ANTIOXIDANTS & REDOX SIGNALING Volume 22, Number 17, 2015 ª Mary Ann Liebert, Inc. DOI: 10.1089/ars.2015.6305

FORUM REVIEW ARTICLE

Myocardial Metabolism in Diabetic Cardiomyopathy: Potential Therapeutic Targets Miranda M. Sung, Shereen M. Hamza, and Jason R.B. Dyck

Abstract

Significance: Cardiovascular complications in diabetes are particularly serious and represent the primary cause of morbidity and mortality in diabetic patients. Despite early observations of cardiac dysfunction in diabetic humans, cardiomyopathy unique to diabetes has only recently been recognized. Recent Advances: Research has focused on understanding the pathogenic mechanisms underlying the initiation and development of diabetic cardiomyopathy. Emerging data highlight the importance of altered mitochondrial function as a major contributor to cardiac dysfunction in diabetes. Mitochondrial dysfunction occurs by several mechanisms involving altered cardiac substrate metabolism, lipotoxicity, impaired cardiac insulin and glucose homeostasis, impaired cellular and mitochondrial calcium handling, oxidative stress, and mitochondrial uncoupling. Critical Issues: Currently, treatment is not specifically tailored for diabetic patients with cardiac dysfunction. Given the multifactorial development and progression of diabetic cardiomyopathy, traditional treatments such as anti-diabetic agents, as well as cellular and mitochondrial fatty acid uptake inhibitors aimed at shifting the balance of cardiac metabolism from utilizing fat to glucose may not adequately target all aspects of this condition. Thus, an alternative treatment such as resveratrol, which targets multiple facets of diabetes, may represent a safe and promising supplement to currently recommended clinical therapy and lifestyle changes. Future Directions: Elucidation of the mechanisms underlying the initiation and progression of diabetic cardiomyopathy is essential for development of effective and targeted treatment strategies. Of particular interest is the investigation of alternative therapies such as resveratrol, which can function as both preventative and mitigating agents in the management of diabetic cardiomyopathy. Antioxid. Redox Signal. 22, 1606–1630.

Introduction

C

ardiovascular disease (CVD) is the principal cause of morbidity and mortality among diabetic patients (38). Although diabetes-related heart failure (HF) was first observed more than 40 years ago (213, 219), it is only recently recognized that diabetics are predisposed to a distinctive cardiomyopathy that is independent of accompanying macro/ microvascular complications typically associated with this disease (16). Indeed, early epidemiological studies showed that independent of age, obesity, or dyslipidemia, diabetes alone was linked with a two-fold increased risk of HF in men and a five-fold increased risk in women (129). Further to this, diabetic patients also comprise 15%–20% of subjects in HF clinical trials (54), and there is evidence that diabetes itself is

an independent predictor of poor outcomes in HF (86, 235). The association between diabetes and cardiac dysfunction in humans is also supported by data from animal models, which demonstrate distinct diabetes-induced cardiac structural, metabolic, and functional changes (99, 122, 288). This specific cardiac phenotype has been defined as diabetic cardiomyopathy and is characterized by early impairment in diastolic function with development of cardiomyocyte hypertrophy, myocardial fibrosis, and cardiomyocyte apoptosis, all of which are in the absence of coronary artery disease (CAD), hypertension, or valvular dysfunction. Although the term ‘‘diabetic cardiomyopathy’’ is now an accepted clinical definition, making a formal diagnosis of diabetic cardiomyopathy remains difficult as what constitutes this condition remains vague and there is no widely accepted

Department of Pediatrics, Cardiovascular Research Centre, University of Alberta, Edmonton, Canada.

1606

MITOCHONDRIAL STRESS IN DIABETIC CARDIOMYOPATHY

diagnostic criteria. Although there is some uncertainty regarding the detailed, step-wise pathogenesis of this condition (223), the earliest signs of diabetic cardiomyopathy are left ventricular hypertrophy (LVH) and diastolic dysfunction in type 1 and type 2 diabetic (T1D and T2D) patients. Although not routinely performed in standard clinical echocardiograms, more rigorous methods of diagnostic echocardiography (65) demonstrate that early mild diastolic dysfunction occurs often in diabetic patients. In fact, impaired left ventricular (LV) relaxation and filling is evident in 21%–75% of asymptomatic diabetic patients who do not have overt CVD or microvascular complications (62, 272, 286, 289, 290), demonstrating that diabetic cardiomyopathy is not a rare condition (212). Moreover, diastolic dysfunction is also observed in rodent models of T1D and T2D (116, 216), in the absence of overt vascular dysfunction or atherosclerosis, suggesting a cardiac-specific response to diabetes. In agreement with this, abnormalities are also evident in cardiomyocytes isolated from T2D rats (88), providing further support that diabetes directly affects the cardiomyocyte. Although it is not entirely clear what initiates the early stages of diabetic cardiomyopathy, possible factors contributing to the development of diastolic dysfunction in diabetes include systemic insulin resistance and hyperglycemia (9), increased cardiac collagen deposition (180), altered cardiomyocytespecific glucose and lipid metabolism (14) and oxidative stress (115, 116, 216), as well as dysregulation of calcium (Ca2 + ) homeostasis (88). Whether these alterations occurring within the cardiomyocyte also contribute to a gradual progression of diabetic cardiomyopathy and the presence of systolic dysfunction (272, 276, 286) that may occur later in the pathogenesis of diabetic cardiomyopathy is currently unknown. Despite the complex and multifactorial pathophysiology of diabetic cardiomyopathy, it is becoming clear that alterations in cardiac energy substrate metabolism are critically involved. Since glucose utilization as a source of energy by the cardiomyocyte is impaired in diabetes (73) and there is an abundant supply of free fatty acids (FFAs), hearts from diabetics have dramatic changes in their metabolic profile (177, 269). As a result, the diabetic heart depends almost completely on fatty acid oxidation (FAO) for adenosine triphosphate (ATP) synthesis (83, 112). Although FAO rates are accelerated, the excess supply and subsequent uptake of lipids into the cardiomyocyte still appears to exceed FA

FIG. 1. Summary of mechanisms contributing to the development of diabetic cardiomyopathy.

1607

utilization, leading to elevated FA storage within the cardiomyocyte. It appears that accumulation of lipid precedes the cardiac dysfunction (177), suggesting that they contribute to the eventual decline in cardiac performance. Consistent with this, the combination of ectopic FA accumulation within the cardiomyocyte and the over-reliance on FAO, consequently, compromises mitochondrial respiratory function and leads to elevated reactive oxygen species (ROS) production by the mitochondria. This contributes to eventual mitochondrial damage (269), thus linking lipotoxicity, oxidative stress, and mitochondrial dysfunction (73). This is extremely important, as impaired mitochondrial function is a key pathophysiological trigger for the development of diabetic cardiomyopathy (124) and may thus represent the final common pathway, leading to cardiac dysfunction (Fig. 1). Although it is recognized that diabetes imposes an extra burden on the heart (due to alterations in cardiac energy metabolism and increased cardiomyocyte oxidative stress and apoptosis that is characteristic of this disease), current therapeutic options do not specifically target diabetes-induced cardiac dysfunction. Clinically, cardiac dysfunction is managed equally between diabetic and non-diabetic patients (67) with no specific therapy recommended at the preclinical stage of diabetic cardiomyopathy. Considering that the perturbations involved in the development of diabetic cardiomyopathy may contribute to the high cardiac mortality in this patient population (63, 89), it is essential to understand the mechanisms underlying this condition to improve disease management and reduce patient morbidity and mortality. As such, we will present evidence that demonstrates the presence of mitochondrial dysfunction in diabetes and how this contributes to the pathogenesis of diabetic cardiomyopathy as well as discuss potential therapeutic targets to prevent or manage this condition. Lastly, we will review exciting recent data supporting the application of metabolic modulators and the natural polyphenol resveratrol in the management of diabetes and discuss the potential use of these therapies in the specific treatment of diabetic cardiomyopathy. Alterations in Myocardial Metabolism and Mitochondrial Function in Diabetes

A growing body of evidence has shown that altered energy metabolism and mitochondrial dysfunction is present in

1608

hearts from patients with insulin resistance and diabetes (202, 224). It has been proposed that altered myocardial energy metabolism resulting from changes in energy substrate supply and utilization, as well as altered mitochondrial function, can ultimately impact cardiac function, leading to the development of diabetic cardiomyopathy (14, 24). Indeed, using magnetic resonance imaging and phosphorus-31-nuclear MR spectroscopy, Diamant et al. (66) showed that diastolic function was impaired and myocardial high-energy phosphate metabolism (as measured by the phosphocreatine [PCr]/ATP ratio) was significantly reduced in asymptomatic normotensive male patients (50–65 years old) with wellcontrolled and uncomplicated T2D compared with control subjects. These findings were confirmed by Clarke and colleagues (224), who showed that T2D patients with normal cardiac mass and function have impaired cardiac energy metabolism as evidenced by a > 30% reduction in PCr/ATP ratio. They also showed that myocardial functional and metabolic decreases were negatively correlated with plasma FFA concentration and positively correlated with plasma glucose in all subjects (224). Furthermore, using positron emission tomography (PET), it was shown that increased body mass index and impaired glucose tolerance were strongly correlated with increased myocardial oxygen consumption (MVO2), reduced cardiac efficiency, and increased myocardial FA metabolism (uptake and utilization) in young, obese, and insulin-resistant women (203). Together, these findings suggest that alterations in cardiac energetics and mitochondrial function occur early in the development of diabetic cardiomyopathy before the onset of overt cardiac dysfunction, thus supporting the concept that altered mitochondrial metabolism may be causative in this pathology. Although a link between impaired cardiac function/metabolism and T2D has been observed in humans, not all human studies have found significant changes in high energy phosphate metabolism or cardiac energy substrate metabolism in obese or diabetic patients (136, 215, 264, 270). However, these studies do not necessarily discount the role of altered mitochondrial metabolism in diabetic cardiomyopathy, as differences in sex, age, diet and lifestyle, endogenous myocardial fat content, and severity of hyperglycemia and diabetes may have influenced the results. Fortunately, this possible conflict as to whether or not cardiac energy substrate metabolism is altered in obese or diabetic patients may have been resolved by recent reports performing direct mitochondrial measurements on atrial tissue samples from diabetic patients. Specifically, Anderson et al. (4) demonstrated that permeabilized atrial myofibers from T2D individuals have significantly decreased capacity for glutamate and FAsupported respiration, as well as increased intra-myocardial triglycerides (TAG), compared with non-diabetic patients. In addition, diabetic patients had increased mitochondrial hydrogen peroxide (H2O2) emission, which may result from a reduced capacity of the thioredoxin 2 system in the mitochondria. Usually, this system neutralizes the natural production of H2O2 by mitochondria; however, when this antioxidant system is compromised, H2O2 production is unchecked and allowed to accumulate (243), thus interfering with mitochondrial and cellular functions. In addition to diabetic patients demonstrating increased mitochondrial H2O2 production, they also display chronically elevated levels of oxidative stress in atrial tissue as shown

SUNG ET AL.

by increased levels of 4-hydroxy-2-nonenal (HNE)—and 3nitrotyrosine—modified proteins. Interestingly, both HNE and tyrosine nitrosylation are known to impair the activity of several complexes of the electron transport chain (ETC) found in the mitochondrial inner membrane and matrix (52), suggesting that impaired ETC activity may contribute to mitochondrial dysfunction. In addition, Montaigne et al. (182) showed, using isolated trabeculae from the atrial appendage, that diabetes was associated with a dramatic increase in myocardial oxidative stress. These authors also used electron microscopy to show that mitochondria from diabetic patients were smaller, had a reduced capacity to retain Ca2 + and reduced expression of mitochondrial fission protein (mitofusin-1), suggesting that there are changes in mitochondrial ultrastructure and dynamics that contribute to the development of cardiomyopathy (182). Furthermore, these impairments in mitochondrial function were associated with impaired contractility (182). Overall, the aforementioned human studies provide direct evidence demonstrating the presence of structural and functional deficiencies in cardiac mitochondria, increased oxidative stress, impaired Ca2 + handling, and myocardial lipid accumulation, all of which have been proposed to be important contributing mechanisms in the pathogenesis of diabetic cardiomyopathy. Since there are inherent limitations to performing mechanistic studies in humans, many of the insights into the underlying mechanisms responsible for diabetic cardiomyopathy have come from studying rodent models. Although rats/mice are relatively resistant to developing atherosclerosis and subsequent CAD, which is commonly found in humans with diabetes (119), this actually allows for more targeted dissection of pathways in the absence of confounding co-morbidities (148). Moreover, human studies of diabetic cardiomyopathy often exclude patients with ischemic heart disease, CAD, or hypertension in attempts to better understand the pathogenesis of diabetic cardiomyopathy in the absence of these confounding diseases (66). Consistent with the concept that diabetic cardiomyopathy is a distinct clinical entity, many of the findings in rodent models of obesity, insulin resistance, and diabetes mimic what has been observed in humans with these same pathologies. In addition, rodent models are amenable to genetic manipulation, allowing researchers to examine the role of specific proteins and signaling pathways in this disease. This not only provides insights into the pathogenesis of diabetic cardiomyopathy but also assists in the identification of potential targets that may be useful for the therapeutic prevention and/or treatment of diabetic cardiomyopathy. Mechanisms of Mitochondrial Dysfunction in Diabetic Cardiomyopathy

The causes of diabetic cardiomyopathy are multifactorial and involve several closely related mechanisms that often feedforward, creating a vicious cycle of cardiomyocyte damage. However, the initiating event appears to be an early maladaptation in cardiac energy metabolism contributing to mitochondrial dysfunction and, eventually, impaired cardiac contractility. For the purpose of this review, we not only will focus mainly on the etiology of diabetic cardiomyopathy in T2D but also will briefly touch on T1D and some of the similarities and differences in their pathogenesis. Although the

MITOCHONDRIAL STRESS IN DIABETIC CARDIOMYOPATHY

initial cause of impaired glucose utilization is different between these two forms of diabetes, ultimately they share similar downstream metabolic consequences of a switch to greater cardiomyocyte FA utilization at the expense of glucose. Altered cardiac energy metabolism

The heart has one of the highest energy demands of any organ in the body and must continually produce ATP to sustain proper contractile function and ionic homeostasis (166, 187, 194). During normal physiological conditions, the healthy adult heart obtains > 95% of its ATP from mitochondrial oxidative phosphorylation (OXPHOS) (195). Therefore, changes in mitochondrial function and ATP production can have negative effects on contractility and it is clear that early derangements in cardiac energy metabolism contribute to the development of impaired cardiac function in diabetes (14, 26, 32, 229). Interestingly, the healthy heart relies on a balance between the oxidation of carbohydrates, FAs, and ketones as fuel sources, and it has the flexibility to switch substrates depending on physiological and pathological conditions (166). However, in the diabetic, the heart is characterized by reduced metabolic flexibility with a decrease in glucose utilization and an increased reliance on FAs via accelerated FAO (166). The shift in fuel preference away from glucose and toward increased FAO in the diabetic heart comes at a higher oxygen (O2) cost and reduces cardiac efficiency (ratio of cardiac work to MVO2) (25, 32, 75, 112, 175). In clinical studies using PET and 11C-palmitate, it has been shown that MVO2, FA uptake, and FAO are increased in obese young women (203) and in patients with T1D (108). Increased MVO2 and reduced cardiac efficiency is also the result of FA-induced mitochondrial uncoupling of ATP synthesis from MVO2, which may lead to generation of ROS, energy depletion, and cardiac dysfunction (25, 26). This increased use of FAs in the hearts of diabetics is dependent on several factors, including (i) FA supply and uptake into the cardiomyocyte, (ii) rate of TAG lipolysis, and (iii) mitochondrial function and FAO [refer to (166) for review]. All of these factors have been shown to be altered in the diabetic heart, leading to an overall chronic increase in FA utilization, which is known to impair cardiac function (11). FA supply and uptake into the cardiomyocyte. Hyperlipidemia, which is commonly observed in obesity and diabetes, arises when excessive lipolysis in expanded adipose tissue leads to elevated circulating plasma levels of FAs and TAG (121). Apart from this, accelerated synthesis of verylow-density lipoprotein TAG from the liver and chylomicrons from dietary absorption also contributes to high plasma FA levels (117, 159, 190). Although FFAs can cross the plasma membrane by passive diffusion, facilitated transport plays a major role in myocardial FA uptake (96). To date, three families/types of fatty acid transport proteins (FATPs) have been identified, including FATP, fatty acid binding protein (FABP), and CD36. Both CD36 (39, 57, 137, 169, 197, 249) and FABP (169) levels have been shown to increase in hearts from diabetic animals, suggesting that these proteins may also be involved in increasing FA uptake in the diabetic heart. Indeed, regardless of the mechanism of entry into the cardiomyocytes, elevated levels of cardiac FA uptake have been observed in rodent models of diabetes, including

1609

streptozotocin (STZ)-induced T1D (168), obese Zucker rats (56, 169), db/db mice (39), and rodents fed a high-fat diet (HFD) (96, 98, 197) [see (33) for review of animal models of diabetic cardiomyopathy], supporting the notion that excess FA uptake is a critical early step in the development of diabetic cardiomyopathy. While FATP, FABP, and CD36 have been shown to play a role in mediating FA transport into the cardiomyocytes (96), CD36 accounts for more than 50% of FA uptake in the heart (144, 185). Translocation of CD36 from intracellular stores to the sarcolemma is necessary for long chain FA uptake to occur (138, 170), and increased sarcolemmal content of CD36 is observed in the diabetic heart (56). Importantly, increased expression of CD36 has also been shown in cardiac myocytes from several T2D models (249), suggesting that CD36 may play an important role in increasing myocardial FA use in the diabetic. Consistent with this, inhibition of FA uptake via genetic ablation of CD36 protects the heart from age- and diet-induced myocardial FA accumulation and cardiac dysfunction (137, 249). In addition, deletion of CD36 rescued hearts in a mouse model of a cardiac-specific diabetic phenotype by reducing lipid accumulation, increasing glucose oxidation, and restoring cardiac function (283). Although preventing FA uptake via CD36 inhibition appears to improve myocardial metabolism by switching metabolism from FA to glucose, there are also other potential benefits discussed in the lipotoxicity section that may contribute to improved cardiac function in the diabetic. Rate of TAG lipolysis. Once inside the cell, FAs are rapidly esterified and converted into fatty acyl-CoA esters by a family of long chain acyl-CoA synthetases (ACSL) (76). Subsequently, FA have three major fates, including direct delivery to the mitochondria for oxidation, esterification to TAG for temporary storage in cytoplasmic lipid droplets, or delivery to the nucleus for the activation of gene transcription (132). Synthesis and hydrolysis of TAG stores are a tightly controlled and dynamic process that can regulate cardiac metabolism and function under basal conditions (133, 134). Therefore, it is not surprising that TAG hydrolases such as adipose triglyceride lipase (ATGL) play important roles in pathological conditions such as diabetic cardiomyopathy (31, 134). As ATGL is the rate-limiting TAG hydrolase in the heart (101, 102, 133), it was originally hypothesized that myocardial ATGL expression would be decreased in diabetes and that decreased ATGL activity would account for increased TAG levels (211). However, we have shown that increased cardiac expression of ATGL is associated with elevated TAG levels in mouse models of T1D (211) and HFD-induced T2D (210). Based on this, it was proposed that increased myocardial ATGL expression during diabetes was a compensatory but insufficient response to the increased FA availability (210, 211). Indeed, mice with a cardiac overexpression of ATGL (myosin heavy chain [MHC]-ATGL) were protected against intra-myocardial TAG accumulation, lipotoxicity, and cardiac dysfunction in STZ and HFD mouse models (210, 211). Since FAs liberated from the TAG pool can activate nuclear peroxisome proliferator-activated receptor-a (PPAR-a) signaling to further promote FA utilization (101, 102, 133, 134, 210, 211), TAG catabolism via ATGL strongly influences FA utilization in the heart by both supplying the heart with FAs for FAO and potentially activating PPAR-a/PGC1

1610

signaling and improving mitochondrial function (see section below). That said, the role that ATGL plays in the regulation of PPAR-a/PGC1 signaling in the heart has yet to be fully defined as different models appear to have some conflicting data (102, 133, 210, 211). Nevertheless, the regulation of TAG catabolism via ATGL may be a novel approach for a therapy aimed at manipulating cardiac energy metabolism. Thus, by targeting ATGL with the aim of mitigating FA supply to the heart, it may be possible to reduce FAO rates and the sequelae that eventually lead to cardiomyocyte and mitochondrial dysfunction—factors critical in the treatment of diabetic cardiomyopathy (Fig. 2). Mitochondrial function and FAO. The significant increase in cardiomyocyte FA utilization in the diabetic can not only

SUNG ET AL.

be influenced by the degree of FA availability and increased FA transport but can also be attributed to increased FAmediated activation of PPAR-a (17, 75, 231). PPAR-a is an important nuclear transcription factor that increases expression of several genes encoding for proteins that promote FA utilization, including carnitine palmitoyal transferase (cpt)1, fatp1, cd36, and uncoupling proteins (UCPs) ucp2, and ucp3. In addition to facilitating FA uptake and oxidation, PPAR-a also increases expression of pyruvate dehydrogenase kinase 4, an inhibitor of pyruvate dehydrogenase (PDH), the key enzyme in glucose oxidation (74, 84, 94, 184). Of relevance, increased PPAR-a expression and activity has been reported in several rodent models of diabetic cardiomyopathy (84). The importance of PPAR-a activation in diabetic cardiomyopathy was demonstrated in mice with cardiac-specific overexpression of PPAR-a (MHC-PPAR-a). These MHCPPAR-a mice displayed dramatically elevated levels of cardiac FA uptake, storage, and oxidation with a concurrent reduction in glucose oxidation and glucose uptake that mimicked a diabetic phenotype. Most importantly, these metabolic changes are also associated with LVH and contractile dysfunction (84). In agreement with the overexpression of PPAR-a, mice that are deficient in PPAR-a have reduced expression of genes involved in FA utilization and reduced myocardial FAO rates (6, 153). Therefore, these two mouse models provide strong evidence that activation of PPAR-a plays an important role in mediating the observed derangements in myocardial energy metabolism and likely plays a key role in the development of cardiac dysfunction in diabetic cardiomyopathy. Excessive lipid accumulation

FIG. 2. Altered FA metabolism in diabetic cardiomyopathy. j FAs are taken up into the cell by passive diffusion across the plasma membrane or facilitated transport via FATPs, including CD36, FATP, and FABP. Once inside the cell, FAs can be k oxidized by the mitochondria to generate ATP, l esterified into the TAG pool to be stored temporarily, or m activate PPAR-a in the nucleus to promote transcription of genes involved in FA utilization, leading to increased FA uptake, FAO, and mitochondrial biogenesis. n Alternatively, FAs can also be liberated from the TAG pool by the hydrolase activity of ATGL that can be delivered to the mitochondria for oxidation or nucleus to activate PPAR-a. ATGL, adipose triglyceride lipase; ATP, adenosine triphosphate; FA, fatty acid; FABP, fatty acid binding protein; FAO, fatty acid oxidation; FATP, fatty acid transport protein; PPAR-a, peroxisome proliferator-activated receptor-a; TAG, triacylglycerol/triglycerides.

Although hearts from obese and diabetic animals are known to have accelerated rates of FAO, many of the animal models studied also have excessive intra-myocardial lipid accumulation (56). Therefore, it appears that, despite increased rates of FAO in the diabetic heart, FA uptake is well in excess of its capacity to be utilized. This imbalance between FA uptake and utilization results in the storage of FAs in the cardiomyocyte in the form of TAGs and/or conversion to other potentially detrimental lipid intermediates (31). Several studies have shown that excessive myocardial lipid accumulation, also known as cardiac steatosis, is deleterious to the heart, leading to numerous cardiac pathologies, including increased cardiomyocyte apoptosis, myocardial fibrosis, LVH, impaired diastolic filling, and contractile dysfunction (83, 84, 297). To establish whether or not perturbations in cardiomyocyte lipid uptake and metabolism directly contribute to cardiac dysfunction, several mouse models were generated with overexpression of FA-handling proteins (i.e., acyl-CoA synthetase 1, FATP1 and PPAR-a). All of these mouse models displayed a cardiac phenotype that was characterized by excessive lipid uptake, resulting in cardiac steatosis and impaired systolic function (50, 51, 133, 282). Interestingly, even in the absence of obesity or systemic metabolic disturbances such as hyperglycaemia, perturbed myocardial FA handling in these mice recapitulates the cardiac phenotype observed in diabetic cardiomyopathy. Overall, these findings suggest that excessive FA uptake and accumulation in the cardiomyocyte has detrimental effects on the heart, leading to diabetes-induced deterioration of function.

MITOCHONDRIAL STRESS IN DIABETIC CARDIOMYOPATHY

Elevated myocardial TAG levels have been associated with cardiac dysfunction in several genetic- and diet-induced models of diabetes (2, 31, 39, 91, 101, 133, 210). However, it is likely that lipotoxicity does not arise from an accumulation of TAG per se, but from the increased generation of potentially detrimental lipid intermediates, including ceramides, diacylglyercides (DAG), long chain acyl CoAs, and/or acylcarnitines (31). As such, there are several mechanisms by which excessive FA storage or cardiac steatosis can be potentially lipotoxic and lead to adverse effects on cardiac function (31). Moreover, ceramides and DAG have emerged as some of the more toxic lipid species that can lead to impaired insulin signaling (128) as well as to apoptotic cell death (109). Although ceramides are a key component making up cellular membranes, ceramides also act as a signaling molecule that can trigger apoptosis by inducing the release of cytochrome C from mitochondria (93). In addition, ceramide and DAG can lead to impaired cardiac insulin signaling by suppressing tyrosine phosphorylation of insulin receptors and insulin receptor substrates, thereby inhibiting activation of the PI3K/Akt pathway involved in insulin signaling (128). Increased intra-myocardial ceramide and DAG levels have been observed in numerous rodent models of diabetic cardiomyopathy and are associated with cardiac dysfunction (84, 267, 292, 297). Moreover, pharmacological and genetic inhibition of ceramide synthesis reduced cardiac ceramide levels and palmitate oxidation rates, as well as restored glucose oxidation rates and cardiac function in mouse models of lipotoxicity (199, 267). Similarly, mice with cardiac-specific overexpression of hormone-sensitive lipase, an important lipase catalyzing the hydrolysis of DAG, protected hearts from STZ-induced cardiac lipid accumulation and mortality (265). Although cardiac FAO is known to be accelerated in diabetes, some studies in genetically engineered mice have suggested that a further increase of FAO may aid in reducing the degree of cardiac lipid accumulation and in improving

1611

cardiac function (162, 241). As a whole, these data in animal models support the idea that accumulation of ceramides, DAG, or acylcarnitines is potentially lipotoxic, resulting in the worsening of insulin resistance and cardiac dysfunction in diabetes. Alternatively, Koves et al. (142) have shown that improving complete FAO in the mitochondria may reduce skeletal muscle insulin resistance. Therefore, potentially a similar approach to increase complete FAO in the heart may also be beneficial to improve insulin resistance. As such, therapies aimed at reducing the accumulation of ceramides and DAG by inhibiting their production or reducing FA uptake into the cardiomyocyte could prove to be protective in diabetic cardiomyopathy (Table 1). Impaired cardiac insulin signaling and glucose metabolism

It is known that the heart also becomes insulin resistant in T2D, and this cardiac insulin resistance is associated with cardiac contractile dysfunction (32). However, in contrast to the considerable amount of work involved in investigating the effects of excess myocardial lipids in diabetes, less is known about the mechanisms that mediate insulin resistanceinduced cardiac dysfunction. As such, it is still not completely clear whether/how impaired cardiac insulin signaling and/or glucose utilization contribute to the development of cardiac dysfunction in the diabetic heart (34). In addition, since animal models of T2D and diet-induced models also commonly display hyperglycemia, hyperlipidemia, obesity, and fluctuations in hormones, it is difficult to tease out the direct effects of impaired cardiac insulin signaling. However, a mouse model with a cardiomyocyte-specific deletion of the insulin receptor (CIRKO mice) has been generated that has allowed testing of whether impaired insulin signaling in the absence of systemic alterations in metabolism influences cardiac function (24). Studies using the CIRKO mice have shown that these mice have reduced insulin-stimulated glucose uptake and gradually

Table 1. Mitochondrial Dysfunction and Altered Metabolism in Animal Models of Type 1 and Type 2 Diabetes and Selected Transgenic Models of Diabetes Cardiac function Type 1 diabetes Streptozotocin Y (85, 211) Ins2 + / - Akita Y (35) OVE26 Y (234, 284) Type 2 diabetes db/db Y (26, 32, 70) ob/ob [/Y (32, 175) UCP-DTA ZDF Y (97, 297) Genetically modified mice MHC-PPARa Y (84) CIRKO Y (13, 24) icATGL KO Y (133) FATP1 Y (50) ACSL1 Y (51) hLpLGPI Y (282)

Mitochondrial function Y (85, 149) Y (35, 36) Y (234) Y (26) Y (25) Y (75)

Y (24)

Cardiac efficiency Y (112) — (35) Y (26, 32) Y (32, 175)

— (24)

Fatty acid oxidation

Glucose oxidation

Lipid accumulation

[ (112) [ (35)

Y (112) Y (35) Y (284)

[ (211) [ (211)

[ (14, 26, 32) [ (32, 175)

Y (26, 32) Y (32, 175)

[ (103) [ (32)

[ (275)

Y (42, 97)

[ (297)

Y (84) Y (13)

[ (84)

[ Y Y [

(84) (13) (133) (50)

[ [ [ [

(133) (50) (51) (282)

ACSL1, acyl-CoA synthetase 1; hLpLGPI, glycosylphosphatidylinositol anchored human lipoprotein lipase; icATGL KO, inducible cardiomyocyte-specific ATGL knockout; MHC, myosin heavy chain; UCP-DTA, uncoupling protein-diphtheria toxin A; ZDF, Zucker diabetic fatty rat; [, increase; Y, decrease; —, no change.

1612

develop a decrease in glucose and FAO, as well as modest cardiac dysfunction with age (13, 24). Furthermore, hearts from CIRKO mice have impaired mitochondrial respiration and ATP synthesis rates, as well as reduced gene expression of OXPHOS mitochondrial proteins, FA handling proteins, and tricarboxylic acid (TCA) cycle proteins (24). CIRKO mice also exhibit increased mitochondrial ROS generation and mitochondrial uncoupling (24), suggesting that impaired cardiac insulin signaling is linked to mitochondrial dysfunction and oxidative stress (Table 1). Conversely, overexpression of human GLUT4 glucose transporter in hearts of db/db mice normalized cardiac glucose and FA metabolism and restored cardiac function in these hearts (229), suggesting that promoting glucose oxidation, in this case by increasing glucose uptake via GLUT4, can restore normal cardiac function. Therefore, targeting the alterations in cardiac energy metabolism specifically by improving insulin sensitivity and promoting glucose metabolism may be effective in preventing the development of diabetic cardiomyopathy. Since cardiac insulin resistance is one of the earliest cardiac manifestations of T2D, impaired cardiac insulin signaling may be an important mechanism in the progression of diabetic cardiomyopathy. Contrary to this, some studies have shown that the hyperinsulinemia present in T2D leads to persistent activation of the insulin receptor that occurs in the absence of increased basal downstream Akt activation (55, 175). Since it has been shown in healthy mice that insulin may directly impair badrenoreceptor-regulated cardiac contractility (87), it has been suggested that this may provide a potential explanation of reduced myocardial inotropic reserve in pathological states of hyperinsulinemia such as T2D. Therefore, further investigation into the role of insulin signaling and adrenergic response in the development of diabetic cardiomyopathy is needed. Data from experimental animal studies suggest that common anti-diabetic treatments targeted at improving systemic insulin resistance and glycemic control such as metformin (53, 77, 285) and troglitazone (214) are also able to prevent moderate cardiac dysfunction in the early stages of T2D. Indeed, some clinical studies, including the UK Prospective Diabetes Study and Hyperinsulinemia: the Outcome of its Metabolic Effects trials, have shown that metformin reduces cardiovascular risk for patients with T2D (1, 139). However, the recent double-blinded, placebo-controlled trial Carotid Atherosclerosis: MEtformin for insulin ResistAnce showed little cardiovascular benefit in non-diabetic patients with coronary heart disease using carotid intima-media thickness as a primary endpoint (209). Apart from this, the Glycometabolic Intervention as Adjunct to Primary Percutaneous Intervention in ST Elevation Myocardial Infarction Trial (GIPS-III; NCT01217307) showed that metformin did not improve LVEF in non-diabetic patients with ST-segment elevation myocardial infarction who underwent percutaneous coronary intervention (154). These findings have called into question whether metformin indeed has cardiovascular benefit. Further evidence from large trials of cardiovascular outcomes such as the Metformin in CABG trial (MetCAB; NCT01438723) will shed more light on this question. Altered cytosolic and mitochondrial calcium handling

Cardiomyocyte contraction is highly dependent on the concentrations of intracellular Ca2 + and ATP. It is becoming

SUNG ET AL.

increasingly evident that alterations in cardiomyocyte Ca2 + handling may be critical in the regulation of cardiac function in diabetic cardiomyopathy (15, 70, 82, 143, 165, 188, 200, 296). Usually, Ca2 + influx into the cell via the L-type Ca2 + channel triggers sarcoplasmic reticulum (SR) Ca2 + release, which raises cytosolic free [Ca2 + ]. This leads to binding of Ca2 + to troponin C (TnC), causing a conformational change and allowing the interaction of actin and myosin to produce contraction. For ventricular relaxation and diastolic filling to occur, cytosolic [Ca2 + ] must decline allowing the dissociation of Ca2 + from TnC. Ca2 + removal from the cytosol is mediated primarily by the actions of the ATP-dependent sarcoplasmic reticulum Ca2 + -ATPase 2a (SERCA2a), which transports Ca2 + back into the SR. Indeed, depressed cardiac function is known to be associated with impaired SRmediated Ca2 + handling in both T1D (143, 165, 188, 200, 296) and T2D (15, 70, 82). Reduced SERCA2a pump activity is the result of decreased expression and/or Ca2 + affinity of SERCA2a or increased expression of its inhibitory protein phospholamban (15, 70, 82, 143, 165, 188, 200, 296). Although the precise mechanism responsible for reduced SERCA2a activity is not known, it has been proposed that oxidative stress results in increased formation of advanced glycation endproducts and O-linked b-N-acetylglucosamine modifications of the SR membrane and proteins, leading to depressed SERCA2a expression and/or function (20, 156). The functional significance of impaired Ca2 + handling is shown in mice with a cardiomyocyte-specific deletion of SERCA2a that develop LV dysfunction (22), whereas overexpression of SERCA2a improves systolic and diastolic function in STZ-induced T1D mice presumably by improving SR Ca2 + sequestration (246, 262). Therefore, perturbations in myocyte Ca2 + handling may lead to cardiac dysfunction and likely contribute to the development of diabetic cardiomyopathy, making it a current therapeutic target of interest. In addition, impaired sympathetic nervous system activity has been observed in T1 and T2 diabetic hearts and has been linked to the progression of diabetic cardiomyopathy (68). Earlier, this was characterized by hyperactivity and elevated norepinephrine (NE) spillover that can progress to sympathetic denervation and reduced NE release, a condition known as cardiac autonomic neuropathy (258). Sympathetic hyperactivation coupled with hyperglycemia in T2D may lead to a rise in ROS levels and impaired mitochondrial respiration, which can exacerbate impaired Ca2 + handling and excitation-contraction coupling, leading to contractile dysfunction in the diabetic heart (261). New research has focused on the role of mitochondrial Ca2 + handling in diabetic cardiomyopathy. Although the primary function of the mitochondria is ATP generation, it can also store Ca2 + , which allows mitochondria to act as a Ca2 + buffer to prevent cytosolic Ca2 + overload (64). Ca2 + is rapidly transported down the electrochemical gradient into the mitochondrial matrix via a Ca2 + uniporter protein complex and extruded via the Na + /Ca2 + exchanger on a beat-tobeat basis in the heart, contributing to fast buffering during the excitation–contraction coupling (73, 171, 250). Consequently, increased cytosolic Na + levels found in myocytes from diabetic animals may also impair mitochondrial Ca2 + uptake (171). Since Ca2 + regulates several important cellular processes, such as muscle contraction, the close proximity of

MITOCHONDRIAL STRESS IN DIABETIC CARDIOMYOPATHY

mitochondria to SR Ca2 + release sites suggests that Ca2 + shuttling between the cytosol and the mitochondria may represent a signaling network coordinating the rate of ATP production with its demand for cardiac contraction (34, 72, 82, 95, 118). There are several mitochondrial enzymes that are Ca2 + sensitive, including TCA cycle enzymes isocitrate dehydrogenase and a-ketoglutarate dehydrogenase and PDH (95). Studies in isolated cardiac mitochondria show that elevations in mitochondrial Ca2 + result in activation of these Ca2 + sensitive dehydrogenases, as well as in increases in ATP production and TCA cycle activity (72, 104, 125, 217). To date, few studies have investigated mitochondrial Ca2 + handling defects in diabetic models (82, 85, 205, 255). However, Flarsheim et al. (85) showed that as early as 2 weeks after STZ treatment in rats (before the development of systolic cardiac dysfunction) there was a significant reduction in the rate of Ca2 + uptake into isolated cardiac mitochondria. This was associated with a reduced capacity to upregulate ATP synthesis via stimulation of PDH and a-ketoglutarate dehydrogenase (85). Therefore, reduced mitochondrial Ca2 + uptake may compromise ATP synthesis rates required for proper cardiac contraction and likely contribute to eventual cytosolic Ca2 + overload, thereby contributing to the development of diabetic cardiomyopathy (Fig. 3).

FIG. 3. Impaired cellular and mitochondrial Ca21 handling in diabetic cardiomyopathy. Cytosolic Ca2 + concentrations increase due to influx of Ca2 + into the cardiomyocyte via L-type Ca2 + channels, which trigger release of stored Ca2 + from the SR. Increased cytosolic Ca2 + is necessary for actin–myosin interaction and cardiac contraction, and removal of intracellular Ca2 + is required for relaxation and ventricular filling, in part, via re-uptake of Ca2 + into the SR via the SERCA2a. In diabetes, a reduction in SERCA2a activity and expression results in reduced Ca2 + uptake into the SR and, eventually, leads to cytosolic Ca2 + overload and impaired cardiac contractile function. Furthermore, mitochondrial Ca2 + uptake is impaired in diabetes, which decreases activation of Ca2 + sensitive hydrogenases PDH and TCA cycle enzymes. Ultimately, this reduces mitochondrial ATP production and contributes to reduced myocardial contractility. Ca2 + , calcium; PDH, pyruvate dehydrogenase; SERCA, sarcoplasmic reticulum Ca2 + -ATPase; SR, sarcoplasmic reticulum; TCA, tricarboxylic acid.

1613

Oxidative stress

ROS are essential for normal function and activity of cell signaling molecules (43). However, ROS production is under tight physiological control to prevent oxidative damage. Oxidative stress results from an imbalance between the generation of ROS and endogenous antioxidant capacity (280). Depending on the accumulated level of ROS, these highly reactive molecules can induce many pathophysiological effects on the cell, including direct oxidation of DNA, proteins, and lipid (280), as well as activation of stresssensitive pathways leading to cell damage (80). ROS production in cardiomyocytes can also trigger and advance the cardiac remodeling process as shown by blunted cardiomyocyte hypertrophy in response to antioxidant treatment (186). Oxidative stress is also a key regulator of myocardial fibrosis (295), which is a feature of cardiac remodeling in diabetic cardiomyopathy. Many studies show increased ROS generation in models of diabetic cardiomyopathy as well as reduced endogenous antioxidant defenses in diabetes (3, 173). Indeed, Tocchetti et al. (261) demonstrated that deficient mitochondrial glutathione and thioredoxin antioxidant systems in the T2D db/db mouse account for excess mitochondrial ROS production. They also showed that this deficiency in mitochondrial antioxidant capacity was a causative factor in impaired excitationcontraction coupling in isolated cardiomyocytes or whole heart preparations (261). As expected, oxidative stress in diabetic cardiomyopathy correlates with excess lipid delivery and elevated mitochondrial FAO rates (269), factors that are characteristic in the setting of obesity and diabetes (73). Although the direct impact of lipid oxidation in mitochondria is unclear, it has been observed that lipids can inhibit or uncouple OXPHOS (279). Hyperglycemia itself induces an overproduction of ROS and is believed to be a causal link between increased circulating glucose and the onset and progression of diabetic complications such as diabetic cardiomyopathy (218). In fact, a hyperglycemia-induced increase in electron transfer donors (such as NADH and FADH2) results in increases in electron flow through the mitochondrial ETC. Subsequently, the ATP/ ADP ratio increases and mitochondrial membrane potential is hyperpolarized. Downstream effects of this include an accumulation of electrons to coenzyme Q, driving the generation of the free radical superoxide. Thus, it has been proposed that these hyperglycemia-induced events result in the generation of ROS that is, in turn, believed to be the fundamental driver of mitochondrial dysfunction that figures so prominently in diabetes-related complications such as diabetic cardiomyopathy (218). Lastly, in the heart, the predominant sources of ROS are from NADPH oxidases, uncoupled nitric oxide synthases (NOS), and, importantly, the mitochondrial respiratory chain itself, which is the main intracellular source of free radicals in the diabetic heart. In addition, recent evidence suggests that monoamine oxidase A and B are contributors to the excessive ROS production characteristic of cardiac dysfunction (126, 127). Indeed, monoamine oxidases as well as the sources of ROS mentioned earlier are upregulated in diabetes and believed to directly contribute to the pathogenesis of diabetic cardiomyopathy (115, 116, 145, 189). NADPH oxidase is a major source of ROS in cardiovascular cells, including cardiomyocytes (228), and is usually

1614

activated to generate superoxide to destroy pathogens. It comprises several Nox isoforms, with the primary cardiac isoforms being Nox 2 and Nox 4 (281). NADPH oxidase activity and expression is upregulated in the setting of LVH and HF (160, 293) and is also increased in diabetes in conjunction with myocardial hypertrophy and fibrosis (115, 230). Another source of ROS, in addition to NADPH oxidase, is uncoupled endothelial nitric oxide synthase (eNOS), which can generate superoxide in the cardiovascular system. Usually, eNOS generates NO in the vasculature via tightly controlled electron transfer (135). However, if there is uncoupling between O2 reduction and NO generation, superoxide is produced. This eNOS uncoupling is present in diabetes and contributes to oxidative stress in this condition (81). Another major source of myocardial ROS is the mitochondrial respiratory chain (71). Mitochondria, as the site of OXPHOS, usually generate superoxide under physiological conditions over the course of the energy-generating pathway (27). This normal ROS generation is quenched by endogenous anti-oxidant mechanisms. However, if the inner mitochondrial membrane potential rises beyond a certain level, extra ROS generation is triggered (140). Such a rise in membrane potential can be a result of increased glucose or FAO (239), the latter of which is certainly increased in diabetic cardiomyopathy. Cardiac ROS accumulation can subsequently cause contractile dysfunction via damage of intracellular organelles and proteins. This generation of ROS is critical in physiological signaling. However, under normal conditions, ROS is quenched by endogenous anti-oxidant mechanisms, thus maintaining a delicate balance between ROS production for normal cellular function and scavenging to prevent damage to cellular components and tissues. The link between mitochondrial respiration and ROS generation is unclear and a topic of some debate in the literature given the recognized role of oxidative stress-induced mitochondrial dysfunction in diseases such as diabetic cardiomyopathy. One theory suggests that once the inner mitochondrial membrane potential rises beyond a certain level, extra ROS generation is triggered (140). Such a rise in membrane potential can be a result of increased FAO (239), which is certainly increased in diabetic cardiomyopathy. Conversely, emerging evidence suggests that a loss of mitochondrial membrane potential may compromise the ROS-scavenging capacity of this organelle, thus contributing to increased ROS accumulation (5). Indeed, mononuclear cells and platelets from T2D patients exhibit reduced mitochondrial membrane potential (100, 277), which is believed to contribute to the vascular dysfunction characteristic of this disease. This recently introduced, novel perspective is termed the Redox-Optimized ROS Balance hypothesis. This hypothesis shifts the focus away from mitochondrial membrane potential as the driving force for ROS generation and instead is centered on the cellular redox environment. In an optimal (physiologically stable) redox environment, ROS production is minimal and supportive of normal cellular signaling processes. By contrast, in either highly reduced or highly oxidized redox environments, ROS accumulation or overproduction can result. Based on this proposal, mitochondrial ROS generation is modulated by the balance between ROS production and scavenging and this balance is disrupted at reduced or oxidized extremes. Thus, in a reduced cellular environment, slow electron flow through the respi-

SUNG ET AL.

ratory chain predisposes the production of superoxide, whereas in an oxidized cellular environment, ROS-scavenging capacity is impaired, resulting in ROS accumulation (5, 58). Although this is an interesting hypothesis that merits further investigation, based on the existing evidence it is difficult to draw a definitive conclusion regarding the mechanism of mitochondrial ROS production in diabetic cardiomyopathy at this time. In addition to increased ROS production, dysfunctional ROS scavenging in the diabetic heart contributes to the severity of oxidative stress. In fact, overexpression of the endogenous antioxidant catalase can restore cardiomyocyte contractility in T1D mice (284). The increase in cardiac oxidative stress also occurs due to cytosolic ROS-induced exacerbation of mitochondrial ROS production (90). This is a key feature in diabetic cardiomyopathy, in which oxidative damage to mitochondria leads to further mitochondrial dysfunction (298, 299). Although the mitochondria are themselves equipped with powerful antioxidant defenses to protect against ROS generation, when this mechanism is compromised, the mitochondrial structures suffer oxidative damage, and mitochondrial DNA are particularly sensitive (274). In diabetic cardiomyopathy, mitochondrial oxidative damage sets in motion a vicious cycle whereupon mitochondrial oxidative damage exacerbates ROS production from these organelles, leading to increased cardiac ROS accumulation and further mitochondrial oxidative damage. Anderson et al. (4) not only demonstrated mitochondrial dysfunction in atrial tissue of diabetic patients, including those with tight glycemic control, but also showed increased mitochondrial peroxide release and depletion of mitochondrial antioxidant molecules (4). The importance of oxidative stress in the pathogenesis of diabetic cardiomyopathy is also highlighted by studies in which endogenous antioxidants (superoxide dismutase, catalase, and glutathione peroxidase) have been overexpressed. These studies demonstrate a significant improvement in contractile function in diabetic rodents (174, 234, 284) and highlight the importance of investigating the consequences of ROS in the diabetic heart and strategies to reduce this oxidative stress as discussed later in this review (Fig. 4). Although ROS can contribute to mitochondrial oxidative damage in hearts from diabetics, ROS can also further compromise mitochondrial function via induction of the mitochondrial permeability transition (106, 113, 114). This phenomenon involves opening of the mitochondrial permeability pore in the inner membrane of mitochondria, which leads to further mitochondrial injury and cell death. Mitochondria utilize electron transport to generate an electrochemical gradient across the inner membrane, which is used by ATP synthase to generate ATP. The inner membrane must be virtually impermeable to ions to sustain this membrane potential and, thus, ATP-generating capacity of the mitochondria. Given the sensitivity of the membrane permeability pore to ROS, the opening of this complex can occur in the setting of diabetes and results in depolarization of the inner membrane potential, causing ATP synthase to operate in reverse and shifts mitochondria from ATP producers to ATP consumers (238). This rapidly depletes cellular energy levels and leads to susceptibility to injury and, eventually, cell death. ROS also sensitize the mitochondrial permeability transition pore to Ca2 + induced opening (8), leading to Ca2 + overload and further

MITOCHONDRIAL STRESS IN DIABETIC CARDIOMYOPATHY

1615

FIG. 4. Pathophysiological mechanisms contributing to mitochondrial dysfunction in diabetes. The characteristic increase in FFAs drives increased cellular FA uptake (via FA transporters such as FATP and CD36) mediated by increased PPAR-a and PGC-1a activity. Increased FA uptake fuels an increase in FAO, precipitating increased ROS production and mitochondrial oxidative damage. The ensuing mitochondrial dysfunction triggers increased mitochondrial ROS production and in conjunction with reduced mitochondrial anti-oxidative capacity and increased cellular ROS, drives a vicious cycle resulting in further mitochondrial oxidative damage. Increased FFA stimulates increased UCP activity, which results in mitochondrial uncoupling, reduced ATP production, and compromised cardiac function. Additional mechanisms contributing to mitochondrial dysfunction include a diabetes-induced increase in mitochondrial biogenesis and altered mitochondrial Ca2 + handling and membrane phospholipid content. AGEs, advanced glycation endproducts; FFA, free fatty acid; O-GlcNAc, O-linked b-N-acetylglucosamine; PGC-1a, peroxisome proliferator-activated receptor c co-activator 1a; ROS, reactive oxygen species; UCP, uncoupling protein. exacerbating the membrane permeability transition (192). In line with this, heart mitochondria from T2D rats accumulate Ca2 + more rapidly than controls (191). Overall, mitochondria from diabetic human and animal hearts are more susceptible to opening of the membrane permeability transition pore (18, 193, 278). This appears to be the case for both types of diabetes, since T1D rat hearts demonstrated a greater propensity for ischemic injury that was attributed to ROS-dependent induction of the mitochondrial permeability transition (240). Thus, not only can ROS lead to contractile dysfunction, but also the high circulating ROS characteristic of diabetes can further impair mitochondrial function by sensitizing the mitochondrial membrane permeability transition pore to opening in response to ROS or circulating Ca2 + , leading to Ca2 + overload, loss of ATP-generating capacity, mitochondrial rupture, and, ultimately, cardiomyocyte death (Fig. 4). Consistent with the importance of the mitochondrial permeability transition pore in diabetes, inhibition of the mitochondrial permeability transition is shown to be cardioprotective in humans (206). Mitochondrial uncoupling

As introduced earlier, a potential mechanism underlying the mismatch between increased MVO2 and reduction in cardiac energy generation and efficiency in diabetes may involve mitochondrial uncoupling and UCPs. In non-cardiac tissues such as brown adipose, UCPs are important for generating heat by uncoupling O2 consumption from ATP production (236). Although the precise role of UCPs in the heart is not fully understood, UCP homologues 2 and 3 are expressed in the myocardium (150) and mitochondrial un-

coupling has been described in diabetic rodent hearts (23, 25, 26). In the mitochondria, UCPs are believed to play a physiological role in the setting of increased mitochondrial substrate flux to decrease the proton gradient and thus limit ROS production (27, 28). Mitochondrial UCPs channel protons from the intermembrane space into the matrix space to bypass F0F1-ATPase, thus reducing ATP production. However, the rise in ROS in diabetes may persistently activate mitochondrial UCPs, resulting in inappropriate dissipation of the proton gradient, chronically reducing ATP production, which ultimately compromises cardiac function. This is the essential basis of the general mechanism leading to mitochondrial uncoupling in diabetic cardiomyopathy and is supported by data demonstrating a restoration of mitochondrial proton gradient to normal levels after inhibition of UCPs (26). However, it must be noted that mitochondrial uncoupling may be a feature unique to T2D, since experimental models of T1D do not show impaired cardiac efficiency or mitochondrial uncoupling although cardiac FAO is increased (35). Although it is currently unclear why this is the case, a new hypothesis describing modulation of mitochondrial ROS production has been proposed and is termed RedoxOptimized ROS Balance (see above for discussion) (5, 58). Another potential differentiating factor may be insulin resistance. Supporting this idea are data demonstrating the development of FA-induced mitochondrial uncoupling in the setting of euglycemia in mice featuring cardiomyocytespecific deletion of insulin receptors (24). Thus, increased UCP expression in hearts from diabetics may have a role in mitochondrial uncoupling that is observed in diabetic cardiomyopathy and, ultimately, contribute to the reduction in

1616

SUNG ET AL.

cardiac energy production and efficiency despite increased FAO rates (Fig. 4). In addition to UCPs, apolipoprotein O (APOO) was recently discovered to be localized to the inner mitochondrial membrane and also causes mitochondrial uncoupling, increased O2 consumption, and ROS production (263). Of importance, APOO was found to be highly expressed in hearts from diabetic patients (147) and in rodent hearts after 9 weeks of HFD (204). Cardiac overexpression of APOO in mice fed an HFD led to mitochondrial structural abnormalities and modestly impaired systolic function (263). Together, these data suggest that APOO-induced proton leak and increased respiration may eventually lead to depletion of NADH and FADH substrates, which the cardiomyocyte may attempt to compensate for by increasing FA uptake. However, this increased FA uptake may eventually be lipotoxic, contributing to oxidative mitochondrial damage as discussed earlier and facilitating the development of cardiac dysfunction in diabetes. Mitochondrial biogenesis

Mitochondrial biogenesis is an intricate and continuous process that controls the number and composition of mitochondria in all tissues (232). Mechanisms that stimulate increased mitochondrial biogenesis also trigger increased mitochondrial DNA replication and assembly of proteins into mitochondrial matrix and membranes (111). Although diabetes increases the initiation and progression of mitochondrial biogenesis (123), the specific targets of diabetes in this process have not been clearly defined. It is known that the transcriptional co-activator peroxisome proliferatoractivated receptor c co-activator 1a (PGC-1a) is a major regulator of mitochondrial biogenesis (131). Indeed, cardiomyocyte-specific overexpression of PGC-1a not only results in a significant increase in cardiomyocyte mitochondrial number but also causes contractile dysfunction (222). Conversely, cardiac-specific knockout of PGC-1a reduces expression of genes related to FAO, TCA cycle, and OXPHOS without altering the density of the mitochondria (7). Since increased mitochondrial biogenesis is characteristic of diabetes (172), it is possible that activation of PGC-1a could be involved in regulating this process. However, this has not been established definitively. Paradoxically, although mitochondrial content is increased in the hearts of insulin resistant and diabetic animals, mitochondrial respiration and ATP synthesis is actually reduced (75). As such, stimulation of mitochondrial biogenesis is a potentially adaptive mechanism in the face of compromised mitochondrial function and cardiac energetics in diabetic cardiomyopathy. However, this adaptation is not matched with an equal compensation of mitochondrial respiration, so it is currently not known whether this is a beneficial/adaptive or detrimental/maladaptive mechanism (Fig. 4). Metabolic Treatments for Diabetic Cardiomyopathy

Since the diabetic heart exhibits major metabolic changes such as increased FAO, reduced glucose oxidation, and impaired insulin sensitivity and these appear to be detrimental to cardiac function, therapies aimed at inhibiting myocardial FA metabolism and reciprocally promoting glucose metabolism may be a promising approach to the treatment of diabetic

FIG. 5. Targets of metabolic inhibitors in the treatment of diabetic cardiomyopathy. FA uptake into the cardiomyocyte can be inhibited via specific CD36 inhibitor AP5055 and AP5258. The CPT1 inhibitors etomoxir, perhexiline, and MCD inhibitors can inhibit FA uptake into the mitochondria for FAO and ATP production. The partial inhibitors of FAO, including trimetazidine and ranolazine, inhibit FAO via inhibition of specific FAO enzymes. These agents shift myocardial energy metabolism in diabetes from FAO to glucose oxidation and, ultimately, reduce cardiac dysfunction associated with diabetic cardiomyopathy. CM, chylomicrons; CPT1, carnitine palmitoyal transferase 1; LPL, lipoprotein lipase; MCD, malonyl-CoA decarboxylase; TG, triacylglyceride; VLDL, very low density lipoprotein.

cardiomyopathy. Existing approaches include inhibition of cellular and mitochondrial FA uptake and partial inhibition of FAO (Fig. 5). Inhibitors of cellular and mitochondrial FA uptake

Since CD36 appears to play a central role in mediating excessive myocardial FA uptake in hearts from diabetic animals (249, 283), small-molecule inhibitors of CD36 that block FA binding to CD36, and thus reduce uptake, have recently been developed (92). Administration of the CD36 inhibitors AP5055 or AP5258 for 3 weeks to the Zucker diabetic fatty rat diabetic rat reduced fasting and dietary plasma glucose levels by 20%–50% (92). CD36 is also well known for its role in the uptake of oxidized low-density lipoprotein into macrophages, and AP5055 also protected against the development of atherosclerosis via formation of smaller plaques with less lipid deposition (92). These CD36 inhibitors are not only yet to be tested in regards to their potential effects in diabetic cardiomyopathy but may also prove to be effective to prevent

MITOCHONDRIAL STRESS IN DIABETIC CARDIOMYOPATHY

excessive FA accumulation in the heart, which could, subsequently, improve function. Several studies have shown that direct inhibition of mitochondrial FA uptake is an effective approach to shift myocardial energy metabolism from FA to glucose utilization (120). CPT1 is an important rate-limiting step that is responsible for long chain FA uptake into the mitochondria for subsequent oxidation and is under the transcriptional control of PPAR-a (176). Inhibitors of CPT1 such as etomoxir and perhexiline, which inhibit mitochondrial FA uptake, have been studied for this purpose. Etomoxir is an irreversible CPT1 inhibitor that has been shown to inhibit FAO and cause reciprocal activation of glucose oxidation via actions on PDH (164, 167). When administered to diabetic rats, etomoxir was shown to increase glucose oxidation rates (273) and ameliorate cardiac dysfunction (107, 226). Interestingly, chronic treatment of rats with etomoxir also induces the expression of SERCA2a in the heart (271) and increases rates of SRmediated Ca2 + uptake (220, 221, 271), which may improve Ca2 + handling, ATP production, and cardiac function in the setting of diabetes. Consistent with animal studies, both etomoxir and perhexiline improve LV ejection fraction in patients with HF (152, 225). However, some caution is warranted with the use of etomoxir, as the Etomoxir for the Recovery of Glucose Oxidation randomized placebocontrolled study had to be stopped early due to several patients in the moderate HF group developing abnormal liver function (110). Although etomoxir and perhexiline have not been tested specifically for the treatment of diabetic cardiomyopathy in patients, both drugs improve cardiac function in patients with HF, a pathology that is also characterized by compromised metabolism and cardiac energy deficiency. Therefore, direct FA inhibitors hold promise to also be useful in patients with diabetic cardiomyopathy. In addition to drugs that directly inhibit CPT1, drugs that elevate malonyl CoA levels have also been designed to inhibit CPT1. One example of this is malonyl-CoA decarboxylase (MCD) inhibitors (45–47, 78). MCD catalyzes the degradation of malonyl-CoA to acetyl-CoA and removes inhibition of CPT1 by malonyl-CoA. MCD expression is highly regulated by PPAR-a transcriptional control (37, 61, 151), and studies have shown that cardiac MCD activity and expression are increased in HFD and STZ-induced diabetes likely via increased PPAR-a activation (37). The use of MCD inhibitors in animal models results in reduced FAO, increased glucose oxidation, and improved insulin sensitivity (244). In addition, mice with MCD deficiency are protected against the development of impaired insulin-stimulated glucose metabolism in response to a HFD (268). Therefore, inhibition of mitochondrial FA uptake via MCD inhibition may represent a novel mechanism for altering the balance of cardiac energy metabolism in diabetic cardiomyopathy. However, it remains to be shown whether MCD inhibitors can result in improved cardiac function in animal models of diabetes and in humans. Partial inhibitors of FAO

Trimetazidine is a partial FAO inhibitor that competitively inhibits mitochondrial long chain 3-ketoacyl-CoA thiolase, the last enzyme in the FA b-oxidation pathway. By inhibiting FAO, trimetazidine shifts substrate preference to glucose

1617

oxidation (130, 163). Currently, trimetazidine is approved as an anti-anginal agent throughout Europe and Asia. Animal studies show that trimetazidine improves recovery after myocardial ischemia-reperfusion injury by reducing proton production and disturbances in ionic homeostasis arising from the uncoupling of glycolysis to glucose oxidation (130, 163). Apart from this, trimetazidine may exert favorable effects on oxidative stress and cardiomyocyte Ca2 + handling by improving SERCA2a activity and preventing intracellular Ca2 + overload (179), which are also important pathological mechanisms known to lead to diabetic cardiomyopathy. Recently, trimetazidine administered to db/db mice (157) and diet-induced obese mice (266) was found to significantly improve cardiac function and protect the heart against lipid accumulation and oxidative stress. Clinically, 6 months of trimetazidine treatment improved cardiac function and physical activity tolerance level in diabetic patients with idiopathic dilated cardiomyopathy (294). Similarly, ranolazine, which is also believed to act by suppressing FAO (227), has been shown to lower fasting glucose and HbA1c (49, 183, 260). Although less is known about the mechanism of action and effectiveness of ranolazine, it has been shown to reduce angina events in patients (141). Therefore, trimetazidine and ranolazine, which have already been clinically approved for the treatment of chronic stable angina, may also have beneficial effects in diabetic cardiomyopathy and could be combined with concurrent anti-diabetic medications, especially since they have the added benefit of improving glycemic control. Alternative treatments for diabetic cardiomyopathy— spotlight on resveratrol

The heart requires optimal mitochondrial function and ATP production to meet the high-energy demands of normal cardiac performance. In diabetes, mitochondrial function is compromised by the multiple mechanisms reviewed earlier, which leads to contractile dysfunction and, ultimately, diabetic cardiomyopathy. From the previous discussion, it is apparent that strategies that improve energy metabolism and substrate preference may have therapeutic benefit in preventing diabetic cardiomyopathy or improving myocardial function in diabetes. One approach that may provide additional benefits may be supplementation of existing treatment strategies with the natural compound resveratrol. Resveratrol (3,5,4¢-trihydroxy-trans-stilbene) is a polyphenol that is produced by plants and is present in a variety of plant-based foods. Many beneficial properties of resveratrol have been demonstrated, including cardioprotection (233), anti-cancer effects (155, 287), and lifespan extension (79, 196). Resveratrol also appears to show promising results in diabetes management (105) and may be worthy of consideration as an adjuvant to traditional diabetes treatment strategies. Although specific data regarding the impact of resveratrol treatment on diabetic cardiomyopathy are quite limited at this time, it is clear that resveratrol has systemic anti-diabetic effects in models of both T1D and T2D (207). Management of T1D and T2D involves blood glucose control, while management of T2D additionally involves preservation of pancreatic b-cells and improved insulin sensitivity. Resveratrol has been shown to improve these three facets of diabetes control (207). Thus, by reducing hyperglycemia, increasing

1618

tissue glucose uptake and preserving endogenous insulin secretory capacity (21, 44, 48, 198, 207, 237, 245, 251–254), resveratrol is able to control the systemic factors that contribute to the development of diabetic cardiomyopathy— without the negative side effects reported for traditionally used anti-diabetic agents such as metformin (59). In addition to systemic effects, resveratrol can enhance GLUT4 translocation via Akt signaling and increase PDH activity in hearts from diabetic animals (40, 201), suggesting that resveratrol may have direct effects to promote myocardial glucose uptake and metabolism in diabetes. As discussed earlier, oxidative stress is a significant contributor to cardiac mitochondrial dysfunction and the pathogenesis of diabetic cardiomyopathy. Evidence shows that resveratrol ameliorates oxidative stress in diabetic cardiomyopathy, in part, by inhibiting ROS production and increasing anti-oxidant capacity, which was associated with improved cardiac function (29, 40, 181, 242, 291). As increased FAO rates also lead to excessive ROS production, it is possible that resveratrol decreases oxidative stress by reducing FAO and increasing glucose utilization (40). In addition to anti-oxidant properties, resveratrol has been shown to improve SERCA2a expression and cardiac function in a mouse model of T1D and restore SERCA2a promoter activity that was otherwise repressed in a high glucose environment (247). The protective effects of resveratrol to restore SERCA2 promoter activity and cardiac function are dependent on inducing silent information regulator (SIRT) 1 expression and activity (247). Several studies have linked activation of SIRT1 and/or SIRT3 by resveratrol to be involved in ameliorating several forms of HF, including diabetic cardiomyopathy (256). Evidence also shows that resveratrol may also reduce oxidative stress through activation of SIRT1 (257). Thus, resveratrol may also prevent or ameliorate diabetic cardiomyopathy by reducing oxidative stress and improving cardiomyocyte Ca2 + -handling via SIRT1 activation and restoring SERCA2a expression and activation. In addition, some studies demonstrate modulation of mitochondrial function by resveratrol in peripheral tissues and suggest that a similar pathway may exist in the heart (10, 146). To date, only a limited number of studies have exam-

SUNG ET AL.

ined the effect of resveratrol on cardiac mitochondrial function, specifically in diabetes. These studies suggest that resveratrol may improve mitochondrial bioenergetics and reduce mitochondrial-derived ROS in diabetes (12). However, further investigation is needed to understand the effects of resveratrol treatment in hearts of obese and diabetic animals, whether resveratrol alters substrate metabolism and mitochondrial function, and ultimately whether this protects against cardiac dysfunction and other features characteristic of diabetic cardiomyopathy. In general, clinical studies of resveratrol for the treatment of CVD are limited, with even fewer focused on diabetes and its complications. To date, only a few clinical studies involving diabetic patients have been completed. These studies report modest reductions in blood glucose levels (19, 29) and improved insulin sensitivity (60, 161, 178). However, this remains controversial due to the wide variance in resveratrol dose, duration of treatment, and age/health of the subjects studied (208, 259). Nevertheless, resveratrol supplementation may be a fruitful avenue for future research using models of diabetic cardiomyopathy and clinical investigation in the diabetic patient population. Recently, it has been demonstrated that resveratrol accumulates in cardiac tissue in a dose and time-dependent manner, which is the first description of its kind (30). In addition, the accumulated tissue level of resveratrol was associated with a long-term improvement in cardiac function in diabetic animals, and almost complete recovery of hemodynamic measures at the highest dose (30). The early stages of diabetic cardiomyopathy characterized by diastolic dysfunction are, in fact, reversible in humans with diabetes who lose weight and normalize metabolism (158). Thus, the pathogenesis of diabetic cardiomyopathy includes this reversible phase, which emphasizes early and aggressive lifestyle modifications in diabetic patients. In this context, there is an opportunity for resveratrol treatment to significantly impact the prognosis of these patients, given that this polyphenol is also characterized as an exercise mimetic (69). While it is not possible to make a definitive conclusion about the clinical application of resveratrol at this time, it is certainly a feasible therapeutic supplement for prevention or management of diabetic cardiomyopathy that merits further investigation (Fig. 6).

FIG. 6. Targets of resveratrol in the treatment of diabetic cardiomyopathy. The naturally occurring polyphenol resveratrol displays multiple systemic and cardiacspecific effects that may be effective in the treatment of diabetic cardiomyopathy, including anti-diabetic, antioxidant, modulation of mitochondrial function, and Ca2 + dynamics.

MITOCHONDRIAL STRESS IN DIABETIC CARDIOMYOPATHY Conclusion

Given that diabetic cardiomyopathy is now recognized to have a high prevalence in the diabetic population, early treatment and screening is optimal to prevent a further decline in cardiac performance. Thorough investigation and understanding of the mechanisms underlying the initiation and progression of diabetic cardiomyopathy has shed considerable light on the important mediators of diabetic cardiomyopathy. It is clear that cardiac insulin resistance and reduced glucose oxidation are early features of diabetes. This, together with increased circulating FA levels result in accelerated FA uptake and FAO and activation of cardiac PPAR-a-mediated gene transcription, which further promotes FA metabolism at the expense of glucose. Over time, mitochondrial dysfunction and impaired cardiac function develops, a phenomenon that is driven by excessive FA uptake, ectopic cardiac lipid accumulation, impaired cytosolic and mitochondrial Ca2 + handling, and increased oxidative stress. All these are key mediators of diabetic cardiomyopathy that contribute to the initiation and exacerbation of this dysfunction. Although drugs that target FA uptake and oxidation have been investigated, current clinical evidence does not yet support approved use of these agents in patients for the treatment of diabetic cardiomyopathy. However, it is possible that the natural polyphenol resveratrol represents an alternative therapeutic strategy to target the multiple pathophysiological factors, which predispose damage to mitochondria as well as targeting facets of mitochondrial function itself. Indeed, experimental data support a role for resveratrol in mitigating mitochondrial damage and the progression of diabetic cardiomyopathy, suggesting that resveratrol may have a promising impact on the future management of diabetic cardiomyopathy by having pleiotropic effects. In summary, diabetic cardiomyopathy remains an important, although sometimes under-appreciated, consequence of obesity and diabetes. Given that CVD is the principal cause of morbidity and mortality among diabetic patients and diabetic cardiomyopathy likely contributes to this, effective therapies for treating diabetic cardiomyopathy may be of tremendous benefit to this growing patient population. Acknowledgments

These authors acknowledge support from Alberta Innovates–Health Solutions (AIHS), Heart and Stroke Foundation of Canada (HSFC), and Canadian Institutes for Health Research (CIHR). References

1. Effect of intensive blood-glucose control with metformin on complications in overweight patients with type 2 diabetes (UKPDS 34). UK Prospective Diabetes Study (UKPDS) Group. Lancet 352: 854–865, 1998. 2. 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 BM 17.0744, a novel PPAR-alpha activator. Am J Physiol Heart Circ Physiol 283: H949–H957, 2002. 3. Aliciguzel Y, Ozen I, Aslan M, and Karayalcin U. Activities of xanthine oxidoreductase and antioxidant enzymes in different tissues of diabetic rats. J Lab Clin Med 142: 172–177, 2003.

1619

4. Anderson EJ, Kypson AP, Rodriguez E, Anderson CA, Lehr EJ, and Neufer PD. Substrate-specific derangements in mitochondrial metabolism and redox balance in the atrium of the type 2 diabetic human heart. J Am Coll Cardiol 54: 1891–1898, 2009. 5. Aon MA, Cortassa S, and O’Rourke B. Redox-optimized ROS balance: a unifying hypothesis. Biochim Biophys Acta 1797: 865–877, 2010. 6. Aoyama T, Peters JM, Iritani N, Nakajima T, Furihata K, Hashimoto T, and Gonzalez FJ. Altered constitutive expression of fatty acid-metabolizing enzymes in mice lacking the peroxisome proliferator-activated receptor alpha (PPARalpha). J Biol Chem 273: 5678–5684, 1998. 7. Arany Z, He H, Lin J, Hoyer K, Handschin C, Toka O, Ahmad F, Matsui T, Chin S, Wu PH, Rybkin II, Shelton JM, Manieri M, Cinti S, Schoen FJ, Bassel-Duby R, Rosenzweig A, Ingwall JS, and Spiegelman BM. Transcriptional coactivator PGC-1 alpha controls the energy state and contractile function of cardiac muscle. Cell Metab 1: 259–271, 2005. 8. Baines CP. The mitochondrial permeability transition pore and ischemia-reperfusion injury. Basic Res Cardiol 104: 181–188, 2009. 9. Bajraktari G, Koltai MS, Ademaj F, Rexhepaj N, Qirko S, Ndrepepa G, and Elezi S. Relationship between insulin resistance and left ventricular diastolic dysfunction in patients with impaired glucose tolerance and type 2 diabetes. Int J Cardiol 110: 206–211, 2006. 10. Baur JA, Pearson KJ, Price NL, Jamieson HA, Lerin C, Kalra A, Prabhu VV, Allard JS, Lopez-Lluch G, Lewis K, Pistell PJ, Poosala S, Becker KG, Boss O, Gwinn D, Wang M, Ramaswamy S, Fishbein KW, Spencer RG, Lakatta EG, Le Couteur D, Shaw RJ, Navas P, Puigserver P, Ingram DK, de Cabo R, and Sinclair DA. Resveratrol improves health and survival of mice on a high-calorie diet. Nature 444: 337–342, 2006. 11. Bayeva M, Sawicki KT, and Ardehali H. Taking diabetes to heart—deregulation of myocardial lipid metabolism in diabetic cardiomyopathy. J Am Heart Assoc 2: e000433, 2013. 12. Beaudoin MS, Perry CG, Arkell AM, Chabowski A, Simpson JA, Wright DC, and Holloway GP. Impairments in mitochondrial palmitoyl-CoA respiratory kinetics that precede development of diabetic cardiomyopathy are prevented by resveratrol in ZDF rats. J Physiol 592: 2519– 2533, 2014. 13. Belke DD, Betuing S, Tuttle MJ, Graveleau C, Young ME, Pham M, Zhang D, Cooksey RC, McClain DA, Litwin SE, Taegtmeyer H, Severson D, Kahn CR, and Abel ED. Insulin signaling coordinately regulates cardiac size, metabolism, and contractile protein isoform expression. J Clin Invest 109: 629–639, 2002. 14. 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. 15. Belke DD, Swanson EA, and Dillmann WH. Decreased sarcoplasmic reticulum activity and contractility in diabetic db/db mouse heart. Diabetes 53: 3201–3208, 2004. 16. Bell DS. Diabetic cardiomyopathy. A unique entity or a complication of coronary artery disease? Diabetes Care 18: 708–714, 1995. 17. Bernal-Mizrachi C, Weng S, Feng C, Finck BN, Knutsen RH, Leone TC, Coleman T, Mecham RP, Kelly DP, and

1620

18.

19. 20.

21. 22.

23. 24.

25.

26.

27.

28. 29.

30.

SUNG ET AL.

Semenkovich CF. Dexamethasone induction of hypertension and diabetes is PPAR-alpha dependent in LDL receptor-null mice. Nat Med 9: 1069–1075, 2003. Bhamra GS, Hausenloy DJ, Davidson SM, Carr RD, Paiva M, Wynne AM, Mocanu MM, and Yellon DM. Metformin protects the ischemic heart by the Akt-mediated inhibition of mitochondrial permeability transition pore opening. Basic Res Cardiol 103: 274–284, 2008. Bhatt JK, Thomas S, and Nanjan MJ. Resveratrol supplementation improves glycemic control in type 2 diabetes mellitus. Nutr Res 32: 537–541, 2012. 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. Bloomgarden ZT. Diabetes complications. Diabetes Care 27: 1506–1514, 2004. Boardman NT, Aronsen JM, Louch WE, Sjaastad I, Willoch F, Christensen G, Sejersted O, and Aasum E. Impaired left ventricular mechanical and energetic function in mice after cardiomyocyte-specific excision of Serca2. Am J Physiol Heart Circ Physiol 306: H1018–H1024, 2014. Boudina S and Abel ED. Mitochondrial uncoupling: a key contributor to reduced cardiac efficiency in diabetes. Physiology (Bethesda) 21: 250–258, 2006. Boudina S, Bugger H, Sena S, O’Neill BT, Zaha VG, Ilkun O, Wright JJ, Mazumder PK, Palfreyman E, Tidwell TJ, Theobald H, Khalimonchuk O, Wayment B, Sheng X, Rodnick KJ, Centini R, Chen D, Litwin SE, Weimer BE, and Abel ED. Contribution of impaired myocardial insulin signaling to mitochondrial dysfunction and oxidative stress in the heart. Circulation 119: 1272–1283, 2009. 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. Boudina S, Sena S, Theobald H, Sheng X, Wright JJ, Hu XX, Aziz S, Johnson JI, Bugger H, Zaha VG, and Abel ED. Mitochondrial energetics in the heart in obesityrelated diabetes: direct evidence for increased uncoupled respiration and activation of uncoupling proteins. Diabetes 56: 2457–2466, 2007. Brand MD, Affourtit C, Esteves TC, Green K, Lambert AJ, Miwa S, Pakay JL, and Parker N. Mitochondrial superoxide: production, biological effects, and activation of uncoupling proteins. Free Radic Biol Med 37: 755–767, 2004. Brand MD and Esteves TC. Physiological functions of the mitochondrial uncoupling proteins UCP2 and UCP3. Cell Metab 2: 85–93, 2005. Brasnyo P, Molnar GA, Mohas M, Marko L, Laczy B, Cseh J, Mikolas E, Szijarto IA, Merei A, Halmai R, Meszaros LG, Sumegi B, and Wittmann I. Resveratrol improves insulin sensitivity, reduces oxidative stress and activates the Akt pathway in type 2 diabetic patients. Br J Nutr 106: 383–389, 2011. Bresciani L, Calani L, Bocchi L, Delucchi F, Savi M, Ray S, Brighenti F, Stilli D, and Del Rio D. Bioaccumulation of resveratrol metabolites in myocardial tissue is dosetime dependent and related to cardiac hemodynamics in diabetic rats. Nutr Metab Cardiovasc Dis 24: 408–415, 2014.

31. Brindley DN, Kok BP, Kienesberger PC, Lehner R, and Dyck JR. Shedding light on the enigma of myocardial lipotoxicity: the involvement of known and putative regulators of fatty acid storage and mobilization. Am J Physiol Endocrinol Metab 298: E897–E908, 2010. 32. 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. 33. Bugger H and Abel ED. Rodent models of diabetic cardiomyopathy. Dis Model Mech 2: 454–466, 2009. 34. Bugger H and Abel ED. Mitochondria in the diabetic heart. Cardiovasc Res 88: 229–240, 2010. 35. Bugger H, Boudina S, Hu XX, Tuinei J, Zaha VG, Theobald HA, Yun UJ, McQueen AP, Wayment B, Litwin SE, and Abel ED. Type 1 diabetic akita mouse hearts are insulin sensitive but manifest structurally abnormal mitochondria that remain coupled despite increased uncoupling protein 3. Diabetes 57: 2924–2932, 2008. 36. Bugger H, Chen D, Riehle C, Soto J, Theobald HA, Hu XX, Ganesan B, Weimer BC, and Abel ED. Tissue-specific remodeling of the mitochondrial proteome in type 1 diabetic akita mice. Diabetes 58: 1986–1997, 2009. 37. Campbell FM, Kozak R, Wagner A, Altarejos JY, Dyck JR, Belke DD, Severson DL, Kelly DP, and Lopaschuk GD. A role for peroxisome proliferator-activated receptor alpha (PPARalpha) in the control of cardiac malonyl-CoA levels: reduced fatty acid oxidation rates and increased glucose oxidation rates in the hearts of mice lacking PPARalpha are associated with higher concentrations of malonyl-CoA and reduced expression of malonyl-CoA decarboxylase. J Biol Chem 277: 4098–4103, 2002. 38. Candido R, Srivastava P, Cooper ME, and Burrell LM. Diabetes mellitus: a cardiovascular disease. Curr Opin Investig Drugs 4: 1088–1094, 2003. 39. Carley AN, Atkinson LL, Bonen A, Harper ME, Kunnathu S, Lopaschuk GD, and Severson DL. Mechanisms responsible for enhanced fatty acid utilization by perfused hearts from type 2 diabetic db/db mice. Arch Physiol Biochem 113: 65–75, 2007. 40. Carolo Dos Santos K, Pereira Braga C, Octavio Barbanera P, Rodrigues Ferreira Seiva F, Fernandes Junior A, and Fernandes AA. Cardiac energy metabolism and oxidative stress biomarkers in diabetic rat treated with resveratrol. PLoS One 9: e102775, 2014. 41. This reference has been deleted. 42. Chatham JC and Seymour AM. Cardiac carbohydrate metabolism in Zucker diabetic fatty rats. Cardiovasc Res 55: 104–112, 2002. 43. Chen YR and Zweier JL. Cardiac mitochondria and reactive oxygen species generation. Circ Res 114: 524–537, 2014. 44. Cheng AY and Fantus IG. Oral antihyperglycemic therapy for type 2 diabetes mellitus. CMAJ 172: 213–226, 2005. 45. Cheng JF, Chen M, Wallace D, Tith S, Haramura M, Liu B, Mak CC, Arrhenius T, Reily S, Brown S, Thorn V, Harmon C, Barr R, Dyck JR, Lopaschuk GD, and Nadzan AM. Synthesis and structure-activity relationship of small-molecule malonyl coenzyme A decarboxylase inhibitors. J Med Chem 49: 1517–1525, 2006. 46. Cheng JF, Huang Y, Penuliar R, Nishimoto M, Liu L, Arrhenius T, Yang G, O’Leary E, Barbosa M, Barr R,

MITOCHONDRIAL STRESS IN DIABETIC CARDIOMYOPATHY

47.

48.

49.

50.

51.

52.

53.

54. 55.

56.

57.

58.

59.

Dyck JR, Lopaschuk GD, and Nadzan AM. Discovery of potent and orally available malonyl-CoA decarboxylase inhibitors as cardioprotective agents. J Med Chem 49: 4055–4058, 2006. Cheng JF, Mak CC, Huang Y, Penuliar R, Nishimoto M, Zhang L, Chen M, Wallace D, Arrhenius T, Chu D, Yang G, Barbosa M, Barr R, Dyck JR, Lopaschuk GD, and Nadzan AM. Heteroaryl substituted bis-trifluoromethyl carbinols as malonyl-CoA decarboxylase inhibitors. Bioorg Med Chem Lett 16: 3484–3488, 2006. Chi TC, Chen WP, Chi TL, Kuo TF, Lee SS, Cheng JT, and Su MJ. Phosphatidylinositol-3-kinase is involved in the antihyperglycemic effect induced by resveratrol in streptozotocin-induced diabetic rats. Life Sci 80: 1713– 1720, 2007. Chisholm JW, Goldfine AB, Dhalla AK, Braunwald E, Morrow DA, Karwatowska-Prokopczuk E, and Belardinelli L. Effect of ranolazine on A1C and glucose levels in hyperglycemic patients with non-ST elevation acute coronary syndrome. Diabetes Care 33: 1163–1168, 2010. 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. Chiu HC, Kovacs A, Ford DA, Hsu FF, Garcia R, Herrero P, Saffitz JE, and Schaffer JE. A novel mouse model of lipotoxic cardiomyopathy. J Clin Invest 107: 813–822, 2001. Choksi KB, Boylston WH, Rabek JP, Widger WR, and Papaconstantinou J. Oxidatively damaged proteins of heart mitochondrial electron transport complexes. Biochim Biophys Acta 1688: 95–101, 2004. Cittadini A, Napoli R, Monti MG, Rea D, Longobardi S, Netti PA, Walser M, Sama M, Aimaretti G, Isgaard J, and Sacca L. Metformin prevents the development of chronic heart failure in the SHHF rat model. Diabetes 61: 944– 953, 2012. Cohen-Solal A, Beauvais F, and Logeart D. Heart failure and diabetes mellitus: epidemiology and management of an alarming association. J Card Fail 14: 615–625, 2008. Cook SA, Varela-Carver A, Mongillo M, Kleinert C, Khan MT, Leccisotti L, Strickland N, Matsui T, Das S, Rosenzweig A, Punjabi P, and Camici PG. Abnormal myocardial insulin signalling in type 2 diabetes and leftventricular dysfunction. Eur Heart J 31: 100–111, 2010. 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. Coort SL, Willems J, Coumans WA, van der Vusse GJ, Bonen A, Glatz JF, and Luiken JJ. Sulfo-N-succinimidyl esters of long chain fatty acids specifically inhibit fatty acid translocase (FAT/CD36)-mediated cellular fatty acid uptake. Mol Cell Biochem 239: 213–219, 2002. Cortassa S, O’Rourke B, and Aon MA. Redox-optimized ROS balance and the relationship between mitochondrial respiration and ROS. Biochim Biophys Acta 1837: 287– 295, 2014. Cottart CH, Nivet-Antoine V, Laguillier-Morizot C, and Beaudeux JL. Resveratrol bioavailability and toxicity in humans. Mol Nutr Food Res 54: 7–16, 2010.

1621

60. Crandall JP, Oram V, Trandafirescu G, Reid M, Kishore P, Hawkins M, Cohen HW, and Barzilai N. Pilot study of resveratrol in older adults with impaired glucose tolerance. J Gerontol A Biol Sci Med Sci 67: 1307–1312, 2012. 61. Cuthbert KD and Dyck JR. Malonyl-CoA decarboxylase is a major regulator of myocardial fatty acid oxidation. Curr Hypertens Rep 7: 407–411, 2005. 62. Dandamudi S, Slusser J, Mahoney DW, Redfield MM, Rodeheffer RJ, and Chen HH. The prevalence of diabetic cardiomyopathy: a population-based study in Olmsted County, Minnesota. J Card Fail 20: 304–309, 2014. 63. Devereux RB, Roman MJ, Paranicas M, O’Grady MJ, Lee ET, Welty TK, Fabsitz RR, Robbins D, Rhoades ER, and Howard BV. Impact of diabetes on cardiac structure and function: the strong heart study. Circulation 101: 2271– 2276, 2000. 64. Dhalla NS, Rangi S, Zieroth S, and Xu YJ. Alterations in sarcoplasmic reticulum and mitochondrial functions in diabetic cardiomyopathy. Exp Clin Cardiol 17: 115–120, 2012. 65. Di Bonito P, Moio N, Cavuto L, Covino G, Murena E, Scilla C, Turco S, Capaldo B, and Sibilio G. Early detection of diabetic cardiomyopathy: usefulness of tissue Doppler imaging. Diabet Med 22: 1720–1725, 2005. 66. 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. 67. Dickstein K, Cohen-Solal A, Filippatos G, McMurray JJ, Ponikowski P, Poole-Wilson PA, Stromberg A, van Veldhuisen DJ, Atar D, Hoes AW, Keren A, Mebazaa A, Nieminen M, Priori SG, and Swedberg K. ESC guidelines for the diagnosis and treatment of acute and chronic heart failure 2008: the Task Force for the Diagnosis and Treatment of Acute and Chronic Heart Failure 2008 of the European Society of Cardiology. Developed in collaboration with the Heart Failure Association of the ESC (HFA) and endorsed by the European Society of Intensive Care Medicine (ESICM). Eur Heart J 29: 2388–2442, 2008. 68. Dimitropoulos G, Tahrani AA, and Stevens MJ. Cardiac autonomic neuropathy in patients with diabetes mellitus. World J Diabetes 5: 17–39, 2014. 69. Dolinsky VW, Jones KE, Sidhu RS, Haykowsky M, Czubryt MP, Gordon T, and Dyck JR. Improvements in skeletal muscle strength and cardiac function induced by resveratrol during exercise training contribute to enhanced exercise performance in rats. J Physiol 590: 2783–2799, 2012. 70. Dong F, Zhang X, Yang X, Esberg LB, Yang H, Zhang Z, Culver B, and Ren J. Impaired cardiac contractile function in ventricular myocytes from leptin-deficient ob/ob obese mice. J Endocrinol 188: 25–36, 2006. 71. Drose S, Brandt U, and Wittig I. Mitochondrial respiratory chain complexes as sources and targets of thiol-based redox-regulation. Biochim Biophys Acta 1844: 1344– 1354, 2014. 72. Duchen MR. Roles of mitochondria in health and disease. Diabetes 53 Suppl 1: S96–S102, 2004. 73. Duncan JG. Mitochondrial dysfunction in diabetic cardiomyopathy. Biochim Biophys Acta 1813: 1351–1359, 2011.

1622

74. Duncan JG and Finck BN. The PPARalpha-PGC-1alpha axis controls cardiac energy metabolism in healthy and diseased myocardium. PPAR Res 2008: 253817, 2008. 75. Duncan JG, Fong JL, Medeiros DM, Finck BN, and Kelly DP. Insulin-resistant heart exhibits a mitochondrial biogenic response driven by the peroxisome proliferatoractivated receptor-alpha/PGC-1alpha gene regulatory pathway. Circulation 115: 909–917, 2007. 76. Durgan DJ, Smith JK, Hotze MA, Egbejimi O, Cuthbert KD, Zaha VG, Dyck JR, Abel ED, and Young ME. Distinct transcriptional regulation of long-chain acyl-CoA synthetase isoforms and cytosolic thioesterase 1 in the rodent heart by fatty acids and insulin. Am J Physiol Heart Circ Physiol 290: H2480–H2497, 2006. 77. Dutta K, Podolin DA, Davidson MB, and Davidoff AJ. Cardiomyocyte dysfunction in sucrose-fed rats is associated with insulin resistance. Diabetes 50: 1186–1192, 2001. 78. 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. 79. 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. 80. Evans JL, Goldfine ID, Maddux BA, and Grodsky GM. Are oxidative stress-activated signaling pathways mediators of insulin resistance and beta-cell dysfunction? Diabetes 52: 1–8, 2003. 81. Faria AM, Papadimitriou A, Silva KC, Lopes de Faria JM, and Lopes de Faria JB. Uncoupling endothelial nitric oxide synthase is ameliorated by green tea in experimental diabetes by re-establishing tetrahydrobiopterin levels. Diabetes 61: 1838–1847, 2012. 82. Fauconnier J, Lanner JT, Zhang SJ, Tavi P, Bruton JD, Katz A, and Westerblad H. Insulin and inositol 1,4,5trisphosphate trigger abnormal cytosolic Ca2 + transients and reveal mitochondrial Ca2 + handling defects in cardiomyocytes of ob/ob mice. Diabetes 54: 2375–2381, 2005. 83. Finck BN, Han X, Courtois M, Aimond F, Nerbonne JM, Kovacs A, Gross RW, and Kelly DP. A critical role for PPARalpha-mediated lipotoxicity in the pathogenesis of diabetic cardiomyopathy: modulation by dietary fat content. Proc Natl Acad Sci U S A 100: 1226–1231, 2003. 84. 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. 85. Flarsheim CE, Grupp IL, and Matlib MA. Mitochondrial dysfunction accompanies diastolic dysfunction in diabetic rat heart. Am J Physiol 271: H192–H202, 1996. 86. From AM, Leibson CL, Bursi F, Redfield MM, Weston SA, Jacobsen SJ, Rodeheffer RJ, and Roger VL. Diabetes in heart failure: prevalence and impact on outcome in the population. Am J Med 119: 591–599, 2006. 87. Fu Q, Xu B, Liu Y, Parikh D, Li J, Li Y, Zhang Y, Riehle C, Zhu Y, Rawlings T, Shi Q, Clark RB, Chen X, Abel ED, and Xiang YK. Insulin inhibits cardiac contractility

SUNG ET AL.

88.

89.

90.

91.

92.

93.

94.

95. 96.

97.

98.

99. 100.

101.

by inducing a Gi-biased beta2-adrenergic signaling in hearts. Diabetes 63: 2676–2689, 2014. Fulop N, Mason MM, Dutta K, Wang P, Davidoff AJ, Marchase RB, and Chatham JC. Impact of Type 2 diabetes and aging on cardiomyocyte function and O-linked Nacetylglucosamine levels in the heart. Am J Physiol Cell Physiol 292: C1370–C1378, 2007. Galderisi M, Anderson KM, Wilson PW, and Levy D. Echocardiographic evidence for the existence of a distinct diabetic cardiomyopathy (the Framingham Heart Study). Am J Cardiol 68: 85–89, 1991. Garciarena CD, Caldiz CI, Correa MV, Schinella GR, Mosca SM, Chiappe de Cingolani GE, Cingolani HE, and Ennis IL. Na + /H + exchanger-1 inhibitors decrease myocardial superoxide production via direct mitochondrial action. J Appl Physiol (1985) 105: 1706–1713, 2008. Ge F, Hu C, Hyodo E, Arai K, Zhou S, Lobdell Ht, Walewski JL, Homma S, and Berk PD. Cardiomyocyte triglyceride accumulation and reduced ventricular function in mice with obesity reflect increased long chain fatty acid uptake and de novo fatty acid synthesis. J Obes 2012: 205648, 2012. Geloen A, Helin L, Geeraert B, Malaud E, Holvoet P, and Marguerie G. CD36 inhibitors reduce postprandial hypertriglyceridemia and protect against diabetic dyslipidemia and atherosclerosis. PLoS One 7: e37633, 2012. 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. 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. Glancy B and Balaban RS. Role of mitochondrial Ca2 + in the regulation of cellular energetics. Biochemistry 51: 2959–2973, 2012. Glatz JF, Luiken JJ, and Bonen A. Membrane fatty acid transporters as regulators of lipid metabolism: implications for metabolic disease. Physiol Rev 90: 367–417, 2010. 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. Greenwalt DE, Scheck SH, and Rhinehart-Jones T. Heart CD36 expression is increased in murine models of diabetes and in mice fed a high fat diet. J Clin Invest 96: 1382–1388, 1995. 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. Guo X, Wu J, Du J, Ran J, and Xu J. Platelets of type 2 diabetic patients are characterized by high ATP content and low mitochondrial membrane potential. Platelets 20: 588–593, 2009. Haemmerle G, Lass A, Zimmermann R, Gorkiewicz G, Meyer C, Rozman J, Heldmaier G, Maier R, Theussl C, Eder S, Kratky D, Wagner EF, Klingenspor M, Hoefler G,

MITOCHONDRIAL STRESS IN DIABETIC CARDIOMYOPATHY

102.

103.

104. 105.

106. 107.

108.

109.

110.

111. 112.

113. 114. 115.

and Zechner R. Defective lipolysis and altered energy metabolism in mice lacking adipose triglyceride lipase. Science 312: 734–737, 2006. Haemmerle G, Moustafa T, Woelkart G, Buttner S, Schmidt A, van de Weijer T, Hesselink M, Jaeger D, Kienesberger PC, Zierler K, Schreiber R, Eichmann T, Kolb D, Kotzbeck P, Schweiger M, Kumari M, Eder S, Schoiswohl G, Wongsiriroj N, Pollak NM, Radner FP, Preiss-Landl K, Kolbe T, Rulicke T, Pieske B, Trauner M, Lass A, Zimmermann R, Hoefler G, Cinti S, Kershaw EE, Schrauwen P, Madeo F, Mayer B, and Zechner R. ATGLmediated fat catabolism regulates cardiac mitochondrial function via PPAR-alpha and PGC-1. Nat Med 17: 1076– 1085, 2011. Hall ME, Maready MW, Hall JE, and Stec DE. Rescue of cardiac leptin receptors in db/db mice prevents myocardial triglyceride accumulation. Am J Physiol Endocrinol Metab 307: E316–E325, 2014. Hansford RG and Zorov D. Role of mitochondrial calcium transport in the control of substrate oxidation. Mol Cell Biochem 184: 359–369, 1998. Hausenblas HA, Schoulda JA, and Smoliga JM. Resveratrol treatment as an adjunct to pharmacological management in type 2 diabetes mellitus - systematic review and meta-analysis. Mol Nutr Food Res 19: 201400173, 2014. Haworth RA and Hunter DR. The Ca2 + -induced membrane transition in mitochondria. II. Nature of the Ca2 + trigger site. Arch Biochem Biophys 195: 460–467, 1979. Hayashi K, Okumura K, Matsui H, Murase K, Kamiya H, Saburi Y, Numaguchi Y, Toki Y, and Hayakawa T. Involvement of 1,2-diacylglycerol in improvement of heart function by etomoxir in diabetic rats. Life Sci 68: 1515– 1526, 2001. Herrero P, Peterson LR, McGill JB, Matthew S, Lesniak D, Dence C, and Gropler RJ. Increased myocardial fatty acid metabolism in patients with type 1 diabetes mellitus. J Am Coll Cardiol 47: 598–604, 2006. Hickson-Bick DL, Buja LM, and McMillin JB. Palmitatemediated alterations in the fatty acid metabolism of rat neonatal cardiac myocytes. J Mol Cell Cardiol 32: 511– 519, 2000. Holubarsch CJ, Rohrbach M, Karrasch M, Boehm E, Polonski L, Ponikowski P, and Rhein S. A double-blind randomized multicentre clinical trial to evaluate the efficacy and safety of two doses of etomoxir in comparison with placebo in patients with moderate congestive heart failure: the ERGO (etomoxir for the recovery of glucose oxidation) study. Clin Sci (Lond) 113: 205–212, 2007. Hoppins S. The regulation of mitochondrial dynamics. Curr Opin Cell Biol 29C: 46–52, 2014. 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. Hunter DR and Haworth RA. The Ca2 + -induced membrane transition in mitochondria. I. The protective mechanisms. Arch Biochem Biophys 195: 453–459, 1979. Hunter DR and Haworth RA. The Ca2 + -induced membrane transition in mitochondria. III. Transitional Ca2 + release. Arch Biochem Biophys 195: 468–477, 1979. Huynh K, Kiriazis H, Du XJ, Love JE, Gray SP, JandeleitDahm KA, McMullen JR, and Ritchie RH. Targeting the upregulation of reactive oxygen species subsequent to

116.

117.

118.

119. 120.

121.

122.

123. 124.

125.

126.

127.

128.

1623

hyperglycemia prevents type 1 diabetic cardiomyopathy in mice. Free Radic Biol Med 60: 307–317, 2013. Huynh K, Kiriazis H, Du XJ, Love JE, Jandeleit-Dahm KA, Forbes JM, McMullen JR, and Ritchie RH. Coenzyme Q10 attenuates diastolic dysfunction, cardiomyocyte hypertrophy and cardiac fibrosis in the db/db mouse model of type 2 diabetes. Diabetologia 55: 1544–1553, 2012. Il’yasova D, Wang F, D’Agostino RB, Jr., Hanley A, and Wagenknecht LE. Prospective association between fasting NEFA and type 2 diabetes: impact of post-load glucose. Diabetologia 53: 866–874, 2010. Jaconi M, Bony C, Richards SM, Terzic A, Arnaudeau S, Vassort G, and Puceat M. Inositol 1,4,5-trisphosphate directs Ca(2 + ) flow between mitochondria and the endoplasmic/sarcoplasmic reticulum: a role in regulating cardiac autonomic Ca(2 + ) spiking. Mol Biol Cell 11: 1845–1858, 2000. Jawien J, Nastalek P, and Korbut R. Mouse models of experimental atherosclerosis. J Physiol Pharmacol 55: 503–517, 2004. Jeffrey FM, Alvarez L, Diczku V, Sherry AD, and Malloy CR. Direct evidence that perhexiline modifies myocardial substrate utilization from fatty acids to lactate. J Cardiovasc Pharmacol 25: 469–472, 1995. Jensen MD, Haymond MW, Rizza RA, Cryer PE, and Miles JM. Influence of body fat distribution on free fatty acid metabolism in obesity. J Clin Invest 83: 1168–1173, 1989. 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. Joseph AM and Hood DA. Relationships between exercise, mitochondrial biogenesis and type 2 diabetes. Med Sport Sci 60: 48–61, 2014. Joshi M, Kotha SR, Malireddy S, Selvaraju V, Satoskar AR, Palesty A, McFadden DW, Parinandi NL, and Maulik N. Conundrum of pathogenesis of diabetic cardiomyopathy: role of vascular endothelial dysfunction, reactive oxygen species, and mitochondria. Mol Cell Biochem 386: 233–249, 2014. Jouaville LS, Pinton P, Bastianutto C, Rutter GA, and Rizzuto R. Regulation of mitochondrial ATP synthesis by calcium: evidence for a long-term metabolic priming. Proc Natl Acad Sci U S A 96: 13807–13812, 1999. Kaludercic N, Carpi A, Nagayama T, Sivakumaran V, Zhu G, Lai EW, Bedja D, De Mario A, Chen K, Gabrielson KL, Lindsey ML, Pacak K, Takimoto E, Shih JC, Kass DA, Di Lisa F, and Paolocci N. Monoamine oxidase B prompts mitochondrial and cardiac dysfunction in pressure overloaded hearts. Antioxid Redox Signal 20: 267–280, 2014. Kaludercic N, Takimoto E, Nagayama T, Feng N, Lai EW, Bedja D, Chen K, Gabrielson KL, Blakely RD, Shih JC, Pacak K, Kass DA, Di Lisa F, and Paolocci N. Monoamine oxidase A-mediated enhanced catabolism of norepinephrine contributes to adverse remodeling and pump failure in hearts with pressure overload. Circ Res 106: 193–202, 2010. Kanety H, Hemi R, Papa MZ, and Karasik A. Sphingomyelinase and ceramide suppress insulin-induced tyrosine phosphorylation of the insulin receptor substrate-1. J Biol Chem 271: 9895–9897, 1996.

1624

129. Kannel WB, Hjortland M, and Castelli WP. Role of diabetes in congestive heart failure: the Framingham study. Am J Cardiol 34: 29–34, 1974. 130. Kantor PF, Lucien A, Kozak R, and Lopaschuk GD. The antianginal drug trimetazidine shifts cardiac energy metabolism from fatty acid oxidation to glucose oxidation by inhibiting mitochondrial long-chain 3-ketoacyl coenzyme A thiolase. Circ Res 86: 580–588, 2000. 131. Kelly DP and Scarpulla RC. Transcriptional regulatory circuits controlling mitochondrial biogenesis and function. Genes Dev 18: 357–368, 2004. 132. Kienesberger PC, Pulinilkunnil T, Nagendran J, and Dyck JR. Myocardial triacylglycerol metabolism. J Mol Cell Cardiol 55: 101–110, 2013. 133. Kienesberger PC, Pulinilkunnil T, Nagendran J, Young ME, Bogner-Strauss JG, Hackl H, Khadour R, Heydari E, Haemmerle G, Zechner R, Kershaw EE, and Dyck JR. Early structural and metabolic cardiac remodelling in response to inducible adipose triglyceride lipase ablation. Cardiovasc Res 99: 442–451, 2013. 134. Kienesberger PC, Pulinilkunnil T, Sung MM, Nagendran J, Haemmerle G, Kershaw EE, Young ME, Light PE, Oudit GY, Zechner R, and Dyck JR. Myocardial ATGL overexpression decreases the reliance on fatty acid oxidation and protects against pressure overload-induced cardiac dysfunction. Mol Cell Biol 32: 740–750, 2012. 135. Kietadisorn R, Juni RP, and Moens AL. Tackling endothelial dysfunction by modulating NOS uncoupling: new insights into its pathogenesis and therapeutic possibilities. Am J Physiol Endocrinol Metab 302: E481–E495, 2012. 136. Knuuti J, Takala TO, Nagren K, Sipila H, Turpeinen AK, Uusitupa MI, and Nuutila P. Myocardial fatty acid oxidation in patients with impaired glucose tolerance. Diabetologia 44: 184–187, 2001. 137. Koonen DP, Febbraio M, Bonnet S, Nagendran J, Young ME, Michelakis ED, and Dyck JR. CD36 expression contributes to age-induced cardiomyopathy in mice. Circulation 116: 2139–2147, 2007. 138. Koonen DP, Glatz JF, Bonen A, and Luiken JJ. Longchain fatty acid uptake and FAT/CD36 translocation in heart and skeletal muscle. Biochim Biophys Acta 1736: 163–180, 2005. 139. Kooy A, de Jager J, Lehert P, Bets D, Wulffele MG, Donker AJ, and Stehouwer CD. Long-term effects of metformin on metabolism and microvascular and macrovascular disease in patients with type 2 diabetes mellitus. Arch Intern Med 169: 616–625, 2009. 140. Korshunov SS, Skulachev VP, and Starkov AA. High protonic potential actuates a mechanism of production of reactive oxygen species in mitochondria. FEBS Lett 416: 15–18, 1997. 141. Kosiborod M, Arnold SV, Spertus JA, McGuire DK, Li Y, Yue P, Ben-Yehuda O, Katz A, Jones PG, Olmsted A, Belardinelli L, and Chaitman BR. Evaluation of ranolazine in patients with type 2 diabetes mellitus and chronic stable angina: results from the TERISA randomized clinical trial (Type 2 Diabetes Evaluation of Ranolazine in Subjects with Chronic Stable Angina). J Am Coll Cardiol 61: 2038–2045, 2013. 142. Koves TR, Ussher JR, Noland RC, Slentz D, Mosedale M, Ilkayeva O, Bain J, Stevens R, Dyck JR, Newgard CB, Lopaschuk GD, and Muoio DM. Mitochondrial overload and incomplete fatty acid oxidation contribute to skeletal muscle insulin resistance. Cell Metab 7: 45–56, 2008.

SUNG ET AL.

143. Kranstuber AL, Del Rio C, Biesiadecki BJ, Hamlin RL, Ottobre J, Gyorke S, and Lacombe VA. Advanced glycation end product cross-link breaker attenuates diabetesinduced cardiac dysfunction by improving sarcoplasmic reticulum calcium handling. Front Physiol 3: 292, 2012. 144. Kuang M, Febbraio M, Wagg C, Lopaschuk GD, and Dyck JR. Fatty acid translocase/CD36 deficiency does not energetically or functionally compromise hearts before or after ischemia. Circulation 109: 1550–1557, 2004. 145. Kumar S, Prasad S, and Sitasawad SL. Multiple antioxidants improve cardiac complications and inhibit cardiac cell death in streptozotocin-induced diabetic rats. PLoS One 8: e67009, 2013. 146. Lagouge M, Argmann C, Gerhart-Hines Z, Meziane H, Lerin C, Daussin F, Messadeq N, Milne J, Lambert P, Elliott P, Geny B, Laakso M, Puigserver P, and Auwerx J. Resveratrol improves mitochondrial function and protects against metabolic disease by activating SIRT1 and PGC1alpha. Cell 127: 1109–1122, 2006. 147. Lamant M, Smih F, Harmancey R, Philip-Couderc P, Pathak A, Roncalli J, Galinier M, Collet X, Massabuau P, Senard JM, and Rouet P. ApoO, a novel apolipoprotein, is an original glycoprotein up-regulated by diabetes in human heart. J Biol Chem 281: 36289–36302, 2006. 148. Lamb HJ, Beyerbacht HP, van der Laarse A, Stoel BC, Doornbos J, van der Wall EE, and de Roos A. Diastolic dysfunction in hypertensive heart disease is associated with altered myocardial metabolism. Circulation 99: 2261–2267, 1999. 149. Lashin OM, Szweda PA, Szweda LI, and Romani AM. Decreased complex II respiration and HNE-modified SDH subunit in diabetic heart. Free Radic Biol Med 40: 886– 896, 2006. 150. Ledesma A, de Lacoba MG, and Rial E. The mitochondrial uncoupling proteins. Genome Biol 3: REVIEWS3015, 2002. 151. Lee GY, Kim NH, Zhao ZS, Cha BS, and Kim YS. Peroxisomal-proliferator-activated receptor alpha activates transcription of the rat hepatic malonyl-CoA decarboxylase gene: a key regulation of malonyl-CoA level. Biochem J 378: 983–990, 2004. 152. Lee L, Campbell R, Scheuermann-Freestone M, Taylor R, Gunaruwan P, Williams L, Ashrafian H, Horowitz J, Fraser AG, Clarke K, and Frenneaux M. Metabolic modulation with perhexiline in chronic heart failure: a randomized, controlled trial of short-term use of a novel treatment. Circulation 112: 3280–3288, 2005. 153. Leone TC, Weinheimer CJ, and Kelly DP. A 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 U S A 96: 7473–7478, 1999. 154. Lexis CP, van der Horst IC, Lipsic E, Wieringa WG, de Boer RA, van den Heuvel AF, van der Werf HW, Schurer RA, Pundziute G, Tan ES, Nieuwland W, Willemsen HM, Dorhout B, Molmans BH, van der Horst-Schrivers AN, Wolffenbuttel BH, ter Horst GJ, van Rossum AC, Tijssen JG, Hillege HL, de Smet BJ, van der Harst P, and van Veldhuisen DJ. Effect of metformin on left ventricular function after acute myocardial infarction in patients without diabetes: the GIPS-III randomized clinical trial. JAMA 311: 1526–1535, 2014. 155. Li CJ, Zhang QM, Li MZ, Zhang JY, Yu P, and Yu DM. Attenuation of myocardial apoptosis by alpha-lipoic acid

MITOCHONDRIAL STRESS IN DIABETIC CARDIOMYOPATHY

156.

157.

158.

159.

160.

161.

162.

163.

164.

165.

166. 167.

168.

169.

through suppression of mitochondrial oxidative stress to reduce diabetic cardiomyopathy. Chin Med J (Engl) 122: 2580–2586, 2009. Li SY, Yang X, Ceylan-Isik AF, Du M, Sreejayan N, and Ren J. Cardiac contractile dysfunction in Lep/Lep obesity is accompanied by NADPH oxidase activation, oxidative modification of sarco(endo)plasmic reticulum Ca2 + ATPase and myosin heavy chain isozyme switch. Diabetologia 49: 1434–1446, 2006. Li YJ, Wang PH, Chen C, Zou MH, and Wang DW. Improvement of mechanical heart function by trimetazidine in db/db mice. Acta Pharmacol Sin 31: 560–569, 2010. Lin CH, Kurup S, Herrero P, Schechtman KB, Eagon JC, Klein S, Davila-Roman VG, Stein RI, Dorn GW, 2nd, Gropler RJ, Waggoner AD, and Peterson LR. Myocardial oxygen consumption change predicts left ventricular relaxation improvement in obese humans after weight loss. Obesity (Silver Spring) 19: 1804–1812, 2011. Lionetti L, Mollica MP, Lombardi A, Cavaliere G, Gifuni G, and Barletta A. From chronic overnutrition to insulin resistance: the role of fat-storing capacity and inflammation. Nutr Metab Cardiovasc Dis 19: 146–152, 2009. Liu J, Zhou J, An W, Lin Y, Yang Y, and Zang W. Apocynin attenuates pressure overload-induced cardiac hypertrophy in rats by reducing levels of reactive oxygen species. Can J Physiol Pharmacol 88: 745–752, 2010. Liu K, Zhou R, Wang B, and Mi MT. Effect of resveratrol on glucose control and insulin sensitivity: a meta-analysis of 11 randomized controlled trials. Am J Clin Nutr 99: 1510–1519, 2014. Liu L, Shi X, Bharadwaj KG, Ikeda S, Yamashita H, Yagyu H, Schaffer JE, Yu YH, and Goldberg IJ. DGAT1 expression increases heart triglyceride content but ameliorates lipotoxicity. J Biol Chem 284: 36312–36323, 2009. Lopaschuk GD, Barr R, Thomas PD, and Dyck JR. Beneficial effects of trimetazidine in ex vivo working ischemic hearts are due to a stimulation of glucose oxidation secondary to inhibition of long-chain 3-ketoacyl coenzyme a thiolase. Circ Res 93: e33–e37, 2003. Lopaschuk GD, McNeil GF, and McVeigh JJ. Glucose oxidation is stimulated in reperfused ischemic hearts with the carnitine palmitoyltransferase 1 inhibitor, Etomoxir. Mol Cell Biochem 88: 175–179, 1989. Lopaschuk GD, Tahiliani AG, Vadlamudi RV, Katz S, and McNeill JH. Cardiac sarcoplasmic reticulum function in insulin- or carnitine-treated diabetic rats. Am J Physiol 245: H969–H976, 1983. Lopaschuk GD, Ussher JR, Folmes CD, Jaswal JS, and Stanley WC. Myocardial fatty acid metabolism in health and disease. Physiol Rev 90: 207–258, 2010. Lopaschuk GD, Wall SR, Olley PM, and Davies NJ. Etomoxir, a carnitine palmitoyltransferase I inhibitor, protects hearts from fatty acid-induced ischemic injury independent of changes in long chain acylcarnitine. Circ Res 63: 1036–1043, 1988. 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

170.

171.

172.

173. 174.

175.

176. 177.

178.

179.

180.

181.

182.

1625

acid transporters in obese Zucker rats. J Biol Chem 276: 40567–40573, 2001. 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. Maack C, Cortassa S, Aon MA, Ganesan AN, Liu T, and O’Rourke B. Elevated cytosolic Na + decreases mitochondrial Ca2 + uptake during excitation-contraction coupling and impairs energetic adaptation in cardiac myocytes. Circ Res 99: 172–182, 2006. Marciniak C, Marechal X, Montaigne D, Neviere R, and Lancel S. Cardiac contractile function and mitochondrial respiration in diabetes-related mouse models. Cardiovasc Diabetol 13: 118, 2014. Maritim AC, Sanders RA, and Watkins JB, 3rd. Diabetes, oxidative stress, and antioxidants: a review. J Biochem Mol Toxicol 17: 24–38, 2003. Matsushima S, Kinugawa S, Ide T, Matsusaka H, Inoue N, Ohta Y, Yokota T, Sunagawa K, and Tsutsui H. Overexpression of glutathione peroxidase attenuates myocardial remodeling and preserves diastolic function in diabetic heart. Am J Physiol Heart Circ Physiol 291: H2237–H2245, 2006. 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. McGarry JD and Brown NF. The mitochondrial carnitine palmitoyltransferase system. From concept to molecular analysis. Eur J Biochem 244: 1–14, 1997. McGavock JM, Lingvay I, Zib I, Tillery T, Salas N, Unger R, Levine BD, Raskin P, Victor RG, and Szczepaniak LS. Cardiac steatosis in diabetes mellitus: a 1H-magnetic resonance spectroscopy study. Circulation 116: 1170– 1175, 2007. Mendez-Del Villar M, Gonzalez-Ortiz M, MartinezAbundis E, Perez-Rubio KG, and Lizarraga-Valdez R. Effect of resveratrol administration on metabolic syndrome, insulin sensitivity, and insulin secretion. Metab Syndr Relat Disord 12: 497–501, 2014. Meng D, Feng L, Chen XJ, Yang D, and Zhang JN. Trimetazidine improved Ca2 + handling in isoprenalinemediated myocardial injury of rats. Exp Physiol 91: 591– 601, 2006. Mizushige K, Yao L, Noma T, Kiyomoto H, Yu Y, Hosomi N, Ohmori K, and Matsuo H. Alteration in left ventricular diastolic filling and accumulation of myocardial collagen at insulin-resistant prediabetic stage of a type II diabetic rat model. Circulation 101: 899–907, 2000. Mohammadshahi M, Haidari F, and Soufi FG. Chronic resveratrol administration improves diabetic cardiomyopathy in part by reducing oxidative stress. Cardiol J 21: 39– 46, 2014. Montaigne D, Marechal X, Coisne A, Debry N, Modine T, Fayad G, Potelle C, El Arid JM, Mouton S, Sebti Y, Duez H, Preau S, Remy-Jouet I, Zerimech F, Koussa M, Richard V, Neviere R, Edme JL, Lefebvre P, and Staels B. Myocardial contractile dysfunction is associated with impaired mitochondrial function and dynamics in type 2 diabetic but not in obese patients. Circulation 130: 554– 564, 2014.

1626

183. Morrow DA, Scirica BM, Chaitman BR, McGuire DK, Murphy SA, Karwatowska-Prokopczuk E, McCabe CH, and Braunwald E. Evaluation of the glycometabolic effects of ranolazine in patients with and without diabetes mellitus in the MERLIN-TIMI 36 randomized controlled trial. Circulation 119: 2032–2039, 2009. 184. 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. 185. Nagendran J, Pulinilkunnil T, Kienesberger PC, Sung MM, Fung D, Febbraio M, and Dyck JR. Cardiomyocytespecific ablation of CD36 improves post-ischemic functional recovery. J Mol Cell Cardiol 63: 180–188, 2013. 186. Nakamura K, Fushimi K, Kouchi H, Mihara K, Miyazaki M, Ohe T, and Namba M. Inhibitory effects of antioxidants on neonatal rat cardiac myocyte hypertrophy induced by tumor necrosis factor-alpha and angiotensin II. Circulation 98: 794–799, 1998. 187. Neely JR and Morgan HE. Relationship between carbohydrate and lipid metabolism and the energy balance of heart muscle. Annu Rev Physiol 36: 413–459, 1974. 188. Netticadan T, Temsah RM, Kent A, Elimban V, and Dhalla NS. Depressed levels of Ca2 + -cycling proteins may underlie sarcoplasmic reticulum dysfunction in the diabetic heart. Diabetes 50: 2133–2138, 2001. 189. Nishikawa T, Edelstein D, Du XL, Yamagishi S, Matsumura T, Kaneda Y, Yorek MA, Beebe D, Oates PJ, Hammes HP, Giardino I, and Brownlee M. Normalizing mitochondrial superoxide production blocks three pathways of hyperglycaemic damage. Nature 404: 787–790, 2000. 190. Normand-Lauziere F, Frisch F, Labbe SM, Bherer P, Gagnon R, Cunnane SC, and Carpentier AC. Increased postprandial nonesterified fatty acid appearance and oxidation in type 2 diabetes is not fully established in offspring of diabetic subjects. PLoS One 5: e10956, 2010. 191. Oliveira PJ, Rolo AP, Seica R, Palmeira CM, Santos MS, and Moreno AJ. Decreased susceptibility of heart mitochondria from diabetic GK rats to mitochondrial permeability transition induced by calcium phosphate. Biosci Rep 21: 45–53, 2001. 192. Oliveira PJ, Rolo AP, Seica R, Sardao V, Monteiro P, Goncalves L, Providencia L, Palmeira CM, Santos MS, and Moreno AJ. Impact of diabetes on induction of the mitochondrial permeability transition. Rev Port Cardiol 21: 759–766, 2002. 193. Oliveira PJ, Seica R, Coxito PM, Rolo AP, Palmeira CM, Santos MS, and Moreno AJ. Enhanced permeability transition explains the reduced calcium uptake in cardiac mitochondria from streptozotocin-induced diabetic rats. FEBS Lett 554: 511–514, 2003. 194. Opie L. The Heart: Physiology and Metabolism. New York: Raven, 1991. 195. Opie L. Heart Physiology: From Cell to Circulation. Philadephia, PA: Lippincott Williams and Wilkins, 2004. 196. 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. 197. Ouwens DM, Diamant M, Fodor M, Habets DD, Pelsers MM, El Hasnaoui M, Dang ZC, van den Brom CE,

SUNG ET AL.

198. 199.

200. 201.

202.

203.

204.

205. 206.

207. 208.

209.

210.

211.

Vlasblom R, Rietdijk A, Boer C, Coort SL, Glatz JF, and Luiken JJ. Cardiac contractile dysfunction in insulin-resistant rats fed a high-fat diet is associated with elevated CD36-mediated fatty acid uptake and esterification. Diabetologia 50: 1938–1948, 2007. Palumbo PJ. Gycemic control, mealtime glucose excursions, and diabetic complications in type 2 diabetes mellitus. Mayo Clin Proc 76: 609–618, 2001. Park TS, Hu Y, Noh HL, Drosatos K, Okajima K, Buchanan J, Tuinei J, Homma S, Jiang XC, Abel ED, and Goldberg IJ. Ceramide is a cardiotoxin in lipotoxic cardiomyopathy. J Lipid Res 49: 2101–2112, 2008. Penpargkul S, Fein F, Sonnenblick EH, and Scheuer J. Depressed cardiac sarcoplasmic reticular function from diabetic rats. J Mol Cell Cardiol 13: 303–309, 1981. Penumathsa SV, Thirunavukkarasu M, Zhan L, Maulik G, Menon VP, Bagchi D, and Maulik N. Resveratrol enhances GLUT-4 translocation to the caveolar lipid raft fractions through AMPK/Akt/eNOS signalling pathway in diabetic myocardium. J Cell Mol Med 12: 2350–2361, 2008. Petersen KF, Dufour S, Befroy D, Garcia R, and Shulman GI. Impaired mitochondrial activity in the insulin-resistant offspring of patients with type 2 diabetes. N Engl J Med 350: 664–671, 2004. 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. Philip-Couderc P, Smih F, Pelat M, Vidal C, Verwaerde P, Pathak A, Buys S, Galinier M, Senard JM, and Rouet P. Cardiac transcriptome analysis in obesity-related hypertension. Hypertension 41: 414–421, 2003. Pierce GN and Dhalla NS. Heart mitochondrial function in chronic experimental diabetes in rats. Can J Cardiol 1: 48–54, 1985. Piot C, Croisille P, Staat P, Thibault H, Rioufol G, Mewton N, Elbelghiti R, Cung TT, Bonnefoy E, Angoulvant D, Macia C, Raczka F, Sportouch C, Gahide G, Finet G, Andre-Fouet X, Revel D, Kirkorian G, Monassier JP, Derumeaux G, and Ovize M. Effect of cyclosporine on reperfusion injury in acute myocardial infarction. N Engl J Med 359: 473–481, 2008. Polonsky KS. The past 200 years in diabetes. N Engl J Med 367: 1332–1340, 2012. Poulsen MK, Henriksen JE, Dahl J, Johansen A, Gerke O, Vach W, Haghfelt T, Hoilund-Carlsen PF, Beck-Nielsen H, and Moller JE. Left ventricular diastolic function in type 2 diabetes mellitus: prevalence and association with myocardial and vascular disease. Circ Cardiovasc Imaging 3: 24–31, 2010. Preiss D, Lloyd SM, Ford I, McMurray JJ, Holman RR, Welsh P, Fisher M, Packard CJ, and Sattar N. Metformin for non-diabetic patients with coronary heart disease (the CAMERA study): a randomised controlled trial. Lancet Diabetes Endocrinol 2: 116–124, 2014. Pulinilkunnil T, Kienesberger PC, Nagendran J, Sharma N, Young ME, and Dyck JR. Cardiac-specific adipose triglyceride lipase overexpression protects from cardiac steatosis and dilated cardiomyopathy following diet-induced obesity. Int J Obes (Lond) 38: 205–215, 2014. Pulinilkunnil T, Kienesberger PC, Nagendran J, Waller TJ, Young ME, Kershaw EE, Korbutt G, Haemmerle G,

MITOCHONDRIAL STRESS IN DIABETIC CARDIOMYOPATHY

212.

213.

214.

215.

216.

217.

218. 219.

220.

221.

222.

223.

Zechner R, and Dyck JR. Myocardial adipose triglyceride lipase overexpression protects diabetic mice from the development of lipotoxic cardiomyopathy. Diabetes 62: 1464–1477, 2013. 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. Regan TJ, Lyons MM, Ahmed SS, Levinson GE, Oldewurtel HA, Ahmad MR, and Haider B. Evidence for cardiomyopathy in familial diabetes mellitus. J Clin Invest 60: 884–899, 1977. Ren J, Dominguez LJ, Sowers JR, and Davidoff AJ. Troglitazone attenuates high-glucose-induced abnormalities in relaxation and intracellular calcium in rat ventricular myocytes. Diabetes 45: 1822–1825, 1996. Rijzewijk LJ, van der Meer RW, Lamb HJ, de Jong HW, Lubberink M, Romijn JA, Bax JJ, de Roos A, Twisk JW, Heine RJ, Lammertsma AA, Smit JW, and Diamant M. Altered myocardial substrate metabolism and decreased diastolic function in nonischemic human diabetic cardiomyopathy: studies with cardiac positron emission tomography and magnetic resonance imaging. J Am Coll Cardiol 54: 1524–1532, 2009. Ritchie RH, Love JE, Huynh K, Bernardo BC, Henstridge DC, Kiriazis H, Tham YK, Sapra G, Qin C, Cemerlang N, Boey EJ, Jandeleit-Dahm K, Du XJ, and McMullen JR. Enhanced phosphoinositide 3-kinase(p110alpha) activity prevents diabetes-induced cardiomyopathy and superoxide generation in a mouse model of diabetes. Diabetologia 55: 3369–3381, 2012. Robb-Gaspers LD, Burnett P, Rutter GA, Denton RM, Rizzuto R, and Thomas AP. Integrating cytosolic calcium signals into mitochondrial metabolic responses. EMBO J 17: 4987–5000, 1998. Rolo AP and Palmeira CM. Diabetes and mitochondrial function: role of hyperglycemia and oxidative stress. Toxicol Appl Pharmacol 212: 167–178, 2006. Rubler S, Dlugash J, Yuceoglu YZ, Kumral T, Branwood AW, and Grishman A. New type of cardiomyopathy associated with diabetic glomerulosclerosis. Am J Cardiol 30: 595–602, 1972. Rupp H, Elimban V, and Dhalla NS. Modification of subcellular organelles in pressure-overloaded heart by etomoxir, a carnitine palmitoyltransferase I inhibitor. FASEB J 6: 2349–2353, 1992. Rupp H, Wahl R, and Hansen M. Influence of diet and carnitine palmitoyltransferase I inhibition on myosin and sarcoplasmic reticulum. J Appl Physiol (1985) 72: 352– 360, 1992. Russell LK, Mansfield CM, Lehman JJ, Kovacs A, Courtois M, Saffitz JE, Medeiros DM, Valencik ML, McDonald JA, and Kelly DP. Cardiac-specific induction of the transcriptional coactivator peroxisome proliferatoractivated receptor gamma coactivator-1alpha promotes mitochondrial biogenesis and reversible cardiomyopathy in a developmental stage-dependent manner. Circ Res 94: 525–533, 2004. Rutter MK, Parise H, Benjamin EJ, Levy D, Larson MG, Meigs JB, Nesto RW, Wilson PW, and Vasan RS. Impact of glucose intolerance and insulin resistance on cardiac structure and function: sex-related differences in the Framingham Heart Study. Circulation 107: 448–454, 2003.

1627

224. 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. 225. Schmidt-Schweda S and Holubarsch C. First clinical trial with etomoxir in patients with chronic congestive heart failure. Clin Sci (Lond) 99: 27–35, 2000. 226. Schmitz FJ, Rosen P, and Reinauer H. Improvement of myocardial function and metabolism in diabetic rats by the carnitine palmitoyl transferase inhibitor etomoxir. Horm Metab Res 27: 515–522, 1995. 227. Scirica BM. Ranolazine in patients with coronary artery disease. Expert Opin Pharmacother 8: 2149–2157, 2007. 228. Selemidis S, Sobey CG, Wingler K, Schmidt HH, and Drummond GR. NADPH oxidases in the vasculature: molecular features, roles in disease and pharmacological inhibition. Pharmacol Ther 120: 254–291, 2008. 229. Semeniuk LM, Kryski AJ, and Severson DL. Echocardiographic assessment of cardiac function in diabetic db/db and transgenic db/db-hGLUT4 mice. Am J Physiol Heart Circ Physiol 283: H976–H982, 2002. 230. Serpillon S, Floyd BC, Gupte RS, George S, Kozicky M, Neito V, Recchia F, Stanley W, Wolin MS, and Gupte SA. Superoxide production by NAD(P)H oxidase and mitochondria is increased in genetically obese and hyperglycemic rat heart and aorta before the development of cardiac dysfunction. The role of glucose-6-phosphate dehydrogenase-derived NADPH. Am J Physiol Heart Circ Physiol 297: H153–H162, 2009. 231. 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. 232. Shaughnessy DT, McAllister K, Worth L, Haugen AC, Meyer JN, Domann FE, Van Houten B, Mostoslavsky R, Bultman SJ, Baccarelli AA, Begley TJ, Sobol R, Hirschey MD, Ideker T, Santos JH, Copeland WC, Tice RR, Balshaw DM, and Tyson FL. Mitochondria, energetics, epigenetics, and cellular responses to stress. Environ Health Perspect 122: 1271–1278, 2014. 233. Shen E, Li Y, Shan L, Zhu H, Feng Q, Arnold JM, and Peng T. Rac1 is required for cardiomyocyte apoptosis during hyperglycemia. Diabetes 58: 2386–2395, 2009. 234. 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. 235. Shindler DM, Kostis JB, Yusuf S, Quinones MA, Pitt B, Stewart D, Pinkett T, Ghali JK, and Wilson AC. Diabetes mellitus, a predictor of morbidity and mortality in the Studies of Left Ventricular Dysfunction (SOLVD) Trials and Registry. Am J Cardiol 77: 1017–1020, 1996. 236. Silva JE. Physiological importance and control of nonshivering facultative thermogenesis. Front Biosci (Schol Ed) 3: 352–371, 2011. 237. Singh CK, Kumar A, Lavoie HA, Dipette DJ, and Singh US. Diabetic complications in pregnancy: is resveratrol a solution? Exp Biol Med 238: 482–490, 2013. 238. Sivitz WI and Yorek MA. Mitochondrial dysfunction in diabetes: from molecular mechanisms to functional significance and therapeutic opportunities. Antioxid Redox Signal 12: 537–577, 2010.

1628

239. Skulachev VP. Role of uncoupled and non-coupled oxidations in maintenance of safely low levels of oxygen and its one-electron reductants. Q Rev Biophys 29: 169–202, 1996. 240. Sloan RC, Moukdar F, Frasier CR, Patel HD, Bostian PA, Lust RM, and Brown DA. Mitochondrial permeability transition in the diabetic heart: contributions of thiol redox state and mitochondrial calcium to augmented reperfusion injury. J Mol Cell Cardiol 52: 1009–1018, 2012. 241. Son NH, Yu S, Tuinei J, Arai K, Hamai H, Homma S, Shulman GI, Abel ED, and Goldberg IJ. PPARgammainduced cardiolipotoxicity in mice is ameliorated by PPARalpha deficiency despite increases in fatty acid oxidation. J Clin Invest 120: 3443–3454, 2010. 242. Soufi FG, Vardyani M, Sheervalilou R, Mohammadi M, and Somi MH. Long-term treatment with resveratrol attenuates oxidative stress pro-inflammatory mediators and apoptosis in streptozotocin-nicotinamide-induced diabetic rats. Gen Physiol Biophys 31: 431–438, 2012. 243. Stanley BA, Sivakumaran V, Shi S, McDonald I, Lloyd D, Watson WH, Aon MA, and Paolocci N. Thioredoxin reductase-2 is essential for keeping low levels of H(2)O(2) emission from isolated heart mitochondria. J Biol Chem 286: 33669–33677, 2011. 244. Stanley WC, Morgan EE, Huang H, McElfresh TA, Sterk JP, Okere IC, Chandler MP, Cheng J, Dyck JR, and Lopaschuk GD. Malonyl-CoA decarboxylase inhibition suppresses fatty acid oxidation and reduces lactate production during demand-induced ischemia. Am J Physiol Heart Circ Physiol 289: H2304–H2309, 2005. 245. Su HC, Hung LM, and Cheng JK. Resveratrol, a red wine antioxidant, possesses an insulin-like effect in streptozotocin-induced diabetic rats. Am J Physiol Endocrinol Metab 290: 1339–1346, 2006. 246. Suarez J, Scott B, and Dillmann WH. Conditional increase in SERCA2a protein is able to reverse contractile dysfunction and abnormal calcium flux in established diabetic cardiomyopathy. Am J Physiol Regul Integr Comp Physiol 295: R1439–R1445, 2008. 247. Sulaiman M, Matta MJ, Sunderesan NR, Gupta MP, Periasamy M, and Gupta M. Resveratrol, an activator of SIRT1, upregulates sarcoplasmic calcium ATPase and improves cardiac function in diabetic cardiomyopathy. Am J Physiol Heart Circ Physiol 298: H833–H843, 2010. 248. This reference has been deleted. 249. Sung MM, Koonen DP, Soltys CL, Jacobs RL, Febbraio M, and Dyck JR. Increased CD36 expression in middleaged mice contributes to obesity-related cardiac hypertrophy in the absence of cardiac dysfunction. J Mol Med (Berl) 89: 459–469, 2011. 250. Szabadkai G and Duchen MR. Mitochondria: the hub of cellular Ca2 + signaling. Physiology (Bethesda) 23: 84– 94, 2008. 251. Szkudelski T. The insulin-suppressive effect of resveratrol an in vitro and in vivo phenomenon. Life Sci 82: 430–435, 2008. 252. Szkudelski T. Resveratrol inhibits insulin secretion from rat pancreatic islets. Eur J Pharmacol 552: 176–181, 2006. 253. Szkudelski T. Resveratrol-induced inhibition of insulin secretion from rat pancreatic islets: evidence for pivotal role of metabolic disturbances. Am J Physiol Endocrinol Metab 293: E901–E907, 2007. 254. Szkudelski T and Szkudelska K. Anti-diabetic effects of resveratrol. Ann N Y Acad Sci 1215: 34–39, 2011.

SUNG ET AL.

255. Tanaka Y, Konno N, and Kako KJ. Mitochondrial dysfunction observed in situ in cardiomyocytes of rats in experimental diabetes. Cardiovasc Res 26: 409–414, 1992. 256. Tanno M, Kuno A, Horio Y, and Miura T. Emerging beneficial roles of sirtuins in heart failure. Basic Res Cardiol 107: 273, 2012. 257. Tanno M, Kuno A, Yano T, Miura T, Hisahara S, Ishikawa S, Shimamoto K, and Horio Y. Induction of manganese superoxide dismutase by nuclear translocation and activation of SIRT1 promotes cell survival in chronic heart failure. J Biol Chem 285: 8375–8382, 2010. 258. Thackeray JT, Beanlands RS, and Dasilva JN. Altered sympathetic nervous system signaling in the diabetic heart: emerging targets for molecular imaging. Am J Nucl Med Mol Imaging 2: 314–334, 2012. 259. Timmers S, Konings E, Bilet L, Houtkooper RH, van de Weijer T, Goossens GH, Hoeks J, van der Krieken S, Ryu D, Kersten S, Moonen-Kornips E, Hesselink MK, Kunz I, Schrauwen-Hinderling VB, Blaak EE, Auwerx J, and Schrauwen P. Calorie restriction-like effects of 30 days of resveratrol supplementation on energy metabolism and metabolic profile in obese humans. Cell Metab 14: 612– 622, 2011. 260. Timmis AD, Chaitman BR, and Crager M. Effects of ranolazine on exercise tolerance and HbA1c in patients with chronic angina and diabetes. Eur Heart J 27: 42–48, 2006. 261. Tocchetti CG, Caceres V, Stanley BA, Xie C, Shi S, Watson WH, O’Rourke B, Spadari-Bratfisch RC, Cortassa S, Akar FG, Paolocci N, and Aon MA. GSH or palmitate preserves mitochondrial energetic/redox balance, preventing mechanical dysfunction in metabolically challenged myocytes/hearts from type 2 diabetic mice. Diabetes 61: 3094–3105, 2012. 262. 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. 263. Turkieh A, Caubere C, Barutaut M, Desmoulin F, Harmancey R, Galinier M, Berry M, Dambrin C, Polidori C, Casteilla L, Koukoui F, Rouet P, and Smih F. Apolipoprotein O is mitochondrial and promotes lipotoxicity in heart. J Clin Invest 124: 2277–2286, 2014. 264. Turpeinen AK, Takala TO, Nuutila P, Axelin T, Luotolahti M, Haaparanta M, Bergman J, Hamalainen H, Iida H, Maki M, Uusitupa MI, and Knuuti J. Impaired free fatty acid uptake in skeletal muscle but not in myocardium in patients with impaired glucose tolerance: studies with PET and 14(R,S)-[18F]fluoro-6-thia-heptadecanoic acid. Diabetes 48: 1245–1250, 1999. 265. Ueno M, Suzuki J, Zenimaru Y, Takahashi S, Koizumi T, Noriki S, Yamaguchi O, Otsu K, Shen WJ, Kraemer FB, and Miyamori I. Cardiac overexpression of hormonesensitive lipase inhibits myocardial steatosis and fibrosis in streptozotocin diabetic mice. Am J Physiol Endocrinol Metab 294: E1109–E1118, 2008. 266. Ussher JR, Fillmore N, Keung W, Mori J, Beker DL, Wagg CS, Jaswal JS, and Lopaschuk GD. Trimetazidine therapy prevents obesity-induced cardiomyopathy in mice. Can J Cardiol 30: 940–944, 2014. 267. Ussher JR, Folmes CD, Keung W, Fillmore N, Jaswal JS, Cadete VJ, Beker DL, Lam VH, Zhang L, and Lopaschuk GD. Inhibition of serine palmitoyl transferase I reduces cardiac ceramide levels and increases glycolysis rates

MITOCHONDRIAL STRESS IN DIABETIC CARDIOMYOPATHY

268.

269. 270.

271. 272.

273. 274. 275.

276.

277.

278.

279. 280. 281.

following diet-induced insulin resistance. PLoS One 7: e37703, 2012. Ussher JR, Koves TR, Jaswal JS, Zhang L, Ilkayeva O, Dyck JR, Muoio DM, and Lopaschuk GD. Insulin-stimulated cardiac glucose oxidation is increased in high-fat dietinduced obese mice lacking malonyl CoA decarboxylase. Diabetes 58: 1766–1775, 2009. van de Weijer T, Schrauwen-Hinderling VB, and Schrauwen P. Lipotoxicity in type 2 diabetic cardiomyopathy. Cardiovasc Res 92: 10–18, 2011. van der Meer RW, Rijzewijk LJ, de Jong HW, Lamb HJ, Lubberink M, Romijn JA, Bax JJ, de Roos A, Kamp O, Paulus WJ, Heine RJ, Lammertsma AA, Smit JW, and Diamant M. Pioglitazone improves cardiac function and alters myocardial substrate metabolism without affecting cardiac triglyceride accumulation and high-energy phosphate metabolism in patients with well-controlled type 2 diabetes mellitus. Circulation 119: 2069–2077, 2009. Vetter R and Rupp H. CPT-1 inhibition by etomoxir has a chamber-related action on cardiac sarcoplasmic reticulum and isomyosins. Am J Physiol 267: H2091–H2099, 1994. Vinereanu D, Nicolaides E, Tweddel AC, Madler CF, Holst B, Boden LE, Cinteza M, Rees AE, and Fraser AG. Subclinical left ventricular dysfunction in asymptomatic patients with type II diabetes mellitus, related to serum lipids and glycated haemoglobin. Clin Sci (Lond) 105: 591–599, 2003. Wall SR and Lopaschuk GD. Glucose oxidation rates in fatty acid-perfused isolated working hearts from diabetic rats. Biochim Biophys Acta 1006: 97–103, 1989. Wallace DC. Mitochondrial genetics: a paradigm for aging and degenerative diseases? Science 256: 628–632, 1992. 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. Wichi R, Malfitano C, Rosa K, De Souza SB, Salemi V, Mostarda C, De Angelis K, and Irigoyen MC. Noninvasive and invasive evaluation of cardiac dysfunction in experimental diabetes in rodents. Cardiovasc Diabetol 6: 14, 2007. Widlansky ME, Wang J, Shenouda SM, Hagen TM, Smith AR, Kizhakekuttu TJ, Kluge MA, Weihrauch D, Gutterman DD, and Vita JA. Altered mitochondrial membrane potential, mass, and morphology in the mononuclear cells of humans with type 2 diabetes. Transl Res 156: 15–25, 2010. Williamson CL, Dabkowski ER, Baseler WA, Croston TL, Alway SE, and Hollander JM. Enhanced apoptotic propensity in diabetic cardiac mitochondria: influence of subcellular spatial location. Am J Physiol Heart Circ Physiol 298: H633–H642, 2010. Wojtczak L and Schonfeld P. Effect of fatty acids on energy coupling processes in mitochondria. Biochim Biophys Acta 1183: 41–57, 1993. Wold LE, Ceylan-Isik AF, and Ren J. Oxidative stress and stress signaling: menace of diabetic cardiomyopathy. Acta Pharmacol Sin 26: 908–917, 2005. Xu Q, Dalic A, Fang L, Kiriazis H, Ritchie RH, Sim K, Gao XM, Drummond G, Sarwar M, Zhang YY, Dart AM, and Du XJ. Myocardial oxidative stress contributes to transgenic beta(2)-adrenoceptor activation-induced cardiomyopathy and heart failure. Br J Pharmacol 162: 1012–1028, 2011.

1629

282. 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. 283. Yang J, Sambandam N, Han X, Gross RW, Courtois M, Kovacs A, Febbraio M, Finck BN, and Kelly DP. CD36 deficiency rescues lipotoxic cardiomyopathy. Circ Res 100: 1208–1217, 2007. 284. 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. 285. Yin M, van der Horst IC, van Melle JP, Qian C, van Gilst WH, Sillje HH, and de Boer RA. Metformin improves cardiac function in a nondiabetic rat model of post-MI heart failure. Am J Physiol Heart Circ Physiol 301: H459– H468, 2011. 286. Yu CM, Lin H, Yang H, Kong SL, Zhang Q, and Lee SW. Progression of systolic abnormalities in patients with ‘‘isolated’’ diastolic heart failure and diastolic dysfunction. Circulation 105: 1195–1201, 2002. 287. Yu T, Sheu SS, Robotham JL, and Yoon Y. Mitochondrial fission mediates high glucose-induced cell death through elevated production of reactive oxygen species. Cardiovasc Res 79: 341–351, 2008. 288. Yue P, Arai T, Terashima M, Sheikh AY, Cao F, Charo D, Hoyt G, Robbins RC, Ashley EA, Wu J, Yang PC, and Tsao PS. Magnetic resonance imaging of progressive cardiomyopathic changes in the db/db mouse. Am J Physiol Heart Circ Physiol 292: H2106–H2118, 2007. 289. Zabalgoitia M, Ismaeil MF, Anderson L, and Maklady FA. Prevalence of diastolic dysfunction in normotensive, asymptomatic patients with well-controlled type 2 diabetes mellitus. Am J Cardiol 87: 320–323, 2001. 290. Zarich SW, Arbuckle BE, Cohen LR, Roberts M, and Nesto RW. Diastolic abnormalities in young asymptomatic diabetic patients assessed by pulsed Doppler echocardiography. J Am Coll Cardiol 12: 114–120, 1988. 291. Zhang H, Morgan B, Potter BJ, Ma L, Dellsperger KC, Ungvari Z, and Zhang C. Resveratrol improves left ventricular diastolic relaxation in type 2 diabetes by inhibiting oxidative/nitrative stress: in vivo demonstration with magnetic resonance imaging. Am J Physiol Heart Circ Physiol 299: H985–H994, 2010. 292. Zhang L, Ussher JR, Oka T, Cadete VJ, Wagg C, and Lopaschuk GD. Cardiac diacylglycerol accumulation in high fat-fed mice is associated with impaired insulinstimulated glucose oxidation. Cardiovasc Res 89: 148– 156, 2011. 293. Zhang P, Hou M, Li Y, Xu X, Barsoum M, Chen Y, and Bache RJ. NADPH oxidase contributes to coronary endothelial dysfunction in the failing heart. Am J Physiol Heart Circ Physiol 296: H840–H846, 2009. 294. Zhao P, Zhang J, Yin XG, Maharaj P, Narraindoo S, Cui LQ, and Tang YS. The effect of trimetazidine on cardiac function in diabetic patients with idiopathic dilated cardiomyopathy. Life Sci 92: 633–638, 2013. 295. Zhao W, Zhao T, Chen Y, Ahokas RA, and Sun Y. Oxidative stress mediates cardiac fibrosis by enhancing transforming growth factor-beta1 in hypertensive rats. Mol Cell Biochem 317: 43–50, 2008. 296. Zhong Y, Ahmed S, Grupp IL, and Matlib MA. Altered SR protein expression associated with contractile dysfunction

1630

SUNG ET AL.

in diabetic rat hearts. Am J Physiol Heart Circ Physiol 281: H1137–H1147, 2001. 297. 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 U S A 97: 1784–1789, 2000. 298. Zorov DB, Filburn CR, Klotz LO, Zweier JL, and Sollott SJ. Reactive oxygen species (ROS)-induced ROS release: a new phenomenon accompanying induction of the mitochondrial permeability transition in cardiac myocytes. J Exp Med 192: 1001–1014, 2000. 299. Zorov DB, Juhaszova M, and Sollott SJ. Mitochondrial reactive oxygen species (ROS) and ROS-induced ROS release. Physiol Rev 94: 909–950, 2014.

Address correspondence to: Dr. Jason R.B. Dyck 458 Heritage Medical Research Centre Department of Pediatrics Cardiovascular Research Centre University of Alberta Edmonton, Alberta T6G 2S2 Canada E-mail: [email protected] Date of first submission to ARS Central, February 24, 2015; date of final revised submission, March 12, 2015; date of acceptance, March 23, 2015.

Abbreviations Used ACSL1 ¼ acyl-CoA synthetase 1 AGEs ¼ advanced glycation endproducts APOO ¼ apolipoprotein O ATGL ¼ adipose triglyceride lipase ATP ¼ adenosine triphosphate Ca2+ ¼ calcium CAD ¼ coronary artery disease CM ¼ chylomicron CPT1 ¼ carnitine palmitoyal transferase 1 CVD ¼ cardiovascular disease DAG ¼ diacylglyercides

eNOS ¼ endothelial nitric oxide synthase ETC ¼ electron transport chain FABP ¼ fatty acid binding protein FAO ¼ fatty acid oxidation FATP1 ¼ fatty acid transport protein 1 FFA ¼ free fatty acid H2 O2 ¼ hydrogen peroxide HF ¼ heart failure HFD ¼ high fat diet hLpLGPI ¼ glycosylphosphatidylinositol anchored human lipoprotein lipase HNE ¼ 4-hydroxy-2-nonenal LPL ¼ lipoprotein lipase LV ¼ left ventricular LVH ¼ left ventricular hypertrophy MCD ¼ malonyl-CoA decarboxylase MHC ¼ myosin heavy chain MVO2 ¼ myocardial oxygen consumption NE ¼ norepinephrine NOS ¼ nitric oxide synthase O2 ¼ oxygen O-GlcNAc ¼ O-linked b-N-acetylglucosamine OXPHOS ¼ oxidative phosphorylation PCr ¼ phosphocreatine PDH ¼ pyruvate dehydrogenase PET ¼ positron emission tomography PGC-1a ¼ peroxisome proliferator-activated receptor c co-activator 1a PPAR-a ¼ peroxisome proliferator-activated receptor-a ROS ¼ reactive oxygen species SERCA ¼ sarcoplasmic reticulum Ca2+-ATPase SIRT ¼ silent information regulator SR ¼ sarcoplasmic reticulum STZ ¼ streptozotocin T1D ¼ type 1 diabetes T2D ¼ type 2 diabetes TAG ¼ triacylglycerol/triglycerides TCA ¼ tricarboxylic acid TnC ¼ troponin C UCP ¼ uncoupling protein UCP-DTA ¼ uncoupling protein-diphtheria toxin A VLDL ¼ very-low-density lipoprotein ZDF ¼ Zucker diabetic fatty rat

Myocardial metabolism in diabetic cardiomyopathy: potential therapeutic targets.

Cardiovascular complications in diabetes are particularly serious and represent the primary cause of morbidity and mortality in diabetic patients. Des...
752KB Sizes 2 Downloads 10 Views