Next Article in Journal
LC-ESI/QTOF-MS Profiling of Chicory and Lucerne Polyphenols and Their Antioxidant Activities
Next Article in Special Issue
Implications of Oxidative Stress in Glioblastoma Multiforme Following Treatment with Purine Derivatives
Previous Article in Journal
Red Cabbage Rather Than Green Cabbage Increases Stress Resistance and Extends the Lifespan of Caenorhabditis elegans
Previous Article in Special Issue
Effect of Reactive Oxygen Species on the Endoplasmic Reticulum and Mitochondria during Intracellular Pathogen Infection of Mammalian Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Oxidative Stress as A Mechanism for Functional Alterations in Cardiac Hypertrophy and Heart Failure

by
Anureet K. Shah
1,
Sukhwinder K. Bhullar
2,
Vijayan Elimban
2 and
Naranjan S. Dhalla
2,*
1
School of Kinesiology, Nutrition and Food Science, California State University, Los Angeles, CA 90032, USA
2
Institute of Cardiovascular Sciences, St. Boniface Hospital Albrechtsen Research Centre, Department of Physiology and Pathophysiology, Max Rady College of Medicine, University of Manitoba, Winnipeg, MB R2H 2A6, Canada
*
Author to whom correspondence should be addressed.
Antioxidants 2021, 10(6), 931; https://doi.org/10.3390/antiox10060931
Submission received: 4 May 2021 / Revised: 1 June 2021 / Accepted: 6 June 2021 / Published: 8 June 2021
(This article belongs to the Special Issue Reactive Oxygen Species in Different Biological Processes)

Abstract

:
Although heart failure due to a wide variety of pathological stimuli including myocardial infarction, pressure overload and volume overload is associated with cardiac hypertrophy, the exact reasons for the transition of cardiac hypertrophy to heart failure are not well defined. Since circulating levels of several vasoactive hormones including catecholamines, angiotensin II, and endothelins are elevated under pathological conditions, it has been suggested that these vasoactive hormones may be involved in the development of both cardiac hypertrophy and heart failure. At initial stages of pathological stimuli, these hormones induce an increase in ventricular wall tension by acting through their respective receptor-mediated signal transduction systems and result in the development of cardiac hypertrophy. Some oxyradicals formed at initial stages are also involved in the redox-dependent activation of the hypertrophic process but these are rapidly removed by increased content of antioxidants in hypertrophied heart. In fact, cardiac hypertrophy is considered to be an adaptive process as it exhibits either normal or augmented cardiac function for maintaining cardiovascular homeostasis. However, exposure of a hypertrophied heart to elevated levels of circulating hormones due to pathological stimuli over a prolonged period results in cardiac dysfunction and development of heart failure involving a complex set of mechanisms. It has been demonstrated that different cardiovascular abnormalities such as functional hypoxia, metabolic derangements, uncoupling of mitochondrial electron transport, and inflammation produce oxidative stress in the hypertrophied failing hearts. In addition, oxidation of catecholamines by monoamine oxidase as well as NADPH oxidase activation by angiotensin II and endothelin promote the generation of oxidative stress during the prolonged period by these pathological stimuli. It is noteworthy that oxidative stress is known to activate metallomatrix proteases and degrade the extracellular matrix proteins for the induction of cardiac remodeling and heart dysfunction. Furthermore, oxidative stress has been shown to induce subcellular remodeling and Ca2+-handling abnormalities as well as loss of cardiomyocytes due to the development of apoptosis, necrosis, and fibrosis. These observations support the view that a low amount of oxyradical formation for a brief period may activate redox-sensitive mechanisms, which are associated with the development of cardiac hypertrophy. On the other hand, high levels of oxyradicals over a prolonged period may induce oxidative stress and cause Ca2+-handling defects as well as protease activation and thus play a critical role in the development of adverse cardiac remodeling and cardiac dysfunction as well as progression of heart failure.

Graphical Abstract

1. Introduction

Heart failure due to several pathological conditions such as myocardial infarction, hypertension, valvular defects, diabetes, atherosclerosis, and different types of cardiomyopathies, is invariably associated with cardiac hypertrophy [1,2,3,4,5,6]. Extensive research regarding hemodynamic, cellular, and biochemical mechanisms have revealed that heart failure may be due to loss of cardiomyocytes, adverse cardiac remodeling, defects in subcellular activities, Ca2+-handling abnormalities, alterations in myocardial metabolism and elevation of different hormones in the circulation [7,8,9,10,11,12]. Some of these cardiovascular alterations and mechanisms associated with the development of cardiac hypertrophy and subsequent heart failure are shown in Figure 1. It is becoming clear that all these mechanisms for the induction of contractile defects are inter-related and cardiac dysfunction is the hallmark for identifying the development of heart failure [13,14,15]. On the other hand, cardiac function of hypertrophied heart is either normal or increased at early stages whereas cardiac performance is impaired in heart failure [16,17,18,19]. Although excellent review articles regarding the molecular and cellular mechanism for the development of cardiac hypertrophy and heart failure are available in the literature [20,21,22,23,24], exact reasons for the occurrence of cardiac dysfunction in hypertrophied heart are not fully understood [25,26]. It is therefore important to gain some information to understand the mechanisms involved in the transition of cardiac hypertrophy to heart failure as well as pathophysiology of cardiac dysfunction during the progression of heart failure.
Although elevated levels of several vasoactive hormones in the circulation due to activation of sympathetic nervous system, renin–angiotensin system, and other neuro-endocrine systems have been shown to occur in different types of heart failure [27,28,29,30], the mechanisms for their beneficial actions for the development of cardiac hypertrophy and adverse effects for the occurrence heart failure are not well understood. It is generally held that the acute effects of elevated vasoactive hormones increase cardiac muscle mass, add contractile units, and produce cardiac hypertrophy upon stimulating their respective receptor-mediated signal transduction pathways. Furthermore, these acute effects are associated with the formation of low amounts of oxyradicals and activation of redox-sensitive signal transduction for the occurrence of cardiac hypertrophy [7,12,15,16]. Myocardial hypertrophy in response to diverse stimuli is an adaptive process where cardiac performance and subcellular organ function are either normal or augmented to maintain hemodynamic homeostasis indicating the beneficial effects of vasoactive hormones [13,16,18,26]. On the other hand, prolonged exposure of hypertrophied heart to elevated levels of vasoactive hormones due to pathological stimuli for a prolonged period has been shown to induce metabolic derangements, Ca2+-handling abnormalities, protease activation, subcellular defects, and cardiac dysfunction leading to the development of heart failure [11,12,13,14,15,17,26]. Thus, it appears that cardiac hypertrophy and heart failure represent two different stages of effects, namely adaptive cardiac remodeling and adverse (maladaptive) cardiac remodeling, initiated by diverse pathological stimuli. The present article is therefore intended to describe some salient features for the development of both cardiac hypertrophy and heart failure as a consequence of some pathological situations including myocardial infarction, pressure overload and volume overload. It is also planned to discuss different mechanisms involved in the generation of oxyradicals for the activation of redox-sensitive hypertrophic process as well as the development of oxidative stress. The consequence of oxidative stress for the transition of cardiac hypertrophy to heart failure and pathophysiology of hypertrophied heart during progression of heart failure due to adverse effects of elevated levels of some vasoactive hormones for a prolonged period will be highlighted. In addition, efforts will be made to describe some of the subcellular and metabolic abnormalities under some experimental conditions, which are known to promote the occurrence of oxidative stress and induce cardiac dysfunction in non-hypertrophied hearts. The effectiveness of different oxyradical scavengers and antioxidants on the experimentally induced cardiac dysfunction as well as subcellular defects will be examined. Evidence will also be presented to show both direct and indirect effects of oxidative stress on subcellular organelles and Ca2+-handling abnormalities associated with heart dysfunction.

2. Development of Cardiac Hypertrophy and Heart Failure

Over the past 30 years various types of pathological stimuli have been shown to involve different signal transduction pathways as well as cellular and molecular mechanisms for the genesis of cardiac hypertrophy and heart failure [5,6,7,8,9,10,20,21,22,23,24,25,26]. Despite difference in the patterns of signal transduction mechanisms, there are several similarities in cardiovascular alterations, which occur during the initial and later stages of cardiac hypertrophy as well as heart failure. Various vasoactive hormones and growth factors are elevated not only for stimulating cardiovascular function and maintaining blood supply to all organs of the body but also for the induction of cardiac hypertrophy [27,28,29,30,31,32,33]. Although cardiac hypertrophy as a consequence of increased muscle mass has been shown to be of hypertrophic type or dilated type depending upon the pathological stimulus, both forms of cardiac growth have been reported to be either physiological or pathological in nature depending upon the type as well as duration and magnitude of the stimulus [13,14,15,16,17,25,26]. It appears that physiological hypertrophy is concerned with improving cardiac performance due to increased number of contractile units and augmented function of subcellular organelles whereas pathological hypertrophy associated with cardiac dysfunction may represent a pre-failure stage or reflect the transition of hypertrophied myocardium to heart failure. It should also be mentioned that there occurs a progressive increase in the levels of some vasodilatory natriuretic peptides (ANP and BNP) as well as endothelial nitric oxide (NO) in the circulation to maintain hemodynamic homeostasis; in fact, both ANP and BNP are commonly used as biomarkers for characterization of the heart failure stage [34,35,36]. On the other hand, prolonged exposure of hypertrophied heart to elevated levels of vasoactive hormones is considered to result in the progression of adverse cardiac remodeling and heart failure. Thus, it appears that cardiac hypertrophy and heart failure due to diverse pathological situations are associated with acute and chronic effects of the elevated levels of vasoactive hormones, respectively.

2.1. Development of Heart Failure Due to Myocardial Infarction

Myocardial infarction is known to produce loss of a portion of the ventricular tissue due to ischemia and is known to be the major cause of heart failure. The development of both cardiac hypertrophy and heart failure due to myocardial infarction are dependent upon the duration and size of infarct in the heart. The hemodynamic alterations such as decreased blood pressure and reduced cardiac output at initial stages activate the sympathetic nervous system and the renin–angiotensin system mainly to increase the circulating levels of catecholamines and angiotensin II, respectively. These vasoactive hormones not only elevate blood pressure but also promote the function of subcellular organelles, augment contractile activity, and induce cardiac hypertrophy [5,6,7,37]. Such beneficial effects of these hormones are mediated through the activation of both α- and β-adrenoceptors as well as angiotensin II receptors and involve the activation of various kinases such as protein kinase A, Ca2+-calmodulin dependent kinase, protein kinase C and mitogen-activated protein kinase to promote protein synthesis in the myocardium [38,39,40,41]. There also occurs an increase in ventricular diastolic pressure as well as ventricular wall tension, which activate macrophages, fibroblasts, and non-myocyte cells in the myocardial interstitium to release different cytokines and growth factors [42,43,44,45,46,47,48]. In addition, several other neuro-endocrine systems including pituitary, endothelium and platelets are also activated to release vasoactive hormones such as vasopressin, endothelin, and serotonin in the circulation [49,50,51,52]. Thus, different vasoactive hormones at early stages of myocardial infarction can be seen to induce adaptive (physiological) cardiac hypertrophy, stimulate cardiac metabolism and improve cardiac function through their respective receptor mediated signal transduction mechanisms.
Vasoactive hormones generate some amount of oxyradicals in hypertrophied myocardium but the presence of high levels of endogenous antioxidants does not permit the occurrence of oxidative stress [53,54]. However, when the activities of antioxidants become saturated with excessive amounts of oxyradicals or the levels of antioxidants become depressed, there occurs oxidative stress for the development of cardiac dysfunction [55,56,57,58,59]. It should be noted that the increased oxyradical formation may occur due to the activation of NADPH oxidase by angiotensin II and endothelin as well as during the oxidation of catecholamines and serotonin by monoamine oxidase. The combination of oxyradicals with NO, produced by elevated levels of endothelial nitric oxidase in hypertrophied hearts [55], has also been reported to produce nitrosative stress which is known to exert adverse effects on the heart. In addition, defects in mitochondrial electron transport as a consequence of metabolic derangements as well as functional hypoxia, upon prolonged exposure of hypertrophied myocardium to vasoactive hormones, have been shown to contribute to the development of oxidative stress [60,61,62,63,64,65,66]. Since oxidative stress and nitrosative stress have been demonstrated to increase Ca2+-influx, activate different proteases and produce alterations in subcellular proteins gene expression directly or indirectly, these pathological entities are considered to induce subcellular remodeling, Ca2+-handling abnormalities and cardiac dysfunction in hypertrophied hearts due to myocardial infarction [7,10,67,68,69,70]. Thus, the development of oxidative stress has been suggested to play a critical role in the transition of cardiac hyper-trophy to heart failure due to myocardial infarction. Some events depicting different mechanisms in this regard are shown in Figure 2.

2.2. Development of Heart Failure Due to Pressure Overload and Volume Overload

Both pressure overload and volume overload are known to induce cardiac hypertrophy and heart failure upon increasing ventricular pressure as well as ventricular wall tension as a consequence of elevating afterload and preload on the heart, respectively [14,15,25,26,71,72,73,74]. It is now well known that pressure overload occurs in some pathological conditions such as hypertension and aortic or mitral valve stenosis, where the heart develops concentric hypertrophy [71,72,73]. On the other hand, volume overload is seen in some clinical situations including mitral valve or aortic valve regurgitation as well as ventricular septal defect, where the heart develops eccentric hypertrophy [71,72,75,76]. The increase in ventricular wall tension is considered to activate the sympathetic nerve endings, cardiac (local) renin–angiotensin system, endothelium, and several other non-myocytes, present in the myocardial interstitium, to release different vasoactive hormones such as norepinephrine, angiotensin II, and endothelin as well as cytokines and growth factors [31,42,43,44,45,51,77]. Although both the sympathetic nervous system and the peripheral renin–angiotensin system are also activated under situations simulating pressure overload or volume overload [77,78,79,80,81,82,83], exact mechanisms for the release of catecholamines and angiotensin II by these interventions are not clear at present. Nonetheless, both catecholamines and angiotensin II as well as endothelin have been reported to induce cardiac hypertrophy through their respective receptor-mediated signal transduction mechanisms [84,85,86,87]. In this regard, cardiac hypertrophy due to catecholamines is elicited by the activation of both β-adrenoceptor–Gs protein–adenylyl cyclase and α-adrenoceptor–Gq protein–phospholipase C pathways whereas that induced by angiotensin II involves Ang II receptor–Gq protein–phospholipase C system. Furthermore, the involvement of Gq protein–phospholipase C pathway has also been shown to occur due to the activation of endothelin receptors. It is noteworthy that heart function has been reported to be unaltered due to volume overload but is augmented because of pressure overload at early stages of cardiac hypertrophy, indicating differences in the regulatory mechanism participating upon the induction of these pathological stimuli [25].
Despite differences in cardiac remodeling (concentric versus eccentric cardiac hypertrophy) upon the induction of pressure overload or volume overload, both forms of hemodynamic overload exhibit cardiac dysfunction and heart failure over a prolonged period [71,72,73,74,88,89,90,91]. Several mechanisms including adverse cardiac remodeling, subcellular defects, metabolic derangements, Ca2+-handling defects, inflammation, and oxidative stress have been proposed to explain the transition of adaptive (compensated or physiological) hypertrophy to maladaptive (decompensated or pathological) hypertrophy as well as the progression of cardiac hypertrophy to heart failure due to pressure overload or volume overload [15,26,32,48,76]. Furthermore, contractile dysfunction in the failing hearts has been shown to be associated with defects in subcellular organelles for Ca2+-handling in cardiomyocytes [7,10,75,82,85,92]. The transition of cardiac hypertrophy to heart failure was also observed due to abnormalities in extracellular matrix proteins as a consequence of the activation of metallomatrix proteases [7,93,94]. Development of apoptosis in cardiomyocytes due to elevated levels of pro-inflammatory cytokines such as TNF-α has been reported to serve as a mechanism of heart failure due to pressure or volume overload [60,95,96,97]. Defects in the β-adrenoceptor signal transduction have also been demonstrated to be associated with the development of heart failure due to these pathological situations [25,98,99,100,101]. In addition, the occurrence of oxidative stress due to elevated levels of vasoactive hormones has been shown to play a major role for the induction of contractile dysfunction in hypertrophied heart [55,56,57,58,59,102,103,104,105]. Thus, it is evident that complex events may be participating in the genesis of cardiac hypertrophy and heart failure due to pressure overload or volume overload. However, a simplified scheme representing these events is shown in Figure 3.

3. Generation of Oxyradicals, Redox Signaling and Consequences of Oxidative Stress in Failing Hearts

An in-depth analysis of the above observations indicates that various changes in subcellular and metabolic mechanisms during the development of cardiac function and heart failure due to myocardial infarction or hemodynamic overload are associated with the generation of oxyradicals. Several other investigators have suggested that different molecular and cellular alterations occur in cardiac hypertrophy and heart failure as a consequence of oxidative stress [106,107,108,109,110,111,112,113]. Since cardiac hypertrophy is an adaptive process; it appears that the formation of a small amount of oxyradicals may not be sufficient for the development of oxidative stress at early stages of pathological stimulus. However, it may generate redox-sensitive signaling to activate the hypertrophic process in the myocardium. On the other hand, high levels of oxyradicals generated due to a prolonged period of pathological stimulus can be seen to result in oxidative stress, adverse cardiac remodeling, cardiac dysfunction, and heart failure. Such a dual role of oxyradical generation is consistent with other pathogenic mechanisms underlying other cardiovascular diseases including cerebral cavernous malformation disease [114]. The small amount of oxyradicals and oxidants, which are formed due to the activation of NADPH oxidase as well as metabolic stimulation and subsequent mitochondrial electron-transport uncoupling as a consequence of the elevated plasma levels of angiotensin II at early stages of pathological stimulus, are removed by different oxyradical scavengers, superoxide dismutase, and catalase, as well as antioxidants [108,110,112]. This condition initiates the redox-sensitive signaling for the activation of hypertrophic process as well as modulation of subcellular activities in cardiomyocytes. A small amount of superoxide anion is also removed by its interaction with NO, which is produced by endothelial nitric oxide synthase at early stages, but this reaction then results in the formation of peroxynitrite at later stages and exert adverse effects [115]. Likewise, low levels of oxyradicals activate nuclear factor erythroid 2-related factor 2 (Nrf2) antioxidant pathway for producing adaptive responses at early period but this master Nrf2 defense pathway has been shown to sensitize cells to oxidative challenges at later stages [114]. Thus, it appears that the adaptive responses of redox-sensitive signaling during the development of cardiac hypertrophy are dependent upon the type, duration, and magnitude of pathological stimulus.
Both experimental and clinical observations have suggested that oxidative stress in hypertrophied myocardium is increased in heart failure [56,57,58,102,103,104,105,106]. Although the level of antioxidants is also increased in hypertrophied hearts [53,74], it appears that elevated levels of antioxidants at late stages of hypertrophy may not be sufficient to prevent the occurrence of oxidative stress and development of cardiac dysfunction. A large amount of oxyradicals is considered to be formed by a wide variety of mechanisms at late stages of cardiac hypertrophy. In this regard, it is noteworthy that elevated levels of both angiotensin II and endothelins have been shown to activate NADPH oxidase [63,64] whereas catecholamines are oxidized by monoamine oxidase [61,62] to produce oxyradicals in the myocardium. The expression of both NADPH oxidase and monoamine oxidase has been shown to be increased in failing hearts [55]. The vasoactive hormones are also known to impair blood flow to the heart and induce functional hypoxia due to constriction of the coronary arteries as well as reduction in capillary density in hypertrophied hearts [42,43,44,45]. The hypoperfusion thus produced can be seen to increase the production of oxyradicals upon inducing defects in the mitochondrial electron transport due to hypoxic insult and contribute to the development of oxidative stress [37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55]. Furthermore, elevated levels of vasoactive hormones have also been reported to activate fibroblasts in the cardiac interstitium to release different growth factors and metallomatrix proteases [31,43,44]. While the growth factors promote the accumulation of collagenous proteins in the extracellular matrix for providing support to hypertrophied hearts [37,42,46], the activation of metallomatrix proteases by oxidative stress results in degradation of the glycocalyx proteins and subsequent cardiac dysfunction [37,44,60]. Accordingly, oxidative stress upon activating metallomatrix proteases has been suggested to play a critical role in the transition of stable cardiac hypertrophy to heart failure [32,60,93]. Some of the events involved in this process are depicted in Figure 4.
The participation of oxidative stress in the genesis of cardiac dysfunction is attributed to the development of several cardiac abnormalities in hypertrophied heart. Although adverse cardiac remodeling is generally considered to explain the occurrence of heart failure [5], it has been argued that a wide variety of changes in subcellular organelles including sarcolemma, sarcoplasmic reticulum, mitochondria, and myofibrils may be more intimately related to the development of contractile dysfunction during the progression of heart failure [7,10,116,117]. Such subcellular defects during the development of heart failure have been shown to occur because of alterations in cation homeostasis, increased concentration of intracellular Ca2+, activation of proteases, and changes in cardiac gene expressions [7,70,118]. Particularly, defects in sarcolemma and sarcoplasmic reticulum may results in Ca2+-handling abnormalities in myocytes whereas those in myofibrils and mitochondria are associated with changes in contractile properties and energy production in the failing hearts, respectively. It should be mentioned that increased concentration of Ca2+ in the failing heart is known to result in mitochondrial Ca2+-overload and impair the generation of ATP. In fact, oxidative stress has also been associated with marked alterations in myocardial metabolism and mitochondrial electron transport system for depression in energy stores in the failing heart [37,55,60,65,66]. In addition, oxidative stress has been demonstrated to induce loss of cardiomyocytes in the heart by inducing apoptosis, necrosis, and fibrosis as a consequence of myocardial inflammation due to activation of macrophages in the cardiac interstitium and release of different cytokines [31,37,48,55,97,119,120,121,122]. Accordingly, oxidative stress generated in hypertrophied heart can be seen to induce cardiac dysfunction through a complex set of mechanisms and may result in the progression of heart failure. A schematic representation of these events is shown in Figure 5. In view of the critical role of oxidative stress in the pathophysiology of cardiac dysfunction in heart failure, different antioxidants have been suggested to exert beneficial effects for the treatment of this devastating disease [54,55,56,57,58]. Although some clinical studies have been supportive of this concept, other clinical trials with antioxidants have failed to show any conclusive benefit of these interventions for the treatment of heart failure [123,124]. Several investigators have discussed in detail the inability of different non-specific antioxidants to exert beneficial effects in heart failure [107,108,109,115]. Thus, better-targeted and more effective antioxidants need to be developed for improved therapy of this disease. Since, oxidative stress in heart failure is also accompanied by nitrosative stress [115] and inflammation [120,121], it is likely that a combination therapy with antioxidants may prove more appropriate.

4. Evidence for the Implications of Oxidative Stress in Cardiac Dysfunction and Subcellular Remodeling

In view of the association of cardiac dysfunction and oxidative stress during the development of heart failure, various investigators have emphasized the role of oxidative stress in the genesis of subcellular and metabolic defects for the occurrence of contractile abnormalities in hypertrophied hearts [6,10,37,55,104,106]. However, it is not clear whether the occurrence of cardiac dysfunction is a consequence of events associated with cardiac hypertrophy or is due to some direct action of oxidative stress on cardiomyocytes per se. Further discussion in this report is thus focused to provide evidence that the generation of oxidative stress in non-hypertrophied heart leads to the development of subcellular alterations and cardiac dysfunction [125,126,127,128,129]. Different oxygen reactive species, which result in the development of oxidative stress, have been reported to be involved not only in inducing changes in cardiac contractile activity but are also considered to be mediators of the myocardial cell injury [130,131,132,133,134]. This article will examine the effects on oxidative stress generation on changes in subcellular activities and cardiac function in ischemic reperfused hearts in the absence and presence of different oxyradical scavengers and antioxidant interventions. The existing literature for changes in cardiac function and subcellular activities will also be analyzed upon perfusing the hearts with some oxidative stress generating systems. Furthermore, different oxyradicals will be shown to exert direct actions on the activities of cardiac subcellular organelles. Such a detailed examination of the effects of oxidative generating systems on cardiac contractile activities, subcellular remodeling and Ca2+-handling in cardiomyocytes will further support the role of oxidative stress in the development and progression of heart failure.

4.1. Alterations in Cardiac Function and Subcellular Activities in Ischemic Reperfused Hearts

Since ischemia-reperfusion is well known to generate oxyradicals [126,135], some studies have examined the relationship between changes in cardiac function and subcellular alterations upon subjecting the heart to ischemia-reperfusion in the absence and presence of different oxyradical scavengers or antioxidants. Various parameters such as left ventricular systolic pressure, rate of contraction, and rate of relaxation were markedly depressed whereas left ventricular end diastolic pressure was increased in the ischemic reperfused hearts [136,137,138,139]. These alterations in cardiac function were associated with marked depressions in the sarcolemmal Na+-K+ ATPase activity [136,140], Ca2+-pump activity, and Na+-Ca2+ exchange activity [141], as well as β-adrenoceptor–adenylyl cyclase mediated pathway [142]. Dramatic reduction in the sarcoplasmic reticulum Ca2+-uptake, Ca2+-pump ATPase, and Ca2+-release [137] as well as Ca2+/calmodulin protein kinase activities [143] were observed in the ischemic reperfused hearts. Furthermore, depressed cardiac function was seen to be associated with marked alterations in mitochondrial oxidative phosphorylation [138] as well as myofibrillar ATPase activities [139]. All these changes in cardiac function as well as sarcolemma, sarcoplasmic reticulum, mitochondria, and myofibrils due to ischemia-reperfusion were attenuated by the presence of an oxyradicals scavenging mixture (superoxide dismutase plus catalase) in the perfusion medium [136,137,138,139,140,141,142,143]. Furthermore, treatments of the hearts with antioxidants such as N-acetylcysteine (NAC) and N-mercaptopropionylglycine (MPG) were also found to partially or fully prevent the ischemia-reperfusion induced alterations in cardiac function as well as different subcellular organelles [138,139,144].
Hearts subjected to ischemia-reperfusion were observed to exhibit varying degrees of depressions in mRNA levels for sarcoplasmic reticular Ca2+-pump ATPase and Ca2+-release channels [137], sarcolemmal Na+-K+ ATPase α2, α3, and β1 isoforms [140], and myofibrillar myosin heavy chain α- and β-isoforms as well as myosin light chain 1 [139]. All theses changes in gene expression for subcellular proteins due to ischemia-reperfusion were attenuated by superoxide dismutase plus catalase indicating the involvement of oxidative stress for subcellular remodeling [137,139,140]. The ischemia-reperfusion induced alterations in cardiac function, subcellular activities, and subcellular gene expressions are also attenuated by ischemic preconditioning [145,146,147,148], which is known to depress the development of oxidative stress in the myocardium [149,150]. Furthermore, the observed depressions in sarcolemmal and sarcoplasmic reticular enzyme activities in the ischemic reperfused hearts have been reported to be due to the activation of Ca2+-dependent protease, calpain [151,152,153,154]. Since oxidative stress due to ischemia-reperfusion has been demonstrated to produce intracellular Ca2+-overload and activate different proteolytic enzymes [55,155], it is likely that changes in subcellular activities are a consequence of an indirect effect of oxidative stress.

4.2. Alterations in Cardiac Function and Subcellular Activities Due to Oxyradical Generating System or H2O2

The effects of oxidative stress on cardiac function and subcellular activities have been examined by perfusing the heart with either xanthine plus xanthine oxidase (a well known oxyradical generating system) or H2O2, an oxidant [136,137,138,139,140,141,144,156,157,158,159]. The depression in cardiac function upon perfusion with xanthine plus xanthine oxidase or H2O2 was associated with decreased activities of sarcolemmal Na+-K+ ATPase, Na+-Ca2+ exchange, and Ca2+-pump ATPase as well as β-adrenoceptor–adenylyl cyclase system [136,140,141,156,157,158,159]. Likewise, sarcoplasmic reticular Ca2+-uptake and release activities, myofibrillar ATPase and mitochondrial oxidative phosphorylation activities were also reduced upon perfusing the hearts with H2O2 or xanthine plus xanthine oxidase [137,138,139]. These alterations in sarcolemma, sarcoplasmic reticulum, myofibrils, and mitochondria due to oxyradicals and oxidants were attenuated by the presence of superoxide dismutase plus catalase in the perfusion medium [136,137,138,141,156,157,158,159]. Furthermore, treatments of hearts with antioxidants, NAC and MPG, were observed to attenuate xanthine plus xanthine oxidase induce depressions in cardiac function as well as sarcoplasmic reticular Ca2+-uptake and Ca2+-release activities [144]. These observations provide evidence that oxidative stress is intimately involved in inducing cardiac dysfunction and subcellular defects.

4.3. Effects of Oxyradical Generating System and H2O2 on Subcellular Activities

To examine whether oxidative stress induces subcellular alterations directly, various isolated organelles were incubated with systems known to generate different species of reactive oxygen. Both H2O2 and oxyradical generating systems were found to depress sarcolemmal Na+-K+ ATPase, Na+-Ca2+ exchange, and Ca2+-pump activities by promoting lipid peroxidation and modifying the sulfhydryl groups [160,161,162,163]. Likewise, Ca2+-uptake and Ca2+-pump ATPase activities in the sarcoplasmic reticulum were decreased by superoxide and hydroxyl radicals as well as H2O2 [164,165,166,167]. It is noteworthy that different reactive oxygen species were observed to depress the sarcolemmal Ca2+-channel binding activity; the effect by superoxide radical was prevented by superoxide dismutase whereas that by H2O2 and hydroxyl radicals was prevented by catalase and mannitol, respectively [168]. The effects of oxyradicals and H2O2 on sarcolemmal Ca2+-ecto ATPase, ATP-independent Ca2+-binding, β-adrenergic density, and adenylyl cyclase were of biphasic nature and oxyradical species specific [169,170,171]. Various oxyradical generating systems were observed to impair mitochondrial oxidative phosphorylation and reduce myofibrillar ATPase activity [138,172]. These observations are consistent with the view that various oxyradicals and oxidants modify the activities of different subcellular organelles directly in the heart. In addition, as described above, oxidative stress may also alter subcellular activities indirectly by affecting cardiac gene expression as well as activating different proteolytic enzymes as a consequence of increased concentration of intracellular Ca2+ in cardiomyocytes.

5. Conclusions

It is evident that heart failure due to myocardial infarction, pressure overload, or volume overload is mainly associated with elevated levels of plasma catecholamines, oangiotensin II and endothelin. These vasoactive hormones stimulate their receptor-mediated signal transduction pathways and induce cardiac hypertrophy, which is a beneficial mechanism for maintaining or augmenting heart function at initial stages. A small amount of oxyradicals is also generated during early periods of hypertrophic process; these radicals are readily removed by the endogenous scavengers for maintaining redox homeostasis. However, prolonged exposure of hypertrophied heart to pathological stimuli and subsequent high levels of circulating hormones has been demonstrated to promote the development of oxidative stress as a consequence of functional hypoxia due to constriction of the coronary arteries, reduction in the capillary density, myocyte inflammation, metabolic derangements, and mitochondrial dysfunction. Furthermore, activation of NADPH oxidase by hormones such as angiotensin II and endothelins, and oxidation of catecholamines by monoamine oxidase also participate in the generation of oxidative stress. Oxidative stress has been suggested to cause Ca2+-handling abnormalities in association with subcellular remodeling, defect in energy production, inflammation, apoptosis, fibrosis, and loss of cardiomyocytes; these abnormalities are considered to result in cardiac dysfunction and heart failure. Such events showing adverse cardiovascular effects of diverse pathological stimuli for the generation of oxidative stress and subsequent myocardial abnormalities are shown in Figure 6. Evidence has also shown that depression in cardiac function and associated subcellular defects upon exposure of the heart to some oxidative stress generating systems were attenuated by oxyradical scavengers and antioxidants. These observations support the view that the development of oxidative stress in hypertrophied heart is an important mechanism for transition of cardiac hypertrophy to heart failure. Furthermore, the details regarding the molecular and cellular effects as well as Ca2+-handling abnormalities due to oxidative stress provide compelling evidence for the potential use of antioxidants for the treatment of heart failure. Thus, various antioxidants or interventions for increasing the antioxidant reserve in hypertrophied myocardium may produce beneficial effects in preventing the occurrence of cardiac dysfunction and delaying the progression of heart failure. However, the results from some clinical trials of antioxidants in heart failure patients have been disappointing, perhaps due to non-specific nature of these agents. Thus, there is real challenge for the cardiovascular community to develop target-orientated specific antioxidants for improved therapy of heart failure.

Author Contributions

A.K.S. searched the literature, analyzed the data and wrote the first draft; S.K.B. and V.E. analyzed the data and wrote the manuscript; N.S.D. conceived, designed and edited the article. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Data is contained within the article.

Acknowledgments

The infrastructural support for the preparation of this article was provided by the St. Boniface Hospital Albrechtsen Research Centre. Thanks, are also due to Andrea Opsima for typing this manuscript.

Conflicts of Interest

The authors have no conflict of interest. All authors have contributed in preparing and editing this manuscript and have approved its submission for publication.

References

  1. Jessup, M.; Brozena, S. Heart failure. N. Engl. J. Med. 2003, 348, 2007–2018. [Google Scholar] [CrossRef]
  2. Parmley, W.W. Pathophysiology and current therapy of congestive heart failure. J. Am. Coll. Cardiol. 1989, 13, 771–785. [Google Scholar] [CrossRef] [Green Version]
  3. McMurray, J.J.; Stewart, S. Epidemiology, aetiology, and prognosis of heart failure. Heart 2000, 83, 596–602. [Google Scholar] [CrossRef] [Green Version]
  4. Lloyd-Jones, D.M.; Larson, M.G.; Leip, E.P.; Beiser, A.; D’Agostino, R.B.; Kannel, W.B.; Murabito, J.M.; Vasan, R.S.; Benjamin, E.J. Daniel Levy Lifetime risk for developing congestive heart failure: The Framingham Heart Study. Circulation 2002, 106, 3068–3072. [Google Scholar] [CrossRef] [Green Version]
  5. Cohn, J.N.; Ferrari, R.; Sharpe, N. Cardiac remodeling–concepts and clinical implications: A consensus paper from an international forum on cardiac remodeling. Behalf of an International Forum on Cardiac Remodeling. J. Am. Coll. Cardiol. 2000, 35, 569–582. [Google Scholar] [CrossRef] [Green Version]
  6. Dhalla, N.S.; Afzal, N.; Beamish, R.E.; Naimark, B.; Takeda, N.; Nagano, M. Pathophysiology of cardiac dysfunction in congestive heart failure. Can. J. Cardiol. 1993, 9, 873–887. [Google Scholar] [PubMed]
  7. Dhalla, N.S.; Saini–Chohan, H.K.; Rodriguez–Leyva, D.; Elimban, V.; Dent, M.R.; Tappia, P.S. Subcellular remodelling may induce cardiac dysfunction in congestive heart failure. Cardiovasc. Res. 2009, 81, 429–438. [Google Scholar] [CrossRef] [Green Version]
  8. Packer, M. Neurohormonal interactions and adaptations in congestive heart failure. Circulation 1988, 77, 721–730. [Google Scholar] [CrossRef] [Green Version]
  9. Mudd, J.O.; Kass, D.A. Tackling heart failure in the twenty–first century. Nature 2008, 451, 919–928. [Google Scholar] [CrossRef] [PubMed]
  10. Dhalla, N.S.; Dent, M.R.; Tappia, P.S.; Sethi, R.; Barta, J.; Goyal, R.K. Subcellular remodeling as a viable target for the treatment of congestive heart failure. J. Cardiovasc. Pharmacol. Ther. 2006, 11, 31–45. [Google Scholar] [CrossRef] [PubMed]
  11. Xie, M.; Burchfield, J.S.; Hill, J.A. Pathological ventricular remodeling: Mechanisms: Part 1 of 2. Circulation 2013, 128, 388–400. [Google Scholar] [CrossRef] [PubMed]
  12. Lyon, R.C.; Zanella, F.; Omens, J.H.; Sheikh, F. Mechanotransduction in cardiac hypertrophy and failure. Circ. Res. 2015, 116, 1462–1476. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Wikman–Coffelt, J.; Parmley, W.W.; Mason, D.T. The cardiac hypertrophy process: Analyses of factors determining pathological vs. physiological development. Circ. Res. 1979, 45, 697–707. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Dhalla, N.S.; Heyliger, C.E.; Beamish, R.E.; Innes, I.R. Pathophysiological aspects of myocardial hypertrophy. Can. J. Cardiol. 1987, 3, 183–196. [Google Scholar] [PubMed]
  15. Shimizu, I.; Minamino, T. Physiological and pathological cardiac hypertrophy. J. Mol. Cell. Cardiol. 2016, 97, 245–262. [Google Scholar] [CrossRef]
  16. Maillet, M.; van Berlo, J.H.; Molkentin, J.D. Molecular basis of physiological heart growth: Fundamental concepts and new players. Nat. Rev. Mol. Cell. Biol. 2013, 14, 38–48. [Google Scholar] [CrossRef] [Green Version]
  17. Bernardo, B.C.; Weeks, K.L.; Pretorius, L.; McMullen, J.R. Molecular distinction between physiological and pathological cardiac hypertrophy: Experimental findings and therapeutic strategies. Pharmacol. Ther. 2010, 128, 191–227. [Google Scholar] [CrossRef]
  18. Molkentin, J.D.; Dorn, G.W., II. Cytoplasmic signaling pathways that regulate cardiac hypertrophy. Annuv. Rev. Physiol. 2001, 63, 391–426. [Google Scholar] [CrossRef]
  19. Heineke, J.; Molkentin, J.D. Regulation of cardiac hypertrophy by intracellular signalling pathways. Nat. Rev. Mol. Cell Biol. 2006, 7, 589–600. [Google Scholar] [CrossRef]
  20. Heger, J.; Schulz, R.; Euler, G. Molecular switches under TGFβ signaling during progression from cardiac hypertrophy to heart failure. Br. J. Pharmacol. 2016, 173, 3–14. [Google Scholar] [CrossRef] [Green Version]
  21. Schirone, L.; Forte, M.; Palmerio, S.; Yee, D.; Nocella, C.; Angelini, F.; Pagano, F.; Schiavon, S.; Bordin, A.; Carrizzo, A.; et al. A review of the molecular mechanisms underlying the development and progression of cardiac remodeling. Oxid. Med. Cell. Longev. 2017, 2017, 3920195. [Google Scholar] [CrossRef]
  22. Nakamura, M.; Sadoshima, J. Mechanisms of physiological and pathological cardiac hypertrophy. Nat. Rev. Cardiol. 2018, 15, 387–407. [Google Scholar] [CrossRef] [PubMed]
  23. Sygitowicz, G.; Maciejak-Jastrzebska, A.; Sitkiewicz, D. MicroRNAs in the development of left ventricular remodeling and postmyocardial infarction heart failure. Pol. Arch. Intern. Med. 2020, 130, 59–65. [Google Scholar] [PubMed] [Green Version]
  24. Mishra, S.; Kass, K.A. Cellular and molecular pathobiology of heart failure with preserved ejection fraction. Nat. Rev. Cardiol. 2021, 18, 400–423. [Google Scholar] [CrossRef] [PubMed]
  25. Sethi, R.; Saini, H.K.; Guo, X.; Wang, X.; Elimban, V.; Dhalla, N.S. Dependence of changes in beta–adrenoceptor signal transduction on type and stage of cardiac hypertrophy. J. Appl. Physiol. 2007, 102, 978–984. [Google Scholar] [CrossRef] [PubMed]
  26. Oldfield, C.J.; Duhamel, T.A.; Dhalla, N.S. Mechanisms for the transition from physiological to pathological cardiac hypertrophy. Can. J. Physiol. Pharmacol. 2019. [Google Scholar] [CrossRef] [Green Version]
  27. Nicholls, D.P.; Onuoha, G.N.; McDowell, G.; Elborn, J.S.; Riley, M.S.; Nugent, A.-M.; Steele, I.C.; Shaw, C.; Buchanan, K.D. Neuroendocrine changes in chronic cardiac failure. Basic Res. Cardiol. 1996, 91 (Suppl. 1), 13–20. [Google Scholar]
  28. Rouleau, J.L. The neurohormonal hypothesis and the treatment of heart failure. Can. J. Cardiol. 1996, 12 (Suppl. l), 3F–8F. [Google Scholar]
  29. Packer, M. The neurohormonal hypothesis: A theory to explain the mechanism of disease progression in heart failure. J. Am. Coll. Cardiol. 1992, 20, 248–254. [Google Scholar] [CrossRef] [Green Version]
  30. Francis, G.S.; Benedict, C.; Johnstone, D.E.; Kirlin, P.C.; Nicklas, J.; Liang, C.S.; Kubo, S.H.; Rudin-Toretsky, E.; Yusuf, S. Comparison of neuroendocrine activation in patients with left ventricular dysfunction with and without congestive heart failure. A substudy of the studies of left ventricular dysfunction (SOLVD). Circulation 1990, 82, 1724–1729. [Google Scholar] [CrossRef] [Green Version]
  31. Weber, K.T.; Brilla, C.G. Pathological hypertrophy and cardiac interstitium. Fibrosis and renin–angiotensin–aldosterone system. Circulation 1991, 83, 1849–1865. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Janicki, J.S.; Brower, G.L.; Gardner, J.D.; Chancey, A.L.; Stewart, J.A., Jr. The dynamic interaction between matrix metalloproteinase activity and adverse myocardial remodeling. Heart Fail. Rev. 2004, 9, 33–42. [Google Scholar] [CrossRef]
  33. Felker, G.M.; Thompson, R.E.; Hare, J.M.; Hruban, R.H.; Clemetson, D.E.; Howard, D.; Baughman, K.L.; Kasper, E.K. Underlying causes and long–term survival in patients with initially unexplained cardiomyopathy. N. Engl. J. Med. 2000, 342, 1077–1084. [Google Scholar] [CrossRef] [PubMed]
  34. Yu, C.M.; Sanderson, J.E.; Shum, I.O.; Chan, S.; Yeung, L.Y.; Hung, Y.T.; Cockram, C.S.; Woo, K.S. Diastolic dysfunction and natriuretic peptides in systolic heart failure. Higher ANP and BNP levels are associated with the restrictive filling pattern. Eur. Heart J. 1996, 17, 1694–1702. [Google Scholar] [CrossRef] [Green Version]
  35. Clerico, A.; Iervasi, G.; Del Chicca, M.G.; Emdin, M.; Maffei, S.; Nannipieri, M.; Sabatino, L.; Forini, F.; Manfredi, C.; Donato, L. Circulating levels of cardiac natriuretic peptides (ANP and BNP) measured by highly sensitive and specific immunoradiometric assays in normal subjects and in patients with different degrees of heart failure. J. Endocrinol. Investig. 1998, 21, 170–179. [Google Scholar] [CrossRef]
  36. Ruetten, H.; Dimmeler, S.; Gehring, D.; Ihling, C.; Zeiher, A.M. Concentric left ventricular remodeling in endothelial nitric oxide synthase knockout mice by chronic pressure overload. Cardiovasc. Res. 2005, 66, 444–453. [Google Scholar] [CrossRef] [Green Version]
  37. Dhalla, N.S.; Elimban, V.; Shah, A.K.; Nusier, M. Mechanisms of cardiac dysfunction in heart failure due to myocardial infarction. J. Integr. Cardiol. 2019, 2, 2–7. [Google Scholar] [CrossRef] [Green Version]
  38. Wang, J.; Liu, X.; Arneja, A.S.; Dhalla, N.S. Alterations in protein kinase A and protein kinase C levels in heart failure due to genetic cardiomyopathy. Can. J. Cardiol. 1999, 15, 683–690. [Google Scholar] [PubMed]
  39. English, J.M.; Cobb, M.H. Pharmacological inhibitors of MAPK pathways. Trends Pharmacol. Sci. 2002, 23, 40–45. [Google Scholar] [CrossRef]
  40. Zhang, W.; Elimban, V.; Nijjar, M.S.; Gupta, S.K.; Dhalla, N.S. Role of mitogen-activated protein kinase in cardiac hypertrophy and heart failure. Exp. Clin. Cardiol. 2003, 8, 173–183. [Google Scholar]
  41. Kirchhefer, U.; Schmitz, W.; Scholz, H.; Neumann, J. Activity of cAMP-dependent protein kinase and Ca2+/calmodulin-dependent protein kinase in failing and nonfailing human hearts. Cardiovasc. Res. 1999, 42, 254–261. [Google Scholar] [CrossRef] [Green Version]
  42. Weber, K.T. Cardiac interstitium in health and disease. J. Am. Coll. Cardiol. 1989, 13, 1637–1652. [Google Scholar] [CrossRef] [Green Version]
  43. Ju, H.; Zhane, S.; Jassal, D.S.; Dixon, I.M.C. Effect of AT1 receptor blockade on cardiac collagen remodeling after myocardial infarction. Cardiovasc. Res. 1997, 35, 223–232. [Google Scholar] [CrossRef]
  44. Briest, W.; Holzl, A.; Rabler, B.; Deten, A.; Beba, H.A.; Zimmer, H. Significance of matrix metalloproteinases in norepinephrine-induced remodeling of rat hearts. Cardiovasc. Res. 2003, 57, 379–387. [Google Scholar] [CrossRef] [Green Version]
  45. Zak, R. Cell proliferation during cardiac growth. Am. J. Cardiol. 1973, 31, 211–219. [Google Scholar] [CrossRef]
  46. Westermann, D.; Kasner, M.; Steendijk, P.; Spillmann, F.; Riad, A.; Weitmann, K.; Hoffmann, W.; Poller, W.; Pauschinger, M.; Spillmann, F.; et al. Role of left ventricular stiffness in heart failure with normal ejection fraction. Circulation 2008, 117, 2051–2060. [Google Scholar] [CrossRef]
  47. Borbely, A.; van der Velden, J.; Papp, Z.; Bronzwaer, J.G.; Edes, I.; Stienen, G.; Paulus, W.J. Cardiomyocyte stiffness in diastolic heart failure. Circulation 2005, 111, 774–781. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Bartekova, M.; Radosinska, J.; Jelemensky, M.; Dhalla, N.S. Role of cytokines and inflammation in heart function during health and disease. Heart Fail. Rev. 2018, 23, 733–758. [Google Scholar] [CrossRef]
  49. Sanganalmath, S.K.; Barta, J.; Takeda, N.; Kumamoto, H.; Dhalla, N.S. Antiplatelet therapy mitigates cardiac remodeling and dysfunction in congestive heart failure due to myocardial infarction. Can. J. Physiol. Pharmacol. 2008, 86, 180–189. [Google Scholar] [CrossRef]
  50. Sanganalmath, S.K.; Babick, A.; Barta, J.; Kumamoto, H.; Takeda, N.; Dhalla, N.S. Antiplatelet therapy attenuates subcellular remodelling in congestive heart failure. J. Cell. Mol. Med. 2008, 12, 1728–1738. [Google Scholar] [CrossRef] [Green Version]
  51. Vanhoutte, P.M. Endothelium and control of vascular function. Hypertension 1989, 13, 658–667. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Rehsia, N.S.; Dhalla, N.S. Potential of endothelin-1 and vasopressin antagonists for the treatment of congestive heart failure. Heart Fail. Rev. 2010, 15, 85–101. [Google Scholar] [CrossRef] [PubMed]
  53. Gupta, M.; Singal, P.K. Higher antioxidant capacity during a chronic stable heart hypertrophy. Circ. Res. 1989, 64, 398–406. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Singal, P.K.; Gupta, M.; Randhawa, A.K. Reduced myocardial injury due to exogenous oxidants in pressure induced heart hypertrophy. Basic Res. Cardiol. 1991, 86, 273–282. [Google Scholar] [CrossRef] [PubMed]
  55. Dhalla, N.S.; Temsah, R.M.; Netticadan, T. Role of oxidative stress in cardiovascular diseases. J. Hypertens. 2000, 18, 655–673. [Google Scholar] [CrossRef]
  56. Belch, J.J.; Bridges, A.B.; Scott, N.; Chopra, M. Oxygen free radicals and congestive heart failure. Br. Heart J. 1991, 65, 245–248. [Google Scholar] [CrossRef] [Green Version]
  57. Diaz-Velez, C.R.; Garcia-Castinieras, S.; Mendoza Ramos, E.; Hernandez Lopez, E. Increased malondialdehyde in peripheral blood of patients with congestive heart failure. Am. Heart J. 1996, 131, 146–152. [Google Scholar] [CrossRef]
  58. Ghatak, A.; Brar, M.J.; Agarwal, A.; Goel, N.; Rastogi, A.K.; Vaish, A.K.; Sircar, A.R.; Chandra, M. Oxy free radical system in heart failure and therapeutic role of vitamin E. Int. J. Cardiol. 1996, 57, 119–127. [Google Scholar] [CrossRef]
  59. Dhalla, A.K.; Hill, M.F.; Singal, P.K. Role of oxidative stress in transition of hypertrophy to heart failure. Am. J. Coll. Cardiol. 1996, 28, 506–514. [Google Scholar] [CrossRef] [Green Version]
  60. Duhamel, T.A.; Dhalla, N.S. New insight into the causes of heart failure. Drug Discov. Today Dis. Mech. 2017, 4, 175–184. [Google Scholar] [CrossRef]
  61. Dhalla, N.S.; Ademeova, A.; Kaur, M. Role of catecholamine oxidation in sudden cardiac death. Fund. Clin. Pharmacol. 2010, 24, 539–546. [Google Scholar] [CrossRef]
  62. Mialet-Perez, J.; Santin, Y.; Parini, A. Monoamine oxidase-A, serotonin and norepinephrine: Synergistic players in cardiac physiology and pathology. J. Neural Transm. 2018, 125, 1627–1634. [Google Scholar] [CrossRef]
  63. Li, B.; Tian, J.; Sun, Y.; Xu, T.-R.; Chi, R.-F.; Zhang, X.-L.; Hu, X.-L.; Zhang, Y.; Qin, F.-Z.; Zhang, W.-F. Activation of NADPH oxidase mediates increased endoplasmic reticulum stress and left ventricular remodeling after myocardial infarction in rabbits. Biochim. Biophys. Acta 2015, 1852, 805–815. [Google Scholar] [CrossRef] [Green Version]
  64. Nguyen Dinh Cat, A.; Montezano, A.C.; Burger, D.; Touyz, R.M. Angiotensin II, NADPH oxidase, and redox signaling in the vasculature. Antioxid. Redox. Signal 2013, 19, 1110–1120. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Ashrafian, H.; Frenneaux, M.P.; Opie, L.H. Metabolic mechanisms in heart failure. Circulation 2007, 116, 434–448. [Google Scholar] [CrossRef]
  66. Lopaschuk, G.D.; Ussher, J.R.; Folmes, C.D.; Jaswal, J.S.; Stanley, W.C. Myocardial fatty acid metabolism in health and disease. Physiol. Rev. 2010, 90, 207–258. [Google Scholar] [CrossRef]
  67. Wang, J.; Guo, X.; Dhalla, N.S. Modification of myosin protein and gene expression in failing hearts due to myocardial infarction by enalapril or losartan. Biochim. Biophys. Acta 2004, 1690, 177–184. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Shao, Q.; Ren, B.; Saini, H.K.; Netticadan, T.; Takeda, N.; Dhalla, N.S. Sarcoplasmic reticulum Ca2+ transport and gene expression in congestive heart failure are modified by imidapril treatment. Am. J. Physiol. Heart Circ. Physiol. 2005, 288, H1674–H1682. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Shao, Q.; Ren, B.; Elimban, V.; Tappia, P.S.; Takeda, N.; Dhalla, N.S. Modification of sarcolemmal Na+-K+-ATPase and Na+/Ca2+ exchanger expression in heart failure by blockade of renin-angiotensin system. Am. J. Physiol. Heart Circ. Physiol. 2005, 288, H2637–H2646. [Google Scholar] [CrossRef] [Green Version]
  70. Muller, A.L.; Dhalla, N.S. Role of various proteases in cardiac remodeling and progression of heart failure. Heart Fail. Rev. 2012, 17, 395–409. [Google Scholar] [CrossRef]
  71. Cantor, E.J.; Babick, A.P.; Vasanji, Z.; Dhalla, N.S.; Netticadan, T. A comparative serial echocardiographic analysis of cardiac structure and function in rats subjected to pressure or volume overload. J. Moll. Cell. Cardiol. 2005, 38, 777–786. [Google Scholar] [CrossRef] [Green Version]
  72. Carabello, B.A. Concentric versus eccentric remodeling. J. Card. Fail. 2002, 8, S258–S263. [Google Scholar] [CrossRef] [PubMed]
  73. Norton, G.R.; Woodiwiss, A.J.; Gaasch, W.H.; Mela, T.; Chung, E.S.; Aurigemma, G.P.; Meyer, T.E. Heart failure in pressure overload hypertrophy. The relative roles of ventricular remodeling and myocardial dysfunction. J. Am. Coll. Cardiol. 2002, 39, 664–671. [Google Scholar] [CrossRef] [Green Version]
  74. Wang, X.; Ren, B.; Liu, S.; Sentex, E.; Tappia, P.S.; Dhalla, N.S. Characterization of cardiac hypertrophy and heart failure due to volume overload in the rat. J. Appl. Physiol. 2003, 94, 752–763. [Google Scholar] [CrossRef] [PubMed]
  75. Dhalla, N.S.; Golfman, L.; Liu, X.; Sasaki, H.; Elimban, V.; Rapp, H. Subcellular remodeling and heart dysfunction in cardiac hypertrophy due to pressure overload. Ann. N. Y. Acad. Sci. 1999, 874, 100–110. [Google Scholar] [CrossRef]
  76. Frey, N.; Olson, E.N. Cardiac hypertrophy: The good, the bad, and the ugly. Annu. Rev. Physiol. 2003, 65, 45–79. [Google Scholar] [CrossRef]
  77. Ruzicka, M.; Leenen, F.H.H. Relevance of blockade of cardiac and circulatory angiotensin-converting enzyme for the prevention of volume overload-induced cardiac hypertrophy. Circulation 1995, 91, 16–19. [Google Scholar] [CrossRef]
  78. Ruzicka, M.; Keeley, F.W.; Leenen, F.H. The renin-angiotensin system and volume overload-induced changes in cardiac collagen and elastin. Circulation 1994, 90, 1989–1996. [Google Scholar] [CrossRef] [Green Version]
  79. Ganguly, P.K.; Lee, S.L.; Beamish, R.E.; Dhalla, N.S. Altered sympathetic system and adrenoceptors during the development of cardiac hypertrophy. Am. Heart J. 1989, 118, 520–525. [Google Scholar] [CrossRef]
  80. Lindpaintner, K.; Ganten, D. The cardiac renin-angiotensin system. An appraisal of present experimental and clinical evidence. Circ. Res. 1991, 68, 905–921. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Kromer, E.P.; Elsner, D.; Riegger, G.A. Role of neurohormonal systems for pressure induced left ventricular hypertrophy in experimental supravalvular aortic stenosis in rats. Am. J. Hypertens. 1991, 4, 521–524. [Google Scholar] [CrossRef] [PubMed]
  82. Liu, X.; Sentex, E.; Golfman, L.; Takeda, S.; Osada, M.; Dhalla, N.S. Modification of cardiac subcellular remodeling due to pressure overload by captopril and losartan. Clin. Exp. Hypertens. 1999, 21, 145–156. [Google Scholar] [CrossRef] [PubMed]
  83. Tarazi, R.C.; Sen, S.; Saragoca, M.; Khairallah, P. The multifactorial role of catecholamines in hypertensive cardiac hypertrophy. Eur. Heart J. 1982, 3 (Suppl. A), 103–110. [Google Scholar] [CrossRef]
  84. Kallfelt, B.J.; Hjalmarson, A.C.; Isaksson, O.G. In vitro effects of catecholamines on protein synthesis in perfused rat heart. J. Mol. Cell. Cardiol. 1976, 8, 787–802. [Google Scholar] [CrossRef]
  85. Takeo, S.; Elmoselhi, A.B.; Goel, R.; Sentex, E.; Wang, J.; Dhalla, N.S. Attenuation of changes in sarcoplasmic reticular and gene expression in cardiac hypertrophy by propranolol and verapamil. Mol. Cell. Biochem. 2000, 213, 111–118. [Google Scholar] [CrossRef] [PubMed]
  86. Singal, T.; Dhalla, N.S.; Tappia, P.S. Phospholipase C may be involved in norepinephrine-induced cardiac hypertrophy. Biochem. Biophys. Res. Commun. 2004, 320, 1015–1019. [Google Scholar] [CrossRef] [PubMed]
  87. Ju, H.; Scammell-LaFleur, T.; Dixon, I.M.C. Altered mRNA abundance of calcium transport genes in cardiac myocytes in induced by angiotensin II. J. Mol. Cell. Cardiol. 1996, 28, 1119–1128. [Google Scholar] [CrossRef] [PubMed]
  88. Carabello, B.A.; Zile, M.R.; Tanaka, R.; Cooper, G. 4th. Left ventricular hypertrophy due to volume overload versus pressure overload. Am. J. Physiol. Heart Circ. Physiol. 1992, 263, H1137–H1144. [Google Scholar] [CrossRef]
  89. Litwin, S.E.; Katz, S.E.; Weinberg, E.O.; Lorell, B.H.; Aurigemma, G.P.; Douglas, P.S. Serial echocardiographic-Doppler assessment of left ventricular geometry and function in rats with pressure-overload hypertrophy. Chronic angiotensin-converting enzyme inhibition attenuates the transition to heart failure. Circulation 1995, 91, 2642–2654. [Google Scholar] [CrossRef]
  90. Mann, D.L.; Spann, J.F.; Cooper, G. Basic mechanisms and models in cardiac hypertrophy: Pathophysiological models. Mod. Concepts Cardiovasc. Dis. 1988, 57, 7–11. [Google Scholar]
  91. Modesti, P.A.; Vanni, S.; Bertolozzi, I.; Cecioni, I.; Polidori, G.; Paniccia, R.; Bandinelli, B.; Perna, A.; Liguori, P.; Boddi, M.; et al. Early sequence of cardiac adaptations and growth factor formation in pressure- and volume-overload hypertrophy. Am. J. Physiol. Heart Circ. Physiol. 2000, 279, H976–H985. [Google Scholar] [CrossRef] [PubMed]
  92. Ding, Y.F.; Brower, G.L.; Zhong, Q.; Murray, D.; Holland, M.; Janicki, J.S.; Zhong, J. Defective intracellular Ca2+ homeostasis contributes to myocytes dysfunction during ventricular remodeling induced by chronic volume overload in rats. Clin. Exp. Pharmacol. Physiol. 2008, 35, 827–835. [Google Scholar] [CrossRef] [PubMed]
  93. Spinale, F.G. Myocardial matrix remodeling and the matrix metalloproteinases: Influence on cardiac form and function. Physiol. Rev. 2007, 87, 1285–1342. [Google Scholar] [CrossRef] [PubMed]
  94. Dolgilevich, S.M.; Siri, F.M.; Atlas, S.A.; Eng, C. Changes in collagenase and collagen gene experession after induction of aortocaval fistula in rats. Am. J. Physiol. Heart Circ. Physiol. 2001, 281, H207–H214. [Google Scholar] [CrossRef] [PubMed]
  95. Chen, Q.M.; Tu, V.C. Apoptosis and heart failure: Mechanisms and therapeutic implications. Am. J. Cardiovasc. Drugs 2002, 2, 43–57. [Google Scholar] [CrossRef] [PubMed]
  96. Li, Z.; Bing, O.H.; Long, X.; Robinson, K.G.; Lakatta, E.G. Increased cardiomyocyte apoptosis during the transition to heart failure in the spontaneously hypertensive rat. Am. J. Physiol. 1997, 272, H2213–H2319. [Google Scholar] [CrossRef]
  97. Dent, M.R.; Das, S.; Dhalla, N.S. Alterations in both death and survival signals for apoptosis in heart failure due to volume overload. J. Mol. Cell. Cardiol. 2007, 43, 726–732. [Google Scholar] [CrossRef]
  98. Vatner, D.E.; Asai, K.; Iwase, M.; Ishikawa, Y.; Shannon, R.P.; Homcy, C.J.; Vatner, S.F. β-Adrenergic receptors-G protein-adenylyl cyclase signal transduction in the failing heart. Am. J. Cardiol. 1999, 83, 80H–85H. [Google Scholar] [CrossRef]
  99. Vatner, D.E.; Vatner, S.F.; Fuji, A.M.; Homcy, C.J. Loss of high affinity cardiac β-Adrenergic receptors in dogs with heart failure. J. Clin. Investig. 1985, 76, 2259–2264. [Google Scholar] [CrossRef]
  100. Wang, X.; Sentex, E.; Chapman, D.; Dhalla, N.S. Alterations of adenylyl cyclase and G proteins in aortocaval shunt-induced heart failure. An. J. Physiol. Heart Circ. Physiol. 2004, 287, H118–H125. [Google Scholar] [CrossRef]
  101. Wang, X.; Sentex, E.; Saini, H.K.; Chapman, D.; Dhalla, N.S. Upregulation of β-Adrenergic receptors in heart failure due to volume overload. Am. J. Physiol. Heart Circ. Physiol. 2005, 289, H151–H159. [Google Scholar] [CrossRef] [Green Version]
  102. Prasad, K.; Gupta, J.B.; Kalra, J.; Lee, P.; Mantha, S.V.; Bharadwaj, B. Oxidative stress as a mechanism of cardiac failure in chronic volume overload in canine model. J. Mol. Cell. Cardiol. 1996, 28, 375–385. [Google Scholar] [CrossRef]
  103. Keith, M.; Geranmayegan, A.; Sole, M.J.; Kurian, R.; Robinson, A.; Omran, A.S.; Jeejeebhoy, K.N. Increased oxidative stress in patients with congestive heart failure. J. Am. Cell. Cardiol. 1998, 31, 1352–1356. [Google Scholar] [CrossRef] [Green Version]
  104. Singal, P.K.; Khaper, N.; Palace, V.; Kumar, D. The role of oxidative stress in the genesis of heart disease. Cardiovasc. Res. 1998, 3, 426–432. [Google Scholar] [CrossRef]
  105. Prasad, K.; Gupta, J.B.; Kalra, J.; Bharadwaj, B. Oxygen free radicals in volume overload heart failure. Mol. Cell. Biochem. 1992, 111, 55–59. [Google Scholar] [CrossRef] [PubMed]
  106. Singal, P.K.; Khaper, N.; Farahmand, F.; Bello-Klein, A. Oxidative stress in congestive heart failure. Curr. Cardiol. Rep. 2000, 2, 206–211. [Google Scholar] [CrossRef] [PubMed]
  107. Takimoto, E.; Kass, D.A. Role of oxidative stress in cardiac hypertrophy and remodeling. Hypertension 2007, 49, 241–248. [Google Scholar] [CrossRef]
  108. Seddon, M.; Looi, Y.H.; Shah, A.M. Oxidative stress and redox signaling in cardiac hypertrophy and heart failure. Heart 2007, 93, 903–907. [Google Scholar] [CrossRef] [Green Version]
  109. Tsutsui, H.; Kinugawa, S.; Matsushima, S. Oxidative stress and heart failure. Am. J. Physiol. Heart Circ. Physiol. 2011, 301, H2181–H2190. [Google Scholar] [CrossRef] [Green Version]
  110. Maulik, S.K.; Kumar, S. Oxidative stress and cardiac hypertrophy: A review. Toxicol. Mech. Methods 2012, 22, 359–366. [Google Scholar] [CrossRef]
  111. Tsutsui, H. Mitochondrial oxidative stress and heart failure. Intern. Med. 2006, 45, 809–813. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Ramachandra, C.J.A.; Cong, S.; Chan, X. Oxidative stress in cardiac hypertrophy: From molecular mechanisms to novel therapeutic targets. Free Radic. Biol. Med. 2021, 166, 297–312. [Google Scholar] [CrossRef]
  113. Tsutsui, H.; Kinugawa, S.; Matsushima, S. Oxidative stress and mitochondrial DNA damage in heart failure. Circ. J. 2008, 72 (Suppl. A), A31–A37. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Antognelli, C.; Trapani, E.; Delle Monache, S.; Perrelli, A.; Daga, M.; Pizzimenti, S.; Barrera, G.; Cassoni, P.; Angelucci, A.; Trabalzini, L.; et al. KRIT1 loss-of-function induces a chronic Nrf2-mediated adaptive homeostasis that sensitizes cells to oxidative stress: Implication for cerebral cavernous malformation disease. Free Radic. Biol. Med. 2018, 115, 202–218. [Google Scholar] [CrossRef] [PubMed]
  115. Ziolo, M.T.; Houser, S.R. Abnormal Ca2+ cycling in failing ventricular myocytes: Role of NOS1-mediated Nitroso-Redox balance. Antioxid. Redox Signal. 2014, 21, 2044–2059. [Google Scholar] [CrossRef] [Green Version]
  116. Dhalla, N.S.; Das, P.K.; Sharma, G.P. Subcellular basis of cardiac contractile failure. J. Mol. Cell. Cardiol. 1978, 10, 363–385. [Google Scholar] [CrossRef]
  117. Dhalla, N.S.; Shao, Q.; Panagia, V. Remodeling of cardiac membranes during the development of congestive heart failure. Heart Fail. Rev. 1998, 2, 261–272. [Google Scholar] [CrossRef]
  118. Dhalla, N.S.; Rangi, S.; Babick, A.P.; Zieroth, S.; Elimban, V. Cardiac remodeling and subcellular defects in heart failure due to myocardial infarction and aging. Heart Fail. Rev. 2012, 17, 671–681. [Google Scholar] [CrossRef]
  119. Dhingra, S.; Sharma, A.K.; Singla, D.K.; Singal, P.K. P38 and ERK ½ MAPKs mediate the interplay of TNF-alpha and IL-10 in regulating oxidative stress and cardiac myocyte apoptosis. Am. J. Physiol. Heart Circ. Physiol. 2007, 293, H3524–H3531. [Google Scholar] [CrossRef]
  120. Milinkovic, I.; Polovina, M.; Simeunovic, D.S.; Ašanin, M.; Seferović, P.M. Oxidative stress and inflammation in heart failure: The best is yet to come. Eur. J. Prev. Cardiol. 2020, 27, 490–493. [Google Scholar] [CrossRef] [Green Version]
  121. Neri, M.; Fineschi, V.; Di Paolo, M.; Pomara, C.; Riezzo, I.; Turillazzi, E.; Cerretani, D. Cardiac oxidative stress and inflammatory cytokines response after myocardial infarction. Curr. Vasc. Pharmacol. 2015, 13, 26–36. [Google Scholar] [CrossRef]
  122. Ayoub, K.F.; Pothineni, N.V.K.; Rutland, J.; Ding, Z.; Mehta, J.L. Immunity, inflammation, and oxidative stress in heart failure: Emerging molecular targets. Cardiovasc. Drugs Ther. 2017, 31, 593–608. [Google Scholar] [CrossRef]
  123. van der Pol, A.; van Gilst, W.H.; Voors, A.A.; van der Meer, P. Treating oxidative stress in heart failure: Past, present and future. Eur. J. Heart Fail. 2019, 21, 425–435. [Google Scholar] [CrossRef] [PubMed]
  124. Bartekova, M.; Adameova, A.; Gorbe, A.; Ferenczyová, K.; Pecháňová, O.; Lazou, A.; Dhalla, N.S.; Ferdinandy, P.; Giricz, Z. Natural and synthetic antioxidants targeting cardiac oxidative stress and redox signaling in cardiometabolic diseases. Free Radic. Biol. Med. 2021, 169, 446–477. [Google Scholar] [CrossRef] [PubMed]
  125. Dhalla, N.S.; Saini, H.K.; Tappia, P.S.; Sethi, R.; Mengi, S.A.; Gupta, S.K. Potential role and mechanism of subcellular remodeling in cardiac function due to ischemic heart disease. J. Cardiovasc. Med. 2007, 8, 238–250. [Google Scholar] [CrossRef] [PubMed]
  126. Dhalla, N.S.; Shah, A.K.; Tappia, P.S. Role of oxidative stress in metabolic and subcellular abnormalities in diabetic cardiomyopathy. Int. J. Mol. Sci. 2020, 21, 2413. [Google Scholar] [CrossRef] [Green Version]
  127. Ducas, A.; Bartekova, M.; Dhalla, N.S. Ischemia-reperfusion injury of the heart: Moving forward with our knowledge. J. Heart Health 2015, 1. [Google Scholar] [CrossRef]
  128. Dhalla, N.S.; Golfman, L.; Takeda, S.; Takeda, N.; Nagano, M. Evidence for the role of oxidative stress in acute ischemic heart disease: A brief review. Can. J. Cardiol. 1999, 15, 587–593. [Google Scholar]
  129. Dhalla, N.S.; Panagia, V.; Singal, P.K.; Makino, N.; Dixon, I.M.C.; Eyolfson, D.A. Alterations in heart membrane calcium transport during the development of ischemia-reperfusion injury. J. Moll. Cell. Cardiol. 1988, 20 (Suppl. 2), 3–13. [Google Scholar] [CrossRef]
  130. Burton, K.P.; McCord, J.M.; Ghai, G. Myocardial alterations due to free-radical generation. Am. J. Physiol. 1984, 246, H776–H783. [Google Scholar] [CrossRef]
  131. Gupta, M.; Singal, P.K. Oxygen radical injury in the presence of desferal, a specific iron-chelating agent. Biochem. Pharmacol. 1987, 36, 3774–3777. [Google Scholar] [CrossRef]
  132. Hammond, B.; Hess, M.L. The oxygen free radical system: Potential mediator of myocardial injury. J. Am. Coll. Cardiol. 1985, 6, 215–220. [Google Scholar] [CrossRef] [Green Version]
  133. Hess, M.L.; Manson, N.H.; Okabe, E. Involvement of free radicals in pathophysiology of ischemic heart disease. Can. J. Physiol. Pharmacol. 1982, 60, 1382–1389. [Google Scholar] [CrossRef] [PubMed]
  134. Kim, M.S.; Akera, T.O. O2 free radicals: Cause of ischemia-reperfusion injury to cardiac Na+-K+ ATPase. Am. J. Physiol. 1987, 256, H368–H374. [Google Scholar] [CrossRef]
  135. Zweier, J.L.; Flaherty, J.T.; Weisfeldt, M.L. Direct measurement of free radical generation following reperfusion of ischemic myocardium. Proc. Nat. Acad. Sci. USA 1987, 84, 1404–1407. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Ostadal, P.; Elmoselhi, A.B.; Zdobnicka, I.; Lukas, A.; Elimban, V.; Dhalla, N.S. Role of oxidative stress in ischemia-reperfusion-induced changes in, Na+,K+-ATPase isoform expression in rat heart. Antioxid. Redox. Signal. 2007, 6, 914–923. [Google Scholar] [CrossRef] [PubMed]
  137. Temsah, R.M.; Netticadan, T.; Chapman, D.; Takeda, S.; Mochizuki, S.; Dhalla, N.S. Alterations in sarcoplasmic reticulum function and gene expression in ischemic-reperfused rat heart. Am. J. Physiol. 1999, 277, H584–H594. [Google Scholar] [CrossRef] [PubMed]
  138. Makazan, Z.; Saini, H.K.; Dhalla, N.S. Role of oxidative stress in alterations of mitochondrial function in ischemic-reperfused hearts. Am. J. Physiol. Heart Circ. Physiol. 2007, 292, H1986–H1994. [Google Scholar] [CrossRef] [PubMed]
  139. Madikka, S.; Elimban, V.; Chapman, D.; Dhalla, N.S. Role of oxidative stress in ischemia-reperfusion-induced alterations in myofibrillar ATPase activities and gene expression in the heart. Can. J. Physiol. Pharmacol. 2009, 87, 120–129. [Google Scholar] [CrossRef] [PubMed]
  140. Ostadal, P.; Elmoselhi, A.B.; Zdbonicka, I.; Lukas, A.; Chapman, D.; Dhalla, N.S. Ischemia-reperfusion alters gene expression of Na+-K+ ATPase isoforms in rat heart. Biochem. Biophys. Res. Commun. 2003, 306, 457–462. [Google Scholar] [CrossRef]
  141. Dixon, I.M.C.; Kaneko, M.; Hata, T.; Panagia, V.; Dhalla, N.S. Alterations in cardiac membrane Ca2+ transport during oxidative stress. Mol. Cell. Biochem. 1990, 99, 125–133. [Google Scholar] [CrossRef]
  142. Persad, S.; Takeda, S.; Panagia, V.; Dhalla, N.S. β-adrenoceptor-linked signal transduction in ischemic-reperfused heart and scavenging of oxyradicals. J. Moll. Cell. Cardiol. 1997, 29, 545–558. [Google Scholar] [CrossRef]
  143. Netticadan, T.; Temsah, R.; Osada, M.; Dhalla, N.S. Status of Ca2+/calmodulin protein kinase phosphorylation of cardiac SR proteins in ischemia-reperfusion. Am. J. Physiol. 1999, 277, C384–C391. [Google Scholar] [CrossRef]
  144. Bartekova, M.; Barancik, M.; Ferenczyova, K.; Dhalla, N.S. Beneficial effects of N-acetylcysteine and N-mercaptopropionylglycine in ischemia reperfusion injury in the heart. Curr. Med. Chem. 2018, 25, 355–366. [Google Scholar] [CrossRef] [PubMed]
  145. Elmoselhi, A.B.; Lukas, A.; Ostadal, P.; Dhalla, N.S. Preconditioning attenuates ischemia-reperfusion-induced remodeling of Na+-K+-ATPase in hearts. Am. J. Physiol. Heart Circ. Physiol. 2003, 285, H1055–H1063. [Google Scholar] [CrossRef] [Green Version]
  146. Osada, M.; Netticadan, T.; Kawabata, K.; Tamura, K.; Dhalla, N.S. Ischemic preconditioning prevents I/R-induced alterations in SR calcium-calmodulin protein kinase II. Am. J. Physiol. Heart Circ. Physiol. 2000, 278, H1791–H1798. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Osada, M.; Netticadan, T.; Tamura, K.; Dhalla, N.S. Modification of ischemia-reperfusion-induced changes in cardiac sarcoplasmic reticulum by preconditioning. Am. J. Physiol. 1998, 274, H2025–H2034. [Google Scholar] [CrossRef]
  148. Temsah, R.M.; Kawabata, K.; Chapman, D.; Dhalla, N.S. Preconditioning prevents alterations in cardiac SR gene expression due to ischemia-reperfusion. Am. J. Physiol. Heart Circ. Physiol. 2002, 282, H1461–H1466. [Google Scholar] [CrossRef] [Green Version]
  149. Saini, H.K.; Machackova, J.; Dhalla, N.S. Role of reactive oxygen species in ischemic preconditioning of subcellular organelles in the heart. Antioxid. Redox Signal. 2004, 6, 393–404. [Google Scholar] [CrossRef]
  150. Muller, A.L.; Dhalla, N.S. Mechanisms of the beneficial actions of ischemic preconditioning on subcellular remodeling in ischemic-reperfused heart. Curr. Cardiol. Rev. 2010, 6, 255–264. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  151. Singh, R.B.; Hryshko, L.; Freed, D.; Dhalla, N.S. Activation of proteolytic enzymes and depression of the sarcolemmal Na+/K+-ATPase in ischemia-reperfused heart may be mediated through oxidative stress. Can. J. Physiol. Pharmacol. 2012, 90, 249–260. [Google Scholar] [CrossRef] [PubMed]
  152. Muller, A.L.; Freed, D.; Dhalla, N.S. Activation of proteases and changes in Na+-K+-ATPase subunits in hearts subjected to ischemia-reperfusion. J. Appl. Physiol. 2013, 114, 351–360. [Google Scholar] [CrossRef] [Green Version]
  153. Singh, R.B.; Dhalla, N.S. Ischemia-reperfusion-induced changes in sarcolemmal Na+/K+-ATPase are due to the activation of calpain in the heart. Can. J. Physiol. Pharmacol. 2010, 88, 388–397. [Google Scholar] [CrossRef] [PubMed]
  154. Singh, R.B.; Chohan, P.K.; Dhalla, N.S.; Netticadan, T. The sarcoplasmic reticulum proteins are targets for calpain action in the ischemic-reperfused heart. J. Moll. Cell. Cardiol. 2004, 37, 101–110. [Google Scholar] [CrossRef] [PubMed]
  155. Dhalla, N.S.; Elmoselhi, A.B.; Hata, T.; Makino, N. Status of myocardial antioxidants in ischemia-reperfusion injury. Cardiovsac. Res. 2000, 47, 446–456. [Google Scholar] [CrossRef]
  156. Matsubara, T.; Dhalla, N.S. Relationship between mechanical dysfunction and depression of sarcolemmal Ca2+-pump activity in hearts perfused with oxygen free radicals. Mol. Cell. Biochem. 1996, 160–161, 179–185. [Google Scholar] [CrossRef] [PubMed]
  157. Matsubara, T.; Dhalla, N.S. Effect of oxygen free radicals on cardiac contractile activity and sarcolemmal Na+-Ca2+ exchange. J. Cardiovasc. Pharmacol. Therapeut. 1996, 1, 211–218. [Google Scholar] [CrossRef]
  158. Persad, S.; Panagia, V.; Dhalla, N.S. Role of H2O2 in changing β-adrenoceptor and adenylyl cyclase in ischemia-reperfused hearts. Mol. Cell. Biochem. 1998, 186, 99–106. [Google Scholar] [CrossRef]
  159. Persad, S.; Takeda, S.; Dhalla, N.S. Alterations in β-adrenoceptor mechanisms in hearts perfused with xanthine plus xanthine oxidase. J. Cardiovasc. Pharmacol. Therapeut. 1997, 2, 115–124. [Google Scholar] [CrossRef]
  160. Kaneko, M.; Beamish, R.E.; Dhalla, N.S. Depression of heart sarcolemmal Ca2+-pump activity by oxygen free radicals. Am. J. Physiol. 1989, 256, H368–H374. [Google Scholar] [CrossRef]
  161. Kaneko, M.; Elimban, V.; Dhalla, N.S. Mechanism for depression of heart sarcolemmal Ca2+ pump by oxygen free radicals. Am. J. Physiol. 1989, 257, H804–H811. [Google Scholar] [CrossRef]
  162. Hata, T.; Kaneko, M.; Beamish, R.E.; Dhalla, N.S. Influence of oxygen free radicals on heart sarcolemmal Na+-Ca2+ exchange. Coron. Artery Dis. 1991, 2, 397–407. [Google Scholar] [CrossRef]
  163. Shao, Q.; Matsubara, T.; Bhatt, S.K.; Dhalla, N.S. Inhibition of cardiac sarcolemma Na+-K+ ATPase by oxyradical generating systems. Mol. Cell. Biochem. 1995, 147, 139–144. [Google Scholar] [CrossRef]
  164. Kramer, J.H.; Mak, I.T.; Weglicki, W.B. Differential sensitivity of canine cardiac sarcolemmal and microsomal enzymes to inhibition by free radical-induced lipid peroxidation. Circ. Res. 1984, 55, 120–124. [Google Scholar] [CrossRef] [Green Version]
  165. Okabe, E.; Hess, M.L.; Oyama, M.; Ito, H. Characterization of free radical-mediated damage of canine cardiac sarcoplasmic reticulum. Arch. Biochem. Biophys. 1983, 225, 164–177. [Google Scholar] [CrossRef]
  166. Rowe, G.T.; Manson, N.H.; Caplan, M.; Hess, M.L. Hydrogen peroxide and hydroxyl radical mediation of activated leukocyte depression of cardiac sarcoplasmic reticulum. Circ. Res. 1983, 53, 584–591. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Scherer, N.M.; Deamer, D.W. Oxidative stress impairs the function of sarcoplasmic reticulum by oxidation of sulfhydryl groups in the Ca2+-ATPase. Arch. Biochem. Biophys. 1986, 246, 589–610. [Google Scholar] [CrossRef]
  168. Kaneko, M.; Lee, S.L.; Wolf, C.M.; Dhalla, N.S. Reduction of calcium channel antagonist binding sites by oxygen free radicals in rat heart. J. Moll. Cell. Cardiol. 1989, 21, 935–943. [Google Scholar] [CrossRef]
  169. Kaneko, M.; Singal, P.K.; Dhalla, N.S. Alterations in heart sarcolemmal Ca2+-ATPase and Ca2+-binding activities due to oxygen free radicals. Basic Res. Cardiol. 1990, 85, 45–54. [Google Scholar] [CrossRef]
  170. Kaneko, M.; Chapman, D.C.; Ganguly, P.K.; Beamish, R.E.; Dhalla, N.S. Modification of cardiac adrenergic receptors by oxygen free radicals. Am. J. Physiol. 1991, 260, H821–H826. [Google Scholar] [CrossRef]
  171. Persad, S.; Rupp, H.; Jindal, R.; Arneja, J.; Dhalla, N.S. Modification of cardiac β-adrenoceptor mechanisms by H2O2. Am. J. Physiol. 1998, 274, H416–H423. [Google Scholar] [CrossRef] [PubMed]
  172. Suzuki, S.; Kaneko, M.; Chapman, D.C.; Dhalla, N.S. Alterations in cardiac contractile proteins due to oxygen free radicals. Biochim. Biophys. Acta 1991, 1074, 95–100. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Schematic representation for the development of cardiac hypertrophy and heart failure as well as associated cardiovascular abnormalities due to diverse pathological stimuli.
Figure 1. Schematic representation for the development of cardiac hypertrophy and heart failure as well as associated cardiovascular abnormalities due to diverse pathological stimuli.
Antioxidants 10 00931 g001
Figure 2. Schematic representation of events for the development of cardiac hypertrophy and oxidative stress as well as the occurrence of cardiac dysfunction in heart failure due to myocardial infarction. SNS—sympathetic nervous system; RAS—renin-angiotensin system; Ang II—angiotensin II; MAO—monoamine oxidase.
Figure 2. Schematic representation of events for the development of cardiac hypertrophy and oxidative stress as well as the occurrence of cardiac dysfunction in heart failure due to myocardial infarction. SNS—sympathetic nervous system; RAS—renin-angiotensin system; Ang II—angiotensin II; MAO—monoamine oxidase.
Antioxidants 10 00931 g002
Figure 3. Schematic representation of events for the development of cardiac hypertrophy and oxidative stress as well as the occurrence of cardiac dysfunction in heart failure due to pressure overload or volume overload. RAS—renin–angiotensin system.
Figure 3. Schematic representation of events for the development of cardiac hypertrophy and oxidative stress as well as the occurrence of cardiac dysfunction in heart failure due to pressure overload or volume overload. RAS—renin–angiotensin system.
Antioxidants 10 00931 g003
Figure 4. Schematic representation of events depicting the involvement of reduction in capillary density, constriction of coronary arteries and activation of fibroblasts in the hypertrophied myocardium for the transition of cardiac hypertrophy to heart failure due to prolonged elevated levels of circulating hormones. ECM—extracellular matrix; MMPs—metallomatrix proteases.
Figure 4. Schematic representation of events depicting the involvement of reduction in capillary density, constriction of coronary arteries and activation of fibroblasts in the hypertrophied myocardium for the transition of cardiac hypertrophy to heart failure due to prolonged elevated levels of circulating hormones. ECM—extracellular matrix; MMPs—metallomatrix proteases.
Antioxidants 10 00931 g004
Figure 5. Schematic representation of events depicting the involvement of alterations in myocardial metabolism, cation homeostasis as well as activation of macrophages for the occurrence of cardiac dysfunction and progression of heart failure due to the development of oxidative stress hormones in hypertrophied heart for a prolonged period.
Figure 5. Schematic representation of events depicting the involvement of alterations in myocardial metabolism, cation homeostasis as well as activation of macrophages for the occurrence of cardiac dysfunction and progression of heart failure due to the development of oxidative stress hormones in hypertrophied heart for a prolonged period.
Antioxidants 10 00931 g005
Figure 6. Schematic representation of events depicting the generation of oxidative stress due to prolonged exposure of hypertrophied heart to vasoactive hormones as well as the occurrence of cardiac abnormalities due to oxidative stress for the development of heart failure in hypertrophied myocardium.
Figure 6. Schematic representation of events depicting the generation of oxidative stress due to prolonged exposure of hypertrophied heart to vasoactive hormones as well as the occurrence of cardiac abnormalities due to oxidative stress for the development of heart failure in hypertrophied myocardium.
Antioxidants 10 00931 g006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Shah, A.K.; Bhullar, S.K.; Elimban, V.; Dhalla, N.S. Oxidative Stress as A Mechanism for Functional Alterations in Cardiac Hypertrophy and Heart Failure. Antioxidants 2021, 10, 931. https://doi.org/10.3390/antiox10060931

AMA Style

Shah AK, Bhullar SK, Elimban V, Dhalla NS. Oxidative Stress as A Mechanism for Functional Alterations in Cardiac Hypertrophy and Heart Failure. Antioxidants. 2021; 10(6):931. https://doi.org/10.3390/antiox10060931

Chicago/Turabian Style

Shah, Anureet K., Sukhwinder K. Bhullar, Vijayan Elimban, and Naranjan S. Dhalla. 2021. "Oxidative Stress as A Mechanism for Functional Alterations in Cardiac Hypertrophy and Heart Failure" Antioxidants 10, no. 6: 931. https://doi.org/10.3390/antiox10060931

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop