sci USA 2002;99:907 12. Taegtmeyer H. Glycogen in the heart an expanded view. J Mol Cell Cardiol 2004;37:7 10. Johnson MH, Everitt B. The high concentration of glycogen in fetal cardiac muscle probably explains why the heart can maintain its contractile activity in the face of severe hypoxia. In: Johnson MH, editor. Essential reproduction. Oxford: Blackwell Scientific;1988. p. 275. Entman ML, Kanike K, Goldstein MA, Nelson TE, Bornet EP, Futch TW, et al. Association of glycogenolysis with cardiac sarcoplasmic reticulum. J Biol Chem 1976;251:3140 6. Chin E, Allen D. Effects of reduced muscle glycogen concentration on force, Ca21 release and contractile protein function in intact mouse skeletal muscle. J Physiol (Lond) 1997;498:17 29. Taegtmeyer H, Peterson MB, Ragavan VV, Ferguson AG, Lesch M. De novo alanine synthesis in isolated oxygen-deprived rabbit myocardium. J Biol Chem 1977;252:5010 8.
199
109. Gertz EW, Wisneski JA, Stanley WC, Neese RA. Myocardial substrate utilization during exercise in humans. J Clin Invest 1988;82:2017 25. 110. Taegtmeyer H, Overturf ML. Effects of moderate hypertension on cardiac function and metabolism in the rabbit. Hypertension 1988;11:416 26. 111. Allard MF, Emanuel PG, Russell JA, Bishop SP, Digerness SB, Anderson PG. Preischemic glycogen reduction or glycolytic inhibition improves postischemic recovery of hypertrophied rat heart. Am J Physiol 1994;267:H66 74. 112. Doenst T, Goodwin GW, Cedars AM, Wang M, Stepkowski S, Taegtmeyer H. Load-induced changes in vivo alter substrate fluxes and insulin responsiveness of rat heart in vitro. Metabolism 2001;50:1083 90. 113. Neely JR, Grotyohann LW. Role of glycolytic products in damage to myocardium: Dissociation of adenosine triphosphate levels and recovery of function of reperfused canine myocardium. Circ Res 1984;55:816 24. 114. Kersten JR, Schmeling TJ, Orth KG, Pagel PS, Warltier DC. Acute hyperglycemia abolishes ischemic preconditioning in vivo. Am J Physiol Heart Circ Physiol 1998;275:H721 5. 115. Gradinac S, Coleman GM, Taegtmeyer H, Sweeney MS, Frazier OH. Improved cardiac function with glucose-insulin-potassium after aortocoronary bypass grafting. Ann Thorac Surg 1989;48:484 9. 116. Taegtmeyer H. The use of hypertonic glucose, insulin, and potassium (GIK) in myocardial preservation. J Appl Cardiol 1991;6:255 9. 117. Aasum E, Lathrop DA, Henden T, Sundset R, Larsen TS. The role of glycolysis in myocardial calcium control. J Mol Cell Cardiol 1998;30:1703 12. 118. Howell NJ, Ashrafian H, Drury NE, Ranasinghe AM, Contractor H, Isackson H, et al. Glucose-insulin-potassium reduces the incidence of low cardiac output episodes after aortic valve replacement for aortic stenosis in patients with left ventricular hypertrophy: results from the Hypertrophy, Insulin, Glucose, and Electrolytes (HINGE) trial. Circulation 123:170 7. 119. Passonneau JV, Lowry OH. Phosphofructokinase and the Pasteur effect. Biochem Biophys Res Comm 1962;7:10 5. 120. Uyeda K. Phosphofructokinase. Adv Enzymol 1979;48:193 244. 121. Kobayashi K, Neely J. Control of maximum rates of glycolysis in rat cardiac muscle. Circ Res 1979;44:166 75. 122. Hue L, Rider MH. Role of fructose 2,6-bisphosphate in the control of glycolysis in mammalian tissues. Biochem J 1987;245:313 24. 123. Rider MH, Bertrand L, Vertommen D, Michels PA, Rousseau GG, Hue L. 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatase: head-to-head with a bifunctional enzyme that controls glycolysis. Biochem J 2004;381:561 79. 124. Rovetto M, Lamberton W, Neely J. Mechanisms of glycolytic inhibition in ischemic rat heart. Circ Res 1975;37:742 51. 125. Nuutila P, Koivisto VA, Knuuti J, Ruotsalainen U, Teras M, Haaparanta M, et al. Glucose-free fatty acid cycle operates in human heart and skeletal muscle in vivo. J Clin Invest 1992;89:1767 74. 126. Peuhkurinen KJ, Hassinen IE. Pyruvate carboxylation as an anaplerotic mechanism in the isolated perfused rat heart. Biochem J 1982;202:67 76.
200
127. Russell RR, Taegtmeyer H. Pyruvate carboxylation prevents the decline in contractile function of rat hearts oxidizing acetoacetate. Am J Physiol 1991;261:H1756 62. 128. Denton RM, Halestrap AP. Regulations of pyruvate metabolism in mammalian tissues. Essays Biochem 1979;15:37 77. 129. Randle P. Regulation of glycolysis and pyruvate oxidation in cardiac muscle. Circ Res 1976;38:I8 15. 130. Olson MS, Dennis SC, DeBuysere MS, Padma A. The regulation of pyruvate dehydrogenase in the isolated perfused rat heart. J Biol Chem 1978;253:7369 75. 131. Kerbey AL, Randle PJ, Cooper RH, Whitehouse S, Pask HT, et al. Regulation of pyruvate dehydrogenase in rat heart. Biochem J 1976;154:327 48. 132. Bing RJ, Siegel A, Ungar I, Gilbert M. Metabolism of the human heart. II. Studies on fat, ketone and amino acid metabolism. Am J Med 1954;16:504 15. 133. Bing RJ. The metabolism of the heart. Trans Am Coll Cardiol 1955;5:8 14. 134. Borst P, Loos JA, Christ EJ, Slater EC. Uncoupling activity of long chain fatty acids. Biochim Biophys Acta 1962;62:509 17. 135. Young ME, Patil S, Ying J, Depre C, Ahuja HS, Shipley GL, et al. Uncoupling protein 3 transcription is regulated by peroxisome proliferator-activated receptor (alpha) in the adult rodent heart. Faseb J 2001;15:833 45. 136. Schaffer JE. Fatty acid transport: the roads taken. Am J Physiol Endocrinol Metab 2002;282:E239 46. 137. Bass NM. The cellular fatty acid binding proteins: aspects of structure, regulation, and function. Int Rev Cytol 1988;111:143 84. 138. van der Vusse G, Groot M. Interrelationship between lactate and cardiac fatty acid metabolism. Mol Cell Biochem 1992;116:11 7. 139. Binas B, Danneberg H, McWhir J, Mullins L, Clark AJ. Requirement for the heart-type fatty acid binding protein in cardiac fatty acid utilization. FASEB J 1999;13:805 12. 140. Erol E, Cline GW, Kim JK, Taegtmeyer H, Binas B. Nonacute effects of H-FABP deficiency on skeletal muscle glucose uptake in vitro. Am J Physiol Endocrinol Metab 2004;287:E977 82. 141. Ellis JM, Mentock SM, Depetrillo MA, Koves TR, Sen S, Watkins SM, et al. Mouse cardiac acyl coenzyme a synthetase 1 deficiency impairs fatty acid oxidation and induces cardiac hypertrophy. Mol Cell Biol 2011;31:1252 62. 142. Sharma S, Adrogue JV, Golfman L, Uray I, Lemm J, Youker K, et al. Intramyocardial lipid accumulation in the failing human heart resembles the lipotoxic rat heart. FASEB J 2004;18:1692 700. 143. Schaffer JE. Lipotoxicity: when tissues overeat. Curr Opin Lipidol 2003;14:281 7. 144. McGarry JD, Mannaerts GP, Foster DW. A possible role for malonyl-CoA in the regulation of hepatic fatty acid oxidation and ketogenesis. J Clin Invest 1977;60:265 70. 145. McGarry JD, Mills SE, Long CS, Foster DW. Observations on the affinity for carnitine and malonyl-CoA sensitivity of carnitine palmitoyl transferase I in animal and human tissues. Demonstration of the presence of malonyl-CoA in non-hepatic tissues of the rat. Biochem J 1983;214:21 8. 146. Paulson D, Ward K, Shug A. Malonyl-CoA inhibition of carnitine palmitoyl transferase. FEBS Lett 1984;176:381 4. 147. Bianchi A, Evans J, Iverson A, Nordlund A, Watts T, Wilters L. Identification of an isozymic form of acetyl-CoA carboxylase. J Biol Chem 1990;265:1502 9.
PART | II
Cardiac Muscle
148. Dyck JR, Barr AJ, Barr RL, Kolattukudy PE, Lopaschuk GD. Characterization of cardiac malonyl-CoA decarboxylase and its putative role in regulating fatty acid oxidation. Am J Physiol Heart Circ Physiol 1998;275:H2122 9. 149. Goodwin GW, Taegtmeyer H. Regulation of fatty acid oxidation of the heart by MCD and ACC during contractile stimulation. Am J Physiol Endocrinol Metab 1999;277:E772 7. 150. Young ME, Goodwin GW, Ying J, Guthrie P, Wilson CR, Laws FA, et al. Regulation of cardiac and skeletal muscle malonyl-CoA decarboxylase by fatty acids. Am J Physiol Endocrinol Metab 2001;280:E471 9. 151. Hardie DG, Carling D. The AMP-activated protein kinase fuel gauge of the mammalian cell? Eur J Biochem 1997;246:259 73. 152. Bremer J, Wojtzak A. Factors controlling the role of fatty acid beta oxidation in rat liver mitochondria. Biochem Biophys Acta 1972;280:515 30. 153. Shrago E, Shug AL, Sul H, Bittar N, Folts JD. Control of energy production in myocardial ischemia. Circ Res 1976;38:75 81. 154. Brandt JM, Djouadi F, Kelly DP. Fatty acids activate transcription of the muscle carnitine palmitoyltransferase I gene in cardiac myocytes via the peroxisome proliferator-activated receptor alpha. J Biol Chem 1998;273:23786 92. 155. Vamecq J, Latruffe N. Medical significance of peroxisome proliferator-activated receptors. Lancet 1999;354:141 8. 156. Dyck JR, Cheng JF, Stanley WC, Barr R, Chandler MP, Brown S, et al. Malonyl coenzyme a decarboxylase inhibition protects the ischemic heart by inhibiting fatty acid oxidation and stimulating glucose oxidation. Circ Res 2004;94:e78 84. 157. Ussher JR, Lopaschuk GD. Targeting malonyl CoA inhibition of mitochondrial fatty acid uptake as an approach to treat cardiac ischemia/reperfusion. Basic Res Cardiol 2009;104:203 10. 158. van Bilsen M, Van der Vusse GJ, Reneman RS. Transcriptional regulation of metabolic processes: implications for cardiac metabolism. Pflugers Arch 1998;437:2 14. 159. Gulick T, Cresci S, Caira T, Moore D, Kelly D. The peroxisome proliferator-activated receptor regulates mitochondrial fatty acid oxidative enzyme gene expression. Proc Natl Acad Sci USA 1994;91:11012 6. 160. van der Lee KA, Vork MM, De Vries JE, Willemsen PH, Glatz JF, Reneman RS, et al. Long-chain fatty acid-induced changes in gene expression in neonatal cardiac myocytes. J Lipid Res 2000;41:41 7. 161. Lehman JJ, Barger PM, Kovacs A, Saffitz JE, Medeiros DM, Kelly DP. Peroxisome proliferator-activated receptor gamma coactivator-1 promotes cardiac mitochondrial biogenesis. J Clin Invest 2000;106:847 56. 162. Kelly DP, Scarpulla RC. Transcriptional regulatory circuits controlling mitochondrial biogenesis and function. Genes Dev 2004;18:357 68. 163. Huss JM, Kelly DP. Mitochondrial energy metabolism in heart failure: a question of balance. J Clin Invest 2005;115:547 55. 164. Ino T, Sherwood WG, Benson LN, Wilson GJ, Freedom RM, Rowe RD. Cardiac manifestations in disorders of fat and carnitine metabolism in infancy. J Am Coll Cardiol 1988;11:1301 8. 165. Wallace DC. Mitochondrial defects in cardiomyopathy and neuromuscular disease. Am Heart J 2000;139:S70 85. 166. Bohles H, Sewell AC. Metabolic cardiomyopathy. rev. 2nd ed. Stuttgart: Medpharm;2004.
Chapter | 15
Cardiomyocyte Metabolism: All Is in Flux
167. Coates PM, Stanley CA. Inherited disorders of mitochondrial fatty acid oxidation. Prog Liver Dis 1992;10:123 38. 168. Hale DE, Thorpe C, Braat K, Wright JH, Roe CR, Coates PM, et al. The L-3-hydroxyacyl-CoA dehydrogenase deficiency. Prog Clin Biol Res 1990;321:503 10. 169. Kelly DP, Strauss AW. Inherited cardiomyopathies. N Engl J Med 1994;330:913 9. 170. Hale DE, Bennett MJ. Fatty acid oxidation disorders: a new class of metabolic diseases. J Pediatr 1992;121:1 11. 171. Young ME, Patil S, Davies PJA, Taegtmeyer H. Anaplerosis in the rat heart: decreased expression of pyruvate arboxylase with diabetes. Diabetes 2000;49(Suppl. 1):A278. 172. Zhou YT, Grayburn P, Karim A, Shimabukuro M, Higa M, Baetens D, et al. Lipotoxic heart disease in obese rats: implications for human obesity. Proc Natl Acad Sci USA 2000;97:1784 9. 173. Golfman LS, Wilson CR, Sharma S, Burgmaier M, Young ME, Guthrie PH, et al. Activation of PPARgamma enhances myocardial glucose oxidation and improves contractile function in isolated working hearts of ZDF rats. Am J Physiol Endocrinol Metab 2005;289:E328 36. 174. Young ME, Laws FA, Goodwin GW, Taegtmeyer H. Reactivation of peroxisome proliferator-activated receptor alpha is associated with contractile dysfunction in hypertrophied rat heart. J Biol Chem 2001;276:44390 5. 175. Dewald O, Sharma S, Adrogue J, Salazar R, Duerr GD, Crapo JD, et al. Downregulation of peroxisome proliferator-activated receptor-alpha gene expression in a mouse model of ischemic cardiomyopathy is dependent on reactive oxygen species and prevents lipotoxicity. Circulation 2005;112:407 15. 176. Taegtmeyer H. Six blind men explore an elephant: aspects of fuel metabolism and the control of tricarboxylic acid cycle activity in heart muscle. Basic Res Cardiol 1984;79:322 36. 177. Rudolph W, Haas D, Richter J, Masinger F, Hofmann H, Dohm P. Uber die Bedeutung von Acetoacetat und ß-Hydroxybutyrat im Stoffwechel der menschlichen Herzens. Klin Wochenscher 1965;43:445 51. 178. Williamson JR, Krebs HA. Acetoacetate as fuel of respiration in the perfused rat heart. Biochem J 1961;80:540 7. 179. Zimmermann ANE, Meijler FL, Hu¨lsmann WC. The inhibitory effect of acetoacetate on myocardial contraction. Lancet 1962;2:757 8. 180. Taegtmeyer H. On the inability of ketone bodies to serve as the only energy providing substrate for rat heart at physiological work load. Basic Res Cardiol 1983;78:435 50. 181. Panchal AR, Comte B, Huang H, Kerwin T, Darvish A, des Rosiers C, et al. Partitioning of pyruvate between oxidation and anaplerosis in swine hearts. Am J Physiol Heart Circ Physiol 2000;279:H2390 8. 182. Krebs HA. Some aspects of the regulation of fuel supply in omnivorous animals. Adv Enzyme Regul 1972;10:397 420. 183. Williamson JR, Safer B, LaNoue KF, Smith CM, Walajtys E. Mitochondrial-cytosolic interactions in cardiac tissue: role of the malate-aspartate cycle in the removal of glycolytic NADH from the cytosol. Symp Soc Exp Biol 1973;27:241 81. 184. Yu X, White LT, Alpert NM, Lewandowski ED. Subcellular metabolite transport and carbon isotope kinetics in the intramyocardial glutamate pool. Biochemistry 1996;35:6963 8. 185. Lewandowski ED, Yu X, LaNoue KF, White LT, Doumen C, O’Donnell JM. Altered metabolite exchange between subcellular
201
186.
187.
188.
189.
190. 191.
192.
193.
194.
195.
196.
197. 198.
199.
200. 201.
202. 203. 204.
compartments in intact postischemic rabbit hearts. Circ Res 1997;81:165 75. O’Donnell JM, Kudej RK, LaNoue KF, Vatner SF, Lewandowski ED. Limited transfer of cytosolic NADH into mitochondria at high cardiac workload. Am J Physiol Heart Circ Physiol 2004;286:H2237 42. Lewandowski ED, Doumen C, White LT, LaNoue KF, Damico LA, Yu X. Multiplet structure of 13C NMR signal from glutamate and direct detection of tricarboxylic acid (TCA) cycle intermediates. Magn Reson Med 1996;35:149 54. Taegtmeyer H, Ferguson AG, Lesch M. Protein degradation and amino acid metabolism in autolyzing rabbit myocardium. Exp Mol Pathol 1977;26:52. Mudge Jr GH, Mills Jr RM, Taegtmeyer H, Gorlin R, Lesch M. Alterations of myocardial amino acid metabolism in chronic ischemic heart disease. J Clin Invest 1976;58:1185 92. Bittl J, Shine K. Protection of ischemic rabbit myocardium by glutamic acid. Am J Physiol 1983;245:H406 12. Taegtmeyer H. Metabolic responses to cardiac hypoxia: increased production of succinate by rabbit papillary muscles. Circ Res 1978;43:808 15. Knapp W, Helus F, Ostertag H, Tilmanns H, Kuebler W. Uptake and turnover of L-[13N] glutamate in the normal human heart and patients with coronary artery disease. Eur J Nucl Med 1982;7:211 5. Robertson J, Vinten-Johansen J, Buckberg G, Rosenkranz E, Maloney J. Safety of prolonged aortic clamping with blood cardioplegia: glutamate enrichment in normal hearts. J Thorac Cardiovas Surg 1984;88:402 10. Young ME, McNulty P, Taegtmeyer H. Adaptation and maladaptation of the heart in diabetes: Part II: potential mechanisms. Circulation 2002;105:1861 70. Cohen P. The hormonal control of glycogen metabolism in mammalian muscle by multisite phosphorylation. Biochem Soc Trans 1979;7:459 80. Dobbins RL, Szczepaniak LS, Bentley B, Esser V, Myhill J, et al. Prolonged inhibition of muscle carnitine palmitoyltransferase-1 promotes intramyocellular lipid accumulation and insulin resistance in rats. Diabetes 2001;50:123 30. Friedman JJ. Fat in all the wrong places. Nature 2002;415:268 9. Chiu HC, Kovacs A, Ford DA, Hsu FF, Garcia R, Herrero P, et al. A novel mouse model of lipotoxic cardiomyopathy. J Clin Invest 2001;107:813 22. Yagyu H, Chen G, Yokoyama M, Hirata K, Augustus A, et al. Lipoprotein lipase (LpL) on the surface of cardiomyocytes increases lipid uptake and produces a cardiomyopathy. J Clin Invest 2003;111:419 26. Shulman GI. Cellular mechanisms of insulin resistance. J Clin Invest 2000;106:171 6. Petersen KF, Dufour S, Befroy D, Garcia R, Shulman GI. Impaired mitochondrial activity in the insulin-resistant offspring of patients with type 2 diabetes. N Engl J Med 2004;350:664 71. Unger RH, Orci L. Diseases of liporegulation: new perspective on obesity and related disorders. FASEB J 2001;15:312 21. Unger RH. Longevity, lipotoxicity and leptin: the adipocyte defense against feasting and famine. Biochimie 2005;87:57 64. Dillmann WH. Diabetes mellitus induces changes in cardiac myosin of the rat. Diabetes 1980;29:579 82.
202
205. Razeghi P, Young ME, Cockrill TC, Frazier OH, Taegtmeyer H. Downregulation of myocardial myocyte enhancer factor 2C and myocyte enhancer factor 2C-regulated gene expression in diabetic patients with nonischemic heart failure. Circulation 2002;106:407 11. 206. Ostrander DB, Sparagna GC, Amoscato AA, McMillin JB, Dowhan W. Decreased cardiolipin synthesis corresponds with cytochrome c release in palmitate-induced cardiomyocyte apoptosis. J Biol Chem 2001;276:38061 7. 207. Bjorkegren J, Veniant M, Kim SK, Withycombe SK, Wood PA, Hellerstein MK, et al. Lipoprotein secretion and triglyceride stores in the heart. J Biol Chem 2001;276:38511 7. 208. Nielsen LB, Bartels ED, Bollano E. Overexpression of apolipoprotein B in the heart impedes cardiac triglyceride accumulation
PART | II
209.
210. 211. 212.
Cardiac Muscle
and development of cardiac dysfunction in diabetic mice. J Biol Chem 2002;277:27014 20. Lee Y, Wang MY, Kakuma T, Wang ZW, Babcock E, McCorkle K, et al. Liporegulation in diet-induced obesity. The antisteatotic role of hyperleptinemia. J Biol Chem 2001;276:5629 35. Ferre´ P. Regulation of gene expression by glucose. Proc Nutr Soc 1999;58:621 3. Vaulont S, Vasseur-Cognet M, Kahn A. Glucose regulation of gene transcription. J Biol Chem 2000;275:31555 8. Zachara NE, Hart GW. O-GlcNAc a sensor of cellular state: the role of nucleocytoplasmic glycosylation in modulating cellular function in response to nutrition and stress. Biochim Biophys Acta 2004;1673:13 28.