Download Role of changes in cardiac metabolism in development - AJP

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Citric acid cycle wikipedia , lookup

Metalloprotein wikipedia , lookup

Lipid signaling wikipedia , lookup

Glucose wikipedia , lookup

Fatty acid synthesis wikipedia , lookup

Basal metabolic rate wikipedia , lookup

Ketosis wikipedia , lookup

Evolution of metal ions in biological systems wikipedia , lookup

Myokine wikipedia , lookup

Biochemistry wikipedia , lookup

Glycolysis wikipedia , lookup

Fatty acid metabolism wikipedia , lookup

Metabolism wikipedia , lookup

Transcript
Am J Physiol Heart Circ Physiol 291: H1489 –H1506, 2006.
First published June 2, 2006; doi:10.1152/ajpheart.00278.2006.
Invited Review
Role of changes in cardiac metabolism in development
of diabetic cardiomyopathy
Ding An and Brian Rodrigues
Division of Pharmacology and Toxicology, Faculty of Pharmaceutical Sciences,
The University of British Columbia, Vancouver, British Columbia, Canada
heart disease; diabetes; insulin resistance; lipoprotein lipase; fatty acid transporters;
PPARs; lipotoxicity; mitochondria dysfunction
HEART DISEASE is a leading cause of death in diabetic patients
(226, 253), with coronary vessel disease and atherosclerosis
being primary reasons for the increased incidence of cardiovascular dysfunction (151, 225, 253). However, a predisposition to heart failure might also reflect the effects of underlying
abnormalities in diastolic function that can be detected in
asymptomatic patients with diabetes alone (24, 61, 74). These
observations suggest a specific impairment of heart muscle
(termed diabetic cardiomyopathy). Rodent models of chronic
diabetes demonstrate abnormalities in diastolic left ventricular
function, with or without systolic left ventricular dysfunction
(73, 205, 219). Because diastolic dysfunction is also seen in
patients with both Type 1 and Type 2 diabetes, it can be
proposed that the diabetic state can directly induce abnormalities in cardiac tissue independent of vascular defects. Several
etiological factors have been put forward to explain the development of diabetic cardiomyopathy, including an increased
stiffness of the left ventricular wall associated with accumulation of connective tissue and insoluble collagen (9, 204, 212)
and abnormalities of various proteins that regulate ion flux,
specifically intracellular calcium (90, 235). More recently, the
view that diabetic cardiomyopathy could also occur as a
consequence of metabolic alterations has been put forward
(146, 147, 149).
Address for reprint requests and other correspondence: B. Rodrigues, Div. of
Pharmacology and Toxicology, Faculty of Pharmaceutical Sciences, The Univ.
of British Columbia, 2146 East Mall, Vancouver, BC, Canada V6T 1Z3
(e-mail: [email protected]).
http://www.ajpheart.org
Because uninterrupted contraction is a unique feature of the
heart, cardiac muscle has a high demand for provision of
energy. Under normal physiological conditions, hearts can
utilize multiple substrates, including fatty acids (FA), carbohydrate, amino acids, and ketones (15). Among these substrates, carbohydrates and FA are the major sources from
which the heart derives most of its energy. In a normal heart,
70% of ATP generation is through FA oxidation, whereas
glucose and lactate account for ⬃30% of energy provided to
the cardiac muscle (86, 173, 210). It should be noted that the
heart can rapidly switch its substrate selection to accommodate
different physiological and pathphysiological conditions involving altered extracellular hormones, substrate availability, and
workload (energy demand) (11, 130, 204, 217, 227). Acutely or
chronically, this regulation occurs through various mechanisms.
Although substrate switching is essential to ensure continuous
ATP generation to maintain heart function, it has also been
associated with deleterious consequences (17, 267).
In Type 1 and Type 2 diabetes mellitus, glucose uptake,
glycolysis, and pyruvate oxidation are impaired. Additionally,
lack of insulin function augments lipolysis and release of FA
from adipose tissue. Under these conditions, the heart rapidly
adapts to using FA exclusively for ATP generation. Chronically, this maladaptation is believed to play a key role in the
development of cardiomyopathy. Hence, it is essential to identify
and understand the mechanisms that control substrate utilization
in the diabetic heart. This review discusses the acute and chronic
cellular changes that modify cardiac metabolism and elucidate
their roles in the development of cardiomyopathy.
0363-6135/06 $8.00 Copyright © 2006 the American Physiological Society
H1489
Downloaded from http://ajpheart.physiology.org/ by 10.220.33.6 on May 5, 2017
An, Ding, and Brian Rodrigues. Role of changes in cardiac metabolism in development of diabetic cardiomyopathy. Am J Physiol Heart Circ Physiol 291: H1489 –
H1506, 2006. First published June 2, 2006; doi:10.1152/ajpheart.00278.2006.—In
patients with diabetes, an increased risk of symptomatic heart failure usually develops
in the presence of hypertension or ischemic heart disease. However, a predisposition to
heart failure might also reflect the effects of underlying abnormalities in diastolic
function that can occur in asymptomatic patients with diabetes alone (termed diabetic
cardiomyopathy). Evidence of cardiomyopathy has also been demonstrated in animal
models of both Type 1 (streptozotocin-induced diabetes) and Type 2 diabetes (Zucker
diabetic fatty rats and ob/ob or db/db mice). During insulin resistance or diabetes, the
heart rapidly modifies its energy metabolism, resulting in augmented fatty acid and
decreased glucose consumption. Accumulating evidence suggests that this alteration of
cardiac metabolism plays an important role in the development of cardiomyopathy.
Hence, a better understanding of this dysregulation in cardiac substrate utilization
during insulin resistance and diabetes could provide information as to potential targets
for the treatment of cardiomyopathy. This review is focused on evaluating the acute and
chronic regulation and dysregulation of cardiac metabolism in normal and insulinresistant/diabetic hearts and how these changes could contribute toward the development of cardiomyopathy.
Invited Review
H1490
ABNORMAL REGULATION OF CARDIAC SUBSTRATE METABOLISM IN DIABETES
DIABETIC CARDIOMYOPATHY
AJP-Heart Circ Physiol • VOL
CARDIAC METABOLISM
Glucose Uptake and Oxidation
Under normal physiological conditions, glucose is one of the
major carbohydrates utilized by the heart. Glucose metabolism
is regulated through multiple steps, including uptake, glycolysis, and pyruvate decarboxylation (Fig. 1). Cardiac glucose
uptake is dependent on the transmembrane glucose gradient
and the content of sarcolemmal glucose transporters (GLUT1
and GLUT4) (132, 154, 189). GLUT1 has a more pronounced
sarcolemmal localization and represents basal cardiac uptake.
When compared with GLUT1, GLUT4 is the dominant transporter in the adult heart, and under basal conditions, a majority
of this transporter is located in an intracellular pool (154). Its
translocation from an intracellular compartment to the sarcolemmal membrane requires insulin (154). Other than recruiting GLUT to the sarcolemmal membrane, insulin also influences glucose transport through its regulation of GLUT gene
expression (137, 183). It should be noted that GLUT-mediated
glucose uptake could also be stimulated through insulin-independent mechanisms. For example, recent studies have demonstrated that AMP-activated protein kinase (AMPK) also
promotes GLUT4 redistribution to the sarcolemmal membrane
(143, 261).
Once inside the cardiomyocyte, glucose is broken down
through glycolysis, a sequence of reactions that convert glucose into pyruvate. Phosphofructokinase-1 (PFK1), the enzyme
that catalyzes the generation of fructose 1,6-bisphosphate from
fructose 6-phosphate, is a rate-limiting enzyme controlling
glycolysis (57, 202). PFK-1 is inhibited by low pH, high
291 • OCTOBER 2006 •
www.ajpheart.org
Downloaded from http://ajpheart.physiology.org/ by 10.220.33.6 on May 5, 2017
The central basis for diabetic heart failure is coronary
disease associated with atherosclerosis (151, 225, 253). Because diabetes is often accompanied by hypertension and/or
ischemic heart disease, which often causes diastolic dysfunction, it is difficult to attribute heart failure in these patients to
diabetes alone (23). On the other hand, Type 1 or Type 2
diabetic patients have also been diagnosed with reduced or
low-normal diastolic function and left ventricular hypertrophy
in the absence of coronary heart disease or hypertension,
identified as cardiomyopathy (24, 81, 192, 195, 200, 231).
These abnormalities in asymptomatic diabetic patients potentially might predispose them to symptomatic heart failure when
they develop hypertension or ischemic heart disease, particularly after myocardial infarction.
In animal models of Type 1 diabetes, either streptozotocin
(STZ)-induced rats or genetic nonobese diabetic mice, systolic
and diastolic dysfunction have been demonstrated using in vivo
or in vitro measurements (177, 185, 232, 239). For example,
systolic and diastolic dysfunction have been detected by using
echocardiography in STZ diabetic animals (123, 177). With the
use of in vivo catheterization of these animals, elevated left
ventricular end-diastolic pressure and reduced left ventricular
systolic pressure with diminished increase or decrease pressure
over time (⫾dP/dt) were observed (124). Reduced peak left
ventricular pressure and ⫾dP/dt were confirmed in ex vivo
perfused STZ hearts (239). Recently, evidence of cardiomyopathy has also been reported in animal models of insulin
resistance and Type 2 diabetes (ZDF rats and ob/ob or db/db
mice) (5, 21, 30, 266). In db/db mice, systolic and diastolic
dysfunction were detected by echocardiography (218). With
the use of ex vivo perfused hearts from db/db mice, augmented
left ventricular end-diastolic pressure and reduced cardiac
output and cardiac power have also been observed (21). Given
that rodents are resistant to atherosclerosis, these models provided strong evidence for the occurrence of diabetic cardiomyopathy. However, in another study, no evidence of impairment
in cardiac function was detected using in ex vivo Langendorff
perfusions of hearts from insulin resistant and obese Zucker
rats (224) and Type 2 diabetic Zucker diabetic fatty (ZDF) rats
(250). It is likely that such inconsistencies arose due to the
rodent models used and the methods employed for evaluating
heart function in the different studies.
It should be acknowledged that the features of cardiomyopathy between human diabetic patients and rodent diabetic
models might be different. The development of diabetic cardiomyopathy is influenced by the severity and duration of
alterations in plasma parameters such as insulin, leptin, glucose, and lipids. Given the discrepancy in changes in these
hormones and substrates between human diabetic patients and
rodent models of Type 1 and Type 2 diabetes, the etiology and
severity of diabetic cardiomyopathy may vary (196). For example, STZ-induced diabetes exhibit severe hypoinsulinemia,
whereas ob/ob mice show leptin deficiency, which are rare in
human diabetes. Additionally, it has been suggested that STZ
can directly impair the cardiac contractile function of isolated
ventricular myocytes (256). Finally, although db/db mice and
Zucker and ZDF rats exhibit hyperinsulinemia, hyperleptinemia, and hyperglycemia, the severity of these alterations vary
between animal models and patients with diabetes (196).
Taken together, experimental data obtained using animals
models of diabetes should be used with caution when extrapolating to the human condition.
Cardiomyopathy is a complicated disorder, and several factors have been associated with its development. These include
increased stiffness of the left ventricular wall associated with
accumulation of connective tissue and insoluble collagen (204,
212), depressed autonomic function (73), impaired endothelium function and sensitivity to various ligands (e.g., ␤-agonists) (23), and abnormalities of various proteins that regulate
ion flux, specifically intracellular calcium (90, 235). More
recently, there is a generalized view that diabetic cardiomyopathy also occurs as a consequence of altered fuel metabolism.
During insulin resistance or diabetes, glucose utilization is
compromised. This alteration, together with increased FA supply, switches cardiac energy generation to utilization of FA.
High FA uptake and metabolism not only augment accumulation of FA intermediates and triglycerides but also increase
oxygen demand and generation of reactive oxygen species
(ROS), leading to cardiac damage. Interestingly, increasing FA
uptake through overexpression of cardiac human lipoprotein
lipase (LPL) (260) or FA transport protein (47), or augmenting
FA oxidation through overexpression of cardiac peroxisome
proliferator-activated receptor (PPAR)-␣ (79) or long-chain
acyl CoA synthase (48), results in a cardiac phenotype resembling diabetic cardiomyopathy. Conversely, normalizing cardiac metabolism in diabetic animals reverses the development
of cardiomyopathy (21, 272). Taken together, these studies
strongly support the role of altered metabolism in the development of diabetic cardiomyopathy.
Invited Review
ABNORMAL REGULATION OF CARDIAC SUBSTRATE METABOLISM IN DIABETES
H1491
(107, 258), which is inhibited by pyruvate and stimulated by
high mitochondrial acetyl-CoA-to-CoA and NADH-toNAD(⫹) ratios (29, 104). Acetyl-CoA then enters the tricarboxylic acid cycle and is eventually broken down to H2O and
CO2 for ATP generation. Oxidation of one glucose provides 30
ATP with 6 oxygen consumed, which combined with 2 glycolytic ATPs produces a phosphate-to-oxygen ratio of 2.58 ATP/
oxygen atom. Although glucose is the focus in this review,
other carbohydrates like lactate also make an important
contribution toward cardiac energy generation (39). In fact,
as greater lactate compared with glucose utilization is observed in the isolated working heart (119, 127, 148), the
inclusion of lactate in the perfusion buffer is recommended
in future studies.
Fig. 1. Glucose utilization in the cardiomyocyte. Glucose uptake into cardiomyocyte occurs through GLUT1 and GLUT4 transporters. Once inside, glucose is broken down through glycolysis, a sequence of reactions that convert
glucose into pyruvate. Phosphofructokinase-1 (PFK1), the enzyme that catalyze the generation of fructose 1,6-bisphosphate from fructose 6-phosphate, is
a rate-limiting enzyme controlling glycolysis. PFK1 is activated by 2,6bisphosphate, which is formed from fructose 6-phosphate catalyzed by PFK-2.
After glycolysis, the pyruvate generated is transported into mitochondria and
decarboxylated to acetyl-CoA through pyruvate dehydrogenase (PDH), a
multienzyme complex. PDH is phosphorylated and inactivated by pyruvate
dehydrogenase kinase (PDK). Acetyl-CoA then enters tricarboxylic acid cycle
(citric acid cycle) and is eventually broken down to H2O and CO2 for ATP
generation.
intracellular citrate, or ATP and is activated by ADP, AMP,
phosphate, and fructose 2,6-bisphosphate (57, 202, 228). Fructose 2,6-bisphosphate is formed from fructose 6-phosphate
catalyzed by PFK-2 (110). Given that PFK-2 is phosphorylated
and activated by hormones, such as insulin (25, 58) and
glucagon (228) or kinases such as AMPK (109, 161), stimulation of PFK-2 through these mechanism increases fructose
2,6-bisphosphate generation, activates PFK-1, and subsequently promotes glycolysis. In an aerobic heart, glycolysis
contributes ⬍10% of total ATP generated, with no oxygen
consumption.
After glycolysis, the pyruvate generated has three destinations: carboxylation to oxaloacetate or malate, reduction to
lactate, and most importantly, decarboxylation to acetyl-CoA.
To be oxidized, pyruvate is transported into the mitochondria
and decarboxylated to acetyl-CoA through pyruvate dehydrogenase (PDH), a multienzyme complex. PDH is phosphorylated and inactivated by pyruvate dehydrogenase kinase (PDK)
AJP-Heart Circ Physiol • VOL
When compared with glucose, FA is the preferred substrate
and accounts for ⬃70% of ATP generated in an aerobic heart.
FA metabolism includes multiple steps and can be regulated by
both acute and chronic mechanisms, with or without modulation of gene expression (Fig. 2).
Lipoprotein lipase. Because the heart has limited capacity to
synthesize and store FA, it relies on continuous exogenous
supply. FA supplied to the heart from the circulation is from
two sources: albumin-bound FA and triglyceride (TG)-rich
lipoproteins. It should be noted that the molar concentration of
FA in lipoprotein-TG is ⬃10-fold larger than FA bound to
albumin (165). Lipoprotein lipase (LPL) is the key enzyme that
hydrolyzes lipoproteins to release FA. Thus, when lipoproteins
are hydrolyzed by LPL, a large amount of FA are released and
are believed to be the principle source of FA supplied to the
heart (14, 237). LPL is produced in cardiomyocytes and subsequently secreted onto heparan sulfate proteoglycan binding
sites on the myocyte cell surface (33, 68, 70). From here, LPL
is transported onto comparable binding sites on the luminal
surface of coronary endothelial cells (193). At these sites, LPL
hydrolyzes TG-rich lipoproteins, such as very-low-density lipoprotein (VLDL) and chylomicrons, to release FA. Perfusion
of working hearts with VLDL or chylomicrons, in the absence
or presence of FA, has revealed that, compared with chylomicrons, VLDL is a poor substrate for LPL (100). In addition,
utilization of chylomicrons was inhibited by free FA, which
failed to affect cardiac VLDL utilization (100). Given the
important role that LPL plays in regulating FA delivery,
alteration in its level is able to change FA delivery and
subsequent oxidation. Indeed, overexpression of LPL in the
heart or skeletal muscle accelerates FA uptake (141, 260).
Conversely, tissue-specific knockout of LPL in the heart
switches the cardiac substrate selection preference to glucose
(12, 13). Other than hydrolysis by LPL, lipoproteins can also
be transported into the cardiac tissue with help of the lipoprotein receptor, with VLDL uptake through this mechanism
being more substantial compared with chylomicrons (178). In
perfused working hearts, inhibition of the lipoprotein receptor
significantly reduced FA oxidation without altering FA uptake,
suggesting that lipoprotein receptor mediated FA uptake channels FA toward oxidation, whereas LPL mediated FA uptake
channels FA toward both oxidation and TG storage (178).
FA transporters. Although FA can translocate into the cell
through passive diffusion across the plasma membrane, FA
291 • OCTOBER 2006 •
www.ajpheart.org
Downloaded from http://ajpheart.physiology.org/ by 10.220.33.6 on May 5, 2017
FA Uptake and Utilization
Invited Review
H1492
ABNORMAL REGULATION OF CARDIAC SUBSTRATE METABOLISM IN DIABETES
uptake shows saturation kinetics and is inhibited by proteases
(7, 158, 159, 230). Thus FA transporters are also likely required to support this process. In the heart, three FA transporters have been identified and these include CD36, FA transport
protein (FATP), and FA binding protein plasma membrane
(FABPpm) (157). Given that 55– 80% of FA transport was
blocked using general transporter inhibitors (131, 158) and that
overexpression of CD36 or FATP has been found to dramatically increase FA metabolism (47, 113), these FA transporters
are believed to play a key role in FA delivery to the cardiac
tissue.
Regulation of FA transport proteins occurs through different
mechanisms. In severe STZ-induced diabetes, expression of
cardiac CD36 and FABPpm were augmented (152), suggesting
transcriptional modification of these transporters. At present,
the mechanisms that induce this change have yet to be elucidated. Additionally, because muscle contraction or acute insulin treatment does not change protein but simply relocates
CD36 from an intracellular pool to the sarcolemmal membrane
(155, 156), posttranslational regulation of this transporter has
also been suggested.
Acyl-CoA synthase. Acyl-CoA synthase (ACS) catalyzes the
esterification of FA to fatty acyl-CoA, the initial step of FA
metabolism. Fatty acyl-CoA can be transported into the mitochondria for oxidation or used for intracellular TG synthesis.
The fate of fatty acyl-CoA is influenced by the location of
different ACS isoforms, energy demand, and the availability of
FA (37). Moreover, ACS not only functions as an enzymecatalyzing esterification but is also actively involved in controlling FA homeostasis (48). Recent studies have demonstrated that ACS is associated with CD36 or FATP on the
cytosolic side of the sarcolemmal membrane (84, 201, 214),
suggesting that ACS also influences FA uptake. Several studies
also suggest that FATP1 is a very long chain ACS (53).
Overexpression of ACS in the heart or fibroblast causes dramatically augmented FA uptake and intracellular TG accumulation (48).
AJP-Heart Circ Physiol • VOL
Under normal conditions, 70 –90% of the esterified FA that
enters cardiomyocytes is oxidized for ATP generation, whereas
10 –30% is converted to TG (228). The TG pool is not static,
with lipolysis and esterification taking place continuously
(240). In a normal heart, intracellular TG level is constant,
indicating a balance of lipogenesis and lipolysis (210). In
situations where FA supply supercedes the cellular oxidative
capacity, such as obesity or diabetes, intracellular TG accumulates and is associated with lipotoxicity (240). Although TG is
unlikely to be a direct mediator of cell apoptosis, its augmented
lipolysis expands fatty acyl-CoA levels, which may be a key
factor mediating cell apoptosis (117, 240). Thus TG is often
used as a marker of lipotoxicity.
PPARs. PPARs are a group of ligand-activated transcriptional factors belonging to the superfamily of nuclear receptors. They are activated by either natural ligands like FA or
numerous pharmacological ligands (80). Once activated,
PPARs form complexes with retinoid X receptors and bind to
the promoter regions of a number of target genes that encode
the proteins involved in controlling FA metabolism (76, 78).
Through regulation of expression of these genes, PPARs modulate FA utilization at the transcriptional level. PPARs have
three isoforms: PPAR-␣, PPAR-␤ (or -␦), and PPAR-␥.
PPAR-␣ is extensively expressed in tissues with high FA
metabolism like the heart (17). Activated by elevated intracellular FA levels, PPAR-␣ promotes expression of genes that
regulate FA oxidation at various steps, such as FA uptake and
binding (LPL, CD36, and FA binding protein), FA esterification (ACS), and FA oxidation (carnitine palmitoyltransferase-1, acyl-CoA oxidase, long-chain acyl-CoA dehydrogenase, and very-long-chain acyl-CoA dehydrogenase) (76, 78,
111). Knocking out cardiac PPAR-␣ abolishes fasting-induced
overexpression of FA metabolic genes and switches substrate
selection from FA to glucose (140, 172). Overexpression of
cardiac PPAR-␣ augments FA uptake and oxidation (77, 79).
Taken together, PPAR-␣ is believed to be the primary regulator of FA metabolism in the heart. Similar to PPAR-␣, PPAR-␤
291 • OCTOBER 2006 •
www.ajpheart.org
Downloaded from http://ajpheart.physiology.org/ by 10.220.33.6 on May 5, 2017
Fig. 2. Control of fatty acid (FA) delivery and
utilization in the cardiomyocyte. FA, either from
adipose tissue or released from triglyceride
(TG)-rich lipoproteins through hydrolysis by lipoprotein lipase (LPL), is taken up into the
cardiomyocyte by three FA transporters: CD36,
FA transport protein (FATP), and FA binding
protein plasma membrane (FABPpm). FA is converted to fatty acyl-CoA, which is transported
into the mitochondria through CPT1/CPT2. Inside the mitochondria, fatty acyl-CoA undergoes
␤-oxidation to generate acetyl-CoA, which is
further oxidized in the tricarboxylic acid cycle.
Utilization of FA is regulated through different
mechanisms. FA, through activation of peroxisome prolferator-activated receptor-␣ (PPAR␣), increases the expression of a number of
enzymes involved in FA oxidation. MalonylCoA, which is generated through carboxylation
of acetyl-CoA catalyzed by acetyl-CoA carboxylase (ACC), inhibits CPT-1 and FA oxidation.
AMP-activated kinase (AMPK) inhibits ACC,
relieves its inhibition on CPT-1, and promotes
FA oxidation. Similarly, malonyl-CoA decarboxylase (MCD), through decreasing malonylCoA by decarboxylating it to acetyl-CoA, enhances CPT-1 and FA oxidation.
Invited Review
ABNORMAL REGULATION OF CARDIAC SUBSTRATE METABOLISM IN DIABETES
lumen LPL activity (8). Interestingly, recent studies using
transgenic mice with a dominant negative form of AMPK has
demonstrated that the lack of AMPK does not affect cardiac
metabolism under physiological conditions (209, 259). At
present, it is unclear as to what compensatory mechanisms are
activated following knockout of AMPK. Additionally, it is
unknown whether overexpression of AMPK could affect cardiac lipid homeostasis and metabolism.
Malonyl-CoA decarboxylase. In addition to AMPK, malonyl-CoA decarboxylase (MCD) is also known to promote FA
oxidation through its lowering of malonyl-CoA. MCD catalyzes the degradation of malonyl-CoA to acetyl-CoA, leading
to reduction of malonyl-CoA (65). This action relieves the
inhibition of CPT-1 by malonyl-CoA and favors FA oxidation.
Recent studies have suggested that inhibition of cardiac MCD
leads to accumulation of malonyl-CoA and reduced FA oxidation (66).
Interaction Between Glucose and FA Metabolism
Regulation of glucose and FA metabolism does not occur
independently, and numerous studies have reported a “crosstalk” between the utilization of these substrates (Fig. 3) (199,
211, 233). Randle et al. (199) demonstrated that FA impairs
basal and insulin-stimulated glucose uptake and oxidation, an
event that is known as the “Randle cycle” (199). FA influences
glucose utilization at multiple levels. Accumulation of FA
impairs insulin-mediated glucose uptake through inhibition of
insulin receptor substrate and protein kinase B (94, 115, 187).
Accumulation of FA leads to augmented intracellular FA
derivatives, such as fatty acyl CoA, diacylglycerol, and ceramide. These FA metabolites activate a serine kinase cascade,
which involves protein kinase C-␪ and I␬B kinase-␤ (inhibitor
of NF-␬B kinase-␤), leading to serine phosphorylation of IRS
(129, 269). Serine phosphorylation of IRS-1 reduces tyrosine
Fig. 3. Inhibition of glucose oxidation by FA
utilization. Accumulation of FA impairs insulin-mediated glucose uptake through inhibition of insulin receptor substrate (IRS) and
protein kinase-B (PKB). Increased intracellular FA also activates PPAR-␣. As a consequence, PPAR-␣ promotes the expression of
genes involved in FA oxidation, as well as
PDK4, known to inhibit PDH and pyruvate
flux. Increased acetyl-CoA and NADH
caused by the high rate of FA oxidation
activate PDK4, leading to further inactivation
of PDH. Augmented acetyl-CoA also causes
accumulation of citrate in the cytosol, which
subsequently inhibits PFK1 and glycolysis.
AJP-Heart Circ Physiol • VOL
291 • OCTOBER 2006 •
www.ajpheart.org
Downloaded from http://ajpheart.physiology.org/ by 10.220.33.6 on May 5, 2017
(or -␦) is expressed abundantly in the cardiac tissue (17).
Activated by elevated intracellular FA (45), PPAR-␤ (or -␦)
augments expression of a group of genes that promote FA
utilization (63, 169). Cardiac-specific knockout of PPAR-␤
also decreases FA oxidative gene expression and FA oxidation
(46). Although the targets of PPAR-␣ and PPAR-␤ are partially overlapping (169), their unique roles and interaction
remains unclear in the heart. PPAR-␥, the third number of the
PPAR family, is highly expressed in adipose tissue. Through
promoting lipogenic gene expression, PPAR-␥ controls lipogenesis. Loss and gain of function experiments have demonstrated that PPAR-␥ is necessary for adipose tissue proliferation and differentiation (16, 238). In isolated cardiomyocytes,
the expression of PPAR-␥ is barely detectable (89), suggesting
a limited direct role for this nuclear receptor in regulating
cardiac metabolism. However, the ability of PPAR-␥ agonists
to normalize elevated plasma concentrations of glucose and FA
in diabetic animals will have a profound indirect effect on
cardiac metabolism.
AMP-activated protein kinase. As an energy sensor, AMPactivated protein kinase (AMPK) is activated following a rise
in the intracellular AMP-to-ATP ratio (97). Once stimulated,
AMPK switches off energy-consuming processes like protein
synthesis, whereas ATP-generating mechanisms, such as FA
oxidation and glycolysis, are turned on (96, 105). In heart and
skeletal muscle, AMPK facilitates FA utilization through its
control of acetyl-CoA carboxylase (ACC) (133, 134). As ACC
catalyzes the conversion of acetyl-CoA to malonyl-CoA,
AMPK by inhibiting ACC is able to decrease malonyl-CoA
and minimize its inhibition of CPT-1, the rate-limiting enzyme-controlling FA oxidation. AMPK has also been implicated in FA delivery to cardiomyocytes through its regulation
of the FA transporter CD36 (155). Additionally, results from
our laboratory have demonstrated a strong correlation between
activation of whole heart AMPK and increases in coronary
H1493
Invited Review
H1494
ABNORMAL REGULATION OF CARDIAC SUBSTRATE METABOLISM IN DIABETES
ALTERATION OF CARDIAC METABOLISM DURING DIABETES
Changes in Plasma Substrates
Glucose and lipids are the major substrates affected by
diabetes. Hyperglycemia is a consequence of decreased glucose clearance and augmented hepatic glucose production,
whereas enhanced lipolysis in adipose tissue and higher lipoprotein synthesis in the liver dramatically increases circulating free FA and TG. Because glucose entry into the cell is
largely dependent on insulin, whereas FA transport across
plasma membrane does not require any hormone, the augmented circulating lipids increase FA delivery to the cardiomyocyte. With the increase of intracellular FA, cardiac tissue
rapidly adapts to promote FA utilization. In addition to diabetes, augmentation of plasma lipids by fasting or intralipidheparin infusion also increases cardiac FA oxidation (52, 266).
Under conditions where FA supply supercedes the oxidative
capacity of the heart, the FA is converted to lipids like TG or
ceramide, with an end result being lipotoxicity (272). In obese
ZDF rats, lowering of plasma lipids with a PPAR-␥ agonist
reduced cardiac TG and ceramide and improved heart function
(272). Overall, these studies suggest that alterations in plasma
lipids may drive changes in cardiac metabolism. It should be
noted that a recent study using 4-wk-old ob/ob and db/db mice
demonstrated altered cardiac metabolism without any change
in plasma substrates (30). Hence, in these genetic mice models,
intrinsic changes in the cardiomyocyte, rather than changes in
circulating substrates, initiate alterations in cardiac metabolism
at this early time point.
Defects in Cardiac Carbohydrate Utilization
In the obese or diabetic heart, myocardial glucose utilization
is compromised at several points. In noninsulin-controlled
Type 1 diabetic animals, reduced GLUT gene and protein
expression compromises cardiac glucose uptake and oxidation
(34). In obese or Type 2 diabetic animals, although there is
hyperglycemia and hyperinsulinemia, cardiac glucose uptake is
reduced as a consequence of reduced GLUT4 protein and
impaired insulin signaling (38, 266). With the use of ob/ob and
db/db mice, a recent study (30) reported that cardiac glucose
oxidation is reduced at 4 wk of age and was associated with
increased FA oxidation. Interestingly, this lower glucose oxidation occurred before the onset of impaired insulin signaling
in the heart and development of hyperglycemia. Thus this early
AJP-Heart Circ Physiol • VOL
reduction in glucose utilization is likely due to suppression by
high FA oxidation rather than impaired cardiac-specific insulin
signaling. In another study that used db/db mice at different
ages, increased cardiac FA oxidation preceded the reduction in
glucose oxidation (5). Higher FA oxidation increases citrate,
which is known to inhibit PFK, the rate-limiting enzyme in
glycolysis. Elevated intracellular FA is also known to increase
PDK-4 expression, which phosphorylates and inhibits PDH.
Finally, high rates of FA oxidation augment acetyl-CoA that
inhibits PDH either allosterically or through activation of PDK.
When compared with animal models, the use of carbohydrates in the human diabetic hearts is more controversial. In
Type 1 diabetic patients, decreased myocardial carbohydrate
uptake is reported (15, 62). With the use of an euglycemic
insulin clamp, another study (179) has shown that cardiac
glucose uptake is normal in these patients, suggesting that
insulin is the major limiting factor that influences cardiac
glucose uptake and no cardiac-specific insulin resistance is
evident, as observed in skeletal muscle. In Type 2 diabetes,
cardiac expression of GLUT4 is compromised likely due to
elevated FA (10). However, a number of studies have also
reported that cardiac tissue from Type 2 diabetic patients
respond normally to insulin and show regular glucose uptake,
suggesting that myocardial insulin resistance is not a common
feature of Type 2 diabetes (118, 160, 242). Interestingly, a
recent study has suggested that impaired myocardial glucose
uptake is only observed in Type 2 diabetic patients with
hypertriglyceridemia, suggesting that myocardial insulin resistance in these patients is associated with hypertriglyceridemia
and augmented plasma FA levels (168).
Cardiac lactate utilization is also compromised following
diabetes. In STZ-induced Type 1 diabetes, cardiac utilization
of lactate is reduced to a greater extent when compared with
glucose oxidation (41, 42). Moreover, hearts from 12-wk-old
ZDF rats also showed lower carbohydrate oxidation, and this
was almost entirely due to a reduction in lactate rather than
glucose oxidation (43, 250). The mechanisms that mediate the
inhibition of lactate utilization are unclear but is likely independent of any alterations in lactate dehyrogenase or lactate
transporter (41).
The shift of cardiac energy substrate utilization from carbohydrate to lipids increases the intracellular glycogen pool,
probably through augmented glycogen synthesis, or impaired
glycogenolysis, or a combination of both processes (103, 136,
167, 223). Emerging evidence indicates that glycogen, in
addition to its role as energy storage, is also able to regulate
metabolism. At least in skeletal muscle, accumulation of glycogen acutely alters glycogen synthesis and glucose metabolism (59, 120 –122). Chronically, glycogen accumulation may
also impair insulin signaling in skeletal muscle (121). Whether
cardiac glycogen accumulation also influences insulin signaling and metabolism is unknown. In addition, recent studies
have identified a glycogen-binding domain in the ␤-subunit of
AMPK and suggests that this domain links AMPK and glycogen (108). Although an association between high glycogen
levels and repressed AMPK activity in human and rat skeletal
muscle have been documented (254, 255), it is still unclear
whether glycogen is able to regulate metabolism through
AMPK.
291 • OCTOBER 2006 •
www.ajpheart.org
Downloaded from http://ajpheart.physiology.org/ by 10.220.33.6 on May 5, 2017
phosphorylation and interferes with its ability to phosphorylate
and activate phosphatidylinositol 3-kinase and protein kinase B
(206). Increased intracellular FA also activates PPAR-␣. As a
consequence, PPAR-␣ promotes the expression of genes involved in FA oxidation, as well as pyruvate dehydrogenase
kinase-4 (PDK4), which is known to inhibit PDH and pyruvate
flux (257). Moreover, increased acetyl-CoA-to-free CoA and
NADH-to-NAD⫹ ratios caused by the high rate of FA oxidation are also known to activate PDK4, leading to inactivation
of PDH (29, 104). Augmented acetyl-CoA-to-free CoA ratio
also causes accumulation of citrate in the cytosol, which
subsequently inhibits PFK and glycolysis (85, 175). Conversely, inhibition of FA oxidation through elevation of malonyl-CoA levels, or using pharmacological inhibitors, favors
glucose oxidation.
Invited Review
ABNORMAL REGULATION OF CARDIAC SUBSTRATE METABOLISM IN DIABETES
Alterations in FA Utilization
AJP-Heart Circ Physiol • VOL
expression of VLDL receptor decreases following STZ-induced diabetes (116).
FA transporters. During obesity and diabetes, the high rate
of FA uptake is facilitated by FA transporters, leading to
augmented FA oxidation and TG storage. In STZ-induced
diabetes, the increase in plasmalemmal CD36 and FABPpm
amplifies FA uptake, an effect resulting from increased CD36
and FABPpm protein expression (152). In Zucker fatty rats,
without a change in total protein, permanent relocation of
CD36 and FABPpm to the cardiomyocyte plasmalemmal membrane augments FA uptake (55, 56, 153). The mechanism for
this permanent repositioning of FA transporters at the plasma
membrane is still unknown. Interestingly, in young (4 wk)
ob/ob or db/db mice, cardiac FA oxidation increases without
any change in circulating substrates (30). Given that the high
rate of FA oxidation requires coordinated FA uptake, augmentation of CD36-mediated FA supply is suggested to be involved in facilitating FA oxidation at this early stage of insulin
resistance.
PPAR-␣. As the principle regulator of cardiac FA metabolism, PPAR-␣ plays an important role in controlling FA oxidation during obesity and diabetes. Activation of this nuclear
receptor in the heart has been reported in almost all obese or
diabetic animal models, including STZ-induced diabetic rats,
ZDF rats, and ob/ob and db/db mice (30, 79, 220). Given that
FA and its derivatives activate PPAR-␣, higher cardiac
PPAR-␣ activity is always observed when circulating lipids are
augmented (30, 171). Hence, in ob/ob and db/db mice, activation of PPAR-␣, evidenced by elevated expression of its
downstream targets, is only observed at 12 wk and is associated with augmented plasma lipid (30). A similar association
between plasma lipids and PPAR-␣ is also seen in hearts from
STZ-induced diabetic rats, with only chronic (6 wk) but not
acute (4 days) diabetes demonstrating activation of PPAR-␣
(unpublished data). Interestingly, in acute STZ diabetes or
4-wk ob/ob or db/db mice, increased FA oxidation is observed
even in the absence of any change in cardiac PPAR-␣ and its
downstream targets, suggesting PPAR-␣ independent control
of FA oxidation. Identification of these mechanisms requires
further studies.
After activation, cardiac PPAR-␣ promotes the expression
of a group of genes involved in various steps of FA oxidation.
Simultaneously, activation of PPAR-␣ also reduces expression
of genes involved in glucose uptake, glycolysis, and oxidation
through direct or indirect mechanisms (77, 79). With the use of
transgenic mice, a recent study (186) demonstrated that knocking out cardiac PPAR-␣ prevented suppression of GLUT4
expression and glucose uptake by elevated plasma FA and
improved myocardial recovery from ischemia following STZ
diabetes induction, high-fat feeding, or fasting. Taken together,
activation of cardiac PPAR-␣ not only favors FA oxidation but
also inhibits glucose uptake and utilization, leading to augmented susceptibility to ischemic damage.
MCD. In STZ-induced diabetes, overexpression of cardiac
MCD protein was observed and could contribute to the high
rate of FA oxidation (213). It is unknown whether MCD also
plays a role in augmenting cardiac FA oxidation in obese and
Type 2 diabetes. Nevertheless, given that PPAR-␣ increases
MCD expression (138, 265), higher MCD protein levels are
expected following PPAR-␣ activation during obesity and
Type 2 diabetes.
291 • OCTOBER 2006 •
www.ajpheart.org
Downloaded from http://ajpheart.physiology.org/ by 10.220.33.6 on May 5, 2017
Utilization of FA by cardiac tissue increases following
obesity and diabetes. This change occurs not only as a consequence of increased FA supply but also through an intrinsic
adaptation/maladaptation to elevated FA (234, 267). It must be
acknowledged that some studies have also observed a reduced,
rather than augmented, FA oxidation in Zucker fatty or ZDF
rats (266, 272). In human diabetic patients, obese women
demonstrate increased FA utilization, associated with augmented cardiac oxygen consumption, and reduced cardiac
efficiency (190). Moreover, elevated cardiac TG and increased
expression of PPAR-␣ target genes have been observed in
patients without ischemic heart disease, very similar to lipotoxic ZDF rat hearts (220).
Lipoprotein lipase. The amount of FA supplied to the heart
through LPL is influenced by multiple factors, including LPL
activity and plasma lipoprotein concentration and composition.
The relative contribution of cardiac LPL activity to the delivery
of free FA to the diabetic heart is inconclusive. Thus LPL
immunoreactive protein or activity has been reported to be
unchanged, increased, or decreased in the diabetic rat heart. In
part, this variability between different studies could be due to
the diversity in the rat strains used, the dosage of STZ used to
induce diabetes, and the duration of the diabetic state (203). In
addition, many of the above investigations utilized procedures
that did not distinguish between functional (i.e., heparin-releasable component localized on capillary endothelial cells that is
implicated in the hydrolysis of circulating TG) and cellular
(i.e., non-heparin-releasable pool that represents a storage form
of the functional enzyme) pools of cardiac LPL because cellular LPL activity or protein levels have largely been obtained
using whole heart homogenates. Rodrigues et al. (203) have
previously reported that in STZ (55 mg/kg)-induced moderate
diabetic Wistar rat hearts, HR-LPL activity is significantly
increased. Induction of more severe diabetes using 100 mg/kg
STZ did not influence HR-LPL (203). In another study using
hearts from 65 mg/kg STZ-induced diabetic rats, HR-LPL
decreased and was associated with reduced VLDL-TG lipolysis (182). Acute treatment of these diabetic rats with insulin
enhanced both HR-LPL and VLDL-TG lipolysis (182). In
mouse models of diabetes, although there is no change of
cardiac HR-LPL activity in either STZ or db/db hearts, utilization of chylomicron-TG increases (174). Hence, the mechanisms that control cardiac LPL during diabetes are complex
and have yet to be completely resolved.
It should be noted that even though LPL may not change
following obesity or diabetes, increased circulating lipoproteins through LPL cleavage could still elevate FA supply to the
cardiac tissue. However, increased concentration could be
offset by changes in lipoprotein composition or lipoprotein
receptor number. Binding of lipoproteins containing apolipoprotein (apo) CII to LPL enhances lipolysis (181), whereas
binding of lipoproteins containing apoCIII or apoE suppresses
LPL activity (1, 2). A recent study has shown that overexpression of HDL-associated apo AV accelerates hydrolysis of
TG-rich lipoproteins, indicating the importance of this apo
lipoprotein in stimulating LPL (166, 188). O’Looney et al.
(181) have shown that induction of diabetes by STZ changes
lipoprotein composition, with reduced apoCII levels, leading to
impaired VLDL-TG lipolysis. A recent study has reported that
H1495
Invited Review
H1496
ABNORMAL REGULATION OF CARDIAC SUBSTRATE METABOLISM IN DIABETES
Limitations of Metabolic Measurements Using
Radiolabeled Substrate
CONSEQUENCES OF ALTERED METABOLISM IN
DIABETIC HEARTS
Impaired Cardiac Function
Emerging evidence supports the concept that alterations in
metabolism contribute toward cardiac contractile dysfunction.
In STZ-induced diabetes, a larger number of studies have
implicated metabolic abnormalities (FA oxidation provides
almost 100% of ATP, with a dramatic decrease of glucose
utilization) in cardiac contractile dysfunction (219, 227, 234,
267). In these animals, contractile failure begins as diastolic
dysfunction, followed by severe systolic dysfunction (196,
219). Normalizing energy metabolism in these hearts reverses
the impaired contractility (40, 176, 248). Animal models of
obesity and Type 2 diabetes also exhibit cardiac dysfunction
associated with altered cardiac metabolism (5, 21, 218). Some
studies have observed both impaired diastolic and systolic
function (3, 218), whereas other studies argue that no change in
systolic function is present (6, 18, 164, 197). This discrepancy
could be due to the severity of diabetes or the methods used to
evaluate cardiac function. In diabetic patients, left ventricular
hypertrophy and impaired isovolumic relaxation and ventricular filling are the most common abnormalities diagnosed (196).
During diabetes, changes in cardiac metabolism occur early
and precede the development of cardiomyopathy. For example,
in STZ-induced diabetes, altered cardiac metabolism is observed as early as 4 days following diabetes induction (88),
whereas evidence of cardiomyopathy is only apparent after
4 – 6 wk. Similarly, in ob/ob and db/db mice, changes in
cardiac metabolism are evident much before confirmation of
cardiac dysfunction (5, 30). To validate the role of altered
metabolism in provoking cardiac dysfunction, several studies
have treated 6- or 9-wk-old db/db mice with either PPAR-␣ or
AJP-Heart Circ Physiol • VOL
291 • OCTOBER 2006 •
www.ajpheart.org
Downloaded from http://ajpheart.physiology.org/ by 10.220.33.6 on May 5, 2017
The majority of metabolic data in the above studies were
obtained by using radiolabeled substrate perfusion of isolated
hearts. In these methods, [3H]glucose and [14C]glucose were
used to evaluate glycolysis and glucose oxidation, whereas
[3H]FA or [14C]FA was used to estimate FA oxidation by
measuring the generation of 3H2O and 14CO2. Although this
measurement allows for the manipulation of substrate concentration, there are limitations. First, the substrates used in
perfused hearts do not truly reflect what the heart receives in
vivo. For example, lipoproteins, lactate, and ketone bodies that
are important substrates for the heart in vivo (39, 98, 165) are
rarely used to examine metabolism in perfused hearts. In
addition, most metabolic data of diabetic hearts are obtained by
using normal concentrations of glucose and/or palmitate,
which differ from the elevated concentrations seen in the
diabetic conditions in vivo. Another drawback is that diabetic
hearts always have an elevated intracellular TG pool. Augmented intracellular TG turnover in diabetic hearts could dilute
utilization of exogenous radiolabeled palmitate oxidation, resulting in an underestimation of FA oxidation (180). Finally,
hormonal effects on energy metabolism are usually not considered (20), and lack of hormones in the in vitro perfusate may
contribute to functional abnormalities that are not recapitulated
in intact models.
PPAR-␥ agonists for 3– 6 wk (3, 4, 36). Even though these
treatments normalized cardiac metabolism, they failed to improve cardiac function in these mice, suggesting that metabolism may be unrelated to heart failure or that the treatment was
not initiated in time. Interestingly, when ZDF rats were treated
with a PPAR-␥ agonist at 6 wk of age [when insulin resistance
compared with 6-wk-old db/db mice is milder (37)], improvements of cardiac metabolism and heart function were observed
(272). In a different study, changing cardiac metabolic profile
by treating ZDF rats with PPAR-␥ agonist also improved
contractile function (91). Additionally, overexpression of
GLUT4 in db/db mice not only normalized cardiac metabolism
but also improved heart function (21, 218). These studies
suggest that acute changes in metabolism likely induce early
and reversible damage to cardiac tissue. Even though these
early alterations are inadequate to produce cardiac functional
changes, early interventions would be favored to provide
protection against development of cardiomyopathy in the later
stages of the disease.
The role of abnormal cardiac metabolism in cardiac dysfunction is also supported by studies using transgenic mice.
Overexpression of cardiac PPAR-␣ increased FA uptake and
oxidation (79). The hearts from these transgenic mice exhibit a
metabolic phenotype similar to diabetic hearts (79). Measurement of heart function revealed systolic dysfunction and ventricular hypertrophy in these hearts, indicating that, in the
absence of systemic metabolic disturbances, alteration of cardiac metabolism is sufficient to induce cardiac contractile
dysfunction (79). This concept is further substantiated by
transgenic mice overexpressing cardiac ACS, FATP1, or LPL
(47, 48, 260). These transgenic mice have shown increased FA
uptake, utilization, or lipid accumulation, and these changes
correlated well with contractile dysfunction.
In contrast, several studies have reported that altered metabolism has no impact on contractile function (224, 250). It
should be noted that, in these studies, heart function was
evaluated using a Langendorff heart perfusion. Given that
diabetic hearts exhibit reduced mitochondrial oxidative capacity and cardiac efficiency (28, 106), it is possible that with
increased workload and energy requirement, cardiac function
would be compromised.
In addition to heart dysfunction per se, diabetic patients also
have a greater incidence and severity of angina and acute
myocardial infarction (147). After a myocardial infarction,
diabetic patients have almost twice the rate of mortality compared with nondiabetics (229). Alterations in myocardial energy metabolism during diabetes are probably an important
contributing factor in explaining this increased susceptibility to
ischemic damage. In obese or diabetic animals, increased
ischemic damage is also observed (5, 60, 93, 147), and normalization of cardiac metabolism in hearts from these animals
has been shown to improve functional recovery following
ischemia-reperfusion (147, 224, 270). Interestingly, a number
of studies have also reported contradictory results, with decreased susceptibility to ischemia-reperfusion damage being
reported in STZ or ZDF hearts (145, 150, 236, 249). Although
the mechanisms for this observation are still unclear, decreased
glycolysis and glycolytic products, a lower Na⫹-Ca2⫹ activity,
and a decreased clearance of protons via the Na⫹/H⫹ exchanger have been proposed to explain this inconsistency (75).
Invited Review
ABNORMAL REGULATION OF CARDIAC SUBSTRATE METABOLISM IN DIABETES
H1497
Although changes in metabolism have been implicated in
diabetic cardiac dysfunction, the mechanisms responsible are
still unclear. Several factors have been proposed, and these
include changes in Ca2⫹ homeostasis, decreased cardiac efficiency, lipotoxicity, and myocardial mitochondrial damage
(Fig. 4).
been documented (215, 271). Irrespective of the mechanism,
suppression of SERCA2a and the Na⫹/Ca2⫹ exchanger results
in poor calcium handling associated with impaired heart function (6, 22, 49). Interestingly, overexpression of SERCA2a in
the diabetic heart improved Ca2⫹ handling (243) and cardiac
function (239).
Effects on Ca2⫹ Homeostasis
Decreased Cardiac Efficiency
When compared with glucose, oxidation of FA consumes
more oxygen (2.58 vs. 2.33 ATP/oxygen atom). Cardiac efficiency, the ratio of cardiac work to myocardial oxygen consumption, changes with the type of substrate. Thus, in perfused
hearts, provision of FA decreases cardiac efficiency compared
with when glucose is the sole substrate (31, 162, 245). Additionally, decreased cardiac efficiency is also observed in human or experimental animals with obesity or diabetes (106,
162, 191). This reduction in cardiac efficiency and increased
oxygen demand makes the heart especially vulnerable to damage following increased workload or ischemia. Interestingly,
even though oxidation of FA requires 10% more oxygen
compared with glucose, cardiac oxygen consumption is over
30% higher in ob/ob or db/db mice compared with control
hearts, suggesting that other mechanisms also contribute to
higher oxygen consumption and lower cardiac efficiency (37).
A recent study reported oxygen wasting for noncontractile
purposes in the diabetic heart (106). The mechanisms for this
oxygen wasting are unknown, and uncoupled proteins (UCPs)
are suggested as a potential target. In hearts from STZ-induced
diabetic rats and ob/ob and db/db mice, augmented gene
expression or protein levels of UCP2 or UCP3 have been
reported (30, 102, 170, 171, 268). Other studies demonstrate no
association between change in cardiac UCP protein levels and
the onset of cardiac inefficiency (28, 30). Hence, additional
studies are required to identify the targets that cause oxygen
wasting observed in hearts from obese and diabetic animals.
Lipotoxicity
Fig. 4. Consequences of changes in cardiac metabolism during diabetes. After
insulin resistance or diabetes, augmented cardiac FA uptake promotes FA
utilization and storage, which contributes toward a reduction in glycolysis and
glucose oxidation. As ATP generated through glycolysis is preferentially used
by ion transporters (like Ca2⫹-ATPase), inhibition of cardiac glycolysis impairs intracellular Ca2⫹ homeostasis. High rate of FA oxidation increases
reactive oxygen species (ROS) generation, along with augmented lipid storage,
leading to lipotoxicity and mitochondrial dysfunction. High rate of FA oxidation also increases oxygen demand and reduces cardiac efficiency. All of these
changes can eventually contribute to the development of diabetic cardiomyopathy.
AJP-Heart Circ Physiol • VOL
A number of studies have suggested that excessive FA
overload induces lipotoxicity and contributes to the initiation
and development of cardiomyopathy (194, 272). With the use
of transgenic mice, studies have shown that elevation of FA
uptake or utilization induces lipotoxicity in the absence of any
systemic metabolic disturbance. Cardiac-specific overexpression of LPL or FATP1 significantly increased FA delivery with
ensuing lipid storage, lipotoxic cardiomyopathy, and contractile dysfunction (47, 260). Moreover, elevating FA utilization
by cardiac-specific overexpression of PPAR-␣ or ACS also
causes cardiomyopathy and cardiac dysfunction, similar to that
seen during diabetes (48, 79). Conversely, reducing FA supply
or utilization prevented the development of cardiomyopathy in
obese or diabetic animals. In ZDF or transgenic mice with
cardiac overexpression of LPL, a PPAR-␥ agonist decreased
plasma and cardiac intracellular lipids and ameliorated cardiomyopathy (244, 272). A recent study also demonstrated that
increasing lipoprotein excretion by overexpressing human
apoB reduced cardiac lipids and improved cardiomyopathy
(264). Taken together, these studies provide convincing evidence that augmented FA supply during obesity or diabetes
impairs cardiac lipid homeostasis and leads to lipotoxicity
cardiomyopathy.
291 • OCTOBER 2006 •
www.ajpheart.org
Downloaded from http://ajpheart.physiology.org/ by 10.220.33.6 on May 5, 2017
Within the cell, enzymes that catalyze glycolysis are located
close to the sarcoplasmic reticulum and sarcolemma (71, 251),
and ATP generated through glycolysis is preferentially used by
ion transporters [like Ca2⫹-ATPase (SERCA2a) and Na⫹-K⫹ATPase] in these membrane fractions (71, 228, 252). Thus
inhibition of cardiac glycolysis by high rates of FA oxidation
during obesity and diabetes may impair intracellular Ca2⫹
homeostasis, a defect that has been proposed to contribute
toward the development of cardiomyopathy. Altered Ca2⫹
homeostasis can also result from glycolysis-independent mechanisms. For example, decreased cardiac expression of
SERCA2a or Na⫹/Ca2⫹ exchanger have been observed in
Type 1 (99, 126) and Type 2 diabetic animals (22). Although
the mechanisms for this reduction in gene expression are
unclear, several studies have suggested that accumulation of
glucose metabolites due to dissociation of glucose influx and
pyruvate oxidation plays a role (51). Conversely, decreasing
glucose metabolites has been shown to prevent the decrease in
SERCA2a following diabetes (207). In other studies, a decrease in the activity of SERCA2a and Na⫹/Ca2⫹ exchanger,
without any change in expression or protein levels, have also
Invited Review
H1498
ABNORMAL REGULATION OF CARDIAC SUBSTRATE METABOLISM IN DIABETES
Hyperglycemia
In addition to hyperlipidemia-induced lipotoxicity, hyperglycemia also provokes glucotoxicity, which could contribute
to cardiac tissue injury. Hyperglycemia promotes the producAJP-Heart Circ Physiol • VOL
tion of reactive oxygen and nitrogen species, leading to cytochrome c-mediated caspase 3 activation and myocardial apoptosis (32). Hyperglycemia-induced oxidative stress also activates poly(ADP-ribose) polymerase-1 (PARP) (64). Mild
activation of PARP regulates multiple cellular reactions such
as DNA repair, gene expression, and cell survival (246).
However, overactivation of PARP could initiate a series of
cellular processes, leading to cellular damage. Through depletion of NAD⫹, an ATP-consuming process, PARP may cause
energy deficit and cell death (69). PARP, through inhibition of
glyceraldehyde phosphate dehydrogenase (GAPDH), diverts
glucose from glycolytic pathways into alternative fates, including advanced glycation end product (AGE) formation, hexosamine, polyol pathway flux, and protein kinase C (PKC)
activation, which are believed to mediate hyperglycemia-induced cardiac tissue damage (64, 196). In this regard, increased
formation of AGE forms irreversible cross-links with many
macromolecules such as collagen, leading to tissue fibrosis,
inactivation of SERCA2a, and ryanodine receptor calciumrelease channel, impaired cardiac relaxation, contractility, and
ventricular stiffness (26, 27, 35, 54). Increased hexosamine
influx reduces SERCA2a protein levels, resulting in prolonged
calcium transients and delayed myocardial relaxation (51).
Increased hexosamine influx also impairs insulin signal transduction and contributes to insulin resistance (163). Increased
polyol flux is associated with a decrease in intracellular glutathione levels and an augmentation in cell apoptosis (83).
Conversely, inhibition of polyol flux protects hearts from
ischemic injury (125, 198). Finally, activation of PKC-␤2 leads
to left ventricular hypertrophy, cardiac myocyte necrosis, multifocal fibrosis, and decreased left ventricular performance,
resulting in cardiomyopathy (247). Taken together, hyperglycemia, through multiple pathways, causes cardiac cellular and
functional changes, possibly contributing to the development
of cardiomyopathy.
Mitochondrial Dysfunction
Mitochondria are the primary source of energy generation
within cells. Acetyl-CoA generated from FA ␤-oxidation or
glycolysis is used by the tricarboxylic acid cycle for production
of NADH and FADH2. These electron carriers transfer electrons to the mitochondria electron transport chain, where ATP
is ultimately generated. Thus cardiac mitochondrial dysfunction is expected to induce deleterious cellular effects, ultimately leading to heart disease. Indeed, humans with inherited
or acquired mitochondrial defects develop cardiomyopathy
(112, 208). With the use of transgenic mice models, mitochondrial dysfunction is also associated with heart disease. For
instance, knock out of Ant1 (which controls exchange of
mitochondrial ATP with cytosolic ADP) or Tfam (a mitochondrial transcription factor that regulates mitochondrial biosynthesis and gene expression) results in a cardiac energy deficit
and cardiomyopathy (92, 142). Mice with cardiac-specific
deletion of mitochondrial genes involved in FA oxidation also
display a cardiomyopathic phenotype (72, 135). Interestingly,
the opposite is also true with overexpression of PPAR-␣
(which regulates the expression of mitochondrial enzymes
involved in FA oxidation) (79) or PGC1 (which controls
mitochondrial biosynthesis) (139) resulting in mitochondrial
dysfunction and cardiomyopathy. In Type 1 or Type 2 diabetic
291 • OCTOBER 2006 •
www.ajpheart.org
Downloaded from http://ajpheart.physiology.org/ by 10.220.33.6 on May 5, 2017
Although limited evidence is available of the existence of
lipotoxicity in human diabetic patients, some studies have
documented similarities between rodent lipotoxicity and the
human metabolic syndrome (240). In a subgroup of patients
with heart failure, Sharma et al. (220) have identified severe
metabolic dysregulation characterized by intracellular TG accumulation and alterations in gene expression, similar to the
lipotoxic rat heart. Increased apoptosis was also detected in
ventricular biopsies from Type 2 diabetic patients (82). Additionally, cardiac PCr-to-ATP ratio was decreased in Type 2
diabetic patients and correlated well with plasma FA concentrations (216).
The mechanisms that mediate cardiac lipotoxicity are still
not completely understood. One potential target is over production of ROS (79). High rate of FA oxidation increases
mitochondrial action potential, leading to augmented ROS
generation. Under normal physiological conditions, ROS is
removed by cellular antioxidants. In the event of excessive
generation of ROS, as observed in STZ-induced diabetic rats,
ZDF rats, and db/db mice (19, 32, 272), it causes cardiomyocyte cell damage and augmented apoptosis. Another potential
mechanism for lipotoxicity is accumulation of lipids, when FA
uptake supercedes its oxidation. Regarding accumulation of
TG, the role of this neutral lipid in inducing contractile dysfunction is still unknown, although a strong association between TG storage and lipotoxicity has been established in both
animal models and human studies (50, 220, 272). Interestingly,
a recent study suggested that TG formation actually provides
protection against the deleterious effects of fatty acyl-CoA
(144). Besides TG, excessive FA also leads to ceramide generation, an intracellular messenger known to trigger apoptosis
(272). Accumulation of ceramide has been found in ZDF rats
(272) or in isolated cardiomyocytes incubated with high fat
(67, 101). Ceramide upregulates inducible nitric oxide synthase
through activation of NF-␬B, leading to increased generation
of nitric oxide and peroxynitrite (241, 272). As a highly
reactive molecule, peroxynitrite causes opening of the mitochondrial permeability transition pore and release of cytochrome c. Additionally, ceramide directly interacts with cytochrome c, leading to its release from the mitochondria (87). As
a consequence, caspase is activated, which initiates the apoptotic pathway in cells. Additionally, with the use of isolated
cardiomyocytes, high FA impairs cardiolipin and leads to
ceramide-independent cell apoptosis (184). Finally, accumulation of FA metabolites has also been associated with insulin
resistance. FA derivatives, such as fatty acyl CoA, diacylglycerol, or ceramide may activate a serine kinase cascade, which
involves protein kinase C-␪ and I␬B kinase-␤ (inhibitor of
NF-␬B kinase-␤), leading to serine phosphorylation of IRS
(44, 114, 128, 129, 269). Following this, tyrosine phosphorylation of IRS and its ability to activate phosphatidylinositol
3-kinase and protein kinase B are compromised (206). Future
studies are required to substantiate the roles of these FA
derivatives in the development of insulin resistance, especially
in the heart.
Invited Review
ABNORMAL REGULATION OF CARDIAC SUBSTRATE METABOLISM IN DIABETES
Summary
In this review, we document that both insulin resistance and
Type 1 and Type 2 diabetes exhibit similar alterations in
cardiac metabolism through extrinsic and intrinsic mechanisms. The predominant change is a suppression of cardiac
glucose utilization and a switch to excessive FA utilization
associated with lipid storage. It should be noted that, in the
majority of these studies, metabolic outcomes were obtained
using ex vivo perfused working hearts. Additionally, only two
substrates (glucose and fatty acid) were present in the perfusion
buffer, and lactate, lipoproteins, ketone bodies, and relevant
hormones were excluded. Hence, cardiac metabolism measured through this method may not reflect the in vivo situation.
Despite these drawbacks, many studies have documented a
strong correlation between altered cardiac metabolism and
cardiac dysfunction observed during insulin resistance or diabetes. This correlation was further substantiated using transgenic mice with lipid oversupply in the absence of systemic
interference. Regarding heart function, both diastolic and systolic dysfunction are observed in Type 1 diabetic animals. This
is likely a consequence of prolonged hypoinsulinemia and
hyperglycemia. With Type 2 diabetic models, there are conflicting reports, with some studies documenting both impaired
systolic and diastolic function, whereas other studies argue
against a change in systolic function. It is possible with Type
2 diabetic models, this discrepancy could be due to the severity
or duration of hyperinsulinemia, and eventual hyperglycemia,
or the techniques used to measure cardiac function. Finally,
during diabetes, changes in cardiac metabolism occur early and
precede the development of cardiomyopathy. Even though
altered metabolism is inadequate to produce cardiac functional
changes at this early time, it is likely that early metabolic
damage is occurring at the cellular or subcellular levels. Overtime, these cumulative defects could be contributing to diabetic
cardiomyopathy. Identification of this early damage could
AJP-Heart Circ Physiol • VOL
facilitate proper interventions and provide protection against
development of diabetic cardiomyopathy in the later stages of
the disease.
GRANTS
This work was supported by operating grants from Heart and Stroke
Foundation of BC and Yukon and Canadian Diabetes Association.
REFERENCES
1. Aalto-Setala K, Fisher EA, Chen X, Chajek-Shaul T, Hayek T,
Zechner R, Walsh A, Ramakrishnan R, Ginsberg HN, and Breslow
JL. Mechanism of hypertriglyceridemia in human apolipoprotein (apo)
C.III transgenic mice Diminished very low density lipoprotein fractional
catabolic rate associated with increased apo CIII and reduced apo E on
the particles. J Clin Invest 90: 1889 –1900, 1992.
2. Aalto-Setala K, Weinstock PH, Bisgaier CL, Wu L, Smith JD, and
Breslow JL. Further characterization of the metabolic properties of
triglyceride-rich lipoproteins from human and mouse apoC-I.I.I. transgenic mice. J Lipid Res 37: 1802–1811, 1996.
3. Aasum E, Belke DD, Severson DL, Riemersma RA, Cooper M,
Andreassen M, and Larsen TS. Cardiac function and metabolism in
Type 2 diabetic mice after treatment with B.M 170744, a novel PPARalpha activator. Am J Physiol Heart Circ Physiol 283: H949 –H957,
2002.
4. Aasum E, Cooper M, Severson DL, and Larsen TS. Effect of BM 17
0744, a PPARalpha ligand, on the metabolism of perfused hearts from
control and diabetic mice. Can J Physiol Pharmacol 83: 183–190, 2005.
5. Aasum E, Hafstad AD, Severson DL, and Larsen TS. Age-dependent
changes in metabolism, contractile function, and ischemic sensitivity in
hearts from db/db mice. Diabetes 52: 434 – 441, 2003.
6. Abe T, Ohga Y, Tabayashi N, Kobayashi S, Sakata S, Misawa H,
Tsuji T, Kohzuki H, Suga H, Taniguchi S, and Takaki M. Left
ventricular diastolic dysfunction in type 2 diabetes mellitus model rats.
Am J Physiol Heart Circ Physiol 282: H138 –H148, 2002.
7. Abumrad NA, Park JH, and Park CR. Permeation of long-chain fatty
acid into adipocytes. Kinetics, specificity, and evidence for involvement
of a membrane protein. J Biol Chem 259: 8945– 8953, 1984.
8. An D, Pulinilkunnil T, Qi D, Ghosh S, Abrahani A, and Rodrigues B.
The metabolic “switch” AMPK regulates cardiac heparin-releasable
lipoprotein lipase. Am J Physiol Endocrinol Metab 288: E246 –E253,
2005.
9. Anguera I, Magrina J, Setoain FJ, Esmatges E, Pare C, Vidal J,
Azqueta M, Garcia A, Grau JM, Vidal-Sicart S, and Betriu A.
Anatomopathological bases of latent ventricular dysfunction in insulindependent diabetics. Rev Esp Cardiol 51: 43–50, 1998.
10. Armoni M, Harel C, Bar-Yoseph F, Milo S, and Karnieli E. Free fatty
acids repress the GLUT4 gene expression in cardiac muscle via novel
response elements. J Biol Chem 280: 34786 –34795, 2005.
11. Atkinson LL, Fischer MA, and Lopaschuk GD. Leptin activates
cardiac fatty acid oxidation independent of changes in the AMP-activated
protein kinase-acetyl-CoA carboxylase-malonyl-CoA axis. J Biol Chem
277: 29424 –29430, 2002.
12. Augustus A, Yagyu H, Haemmerle G, Bensadoun A, Vikramadithyan RK, Park SY, Kim JK, Zechner R, and Goldberg IJ.
Cardiac-specific knock-out of lipoprotein lipase alters plasma lipoprotein
triglyceride metabolism and cardiac gene expression. J Biol Chem 279:
25050 –25057, 2004.
13. Augustus AS, Buchanan J, Park TS, Hirata K, Noh HL, Sun J,
Homma S, D’Armiento J, Abel ED, and Goldberg IJ. Loss of
lipoprotein lipase-derived fatty acids leads to increased cardiac glucose
metabolism and heart dysfunction. J Biol Chem 281: 8716 – 8723, 2006.
14. Augustus AS, Kako Y, Yagyu H, and Goldberg IJ. Routes of FA
delivery to cardiac muscle: modulation of lipoprotein lipolysis alters
uptake of TG-derived FA. Am J Physiol Endocrinol Metab 284: E331–
E339, 2003.
15. Avogaro A, Nosadini R, Doria A, Fioretto P, Velussi M, Vigorito C,
Sacca L, Toffolo G, Cobelli C, Trevisan R, E. Duner, R. Razzolini, F.
Rengo, and G. Crepaldi. Myocardial metabolism in insulin-deficient
diabetic humans without coronary artery disease. Am J Physiol Endocrinol Metab 258: E606 –E618, 1990.
16. Barak Y, Nelson MC, Ong ES, Jones YZ, Ruiz-Lozano P, Chien KR,
Koder A, and Evans RM. PPARgamma is required for placental,
cardiac, and adipose tissue development. Mol Cell 4: 585–595, 1999.
291 • OCTOBER 2006 •
www.ajpheart.org
Downloaded from http://ajpheart.physiology.org/ by 10.220.33.6 on May 5, 2017
rats, impaired mitochondrial function has been demonstrated
(222, 262), an observation similar to hearts from transgenic
mice. Given the similarity between transgenic mice and models
of diabetes, these studies strongly suggest that impairment of
mitochondrial function contributes to the development of cardiomyopathy.
The mechanisms by which mitochondrial dysfunction contributes toward cardiomyopathy are still unclear. One potential
target is ROS. Augmented ROS generation following high FA
oxidation induces oxidative stress and cell damage. Inhibition
of ROS by overexpression of metallothionein, MnSOD, or
catalase has been shown to protect against mitochondrial dysfunction and cardiomyopathy (221, 262, 263). Another target is
ceramide, a sphingolipid that has been shown to accumulate
following FA overload. Ceramide, through inhibition of mitochondrial respiratory chain, or inducing cytochrome c release
from mitochondria and apoptosis, is suggested to provoke the
development of cardiomyopathy (67, 95). Augmented expression of PGC1 in hearts from obese or diabetic animals may
increase mitochondrial proliferation, mitochondrial ultrastructural abnormalities and dysfunction, and ultimately cardiomyopathy (139). Finally, following end stage diabetes, dysfunction of mitochondria causes energetic starvation, leading to
heart failure (208).
H1499
Invited Review
H1500
ABNORMAL REGULATION OF CARDIAC SUBSTRATE METABOLISM IN DIABETES
AJP-Heart Circ Physiol • VOL
37. Carley AN and Severson DL. Fatty acid metabolism is enhanced in type
2 diabetic hearts. Biochim Biophys Acta 1734: 112–126, 2005.
38. Carroll R, Carley AN, Dyck JR, and Severson DL. Metabolic effects
of insulin on cardiomyocytes from control and diabetic db/db mouse
hearts. Am J Physiol Endocrinol Metab 288: E900 –E906, 2005.
39. Chatham JC. Lactate–the forgotten fuel! J Physiol 542: 333, 2002.
40. Chatham JC and Forder JR. Relationship between cardiac function
and substrate oxidation in hearts of diabetic rats. Am J Physiol Heart Circ
Physiol 273: H52–H58, 1997.
41. Chatham JC, Gao ZP, Bonen A, and Forder JR. Preferential inhibition of lactate oxidation relative to glucose oxidation in the rat heart
following diabetes. Cardiovasc Res 43: 96 –106, 1999.
42. Chatham JC, Gao ZP, and Forder JR. Impact of 1 wk of diabetes on
the regulation of myocardial carbohydrate and fatty acid oxidation. Am J
Physiol Endocrinol Metab 277: E342–E351, 1999.
43. Chatham JC and Seymour AM. Cardiac carbohydrate metabolism in
Zucker diabetic fatty rats. Cardiovasc Res 55: 104 –112, 2002.
44. Chavez JA, Knotts TA, Wang LP, Li G, Dobrowsky RT, Florant GL,
and Summers SAA. Role for ceramide, but not diacylglycerol, in the
antagonism of insulin signal transduction by saturated fatty acids. J Biol
Chem 278: 10297–10303, 2003.
45. Chawla A, Lee CH, Barak Y, He W, Rosenfeld J, Liao D, Han J,
Kang H, and Evans RM. PPARdelta is a very low-density lipoprotein
sensor in macrophages. Proc Natl Acad Sci USA 100: 1268 –1273, 2003.
46. Cheng L, Ding G, Qin Q, Huang Y, Lewis W, He N, Evans RM,
Schneider MD, Brako FA, Xiao Y, Chen YE, and Yang Q. Cardiomyocyte-restricted peroxisome proliferator-activated receptor-delta deletion perturbs myocardial fatty acid oxidation and leads to cardiomyopathy. Nat Med 10: 1245–1250, 2004.
47. 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, and Schaffer JE.
Transgenic expression of fatty acid transport protein 1 in the heart causes
lipotoxic cardiomyopathy. Circ Res 96: 225–233, 2005.
48. Chiu HC, Kovacs A, Ford DA, Hsu FF, Garcia R, Herrero P, Saffitz
JE, and Schaffer JEA. novel mouse model of lipotoxic cardiomyopathy.
J Clin Invest 107: 813– 822, 2001.
49. Choi KM, Zhong Y, Hoit BD, Grupp IL, Hahn H, Dilly KW,
Guatimosim S, Lederer WJ, and Matlib MA. Defective intracellular
Ca2⫹ signaling contributes to cardiomyopathy in Type 1 diabetic rats.
Am J Physiol Heart Circ Physiol 283: H1398 –H1408, 2002.
50. Christoffersen C, Bollano E, Lindegaard ML, Bartels ED, Goetze JP,
Andersen CB, and Nielsen LB. Cardiac lipid accumulation associated
with diastolic dysfunction in obese mice. Endocrinology 144: 3483–
3490, 2003.
51. Clark RJ, McDonough PM, Swanson E, Trost SU, Suzuki M,
Fukuda M, and Dillmann WH. Diabetes and the accompanying hyperglycemia impairs cardiomyocyte calcium cycling through increased nuclear O-GlcNAcylation. J Biol Chem 278: 44230 – 44237, 2003.
52. Clerk LH, Rattigan S, and Clark MG. Lipid infusion impairs physiologic insulin-mediated capillary recruitment and muscle glucose uptake
in vivo. Diabetes 51: 1138 –1145, 2002.
53. Coe NR, Smith AJ, Frohnert BI, Watkins PA, and Bernlohr DA. The
fatty acid transport protein (FATP1) is a very long chain acyl-CoA
synthetase. J Biol Chem 274: 36300 –36304, 1999.
54. Cooper ME. Importance of advanced glycation end products in diabetesassociated cardiovascular, and renal disease. Am J Hypertens 17: 31S–
38S, 2004.
55. Coort SL, Hasselbaink DM, Koonen DP, Willems J, Coumans WA,
Chabowski A, van der Vusse GJ, Bonen A, Glatz JF, and Luiken JJ.
Enhanced sarcolemmal FAT/CD36 content and triacylglycerol storage in
cardiac myocytes from obese zucker rats. Diabetes 53: 1655–1663, 2004.
56. Coort SL, Luiken JJ, van der Vusse GJ, Bonen A, and Glatz JF.
Increased FAT (fatty acid translocase)/CD36-mediated long-chain fatty
acid uptake in cardiac myocytes from obese Zucker rats. Biochem Soc
Trans 32: 83– 85, 2004.
57. Depre C, Veitch K, and Hue L. Role of fructose 2,6-bisphosphate in the
control of glycolysis. Stimulation of glycogen synthesis by lactate in the
isolated working rat heart. Acta Cardiol 48: 147–164, 1993.
58. Deprez J, Vertommen D, Alessi DR, Hue L, and Rider MH. Phosphorylation and activation of heart 6-phosphofructo-2-kinase by protein
kinase B. and other protein kinases of the insulin signaling cascades.
J Biol Chem 272: 17269 –17275, 1997.
291 • OCTOBER 2006 •
www.ajpheart.org
Downloaded from http://ajpheart.physiology.org/ by 10.220.33.6 on May 5, 2017
17. Barger PM and Kelly DP. PPARsignaling in the control of cardiac
energy metabolism. Trends Cardiovasc Med 10: 238 –245, 2000.
18. Barouch LA, Berkowitz DE, Harrison RW, O’Donnell CP, and Hare
JM. Disruption of leptin signaling contributes to cardiac hypertrophy
independently of body weight in mice. Circulation 108: 754 –759, 2003.
19. Barouch LA, Gao D, Chen L, Miller KL, Xu W, Phan AC, Kittleson
MM, Minhas KM, Berkowitz DE, Wei C, and Hare JM. Cardiac
myocyte apoptosis is associated with increased DNA damage and decreased survival in murine models of obesity. Circ Res 98: 119 –124,
2006.
20. Barr RL and Lopaschuk GD. Methodology for measuring in vitro/ex
vivo cardiac energy metabolism. J Pharmacol Toxicol Methods 43:
141–152, 2000.
21. Belke DD, Larsen TS, Gibbs EM, and Severson DL. Altered metabolism causes cardiac dysfunction in perfused hearts from diabetic (db/db)
mice. Am J Physiol Endocrinol Metab 279: E1104 –E1113, 2000.
22. Belke DD, Swanson EA, and Dillmann WH. Decreased sarcoplasmic
reticulum activity and contractility in diabetic db/db mouse heart. Diabetes 53: 3201–3208, 2004.
23. Bell DS. Diabetic cardiomyopathy. A unique entity or a complication of
coronary artery disease? Diabetes Care 18: 708 –714, 1995.
24. Bertoni AG, Tsai A, Kasper EK, and Brancati FL. Diabetes and
idiopathic cardiomyopathy: a nationwide case-control study. Diabetes
Care 26: 2791–2795, 2003.
25. Bertrand L, Alessi DR, Deprez J, Deak M, Viaene E, Rider MH, and
Hue L. Heart 6-phosphofructo-2-kinase activation by insulin results from
Ser-466 and Ser-483 phosphorylation and requires 3-phosphoinositidedependent kinase-1, but not protein kinase B. J Biol Chem 274: 30927–
30933, 1999.
26. Bidasee KR, Nallani K, Yu Y, Cocklin RR, Zhang Y, Wang M,
Dincer UD, and Besch HR Jr. Chronic diabetes increases advanced
glycation end products on cardiac ryanodine receptors/calcium-release
channels. Diabetes 52: 1825–1836, 2003.
27. Bidasee KR, Zhang Y, Shao CH, Wang M, Patel KP, Dincer UD, and
Besch HR Jr. Diabetes increases formation of advanced glycation end
products on Sarco(endo)plasmic reticulum Ca2⫹-ATPase. Diabetes 53:
463– 473, 2004.
28. Boudina S, Sena S, O’Neill BT, Tathireddy P, Young ME, and Abel
ED. Reduced mitochondrial oxidative capacity and increased mitochondrial uncoupling impair myocardial energetics in obesity. Circulation
112: 2686 –2695, 2005.
29. Bowker-Kinley MM, Davis WI, Wu P, Harris RA, and Popov KM.
Evidence for existence of tissue-specific regulation of the mammalian
pyruvate dehydrogenase complex. Biochem J 329: 191–196, 1998.
30. Buchanan J, Mazumder PK, Hu P, Chakrabarti G, Roberts MW,
Yun UJ, Cooksey RC, Litwin SE, and Abel ED. Reduced cardiac
efficiency and altered substrate metabolism precedes the onset of hyperglycemia and contractile dysfunction in two mouse models of insulin
resistance and obesity. Endocrinology 146: 5341–5349, 2005.
31. Burkhoff D, Weiss RG, Schulman SP, Kalil-Filho R, Wannenburg T,
and Gerstenblith G. Influence of metabolic substrate on rat heart
function and metabolism at different coronary flows. Am J Physiol Heart
Circ Physiol 261: H741–H750, 1991.
32. Cai L, Li W, Wang G, Guo L, Jiang Y, and Kang YJ. Hyperglycemiainduced apoptosis in mouse myocardium: mitochondrial cytochrome
C-mediated caspase-3 activation pathway. Diabetes 51: 1938 –1948,
2002.
33. Camps L, Reina M, Llobera M, Vilaro S, and Olivecrona T. Lipoprotein lipase: cellular origin and functional distribution. Am J Physiol Cell
Physiol 258: C673–C681, 1990.
34. Camps M, Castello A, Munoz P, Monfar M, Testar X, Palacin M,
and Zorzano A. Effect of diabetes and fasting on GLUT-4 (muscle/fat)
glucose-transporter expression in insulin-sensitive tissues Heterogeneous
response in heart, red and white muscle. Biochem J 282: 765–772, 1992.
35. Candido R, Forbes JM, Thomas MC, Thallas V, Dean RG, Burns
WC, Tikellis C, Ritchie RH, Twigg SM, Cooper ME, and Burrell
LMA. breaker of advanced glycation end products attenuates diabetesinduced myocardial structural changes. Circ Res 92: 785–792, 2003.
36. Carley AN, Semeniuk LM, Shimoni Y, Aasum E, Larsen TS, Berger
JP, and Severson DL. Treatment of type 2 diabetic db/db mice with a
novel PPARgamma agonist improves cardiac metabolism but not contractile function. Am J Physiol Endocrinol Metab 286: E449 –E455,
2004.
Invited Review
ABNORMAL REGULATION OF CARDIAC SUBSTRATE METABOLISM IN DIABETES
AJP-Heart Circ Physiol • VOL
79.
80.
81.
82.
83.
84.
85.
86.
87.
88.
89.
90.
91.
92.
93.
94.
95.
96.
97.
98.
99.
metabolism in the normal and diseased heart. J Mol Cell Cardiol 34:
1249 –1257, 2002.
Finck BN, Lehman JJ, Leone TC, Welch MJ, Bennett MJ, Kovacs A,
Han X, Gross RW, Kozak R, Lopaschuk GD, and Kelly DP. The
cardiac phenotype induced by PPARalpha overexpression mimics that
caused by diabetes mellitus. J Clin Invest 109: 121–130, 2002.
Forman BM, Chen J, and Evans RM. Hypolipidemic drugs, polyunsaturated fatty acids, and eicosanoids are ligands for peroxisome proliferator-activated receptors alpha and delta. Proc Natl Acad Sci USA 94:
4312– 4317, 1997.
Francis GS. Diabetic cardiomyopathy: fact or fiction? Heart 85: 247–
248, 2001.
Frustaci A, Kajstura J, Chimenti C, Jakoniuk I, Leri A, Maseri A,
Nadal-Ginard B, and Anversa P. Myocardial cell death in human
diabetes. Circ Res 87: 1123–1132, 2000.
Galvez AS, Ulloa JA, Chiong M, Criollo A, Eisner V, Barros LF, and
Lavandero S. Aldose reductase induced by hyperosmotic stress mediates
cardiomyocyte apoptosis: differential effects of sorbitol and mannitol.
J Biol Chem 278: 38484 –38494, 2003.
Gargiulo CE, Stuhlsatz-Krouper SM, and Schaffer JE. Localization
of adipocyte long-chain fatty acyl-CoA synthetase at the plasma membrane. J Lipid Res 40: 881– 892, 1999.
Garland PB, Randle PJ, and Newsholme EA. Citrate as an intermediary in the inhibition of phosphofructokinase in rat heart muscle by fatty
acids, ketone bodies, pyruvate, diabetes, and starvation. Nature 200:
169 –170, 1963.
Gertz EW, Wisneski JA, Stanley WC, and Neese RA. Myocardial
substrate utilization during exercise in humans. Dual carbon-labeled
carbohydrate isotope experiments. J Clin Invest 82: 2017–2025, 1988.
Ghafourifar P, Klein SD, Schucht O, Schenk U, Pruschy M, Rocha S,
and Richter C. Ceramide induces cytochrome c release from isolated
mitochondria. Importance of mitochondrial redox state. J Biol Chem 274:
6080 – 6084, 1999.
Ghosh S, An D, Pulinilkunnil T, Qi D, Lau HC, Abrahani A, Innis
SM, and Rodrigues B. Role of dietary fatty acids and acute hyperglycemia in modulating cardiac cell death. Nutrition 20: 916 –923, 2004.
Gilde AJ, van der Lee KA, Willemsen PH, Chinetti G, van der Leij
FR, van der Vusse GJ, Staels B, and van Bilsen M. Peroxisome
proliferator-activated receptor (PPAR) alpha and PPARbeta/delta, but
not PPARgamma, modulate the expression of genes involved in cardiac
lipid metabolism. Circ Res 92: 518 –524, 2003.
Golfman LS, Takeda N, and Dhalla NS. Cardiac membrane Ca(2⫹)transport in alloxan-induced diabetes in rats. Diabetes Res Clin Pract 31,
Suppl: S73–S77, 1996.
Golfman LS, Wilson CR, Sharma S, Burgmaier M, Young ME,
Guthrie PH, Van Arsdall M, Adrogue JV, Brown KK, and Taegtmeyer H. Activation of PPARgamma enhances myocardial glucose
oxidation and improves contractile function in isolated working hearts of
ZDF rats. Am J Physiol Endocrinol Metab 289: E328 –E336, 2005.
Graham BH, Waymire KG, Cottrell B, Trounce IA, MacGregor GR,
and Wallace DCA. Mouse model for mitochondrial myopathy and
cardiomyopathy resulting from a deficiency in the heart/muscle isoform
of the adenine nucleotide translocator. Nat Genet 16: 226 –234, 1997.
Greer JJ, Ware DP, and Lefer DJ. Myocardial infarction and heart
failure in the db/db diabetic mouse. Am J Physiol Heart Circ Physiol 290:
H146 –H153, 2006.
Grundleger ML and Thenen SW. Decreased insulin binding, glucose
transport, and glucose metabolism in soleus muscle of rats fed a high fat
diet. Diabetes 31: 232–237, 1982.
Gudz TI, Tserng KY, and Hoppel CL. Direct inhibition of mitochondrial respiratory chain complex III by cell-permeable ceramide. J Biol
Chem 272: 24154 –24158, 1997.
Hardie DG. Minireview: the AMPactivated protein kinase cascade: the
key sensor of cellular energy status. Endocrinology 144: 5179 –5183,
2003.
Hardie DG and Carling D. The AMPactivated protein kinase–fuel
gauge of the mammalian cell? Eur J Biochem 246: 259 –273, 1997.
Hasselbaink DM, Glatz JF, Luiken JJ, Roemen TH, and Van der
Vusse GJ. Ketone bodies disturb fatty acid handling in isolated cardiomyocytes derived from control and diabetic rats. Biochem J 371: 753–
760, 2003.
Hattori Y, Matsuda N, Kimura J, Ishitani T, Tamada A, Gando S,
Kemmotsu O, and Kanno M. Diminished function and expression of
291 • OCTOBER 2006 •
www.ajpheart.org
Downloaded from http://ajpheart.physiology.org/ by 10.220.33.6 on May 5, 2017
59. Derave W, Hansen BF, Lund S, Kristiansen S, and Richter EA.
Muscle glycogen content affects insulin-stimulated glucose transport and
protein kinase B activity. Am J Physiol Endocrinol Metab 279: E947–
E955, 2000.
60. Di Filippo C, Marfella R, Cuzzocrea S, Piegari E, Petronella P,
Giugliano D, Rossi F, and D’Amico M. Hyperglycemia in streptozotocin-induced diabetic rat increases infarct size associated with low levels
of myocardial HO-1 during ischemia/reperfusion. Diabetes 54: 803– 810,
2005.
61. Diamant M, Lamb HJ, Groeneveld Y, Endert EL, Smit JW, Bax JJ,
Romijn JA, de Roos A, and Radder JK. Diastolic dysfunction is
associated with altered myocardial metabolism in asymptomatic normotensive patients with well-controlled type 2 diabetes mellitus. J Am Coll
Cardiol 42: 328 –335, 2003.
62. Doria A, Nosadini R, Avogaro A, Fioretto P, and Crepaldi G.
Myocardial metabolism in type 1 diabetic patients without coronary
artery disease. Diabet Med 8: S104 –S107, 1991.
63. Dressel U, Allen TL, Pippal JB, Rohde PR, Lau P, and Muscat GE.
The peroxisome proliferator-activated receptor beta/delta agonist,
GW501516, regulates the expression of genes involved in lipid catabolism and energy uncoupling in skeletal muscle cells. Mol Endocrinol 17:
2477–2493, 2003.
64. Du X, Matsumura T, Edelstein D, Rossetti L, Zsengeller Z, Szabo C,
and Brownlee M. Inhibition of GAPDH activity by poly(ADP-ribose)
polymerase activates three major pathways of hyperglycemic damage in
endothelial cells. J Clin Invest 112: 1049 –1057, 2003.
65. Dyck JR, Barr AJ, Barr RL, Kolattukudy PE, and Lopaschuk GD.
Characterization of cardiac malonyl-CoA decarboxylase and its putative
role in regulating fatty acid oxidation. Am J Physiol Heart Circ Physiol
275: H2122–H2129, 1998.
66. Dyck JR, Cheng JF, Stanley WC, Barr R, Chandler MP, Brown S,
Wallace D, Arrhenius T, Harmon C, Yang G, Nadzan AM, and
Lopaschuk GD. Malonyl coenzyme a decarboxylase inhibition protects
the ischemic heart by inhibiting fatty acid oxidation and stimulating
glucose oxidation. Circ Res 94: e78 – e84, 2004.
67. Dyntar D, Eppenberger-Eberhardt M, Maedler K, Pruschy M, Eppenberger HM, Spinas GA, and Donath MY. Glucose and palmitic
acid induce degeneration of myofibrils and modulate apoptosis in rat
adult cardiomyocytes. Diabetes 50: 2105–2113, 2001.
68. Eckel RH. Lipoprotein lipase. A multifunctional enzyme relevant to
common metabolic diseases. N Engl J Med 320: 1060 –1068, 1989.
69. Eliasson MJ, Sampei K, Mandir AS, Hurn PD, Traystman RJ, Bao
J, Pieper A, Wang ZQ, Dawson TM, Snyder SH, and Dawson VL.
Poly(ADP-ribose) polymerase gene disruption renders mice resistant to
cerebral ischemia. Nat Med 3: 1089 –1095, 1997.
70. Enerback S and Gimble JM. Lipoprotein lipase gene expression:
physiological regulators at the transcriptional and post-transcriptional
level. Biochim Biophys Acta 1169: 107–125, 1993.
71. Entman ML, Bornet EP, Van Winkle WB, Goldstein MA, and
Schwartz A. Association of glycogenolysis with cardiac sarcoplasmic
reticulum: I. Effect of glycogen depletion, deoxycholate solubilization
and cardiac ischemia: evidence for a phorphorylase kinase membrane
complex. J Mol Cell Cardiol 9: 515–528, 1977.
72. Exil VJ, Roberts RL, Sims H, McLaughlin JE, Malkin RA, Gardner
CD, Ni G, Rottman JN, and Strauss AW. Very-long-chain acylcoenzyme a dehydrogenase deficiency in mice. Circ Res 93: 448 – 455,
2003.
73. Fang ZY, Prins JB, and Marwick TH. Diabetic cardiomyopathy:
evidence, mechanisms, and therapeutic implications. Endocr Rev 25:
543–567, 2004.
74. Fein FS. Diabetic cardiomyopathy. Diabetes Care 13: 1169 –1179, 1990.
75. Feuvray D and Lopaschuk GD. Controversies on the sensitivity of the
diabetic heart to ischemic injury: the sensitivity of the diabetic heart to
ischemic injury is decreased. Cardiovasc Res 34: 113–120, 1997.
76. Finck BN. The role of the peroxisome proliferator-activated receptor
alpha pathway in pathological remodeling of the diabetic heart. Curr
Opin Clin Nutr Metab Care 7: 391–396, 2004.
77. Finck BN, Han X, Courtois M, Aimond F, Nerbonne JM, Kovacs A,
Gross RW, and Kelly DPA. Critical role for PPARalpha-mediated
lipotoxicity in the pathogenesis of diabetic cardiomyopathy: modulation
by dietary fat content. Proc Natl Acad Sci USA 100: 1226 –1231, 2003.
78. Finck BN and Kelly DP. Peroxisome proliferator-activated receptor
alpha (PPARalpha) signaling in the gene regulatory control of energy
H1501
Invited Review
H1502
100.
101.
102.
103.
104.
106.
107.
108.
109.
110.
111.
112.
113.
114.
115.
116.
117.
118.
119.
120.
121.
the cardiac Na⫹-Ca2⫹ exchanger in diabetic rats: implication in Ca2⫹
overload. J Physiol 527: 85–94, 2000.
Hauton D, Bennett MJ, and Evans RD. Utilisation of triacylglycerol
and non-esterified fatty acid by the working rat heart: myocardial lipid
substrate preference. Biochim Biophys Acta 1533: 99 –109, 2001.
Hickson-Bick DL, Buja ML, and McMillin JB. Palmitate-mediated
alterations in the fatty acid metabolism of rat neonatal cardiac myocytes.
J Mol Cell Cardiol 32: 511–519, 2000.
Hidaka S, Kakuma T, Yoshimatsu H, Sakino H, Fukuchi S, and
Sakata T. Streptozotocin treatment upregulates uncoupling protein 3
expression in the rat heart. Diabetes 48: 430 – 435, 1999.
Higuchi M, Miyagi K, Nakasone J, and Sakanashi M. Role of high
glycogen in underperfused diabetic rat hearts with added norepinephrine.
J Cardiovasc Pharmacol 26: 899 –907, 1995.
Holness MJ and Sugden MC. Regulation of pyruvate dehydrogenase
complex activity by reversible phosphorylation. Biochem Soc Trans 31:
1143–1151, 2003.
Horman S, Browne G, Krause U, Patel J, Vertommen D, Bertrand L,
Lavoinne A, Hue L, Proud C, and Rider M. Activation of AMPactivated protein kinase leads to the phosphorylation of elongation factor 2
and an inhibition of protein synthesis. Curr Biol 12: 1419 –1423, 2002.
How OJ, Aasum E, Severson DL, Chan WY, Essop MF, and Larsen
TS. Increased myocardial oxygen consumption reduces cardiac efficiency in diabetic mice. Diabetes 55: 466 – 473, 2006.
Huang B, Wu P, Popov KM, and Harris RA. Starvation and diabetes
reduce the amount of pyruvate dehydrogenase phosphatase in rat heart
and kidney. Diabetes 52: 1371–1376, 2003.
Hudson ER, Pan DA, James J, Lucocq JM, Hawley SA, Green KA,
Baba O, Terashima T, and Hardie DG. A novel domain in AMPactivated protein kinase causes glycogen storage bodies similar to those
seen in hereditary cardiac arrhythmias. Curr Biol 13: 861– 866, 2003.
Hue L, Beauloye C, Marsin AS, Bertrand L, Horman S, and Rider
MH. Insulin and ischemia stimulate glycolysis by acting on the same
targets through different and opposing signaling pathways. J Mol Cell
Cardiol 34: 1091–1097, 2002.
Hue L and Rider MH. Role of fructose 2,6-bisphosphate in the control
of glycolysis in mammalian tissues. Biochem J 245: 313–324, 1987.
Huss JM and Kelly DP. Nuclear receptor signaling and cardiac energetics. Circ Res 95: 568 –578, 2004.
Huss JM and Kelly DP. Mitochondrial energy metabolism in heart
failure: a question of balance. J Clin Invest 115: 547–555, 2005.
Ibrahimi A, Bonen A, Blinn WD, Hajri T, Li X, Zhong K, Cameron
R, and Abumrad NA. Muscle-specific overexpression of FAT/CD36
enhances fatty acid oxidation by contracting muscle, reduces plasma
triglycerides and fatty acids, and increases plasma glucose and insulin.
J Biol Chem 274: 26761–26766, 1999.
Itani SI, Ruderman NB, Schmieder F, and Boden G. Lipid-induced
insulin resistance in human muscle is associated with changes in diacylglycerol, protein kinase C., and IkappaB-alpha. Diabetes 51: 2005–
2011, 2002.
Itani SI, Zhou Q, Pories WJ, MacDonald KG, and Dohm GL.
Involvement of protein kinase C in human skeletal muscle insulin
resistance and obesity. Diabetes 49: 1353–1358, 2000.
Iwasaki T, Takahashi S, Takahashi M, Zenimaru Y, Kujiraoka T,
Ishihara M, Nagano M, Suzuki J, Miyamori I, Naiki H, Sakai J,
Fujino T, Miller NE, Yamamoto TT, and Hattori H. Deficiency of the
very low-density lipoprotein (VLDL) receptors in streptozotocin-induced
diabetic rats: insulin dependency of the VLDL receptor. Endocrinology
146: 3286 –3294, 2005.
Jagasia D and McNulty PH. Diabetes mellitus and heart failure.
Congest Heart Fail 9: 133–139; quiz 140 –131, 2003.
Jagasia D, Whiting JM, Concato J, Pfau S, and McNulty PH. Effect
of non-insulin-dependent diabetes mellitus on myocardial insulin responsiveness in patients with ischemic heart disease. Circulation 103: 1734 –
1739, 2001.
Jeffrey FM, Diczku V, Sherry AD, and Malloy CR. Substrate selection
in the isolated working rat heart: effects of reperfusion, afterload, and
concentration. Basic Res Cardiol 90: 388 –396, 1995.
Jensen J, Aslesen R, Ivy JL, and Brors O. Role of glycogen concentration and epinephrine on glucose uptake in rat epitrochlearis muscle.
Am J Physiol Endocrinol Metab 272: E649 –E655, 1997.
Jensen J, Jebens E, Brennesvik EO, Ruzzin J, Soos MA, Engebretsen
EM, O’Rahilly S, and Whitehead JP. Muscle glycogen inharmoniously
AJP-Heart Circ Physiol • VOL
122.
123.
124.
125.
126.
127.
128.
129.
130.
131.
132.
133.
134.
135.
136.
137.
138.
regulates glycogen synthase activity, glucose uptake, and proximal insulin signaling. Am J Physiol Endocrinol Metab 290: E154 –E162, 2006.
Jensen J, Ruzzin J, Jebens E, Brennesvik EO, and Knardahl S.
Improved insulin-stimulated glucose uptake and glycogen synthase activation in rat skeletal muscles after adrenaline infusion: role of glycogen
content and PKB phosphorylation. Acta Physiol Scand 184: 121–130,
2005.
Joffe II, Travers KE, Perreault-Micale CL, Hampton T, Katz SE,
Morgan JP, and Douglas PS. Abnormal cardiac function in the streptozotocin-induced non-insulin-dependent diabetic rat: noninvasive assessment with doppler echocardiography and contribution of the nitric
oxide pathway. J Am Coll Cardiol 34: 2111–2119, 1999.
Kajstura J, Fiordaliso F, Andreoli AM, Li B, Chimenti S, Medow
MS, Limana F, Nadal-Ginard B, Leri A, and Anversa P. IGF-1
overexpression inhibits the development of diabetic cardiomyopathy and
angiotensin II-mediated oxidative stress. Diabetes 50: 1414 –1424, 2001.
Kaneko M, Bucciarelli L, Hwang YC, Lee L, Yan SF, Schmidt AM,
and Ramasamy R. Aldose reductase and AGE-RAGE pathways: key
players in myocardial ischemic injury. Ann NY Acad Sci 1043: 702–709,
2005.
Kashihara H, Shi ZQ, Yu JZ, McNeill JH, and Tibbits GF. Effects of
diabetes and hypertension on myocardial Na⫹-Ca2⫹ exchange. Can
J Physiol Pharmacol 78: 12–19, 2000.
Khairallah M, Labarthe F, Bouchard B, Danialou G, Petrof BJ, and
Des Rosiers C. Profiling substrate fluxes in the isolated working mouse
heart using 13C-labeled substrates: focusing on the origin and fate of
pyruvate and citrate carbons. Am J Physiol Heart Circ Physiol 286:
H1461–H1470, 2004.
Kim JK, Fillmore JJ, Chen Y, Yu C, Moore IK, Pypaert M, Lutz EP,
Kako Y, Velez-Carrasco W, Goldberg IJ, Breslow JL, and Shulman
GI. Tissue-specific overexpression of lipoprotein lipase causes tissuespecific insulin resistance. Proc Natl Acad Sci USA 98: 7522–7527, 2001.
Kim JK, Fillmore JJ, Sunshine MJ, Albrecht B, Higashimori T, Kim
DW, Liu ZX, Soos TJ, Cline GW, O’Brien WR, Littman DR, and
Shulman GI. PKCtheta knockout mice are protected from fat-induced
insulin resistance. J Clin Invest 114: 823– 827, 2004.
King KL, Okere IC, Sharma N, Dyck JR, Reszko AE, McElfresh TA,
Kerner J, Chandler MP, Lopaschuk GD, and Stanley WC. Regulation of cardiac malonyl-CoA content and fatty acid oxidation during
increased cardiac power. Am J Physiol Heart Circ Physiol 289: H1033–
H1037, 2005.
Koonen DP, Glatz JF, Bonen A, and Luiken JJ. Long-chain fatty acid
uptake and FAT/CD36 translocation in heart and skeletal muscle. Biochim Biophys Acta 1736: 163–180, 2005.
Kraegen EW, Sowden JA, Halstead MB, Clark PW, Rodnick KJ,
Chisholm DJ, and James DE. Glucose transporters and in vivo glucose
uptake in skeletal and cardiac muscle: fasting, insulin stimulation and
immunoisolation studies of GLUT1 and GLUT4. Biochem J 295: 287–
293, 1993.
Kudo N, Barr AJ, Barr RL, Desai S, and Lopaschuk GD. High rates
of fatty acid oxidation during reperfusion of ischemic hearts are associated with a decrease in malonyl-CoA levels due to an increase in
5⬘-AMP-activated protein kinase inhibition of acetyl-CoA carboxylase.
J Biol Chem 270: 17513–17520, 1995.
Kudo N, Gillespie JG, Kung L, Witters LA, Schulz R, Clanachan AS,
and Lopaschuk GD. Characterization of 5⬘AMP activated protein kinase activity in the heart and its role in inhibiting acetyl-CoA carboxylase
during reperfusion following ischemia. Biochim Biophys Acta 1301:
67–75, 1996.
Kurtz DM, Rinaldo P, Rhead WJ, Tian L, Millington DS, Vockley J,
Hamm DA, Brix AE, Lindsey JR, Pinkert CA, O’Brien WE, and
Wood PA. Targeted disruption of mouse long-chain acyl-CoA dehydrogenase gene reveals crucial roles for fatty acid oxidation. Proc Natl Acad
Sci USA 95: 15592–15597, 1998.
Laughlin MR, Petit WA Jr, Shulman RG and Barrett EJ. Measurement of myocardial glycogen synthesis in diabetic and fasted rats. Am J
Physiol Endocrinol Metab 258: E184 –E190, 1990.
Laybutt DR, Thompson AL, Cooney GJ, and Kraegen EW. Selective
chronic regulation of GLUT1 and GLUT4 content by insulin, glucose,
and lipid in rat cardiac muscle in vivo. Am J Physiol Heart Circ Physiol
273: H1309 –H1316, 1997.
Lee GY, Kim NH, Zhao ZS, Cha BS, and Kim YS. Peroxisomalproliferator-activated receptor alpha activates transcription of the rat
291 • OCTOBER 2006 •
www.ajpheart.org
Downloaded from http://ajpheart.physiology.org/ by 10.220.33.6 on May 5, 2017
105.
ABNORMAL REGULATION OF CARDIAC SUBSTRATE METABOLISM IN DIABETES
Invited Review
ABNORMAL REGULATION OF CARDIAC SUBSTRATE METABOLISM IN DIABETES
139.
140.
141.
142.
144.
145.
146.
147.
148.
149.
150.
151.
152.
153.
154.
155.
156.
157.
158.
AJP-Heart Circ Physiol • VOL
159. Luiken JJ, van Nieuwenhoven FA, America G, van der Vusse GJ,
and Glatz JF. Uptake and metabolism of palmitate by isolated cardiac
myocytes from adult rats: involvement of sarcolemmal proteins. J Lipid
Res 38: 745–758, 1997.
160. Maki M, Nuutila P, Laine H, Voipio-Pulkki LM, Haaparanta M,
Solin O, and Knuuti JM. Myocardial glucose uptake in patients with
NIDDM and stable coronary artery disease. Diabetes 46: 1491–1496,
1997.
161. Marsin AS, Bertrand L, Rider MH, Deprez J, Beauloye C, Vincent
MF, Van den Berghe G, Carling D, and Hue L. Phosphorylation and
activation of heart PFK-2 by AMPK has a role in the stimulation of
glycolysis during ischaemia. Curr Biol 10: 1247–1255, 2000.
162. Mazumder PK, O’Neill BT, Roberts MW, Buchanan J, Yun UJ,
Cooksey RC, Boudina S, and Abel ED. Impaired cardiac efficiency and
increased fatty acid oxidation in insulin-resistant ob/ob mouse hearts.
Diabetes 53: 2366 –2374, 2004.
163. McClain DA and Crook ED. Hexosamines and insulin resistance.
Diabetes 45: 1003–1009, 1996.
164. McDonagh PF and Hokama JY. Microvascular perfusion and transport
in the diabetic heart. Microcirculation 7: 163–181, 2000.
165. Merkel M, Eckel RH, and Goldberg IJ. Lipoprotein lipase: genetics,
lipid uptake, and regulation. J Lipid Res 43: 1997–2006, 2002.
166. Merkel M, Loeffler B, Kluger M, Fabig N, Geppert G, Pennacchio
LA, Laatsch A, and Heeren J. Apolipoprotein AV accelerates plasma
hydrolysis of triglyceride-rich lipoproteins by interaction with proteoglycan-bound lipoprotein lipase. J Biol Chem 280: 21553–21560, 2005.
167. Montanari D, Yin H, Dobrzynski E, Agata J, Yoshida H, Chao J, and
Chao L. Kallikrein gene delivery improves serum glucose and lipid
profiles and cardiac function in streptozotocin-induced diabetic rats.
Diabetes 54: 1573–1580, 2005.
168. Monti LD, Landoni C, Setola E, Galluccio E, Lucotti P, Sandoli EP,
Origgi A, Lucignani G, Piatti P, and Fazio F. Myocardial insulin
resistance associated with chronic hypertriglyceridemia and increased
FFA levels in Type 2 diabetic patients. Am J Physiol Heart Circ Physiol
287: H1225–H1231, 2004.
169. Muoio DM, MacLean PS, Lang DB, Li S, Houmard JA, Way JM,
Winegar DA, Corton JC, Dohm GL, and Kraus WE. Fatty acid
homeostasis and induction of lipid regulatory genes in skeletal muscles of
peroxisome proliferator-activated receptor (PPAR) alpha knock-out mice
Evidence for compensatory regulation by PPAR delta. J Biol Chem 277:
26089 –26097, 2002.
170. Murray AJ, Anderson RE, Watson GC, Radda GK, and Clarke K.
Uncoupling proteins in human heart. Lancet 364: 1786 –1788, 2004.
171. Murray AJ, Panagia M, Hauton D, Gibbons GF, and Clarke K.
Plasma free fatty acids and peroxisome proliferator-activated receptor
{alpha} in the control of myocardial uncoupling protein levels. Diabetes
54: 3496 –3502, 2005.
172. Nagao M, Parimoo B, and Tanaka K. Developmental, nutritional, and
hormonal regulation of tissue-specific expression of the genes encoding
various acyl-CoA dehydrogenases and alpha-subunit of electron transfer
flavoprotein in rat. J Biol Chem 268: 24114 –24124, 1993.
173. Neely JR, Rovetto MJ, and Oram JF. Myocardial utilization of
carbohydrate and lipids. Prog Cardiovasc Dis 15: 289 –329, 1972.
174. Neitzel AS, Carley AN, and Severson DL. Chylomicron and palmitate
metabolism by perfused hearts from diabetic mice. Am J Physiol Endocrinol Metab 284: E357–E365, 2003.
175. Newsholme EA and Randle PJ. Regulation of glucose uptake by
muscle. 7 Effects of fatty acids, ketone bodies and pyruvate, and of
alloxan-diabetes, starvation, hypophysectomy and adrenalectomy, on the
concentrations of hexose phosphates, nucleotides and inorganic phosphate in perfused rat heart. Biochem J 93: 641– 651, 1964.
176. Nicholl TA, Lopaschuk GD, and McNeill JH. Effects of free fatty acids
and dichloroacetate on isolated working diabetic rat heart. Am J Physiol
Heart Circ Physiol 261: H1053–H1059, 1991.
177. Nielsen LB, Bartels ED, and Bollano E. Overexpression of apolipoprotein B in the heart impedes cardiac triglyceride accumulation and development of cardiac dysfunction in diabetic mice. J Biol Chem 277:
27014 –27020, 2002.
178. Niu YG, Hauton D, and Evans RD. Utilization of triacylglycerol-rich
lipoproteins by the working rat heart: routes of uptake and metabolic
fates. J Physiol 558: 225–237, 2004.
179. Nuutila P, Knuuti J, Ruotsalainen U, Koivisto VA, Eronen E, Teras
M, Bergman J, Haaparanta M, Voipio-Pulkki LM, Viikari J, T.
Rönnemaa, U. Wegelius, and H. Yki-Järvinen. Insulin resistance is
291 • OCTOBER 2006 •
www.ajpheart.org
Downloaded from http://ajpheart.physiology.org/ by 10.220.33.6 on May 5, 2017
143.
hepatic malonyl-CoA decarboxylase gene: a key regulation of malonylCoA level. Biochem J 378: 983–990, 2004.
Lehman JJ, Barger PM, Kovacs A, Saffitz JE, Medeiros DM, and
Kelly DP. Peroxisome proliferator-activated receptor gamma coactivator-1 promotes cardiac mitochondrial biogenesis. J Clin Invest 106:
847– 856, 2000.
Leone TC, Weinheimer CJ, and Kelly DPA. critical role for the
peroxisome proliferator-activated receptor alpha (PPARalpha) in the
cellular fasting response: the PPARalpha-null mouse as a model of fatty
acid oxidation disorders. Proc Natl Acad Sci USA 96: 7473–7478, 1999.
Levak-Frank S, Radner H, Walsh A, Stollberger R, Knipping G,
Hoefler G, Sattler W, Weinstock PH, Breslow JL, and Zechner R.
Muscle-specific overexpression of lipoprotein lipase causes a severe
myopathy characterized by proliferation of mitochondria and peroxisomes in transgenic mice. J Clin Invest 96: 976 –986, 1995.
Li H, Wang J, Wilhelmsson H, Hansson A, Thoren P, Duffy J, Rustin
P, and Larsson NG. Genetic modification of survival in tissue-specific
knockout mice with mitochondrial cardiomyopathy. Proc Natl Acad Sci
USA 97: 3467–3472, 2000.
Li J, Hu X, Selvakumar P, Russell RR 3rd, Cushman SW, Holman
GD, and Young LH. Role of the nitric oxide pathway in AMPKmediated glucose uptake and GLUT4 translocation in heart muscle. Am J
Physiol Endocrinol Metab 287: E834 –E841, 2004.
Listenberger LL, Han X, Lewis SE, Cases S, Farese RV Jr, Ory DS,
and Schaffer JE. Triglyceride accumulation protects against fatty acidinduced lipotoxicity. Proc Natl Acad Sci USA 100: 3077–3082, 2003.
Liu Y, Thornton JD, Cohen MV, Downey JM, and Schaffer SW.
Streptozotocin-induced non-insulin-dependent diabetes protects the heart
from infarction. Circulation 88: 1273–1278, 1993.
Lopaschuk GD. Alterations in myocardial fatty acid metabolism contribute to ischemic injury in the diabetic. Can J Cardiol 5: 315–320,
1989.
Lopaschuk GD. Metabolic abnormalities in the diabetic heart. Heart
Fail Rev 7: 149 –159, 2002.
Lopaschuk GD and Barr RL. Measurements of fatty acid and carbohydrate metabolism in the isolated working rat heart. Mol Cell Biochem
172: 137–147, 1997.
Lopaschuk GD, Belke DD, Gamble J, Itoi T, and Schonekess BO.
Regulation of fatty acid oxidation in the mammalian heart in health and
disease. Biochim Biophys Acta 1213: 263–276, 1994.
Lopaschuk GD, Saddik M, Barr R, Huang L, Barker CC, and
Muzyka RA. Effects of high levels of fatty acids on functional recovery
of ischemic hearts from diabetic rats. Am J Physiol Endocrinol Metab
263: E1046 –E1053, 1992.
Lteif AA, Mather KJ, and Clark CM. Diabetes and heart disease an
evidence-driven guide to risk factors management in diabetes. Cardiol
Rev 11: 262–274, 2003.
Luiken JJ, Arumugam Y, Bell RC, Calles-Escandon J, Tandon NN,
Glatz JF, and Bonen A. Changes in fatty acid transport and transporters
are related to the severity of insulin deficiency. Am J Physiol Endocrinol
Metab 283: E612–E621, 2002.
Luiken JJ, Arumugam Y, Dyck DJ, Bell RC, Pelsers MM, Turcotte
LP, Tandon NN, Glatz JF, and Bonen A. Increased rates of fatty acid
uptake and plasmalemmal fatty acid transporters in obese Zucker rats.
J Biol Chem 276: 40567– 40573, 2001.
Luiken JJ, Coort SL, Koonen DP, van der Horst DJ, Bonen A,
Zorzano A, and Glatz JF. Regulation of cardiac long-chain fatty acid
and glucose uptake by translocation of substrate transporters. Pflügers
Arch 448: 1–15, 2004.
Luiken JJ, Coort SL, Willems J, Coumans WA, Bonen A, van der
Vusse GJ, and Glatz JF. Contraction-induced fatty acid translocase/
CD36 translocation in rat cardiac myocytes is mediated through AMPactivated protein kinase signaling. Diabetes 52: 1627–1634, 2003.
Luiken JJ, Koonen DP, Willems J, Zorzano A, Becker C, Fischer Y,
Tandon NN, Van Der Vusse GJ, Bonen A, and Glatz JF. Insulin
stimulates long-chain fatty acid utilization by rat cardiac myocytes
through cellular redistribution of FAT/CD36. Diabetes 51: 3113–3119,
2002.
Luiken JJ, Schaap FG, van Nieuwenhoven FA, van der Vusse GJ,
Bonen A, and Glatz JF. Cellular fatty acid transport in heart and skeletal
muscle as facilitated by proteins. Lipids 34, Suppl: S169 –S175, 1999.
Luiken JJ, Turcotte LP, and Bonen A. Protein-mediated palmitate
uptake and expression of fatty acid transport proteins in heart giant
vesicles. J Lipid Res 40: 1007–1016, 1999.
H1503
Invited Review
H1504
180.
181.
182.
183.
184.
186.
187.
188.
189.
190.
191.
192.
193.
194.
195.
196.
197.
198.
localized to skeletal but not heart muscle in type 1 diabetes. Am J Physiol
Endocrinol Metab 264: E756 –E762, 1993.
O’Donnell JM, Zampino M, Alpert NM, Fasano MJ, Geenen DL,
and Lewandowski ED. Accelerated triacylglycerol turnover kinetics in
hearts of diabetic rats include evidence for compartmented lipid storage.
Am J Physiol Endocrinol Metab 290: E448 –E455, 2006.
O’Looney P, Irwin D, Briscoe P, and Vahouny GV. Lipoprotein
composition as a component in the lipoprotein clearance defect in
experimental diabetes. J Biol Chem 260: 428 – 432, 1985.
O’Looney P, Vander Maten M, and Vahouny GV. Insulin-mediated
modifications of myocardial lipoprotein lipase and lipoprotein metabolism. J Biol Chem 258: 12994 –13001, 1983.
Olson AL and Pessin JE. Transcriptional regulation of the human
GLUT4 gene promoter in diabetic transgenic mice. J Biol Chem 270:
23491–23495, 1995.
Ostrander DB, Sparagna GC, Amoscato AA, McMillin JB, and
Dowhan W. Decreased cardiolipin synthesis corresponds with cytochrome c release in palmitate-induced cardiomyocyte apoptosis. J Biol
Chem 276: 38061–38067, 2001.
Pacher P, Liaudet L, Soriano FG, Mabley JG, Szabo E, and Szabo C.
The role of poly(ADP-ribose) polymerase activation in the development
of myocardial and endothelial dysfunction in diabetes. Diabetes 51:
514 –521, 2002.
Panagia M, Gibbons GF, Radda GK, and Clarke K. PPARalpha
activation required for decreased glucose uptake and increased susceptibility to injury during ischemia. Am J Physiol Heart Circ Physiol 288:
H2677–H2683, 2005.
Paz K, Hemi R, LeRoith D, Karasik A, Elhanany E, Kanety H, and
Zick YA. molecular basis for insulin resistance. Elevated serine/threonine phosphorylation of IRS-1 and IRS-2 inhibits their binding to the
juxtamembrane region of the insulin receptor and impairs their ability to
undergo insulin-induced tyrosine phosphorylation. J Biol Chem 272:
29911–29918, 1997.
Pennacchio LA, Olivier M, Hubacek JA, Cohen JC, Cox DR, Fruchart JC, Krauss RM, and Rubin EM. An apolipoprotein influencing
triglycerides in humans and mice revealed by comparative sequencing.
Science 294: 169 –173, 2001.
Pessin JE and Bell GI. Mammalian facilitative glucose transporter
family: structure and molecular regulation. Annu Rev Physiol 54: 911–
930, 1992.
Peterson LR, Herrero P, Schechtman KB, Racette SB, Waggoner
AD, Kisrieva-Ware Z, Dence C, Klein S, Marsala J, Meyer T, and
Gropler RJ. Effect of obesity and insulin resistance on myocardial
substrate metabolism and efficiency in young women. Circulation 109:
2191–2196, 2004.
Peterson LR, Waggoner AD, Schechtman KB, Meyer T, Gropler RJ,
Barzilai B, and Davila-Roman VG. Alterations in left ventricular
structure and function in young healthy obese women: assessment by
echocardiography and tissue Doppler imaging. J Am Coll Cardiol 43:
1399 –1404, 2004.
Picano E. Diabetic cardiomyopathy the importance of being earliest.
J Am Coll Cardiol 42: 454 – 457, 2003.
Pillarisetti S, Paka L, Sasaki A, Vanni-Reyes T, Yin B, Parthasarathy N, Wagner WD, and Goldberg IJ. Endothelial cell heparanase
modulation of lipoprotein lipase activity. Evidence that heparan sulfate
oligosaccharide is an extracellular chaperone. J Biol Chem 272: 15753–
15759, 1997.
Pillutla P, Hwang YC, Augustus A, Yokoyama M, Yagyu H,
Johnston TP, Kaneko M, Ramasamy R, and Goldberg IJ. Perfusion
of hearts with triglyceride-rich particles reproduces the metabolic abnormalities in lipotoxic cardiomyopathy. Am J Physiol Endocrinol Metab
288: E1229 –E1235, 2005.
Poirier P, Bogaty P, Garneau C, Marois L, and Dumesnil JG.
Diastolic dysfunction in normotensive men with well-controlled type 2
diabetes: importance of maneuvers in echocardiographic screening for
preclinical diabetic cardiomyopathy. Diabetes Care 24: 5–10, 2001.
Poornima IG, Parikh P, and Shannon RP. Diabetic cardiomyopathy:
the search for a unifying hypothesis. Circ Res 98: 596 – 605, 2006.
Raev DC. Left ventricular function, and specific diabetic complications
in other target organs in young insulin-dependent diabetics: an echocardiographic study. Heart Vessels 9: 121–128, 1994.
Ramasamy R, Oates PJ, and Schaefer S. Aldose reductase inhibition
protects diabetic and nondiabetic rat hearts from ischemic injury. Diabetes 46: 292–300, 1997.
AJP-Heart Circ Physiol • VOL
199. Randle PJ, Garland PB, Hales CN, and Newsholme EA. The glucose
fatty-acid cycle. Its role in insulin sensitivity and the metabolic disturbances of diabetes mellitus. Lancet 1: 785–789, 1963.
200. Redfield MM, Jacobsen SJ, Burnett JC Jr, Mahoney DW, Bailey KR,
and Rodeheffer RJ. Burden of systolic and diastolic ventricular dysfunction in the community: appreciating the scope of the heart failure
epidemic. JAMA 289: 194 –202, 2003.
201. Richards MR, Harp JD, Ory DS, and Schaffer JE. Fatty acid transport
protein 1 and long-chain acyl coenzyme A synthetase 1 interact in
adipocytes. J Lipid Res 47: 665– 672, 2006.
202. Rider MH, Bertrand L, Vertommen D, Michels PA, Rousseau GG,
and Hue L. 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatase:
head-to-head with a bifunctional enzyme that controls glycolysis. Biochem J 381: 561–579, 2004.
203. Rodrigues B, Cam MC, Jian K, Lim F, Sambandam N, and Shepherd G. Differential effects of streptozotocin-induced diabetes on cardiac lipoprotein lipase activity. Diabetes 46: 1346 –1353, 1997.
204. Rodrigues B, Cam MC, and McNeill JH. Myocardial substrate metabolism: implications for diabetic cardiomyopathy. J Mol Cell Cardiol 27:
169 –179, 1995.
205. Rodrigues B, Cam MC, and McNeill JH. Metabolic disturbances in
diabetic cardiomyopathy. Mol Cell Biochem 180: 53–57, 1998.
206. Rui L, Aguirre V, Kim JK, Shulman GI, Lee A, Corbould A, Dunaif
A, and White MF. Insulin/IGF1 and TNFalpha stimulate phosphorylation of IRS-1 at inhibitory Ser307 via distinct pathways. J Clin Invest
107: 181–189, 2001.
207. Rupp H, Elimban V, and Dhalla NS. Modification of myosin isozymes
and SR Ca(2⫹)-pump ATPase of the diabetic rat heart by lipid-lowering
interventions. Mol Cell Biochem 132: 69 – 80, 1994.
208. Russell LK, Finck BN, and Kelly DP. Mouse models of mitochondrial
dysfunction and heart failure. J Mol Cell Cardiol 38: 81–91, 2005.
209. Russell RR 3rd, Li J, Coven DL, Pypaert M, Zechner C, Palmeri M,
Giordano FJ, Mu J, Birnbaum MJ, and Young LH. AMP activated
protein kinase mediates ischemic glucose uptake and prevents postischemic cardiac dysfunction, apoptosis, and injury. J Clin Invest 114: 495–
503, 2004.
210. Saddik M and Lopaschuk GD. Myocardial triglyceride turnover and
contribution to energy substrate utilization in isolated working rat hearts.
J Biol Chem 266: 8162– 8170, 1991.
211. Saha AK, Vavvas D, Kurowski TG, Apazidis A, Witters LA, Shafrir
E, and Ruderman NB. Malonyl-CoA regulation in skeletal muscle: its
link to cell citrate and the glucose-fatty acid cycle. Am J Physiol
Endocrinol Metab 272: E641–E648, 1997.
212. Saito F, Kawaguchi M, Izumida J, Asakura T, Maehara K, and
Maruyama Y. Alteration in haemodynamics and pathological changes in
the cardiovascular system during the development of Type 2 diabetes
mellitus in OLETF rats. Diabetologia 46: 1161–1169, 2003.
213. Sakamoto J, Barr RL, Kavanagh KM, and Lopaschuk GD. Contribution of malonyl-CoA decarboxylase to the high fatty acid oxidation
rates seen in the diabetic heart. Am J Physiol Heart Circ Physiol 278:
H1196 –H1204, 2000.
214. Schaffer JE. Fatty acid transport: the roads taken. Am J Physiol Endocrinol Metab 282: E239 –E246, 2002.
215. Schaffer SW, Ballard-Croft C, Boerth S, and Allo SN. Mechanisms
underlying depressed Na⫹/Ca2⫹ exchanger activity in the diabetic
heart. Cardiovasc Res 34: 129 –136, 1997.
216. Scheuermann-Freestone M, Madsen PL, Manners D, Blamire AM,
Buckingham RE, Styles P, Radda GK, Neubauer S, and Clarke K.
Abnormal cardiac and skeletal muscle energy metabolism in patients
with type 2 diabetes. Circulation 107: 3040 –3046, 2003.
217. Schonekess BO. Competition between lactate, and fatty acids as sources
of ATP in the isolated working rat heart. J Mol Cell Cardiol 29:
2725–2733, 1997.
218. Semeniuk LM, Kryski AJ, and Severson DL. Echocardiographic
assessment of cardiac function in diabetic db/db and transgenic db/dbhGLUT4 mice. Am J Physiol Heart Circ Physiol 283: H976 –H982,
2002.
219. Severson DL. Diabetic cardiomyopathy: recent evidence from mouse
models of type 1, and type 2 diabetes. Can J Physiol Pharmacol 82:
813– 823, 2004.
220. Sharma S, Adrogue JV, Golfman L, Uray I, Lemm J, Youker K,
Noon GP, Frazier OH, and Taegtmeyer H. Intramyocardial lipid
accumulation in the failing human heart resembles the lipotoxic rat heart.
FASEB J 18: 1692–1700, 2004.
291 • OCTOBER 2006 •
www.ajpheart.org
Downloaded from http://ajpheart.physiology.org/ by 10.220.33.6 on May 5, 2017
185.
ABNORMAL REGULATION OF CARDIAC SUBSTRATE METABOLISM IN DIABETES
Invited Review
ABNORMAL REGULATION OF CARDIAC SUBSTRATE METABOLISM IN DIABETES
AJP-Heart Circ Physiol • VOL
244. Vikramadithyan RK, Hirata K, Yagyu H, Hu Y, Augustus A,
Homma S, and Goldberg IJ. Peroxisome proliferator-activated receptor
agonists modulate heart function in transgenic mice with lipotoxic
cardiomyopathy. J Pharmacol Exp Ther 313: 586 –593, 2005.
245. Vincent G, Bouchard B, Khairallah M, and Des Rosiers C. Differential modulation of citrate synthesis and release by fatty acids in
perfused working rat hearts. Am J Physiol Heart Circ Physiol 286:
H257–H266, 2004.
246. Virag L and Szabo C. The therapeutic potential of poly(ADP-ribose)
polymerase inhibitors. Pharmacol Rev 54: 375– 429, 2002.
247. Wakasaki H, Koya D, Schoen FJ, Jirousek MR, Ways DK, Hoit BD,
Walsh RA, and King GL. Targeted overexpression of protein kinase
Cbeta2 isoform in myocardium causes cardiomyopathy. Proc Natl Acad
Sci USA 94: 9320 –9325, 1997.
248. Wall SR and Lopaschuk GD. Glucose oxidation rates in fatty acidperfused isolated working hearts from diabetic rats. Biochim Biophys
Acta 1006: 97–103, 1989.
249. Wang P and Chatham JC. Onset of diabetes in Zucker diabetic fatty
(ZDF) rats leads to improved recovery of function after ischemia in the
isolated perfused heart. Am J Physiol Endocrinol Metab 286: E725–
E736, 2004.
250. Wang P, Lloyd SG, Zeng H, Bonen A, and Chatham JC. Impact of
altered substrate utilization on cardiac function in isolated hearts from
Zucker diabetic fatty rats. Am J Physiol Heart Circ Physiol 288: H2102–
H2110, 2005.
251. Weiss J and Hiltbrand B. Functional compartmentation of glycolytic
versus oxidative metabolism in isolated rabbit heart. J Clin Invest 75:
436 – 447, 1985.
252. Weiss JN and Lamp ST. Glycolysis preferentially inhibits ATP-sensitive K⫹ channels in isolated guinea pig cardiac myocytes. Science 238:
67– 69, 1987.
253. Wilson PW. Diabetes mellitus, and coronary heart disease. Endocrinol
Metab Clin North Am 30: 857– 881, 2001.
254. Wojtaszewski JF, Jorgensen SB, Hellsten Y, Hardie DG, and Richter
EA. Glycogen-dependent effects of 5-aminoimidazole-4-carboxamide
(AICA)-riboside on AMP-activated protein kinase and glycogen synthase
activities in rat skeletal muscle. Diabetes 51: 284 –292, 2002.
255. Wojtaszewski JF, MacDonald C, Nielsen JN, Hellsten Y, Hardie DG,
Kemp BE, Kiens B, and Richter EA. Regulation of 5⬘AMPactivated
EA protein kinase activity and substrate utilization in exercising human
skeletal muscle. Am J Physiol Endocrinol Metab 284: E813–E822, 2003.
256. Wold LE and Ren J. Streptozotocin directly impairs cardiac contractile
function in isolated ventricular myocytes via a p38 map kinase-dependent
oxidative stress mechanism. Biochem Biophys Res Commun 318: 1066 –
1071, 2004.
257. Wu P, Inskeep K, Bowker-Kinley MM, Popov KM, and Harris RA.
Mechanism responsible for inactivation of skeletal muscle pyruvate
dehydrogenase complex in starvation and diabetes. Diabetes 48: 1593–
1599, 1999.
258. Wu P, Sato J, Zhao Y, Jaskiewicz J, Popov KM, and Harris RA.
Starvation and diabetes increase the amount of pyruvate dehydrogenase
kinase isoenzyme 4 in rat heart. Biochem J 329: 197–201, 1998.
259. Xing Y, Musi N, Fujii N, Zou L, Luptak I, Hirshman MF, Goodyear
LJ, and Tian R. Glucose metabolism and energy homeostasis in mouse
hearts overexpressing dominant negative alpha2 subunit of AMP-activated protein kinase. J Biol Chem 278: 28372–28377, 2003.
260. Yagyu H, Chen G, Yokoyama M, Hirata K, Augustus A, Kako Y, Seo
T, Hu Y, Lutz EP, Merkel M, Bensadoun A, Homma S, and Goldberg IJ. Lipoprotein lipase (LpL) on the surface of cardiomyocytes
increases lipid uptake and produces a cardiomyopathy. J Clin Invest 111:
419 – 426, 2003.
261. Yang J and Holman GD. Insulin and contraction stimulate exocytosis,
but increased AMP-activated protein kinase activity resulting from oxidative metabolism stress slows endocytosis of GLUT4 in cardiomyocytes. J Biol Chem 280: 4070 – 4078, 2005.
262. Ye G, Metreveli NS, Donthi RV, Xia S, Xu M, Carlson EC, and
Epstein PN. Catalase protects cardiomyocyte function in models of type
1 and type 2 diabetes. Diabetes 53: 1336 –1343, 2004.
263. Ye G, Metreveli NS, Ren J, and Epstein PN. Metallothionein prevents
diabetes-induced deficits in cardiomyocytes by inhibiting reactive oxygen species production. Diabetes 52: 777–783, 2003.
264. Yokoyama M, Yagyu H, Hu Y, Seo T, Hirata K, Homma S, and
Goldberg IJ. Apolipoprotein B production reduces lipotoxic cardiomy-
291 • OCTOBER 2006 •
www.ajpheart.org
Downloaded from http://ajpheart.physiology.org/ by 10.220.33.6 on May 5, 2017
221. Shen X, Zheng S, Metreveli NS, and Epstein PN. Protection of Cardiac
Mitochondria by Overexpression of MnSOD Reduces Diabetic Cardiomyopathy. Diabetes 55: 798 – 805, 2006.
222. Shen X, Zheng S, Thongboonkerd V, Xu M, Pierce WM Jr. Klein JB
and Epstein PN Cardiac mitochondrial damage and biogenesis in a
chronic model of type 1 diabetes. Am J Physiol Endocrinol Metab 287:
E896 –E905, 2004.
223. Shipp JC and Murthy VK. The heart and diabetes. Vascular and
metabolic aspects. Adv Myocardiol 2: 81– 84, 1980.
224. Sidell RJ, Cole MA, Draper NJ, Desrois M, Buckingham RE, and
Clarke K. Thiazolidinedione treatment normalizes insulin resistance and
ischemic injury in the zucker Fatty rat heart. Diabetes 51: 1110 –1117,
2002.
225. Sowers JR, Epstein M, and Frohlich ED. Diabetes, hypertension, and
cardiovascular disease: an update. Hypertension 37: 1053–1059, 2001.
226. Stamler J, Vaccaro O, Neaton JD, and Wentworth D. Diabetes, other
risk factors, and 12-yr cardiovascular mortality for men screened in the
Multiple Risk Factor Intervention Trial. Diabetes Care 16: 434 – 444,
1993.
227. Stanley WC, Lopaschuk GD, and McCormack JG. Regulation of
energy substrate metabolism in the diabetic heart. Cardiovasc Res 34:
25–33, 1997.
228. Stanley WC, Recchia FA, and Lopaschuk GD. Myocardial substrate
metabolism in the normal and failing heart. Physiol Rev 85: 1093–1129,
2005.
229. Stone PH, Muller JE, Hartwell T, York BJ, Rutherford JD, Parker
CB, Turi ZG, Strauss HW, Willerson JT, Robertson T, et al. The
effect of diabetes mellitus on prognosis and serial left ventricular function after acute myocardial infarction: contribution of both coronary
disease and diastolic left ventricular dysfunction to the adverse prognosis
The MILIS Study Group. J Am Coll Cardiol 14: 49 –57, 1989.
230. Stremmel W. Fatty acid uptake by isolated rat heart myocytes represents
a carrier-mediated transport process. J Clin Invest 81: 844 – 852, 1988.
231. Struthers AD and Morris AD. Screening for and treating left-ventricular abnormalities in diabetes mellitus: a new way of reducing cardiac
deaths. Lancet 359: 1430 –1432, 2002.
232. Suarez J, Belke DD, Gloss B, Dieterle T, McDonough PM, Kim YK,
Brunton LL, and Dillmann WH. In vivo adenoviral transfer of sorcin
reverses cardiac contractile abnormalities of diabetic cardiomyopathy.
Am J Physiol Heart Circ Physiol 286: H68 –H75, 2004.
233. Taegtmeyer H, Hems R, and Krebs HA. Utilization of energy-providing substrates in the isolated working rat heart. Biochem J 186: 701–711,
1980.
234. Taegtmeyer H, McNulty P, and Young ME. Adaptation and maladaptation of the heart in diabetes: Part I: general concepts. Circulation 105:
1727–1733, 2002.
235. Tahiliani AG and McNeill JH. Diabetes-induced abnormalities in the
myocardium. Life Sci 38: 959 –974, 1986.
236. Tani M and Neely JR. Hearts from diabetic rats are more resistant to in
vitro ischemia: possible role of altered Ca2⫹ metabolism. Circ Res 62:
931–940, 1988.
237. Teusink B, Voshol PJ, Dahlmans VE, Rensen PC, Pijl H, Romijn JA,
and Havekes LM. Contribution of fatty acids released from lipolysis of
plasma triglycerides to total plasma fatty acid flux and tissue-specific
fatty acid uptake. Diabetes 52: 614 – 620, 2003.
238. Tontonoz P, Hu E, and Spiegelman BM. Stimulation of adipogenesis in
fibroblasts by PPAR gamma 2, a lipid-activated transcription factor. Cell
79: 1147–1156, 1994.
239. Trost SU, Belke DD, Bluhm WF, Meyer M, Swanson E, and
Dillmann WH. Overexpression of the sarcoplasmic reticulum Ca(2⫹)ATPase improves myocardial contractility in diabetic cardiomyopathy.
Diabetes 51: 1166 –1171, 2002.
240. Unger RH. Lipotoxic diseases. Annu Rev Med 53: 319 –336, 2002.
241. Unger RH and Orci L. Lipoapoptosis: its mechanism and its diseases.
Biochim Biophys Acta 1585: 202–212, 2002.
242. Utriainen T, Takala T, Luotolahti M, Ronnemaa T, Laine H, Ruotsalainen U, Haaparanta M, Nuutila P, and Yki-Jarvinen H. Insulin
resistance characterizes glucose uptake in skeletal muscle but not in the
heart in NIDDM. Diabetologia 41: 555–559, 1998.
243. Vetter R, Rehfeld U, Reissfelder C, Weiss W, Wagner KD, Gunther
J, Hammes A, Tschope C, Dillmann W, and Paul M. Transgenic
overexpression of the sarcoplasmic reticulum Ca2⫹ATPase improves
reticular Ca2⫹ handling in normal and diabetic rat hearts. FASEB J 16:
1657–1659, 2002.
H1505
Invited Review
H1506
265.
266.
267.
268.
ABNORMAL REGULATION OF CARDIAC SUBSTRATE METABOLISM IN DIABETES
opathy: studies in heart-specific lipoprotein lipase transgenic mouse.
J Biol Chem 279: 4204 – 4211, 2004.
Young ME, Goodwin GW, Ying J, Guthrie P, Wilson CR, Laws FA,
and Taegtmeyer H. Regulation of cardiac and skeletal muscle malonylCoA decarboxylase by fatty acids. Am J Physiol Endocrinol Metab 280:
E471–E479, 2001.
Young ME, Guthrie PH, Razeghi P, Leighton B, Abbasi S, Patil S,
Youker KA, and Taegtmeyer H. Impaired long-chain fatty acid oxidation and contractile dysfunction in the obese Zucker rat heart. Diabetes
51: 2587–2595, 2002.
Young ME, McNulty P, and Taegtmeyer H. Adaptation and maladaptation of the heart in diabetes: Part II: potential mechanisms. Circulation
105: 1861–1870, 2002.
Young ME, Patil S, Ying J, Depre C, Ahuja HS, Shipley GL,
Stepkowski SM, Davies PJ, and Taegtmeyer H. Uncoupling protein 3
transcription is regulated by peroxisome proliferator-activated receptor
(alpha) in the adult rodent heart. FASEB J 15: 833– 845, 2001.
269. Yuan M, Konstantopoulos N, Lee J, Hansen L, Li ZW, Karin M, and
Shoelson SE. Reversal of obesity- and diet-induced insulin resistance
with salicylates or targeted disruption of Ikkbeta. Science 293: 1673–
1677, 2001.
270. Yue TL, Bao W, Gu JL, Cui J, Tao L, Ma XL, Ohlstein EH, and
Jucker BM. Rosiglitazone treatment in Zucker diabetic Fatty rats is
associated with ameliorated cardiac insulin resistance and protection
from ischemia/reperfusion-induced myocardial injury. Diabetes 54: 554 –
562, 2005.
271. Zarain-Herzberg A, Yano K, Elimban V, and Dhalla NS. Cardiac
sarcoplasmic reticulum Ca(2⫹)-ATPase expression in streptozotocininduced diabetic rat heart. Biochem Biophys Res Commun 203: 113–120,
1994.
272. Zhou YT, Grayburn P, Karim A, Shimabukuro M, Higa M, Baetens
D, Orci L, and Unger RH. Lipotoxic heart disease in obese rats:
implications for human obesity. Proc Natl Acad Sci USA 97: 1784 –1789,
2000.
Downloaded from http://ajpheart.physiology.org/ by 10.220.33.6 on May 5, 2017
AJP-Heart Circ Physiol • VOL
291 • OCTOBER 2006 •
www.ajpheart.org