Next Article in Journal
Oxidative Stress and Antioxidants in Pediatric Asthma’s Evolution and Management
Next Article in Special Issue
Oxidative Stress and Cardiovascular Complications in Type 2 Diabetes: From Pathophysiology to Lifestyle Modifications
Previous Article in Journal
Interactions Between Ferroptosis and Oxidative Stress in Ischemic Stroke
Previous Article in Special Issue
Eicosapentaenoic Acid (EPA) and Docosahexaenoic Acid (DHA) Ameliorate Heart Failure through Reductions in Oxidative Stress: A Systematic Review and Meta-Analysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Mitochondrial Reactive Oxygen Species Dysregulation in Heart Failure with Preserved Ejection Fraction: A Fraction of the Whole

by
Caroline Silveira Martinez
,
Ancheng Zheng
and
Qingzhong Xiao
*
Centre for Clinical Pharmacology and Precision Medicine, William Harvey Research Institute, Faculty of Medicine and Dentistry, Queen Mary University of London, London EC1M 6BQ, UK
*
Author to whom correspondence should be addressed.
Antioxidants 2024, 13(11), 1330; https://doi.org/10.3390/antiox13111330
Submission received: 25 September 2024 / Revised: 19 October 2024 / Accepted: 28 October 2024 / Published: 31 October 2024
(This article belongs to the Special Issue Oxidative Stress in Cardiovascular Diseases (CVDs))

Abstract

:
Heart failure with preserved ejection fraction (HFpEF) is a multifarious syndrome, accounting for over half of heart failure (HF) patients receiving clinical treatment. The prevalence of HFpEF is rapidly increasing in the coming decades as the global population ages. It is becoming clearer that HFpEF has a lot of different causes, which makes it challenging to find effective treatments. Currently, there are no proven treatments for people with deteriorating HF or HFpEF. Although the pathophysiologic foundations of HFpEF are complex, excessive reactive oxygen species (ROS) generation and increased oxidative stress caused by mitochondrial dysfunction seem to play a critical role in the pathogenesis of HFpEF. Emerging evidence from animal models and human myocardial tissues from failed hearts shows that mitochondrial aberrations cause a marked increase in mitochondrial ROS (mtROS) production and oxidative stress. Furthermore, studies have reported that common HF medications like beta blockers, angiotensin receptor blockers, angiotensin-converting enzyme inhibitors, and mineralocorticoid receptor antagonists indirectly reduce the production of mtROS. Despite the harmful effects of ROS on cardiac remodeling, maintaining mitochondrial homeostasis and cardiac functions requires small amounts of ROS. In this review, we will provide an overview and discussion of the recent findings on mtROS production, its threshold for imbalance, and the subsequent dysfunction that leads to related cardiac and systemic phenotypes in the context of HFpEF. We will also focus on newly discovered cellular and molecular mechanisms underlying ROS dysregulation, current therapeutic options, and future perspectives for treating HFpEF by targeting mtROS and the associated signal molecules.

1. Introduction

Heart failure (HF) is a potentially fatal medical condition characterized by a combination of cardiac and non-cardiac symptoms. Recent guidelines classify HF into distinct phenotypes based on left ventricular ejection fraction (LVEF): HF with reduced ejection fraction (HFrEF, LVEF as ≤40%), HF with mildly reduced ejection fraction (HFmrEF, LVEF between 41% and 49%), and HF with preserved ejection fraction (HFpEF, LVEF ≥ 50%) [1]. According to the latest ESC guideline, apart from having a LVEF ≥ 50%, the presence of symptoms and signs of HF (exertional intolerance, breathlessness, extravascular fluid accumulation in the lungs, subcutaneous tissues, and abdominal cavity), evidence of structural and/or functional cardiac abnormalities, and/or raised natriuretic peptides (NPs) are diagnostic indicators of HFpEF [1]. HFpEF is a significant issue in public health that imposes a substantial healthcare burden on both individuals and countries. The expenditure of HF-related costs in 2012 is estimated to be USD 30.7 billion, and the projection is an increase of 127% by 2030 [1,2]. HFpEF is not an isolated cardiac disease but a metabolic syndrome with a high comorbidity burden, with its prevalence increases with age. Common comorbidities include hypertension, diabetes, obesity, atrial fibrillation (AF), anemia, chronic obstructive pulmonary disease (COPD), and chronic kidney disease (CKD). The combination of these comorbidities seems to play a key pathophysiological role in the development of the myocardial stiffening, fibrosis, and diastolic dysfunction via systemic microvascular inflammation and coronary microvascular dysfunction in patients with HFpEF [3].
Increasing evidence has suggested that excessive reactive oxygen species (ROS) generation and oxidative stress seem to play a critical role in the pathogenesis of HFpEF [4,5]. Although ROS are normally linked to harmful effects on cardiac remodeling, small amounts of ROS is required for maintaining mitochondrial homeostasis and cardiac functions. In this review, we will provide an overview and discuss recent findings on mitochondrial ROS (mtROS) production, its threshold for imbalance, as well as consequent dysfunction causing related cardiac and systemic phenotypes in the context of HFpEF, focusing on newly discovered cellular and molecular mechanisms underlying ROS dysregulation, the current therapeutic options, and future perspectives to treat HFpEF by targeting mtROS and associated signal molecules.

2. HFpEF

2.1. Worldwide Prevalence, Morbidity, and Mortality

HF affects more than 64 million people worldwide [6]. In Europe, the prevalence of HF is estimated to be 1% for those aged <55 and >10% for those aged 70 years or over [1]. Epidemiological research with LVEF registries (2000–2016) indicates that HFpEF accounts for as much as 55% of all HF incidence [7]. The rising occurrence of HFpEF may be attributed in part to the aging population, heightened risk factors, and related comorbidities, as well as advancements in diagnostic accuracy [7]. Unfortunately, no drug has been identified to reduce mortality and morbidity in HFpEF so far. The available treatments seem to be more responsive in patients with HFmrEF and HFrEF [1]. Additionally, comorbidities including cardiovascular (CV) and non-CV diseases also significantly influence the treatment response and outcomes of HFpEF. Among patients with HF, there is a rise in non-CV comorbidities as the population ages. In HFpEF, the co-presence of CKD, COPD, depression, sleep disorders, obesity, diabetes mellitus, cancer, and iron deficiency all can hinder the diagnosis and have a great influence on clinical phenotype, response to drugs, hospitalization, prognosis, and health-related quality of life [8]. Alarmingly, the prevalence of HFpEF has risen from 38% to 55% among individuals hospitalized for HF during the past 35 years [9], and hospital admissions for HF are predicted to increase up to 50% in the next 25 years as a result of the aged population and the increased incidence of comorbidities [1,10]. Prognostic studies for individuals with HFpEF can differ based on factors such as the design of the study, the presence of other medical conditions, and the stage of the disease. Observational studies indicate that patients with decompensated HFpEF had mortality rates of approximately 20–30% within one year [9,11,12,13,14] and, after first hospitalization, the short-term mortality of HFpEF is comparable with HFrEF [9].
The primary cause of death for HFpEF patients is variable according to the nature of the study and the characteristics of the population. For instance, it has been reported that CV diseases (CVDs) are the main cause of death varying from 60% to 70% in clinical trials [15], and patients with older age groups show lower rates of CVDs as the cause of death, varying from 50 to 60% [15]. Interestingly, the incidence of non-CV fatalities is typically greater in patients with HFpEF compared to those with HFrEF [16]. Corroborating with this finding, a report from Olmsted County showed a rate of 49% of non-CV deaths in HFpEF patients [12]. Moreover, according to the Framingham Heart Study, the non-CV mortality rates may vary depending on sex, with 61% in men and 51% in women [17]. Sudden death accounts for approximately 25% of all-cause mortality in HFpEF [15]. A recent long-term follow-up study shows that after 2.1–7.9 years of an acute decompensated HFpEF, 68% of patients died, with 47% from CV and 45% from non-CV causes. Specifically, data showed that coronary artery disease (CAD) and tricuspid regurgitation were predictors of CV death, and stroke, kidney disease, lower BMI, and lower sodium were predictors of non-CV death, respectively. Anemia and higher age were related to both events [18]. Similarly, a recent observational and retrospective analysis showed that anemia and iron deficiency significantly increase mortality risk among HFpEF [19]. In this study, anemia, advanced age, iron deficiency, decreased LVEF, CKD, and paroxysmal nocturnal dyspnea were related to all-cause mortality in HFpEF [19].

2.2. The Pathophysiology of HFpEF

HFpEF does not have a cardiocentric physiopathology, with heterogeneous clinical presentation attributing to different comorbidities from several organ/tissue systems [20] (Figure 1). There has been recent recognition that comorbidities such as diabetes, obesity, and renal, pulmonary, or skeletal muscle diseases, often associated with inflammation and visceral fat, contribute to the development and outcomes of HFpEF and stimulate the creation of the term “inflammatory-metabolic HFpEF”, as recently reviewed [21]. Macrophage infiltration induces adipose tissue inflammation, upregulating proinflammatory adipokines (e.g., leptin, TNF-α, IL-6, and resistin) and downregulating anti-inflammatory adipokines (e.g., adiponectin, omentin-1), resulting in chronic and low-grade systemic inflammation [22,23,24]. The activation of these proinflammatory pathways seems to raise systemic inflammation and oxidative and nitrosative stress, which cause mitochondrial and endothelium dysfunctions, contributing to multisystemic abnormalities in HFpEF [21].
Diastolic dysfunction plays a key role in the pathophysiology of HFpEF, but there are also multiple cardiac, vascular, and extracardiac problems that contribute to this condition. The pathophysiology of HFpEF encompasses deficiencies in the relaxation and contraction abilities of the left ventricle (LV), abnormalities in the functioning of the left atrium (LA), high pressure in the lungs (pulmonary hypertension or PH), abnormalities in the exchange of gases, dysfunction in the right side of the heart, irregularities in the autonomic nervous system, stiffening of blood vessels, inadequate blood supply to the heart muscle (myocardial ischemia), impaired functioning of the endothelium, kidney disease, and issues in the peripheral tissues such as skeletal muscle and fat [4]. In the following sections, we will discuss the key pathophysiologic mechanisms of HFpEF, with a primary emphasis on the heart, as most of the available material pertains to this organ.
Diastolic dysfunction is produced by an elevation in passive myocardial stiffness due to changes in the viscoelastic properties of the heart muscle [25]. The stiffness of the LV alters the relationship between end-diastolic pressure and volume, which will be more prominent during physical activity [26,27]. The primary molecular cause of LV stiffness is the protein titin, which acts as a two-way spring in cardiomyocytes (CMs) and is regulated by phosphorylation through cyclic guanosine monophosphate (cGMP) kinases [28]. The phosphorylation process is affected by the presence of nitric oxide (NO), linking the health of blood vessels to the functioning of the heart [29]. Moreover, alterations in the extracellular matrix, specifically the augmentation of fibrillar collagen, play a role in the rigidity of the heart muscle and are linked to elevated myocardial fibrosis [30].
HFpEF patients may present systolic dysfunction, which commonly occurs due to issues associated with calcium control, beta-adrenergic signaling, myocardial energy utilization, or tissue blood flow [25]. Although these systolic dysfunctions may be moderate when the body is at rest, they can worsen when subjected to physical stress, resulting in a decrease in the amount of blood pumped by the heart and an increase in the inability to tolerate exercise [31,32]. Systolic dysfunction not only impairs cardiac efficiency but also worsens diastolic dysfunction by limiting the heart’s capacity to contract effectively and decreasing the ventricular elastic recoil required for adequate diastolic filling. The abovementioned sequence of cardiac malfunctions can result in higher death rates among patients with HFpEF and also contribute to other comorbidities such as PH and LA dysfunction due to decreased atrioventricular coupling, as nicely documented in the literature [33,34,35].
The LA initially counterbalances LV diastolic failure by functioning as a blood storage area, facilitating LV filling without raising LA pressures [20]. Over time, chronic or severe LV dysfunction results in the enlargement and impaired function of the LA, which is strongly associated with the progression of PH, right ventricular dysfunction (RVD), and AF [36,37]. AF is identified as a notable indicator of LA myopathy [37]. It worsens diastolic dysfunction by elevating the burden on the atria and interrupting the synchronization between the atria and ventricles [37]. As a result, AF contributes to mitral regurgitation and overall inefficiency of the heart. This series of events leads to an increase in diastolic ventricular interaction (DVI), where the enlargement of the LA and right side of the heart increases the overall volume of the heart [37,38]. This puts pressure on the pericardium, which disrupts the Frank–Starling mechanism that is crucial for maintaining the output of blood from the heart [38]. Echocardiographic techniques, such as speckle tracking, reveal that LA strain and compliance gradually decline due to persistent LV diastolic dysfunction [36,39,40]. This loss in cardiac function not only distinguishes HFpEF from other factors causing difficulty in breathing but also increases the likelihood of developing AF, resulting in worse clinical results such as higher mortality rates, reduced ability to engage in physical activity, and worsened RVD [36,37,41,42].
In addition, a significant percentage (between 50 and 80%) of patients with HFpEF develop PH [43]. This is mainly caused by increased pressure in the LA, leading to isolated postcapillary PH. Prolonged exposure to these increased pressures exacerbates the resistance and consequent remodeling of the blood vessels in the lungs, resulting in a combination of pre- and postcapillary PH [43]. This condition not only reduces the ability to exercise but also raises the likelihood of right HF, hospitalization, and death [44].
The right ventricle (RV), which is not well adapted to handle high afterload, experiences severe negative effects under these circumstances. This leads to systemic venous congestion and a variety of problems including edema, organ failure, AF, and cardiac cachexia [45,46,47]. Studies emphasize that RVD in HFpEF is substantially associated with poorer outcomes [41,47]. In addition, prolonged elevated LA pressure harms pulmonary capillaries, impairing lung diffusion and the exchange of carbon monoxide, which in turn increases mortality risk in HFpEF patients [48,49]. Moreover, HFpEF is characterized by hypertension in about 90% of patients, with a high prevalence of increased stiffness in the aorta and conduit vessels, especially in individuals with diabetes [50,51]. Blood pressure (BP) instability occurs as a result of the combined effects of ventricular and arterial stiffness [52]. Patients diagnosed with HFpEF also demonstrate impaired vasodilation that depends on the endothelium and is related to systemic inflammation and a decrease in the availability of NO [3,53,54]. Moreover, there is a high incidence of coronary microvascular dysfunction, which impacts both endothelium-dependent and independent mechanisms [55]. This dysfunction is linked to severe disease indicators such as AF, RVD, and limited exercise capacity [56,57,58].
While the cardiac output stays within a normal range during periods of rest, the potential to enhance it during physical effort is compromised, leading to a decrease in aerobic capacity [59]. This can be attributed in part to a reduction in stroke volume (SV) reserve caused by cardiac dysfunction and chronotropic incompetence [60]. The restriction in heart rate reserve appears to be associated with a decrease in adrenergic sensitivity rather than problems with central outflow [60,61].
Research on autonomic function in HFpEF yields conflicting findings; certain studies suggest impaired arterial baroreflex sensitivity [60], while others observe intact parasympathetic and sympathetic responses to stress [62]. In rodent models, a reduction in baroreflex sensitivity results in challenges in regulating variations in volume, which in turn leads to an elevation in left atrial pressure [63]. Sympathetic outflow aids in narrowing large-capacity veins, increasing the flow of blood back to the heart by moving it from less strained to more strained areas. This process, which is especially evident in HFpEF associated with obesity, plays a key role in the elevation of ventricular filling pressures during physical activity [32].
Individuals with HFpEF frequently demonstrate reduced oxygen consumption (VO2) as a result of irregularities in the differential in oxygen content between arteries and veins (AVO2diff). This indicates problems in the delivery and utilization of oxygen in the skeletal muscles [64,65]. Histological examinations reveal a decrease in the number of blood vessels and a change towards a higher proportion of Type II (fast-twitch) muscle fibers, with a lower number of Type I (aerobic) fibers [66]. Additionally, there is evidence of impaired mitochondrial function, which further hampers the utilization of oxygen in the peripheral tissues [67].

3. Animal Models of HFpEF

Multiple animal models have been reported to simulate human HFpEF in the past decades, and each comes with unique advantages and disadvantages (Table 1). The first animal models used to investigate HFpEF focused on hypertension, diastolic dysfunction, or left ventricular hypertrophy (LVH), such as pressure overload in Dahl salt-sensitive rats (DSSRs) [68,69] and spontaneously hypertensive rats (SHRs) [70], as well as models that involve renovascular hypertension, aortic constriction, or hormone-induced hypertension [71]. These models have been used in the investigation of the fundamental mechanisms of HFpEF, with a particular emphasis on hypertension and CV stress. However, there are two major limitations: First is that these animals respond to treatment with angiotensin II or another hormone blockade [72,73], which does not happen in the human syndrome. Second, these models show two different phases that normally do not occur in human disease [74]. Initially, there is a hypercompensated phase characterized by preserved LVEF, delayed diastolic relaxation, and impaired ventricular compliance [75]; after this initial phase, there is a more “dilated period” where both systolic and diastolic BP decrease [69,76]. Despite these limitations, models of pressure–volume overload are still good strategies for testing new treatments for HFpEF [77,78].
With the global spread of cardiometabolic HFpEF, animal models combining hemodynamic and metabolic stress have been developed [21,79]. One good example is the crossing of the spontaneously hypertensive HF rat with the Zucker diabetic rat, carrying a leptin receptor mutation (Lepr), generating the ZSF1 rat [79]. The generated ZSF1 animal model exhibits a combination of diastolic dysfunction, hypertension, LVH, cardiac fibrosis, obesity, hyperlipidemia, renal dysfunction, reduced NO signaling, and aortic stiffening—key characteristics of HFpEF [80]. Despite not developing fluid imbalances, the ZSF1 rat has become widely and successfully adopted in both academic and industry research as a model for studying HFpEF [81,82]. However, there are some limitations to using the ZSF1 rat model for HFpEF. The ZSF1 rat does not present increased LVDP or raised natriuretic peptide until a very later stage and responds to the inhibition of the angiotensin-converting enzyme [83]. Moreover, due to the mixed genetic background, there is a dilution of each phenotype behind it.
Another often-used animal model is the two-hit model, which combines the inhibition of nitric oxide synthase 1 and 3 (NOS1 and NOS3) by N-nitro-l-arginine methyl ester (l-NAME) with a high-fat diet in C57BL/6 mice [5]. The two-hit model is well defined, yielding robust and dependable findings from assessments of both inactivity and physical exertion. This model gives us important information regarding cardiac histology, exercise intolerance, and physiopathology of HFpEF, crucial for studying the human syndrome. However, it has some important limitations. First, female mice are less susceptible to developing the HFpEF phenotype in this model [84], which contrasts with the human situation where postmenopausal women are strongly impacted by HFpEF. Second, the blockage of NOS by l-NAME to induce hypertension does not represent the underlying mechanisms of hypertension in HFpEF patients. Lastly, the two-hit model does not show skeletal muscle alterations [5], therefore limiting its ability to fully represent human disease.
The use of larger mammals such as dogs, pigs, and non-human primates can increase therapeutic applicability and translation in the study of HFpEF. A dog model of secondary hypertension because of kidney disease leads to cardiac responses similar to HFpEF syndrome [85,86]. Similarly, pig models receiving a combination of deoxycorticosterone acetate (DOCA), a Western diet, and salt shows important features of HFpEF [87,88,89]. The pig model of HFpEF shows an enlargement of the LA with decreased contraction capacity, scarring of the heart muscle, increased oxidative and nitrosative stress, and reduced titin phosphorylation [87,88,89]. Nevertheless, these models still fail to completely replicate the intricacies of the human situation. In comparison to HFpEF patients, the pig models have lower diastolic pressure, increased relaxation time, and a smaller LV chamber with compensatory hypertrophy [87,88].
Indeed, the creation of a HFpEF animal model that effectively replicates the entire range of severity and comorbidities of HFpEF as seen in humans is a big challenge. The models that attempt to simulate the cardiometabolic phenotype (e.g., obesity + hemodynamic alterations) have the biggest potential. However, neither of them can properly simulate obesity or hypertension as observed in HFpEF patients. These patients usually show obesity grade 2 or higher (BMI ≥ 35 kg/m2, WHO classification), which is difficult to achieve with animal models. Regarding BP, patients often have treated BP (systolic BP < 130 mmHg) while, in the animal models, the hypertension normally remains untreated. In addition, fluid retention, skeletal muscle alternations, and metabolic dysfunction should be encouraged in animal models of HFpEF. Nonetheless, to evaluate prospective medicines and achieve translational relevance, easily implemented animal models should be used, including large mammals, even with the high cost and absence of the full reproducibility of the syndrome.

4. Mitochondrial Metabolism in Cardiac Hemostasis and Injury-Induced Remodeling

4.1. Mitochondria—The Powerhouse of the Heart

Human individuals produce and consume an amount of adenosine triphosphate (ATP) nearly equivalent to their body weight every single day [90]. The heart, in particular, has high energy demands, accounting for only 0.5% of body weight but consuming roughly 8% of ATP produced. Moreover, mitochondria occupy approximately one-third of the volume of adult CMs and produce the majority of the ATP consumed by the heart (~95%) [91,92,93]. Mitochondria produce abundant ATP by metabolizing a variety of fuels including fatty acids (FAs), glucose, lactate, ketones, pyruvate, and amino acids, primarily by mitochondrial oxidative phosphorylation (OXPHOS) (Figure 2). Any disruptions in the metabolic energy pathways that produce ATP can have catastrophic consequences on cardiac functions, resulting in a variety of CVDs including HFpEF [94]. Mitochondria are organelles with a double membrane structure, containing an inner and an outer membrane (Figure 2). The inner membrane of mitochondria is convoluted into many infoldings called cristae. Mitochondrial fragmentation and cristae destruction were clearly evident, with decreased mitochondrial area in HFpEF [95]. OXPHOS primarily consists of five enzymatic complexes, namely the electron transport chain (ETC), carried out by large multi-subunit protein complexes in the cristae membranes [96], which produce ATP by coupling the oxidation of reducing equivalents in mitochondria to the generation and subsequent dissipation of a proton gradient across the inner mitochondrial membrane [97].
The tricarboxylic acid (TCA) cycle, also known as the citric acid cycle or the Krebs cycle, is localized in the inner mitochondria matrix and is a closed loop that forms a metabolic engine within cells [98]. The TCA cycle itself does not consume molecular oxygen or produce meaningful amounts of ATP (Figure 2). However, it mainly transfers electrons from inputs (e.g., acetyl-CoA) to electron carriers (e.g., NAD+, FAD+), which then go into ETC and produce abundant ATP [99]. From an energy-generating perspective, the primary function of the TCA cycle is to oxidize acetyl-CoA to generate two molecules of CO2, two ATP, and some byproducts including three NADH and one FADH2. These byproducts further feed into the ETC at complex I (NADH dehydrogenase) and complex II (succinate dehydrogenase), respectively, to ultimately produce 30–32 ATP molecules through OXPHOS. Importantly, TCA cycles also produce and release abundant metabolites to control the biosynthesis of macromolecules such as nucleotides, lipids, and proteins, playing a critical role in cellular homeostasis, stress response, and disease pathogenesis [100,101].

4.2. Mitochondrial Metabolism and Metabolites in Cardiac Hemostasis

In the adult heart, ATP produced through OXPHOS in mitochondria accounts for 95% of myocardium ATP requirements, while 5% of ATP comes from anaerobic glycolysis [102] (Figure 2). In oxidative metabolism, the majority of ATP production (around 60%) is from the oxidation of mitochondrial FAs, followed by lesser contributions from glucose, lactate, and ketone bodies [103]. Changes in substrate availability and uptake can happen in the healthy heart to maintain ATP production [104]. For instance, under well-fed conditions, glucose can become a significant fuel for the heart via glycolysis and glucose-derived pyruvate oxidation. However, in fasting conditions, FAs can serve as the primary cardiac fuel.
FAs can enter CMs to produce fatty acyl-CoA, which is then transported into the mitochondria. Fatty acyl-CoA undergoes β-oxidation to produce acetyl-CoA, serving as an important substrate for the TCA cycle [105]. Glucose is also an important fuel for the heart because it can generate ATP both from anaerobic glycolysis and providing pyruvate to the TCA cycle. The majority of pyruvate derived from glycolysis is converted to acetyl-CoA by PDH (pyruvate dehydrogenase), and acetyl-CoA is then further metabolized in the TCA cycle. Additionally, lactate is increasingly being recognized as an important energy substrate of the heart. Lactate can be produced and released to the blood by skeletal muscle [106], then taken up by the heart where it converts into pyruvate via LDH (lactate dehydrogenase). The resultant pyruvate is then converted to acetyl-CoA, which enters into the TCA cycle in a similar way to the pyruvate generated from glycolysis. Recent studies also found that under some circumstances, lactate can provide carbons to the TCA cycles in a pyruvate-independent manner [107]. Moreover, it has been proved that ketone bodies are also important fuel in the heart. The ketone bodies, namely acetoacetate, acetone, and β-hydroxybutyrate (βOHB), are produced in the liver, with βOHB being the predominant ketone body taken up and oxidized in the heart [108,109]. After uptake, βOHB is transported to mitochondria and is oxidized to acetoacetate by BDH1 (β-hydroxybutyrate dehydrogenase 1), which is then converted to acetoacetyl-CoA (Acetoacetate-CoA) and undergoes a thiolysis reaction to produce acetyl-CoA. If circulating levels of ketone bodies are elevated, they can become a major fuel in the heart [110]. Finally, recent evidence shows that amino acids are also a potential source in ATP production in the heart. Branched chain amino acid (BCAA) oxidation, which occurs in the mitochondria, is the best characterized source of amino acid oxidation in the heart. BCAAs will be converted to branched chain α-keto acids (BCKAs) via reversible transamination, followed by irreversible decarboxylation by the branched chain α-keto acid dehydrogenase (BCKDH), eventually generating acetyl-CoA for the TCA cycle and succinyl-CoA for anaplerosis, respectively [111].
Importantly, TCA cycle metabolites appear to be an important mitochondrial messenger, capable of eliciting both genetic and epigenetic reprogramming, as exemplified by acetyl-CoA produced via pyruvate conversion, FA oxidation (FAO), and ketone body oxidation [112] (Figure 2). Acetyl-CoA is a building block for the biosynthesis of lipids, ketone bodies, amino acids, and cholesterol. It also provides acetyl groups for histone acetylation and modulates the activity of metabolic enzymes [113]. Another example is succinate. Under aerobic conditions, succinate is mainly produced from succinyl coenzyme A and converted to fumarate by the TCA cycle enzyme succinate dehydrogenase (SDH). Succinate can act as a messenger between intracellular metabolic status and cellular functions to regulate DNA methylation and lysine succinylation. Under low oxygen conditions, SDH activity is reduced, causing succinate accumulation and decreased fumarate [114,115]. Fumarate converted from succinate also has an important role in chromatin modifications and protein succination [116]. Therefore, TCA cycle metabolites not only function as a part of the TCA cycle and contribute to the production of ATP but also play important roles in biosynthesis of macromolecules, as well as genetic and epigenetic regulation.

4.3. Mitochondrial Metabolism and Metabolites in Injury-Induced Cardiac Remodeling, with a Focus on HFpEF

Once ATP synthesis is stopped, the stored ATP can only sustain the heartbeat for a few seconds [117]. Therefore, the rate of ATP synthesis must match the rate of ATP consumption in the heart, and undoubtedly the most crucial change happening in mitochondrial energy metabolism observed in the failing heart is energy or ATP deficiency [118] (Figure 3). The metabolic changes begin during compensated cardiac hypertrophy and are accentuated in overt HF [119]. As mentioned before, a healthy heart uses FAs as its primary energy fuel to produce ATP, while glucose becomes the main energy fuel in HF, as evidenced by increased glucose uptake and glycolysis but decreased FAO and OXPHOS in HFrEF [120,121]. Similar to HFrEF, HFpEF myocardium has reduced expression levels of proteins involved in FA uptake and oxidation, indicating impaired utilization of FA in the heart [122,123,124]. Unlike in HFrEF, glucose uptake and metabolism are inhibited in HFpEF myocardium [122,125]. Instead, HFpEF patients show increased amino acids in the heart and circulating blood, and the main upregulated amino acids in the HFpEF heart are BCAAs. However, their catabolites are lower compared to healthy myocardium, indicating impaired BCAA oxidation in HFpEF myocardium [122,126]. Ketone body oxidation is considered protective for HFrEF since it provides an important alternative fuel for ATP production in the heart. Compared to the control, no apparent changes in ketone body amount are found in the HFpEF heart [122]. However, it has been shown that the serum concentrations of the ketone bodies acetoacetate, 2-hydroxybutyrate, and 3 hydroxybutyrate are lower in HFrEF patients than both HFpEF patients and non-HF controls [126], implying a greater reliance on ketone bodies as an energy source in HFrEF patients. Interestingly, new evidence from recent clinical trials with sodium-glucose co-transporter-2 (SGLT2) inhibitors, which can increase circulating ketones, proves the beneficial effects of SGLT2 inhibitors on human HFpEF [127,128]; however, the underlying mechanisms remain unknown.
The cardiac expression levels of pyruvate and acetyl-CoA in HF are increased, indicating inadequate TCA cycle intermediate replacement. Lower levels of TCA cycle intermediates such as malate, fumarate, and succinate in HFpEF myocardium also suggest inadequate anaplerosis [122]. PDH converts pyruvate to acetyl-CoA in mitochondria, and its activity varies inversely with protein phosphorylation by PDK4. Both phospho-PDH/total PDH ratio and expression of PDK4 were significantly lower in HFpEF, which would favor enhanced pyruvate conversion to acetyl-CoA [129]. Except for contributing to producing ATP, acetyl-CoA and TCA cycle metabolites also have important roles in post-translational modifications, especially in acetylation. Protein acetylation is increased in failing human and mouse hearts, which is more exacerbated in HFpEF compared with HFrEF [123,130,131]. Hyperacetylation has been linked with decreased enzymatic activities of multiple mitochondrial proteins and impaired FAO [123,132]. Sirtuin enzyme Sirt3 knockout mice even developed more pronounced diastolic dysfunction, suggesting a protective effect of CM Sirt3 in HFpEF pathogenesis [123]. NAD+ repletion via oral supplementation of an NAD+ precursor nicotinamide riboside in mice reversed the hyperacetylation of FAO key enzymes and thus restored mitochondrial function and improved pathological cardiac remodeling in the HFpEF model [123]. This increased protein acetylation could be attributed to the enlargement of the acetyl-CoA pool in the mitochondrial matrix [133]. Interestingly, the administration of ketone body β-hydroxybutyrate can downregulate the acetyl-CoA pool and protein acetylation, partially via activation of citrate synthase and inhibition of FA uptake in HFpEF mice [133].
Genetic studies unravel the genetic links between mitochondrial metabolism and cardiac pathologies. SDH (succinate dehydrogenase) can link the TCA cycle with the ETC, therefore playing a unique role in mitochondrial metabolism. Dilated cardiomyopathy (DCM) can be caused by a G555E mutation in the SDHA gene, which significantly reduces tissue-specific SDH enzymatic activity in cardiac muscle [134]. Mice with cardiac deletion of SDHAF4 (succinate dehydrogenase assembly factor 4), an assembly factor that is responsible for proper assembly of the SDH subunits [135], display SDH activity suppression, metabolic deficiency, and DCM phenotypes. Importantly, supplementation with fumarate, which is depleted due to insufficient SDH activity, in SDHAF4 knockout mice can substantially improve mitochondrial dynamics and cardiac functions, as well as prolong the lifespan of DCM mice [136]. These findings show that the normal function of SDH is important for producing enough fumarate and maintaining normal cardiac physiological functions. However, the potential functions of SDH and fumarate need to be further studied in HFpEF.
A recent study showed that mitochondrial complex disorder may occur from the early stages of HFpEF, evidenced by reduced expression of key proteins required for complex I, ATP synthase, efficient electron transfer, and supercomplex formation [137]. Complex I-mediated mitochondrial respiration is impaired in permeabilized cardiac fibers of Zucker/spontaneously hypertensive F1 hybrid obese (ZSF1) rats that develop an HFpEF-like phenotype. The suppression of mitochondrial respiration could, in part, be caused by increased mitochondrial matrix Ca2+ levels, which would contribute to a lower proton motive force for ATP generation [138,139]. Mitochondrial complex I deficiency by Ndufs4 gene inactivation increases mitochondrial protein acetylation and accelerates HFrEF [140]. But the function of complex I in HFpEF has not been well studied. Limited studies using patients’ cardiac tissues showed that genes involved in OXPHOS are elevated in HFpEF but decreased in HFrEF. However, the higher level of OXPHOS expression in HFpEF was attenuated after adjustment for body mass index, suggesting that obesity may contribute to the elevated OXPHOS gene expression in HFpEF [141]. It is worth mentioning that studying cardiac tissues from patients with HFpEF is extremely limited, likely due to challenges in obtaining cardiac tissue from HFpEF patients, as opposed to patients with HFrEF.
PCSK9 (proprotein convertase subtilisin/kexin type 9) is known to be secreted into the circulation, mainly by the liver. It interacts with low-density lipoprotein receptor (LDLR) homologous and non-homologous receptors, thus favoring their intracellular degradation [142]. Therefore, PCSK9 inhibitors have been widely used to treat hypercholesterolemia and to reduce cardiovascular risk in patients with ischemic heart disease. Interestingly, it has been reported that PCSK9 genetic deficiency promotes HFpEF progression, independent of the modulation of LDLR and of circulating PCSK9 [143]. PCSK9-deficient mice display impaired OXPHOS and mitochondrial metabolism, evidenced by impaired FAO and decreased TCA cycle flux. Such impairments may be attributed to increased cholesterol accumulation in the heart [143]. This new evidence raises caution in terms of the safety issue with the increasing use of PCSK9 inhibitors in clinics.

5. Mitochondrial Quality Control (MQC): The Protective “Mito Orchestra” in the Context of HFpEF

As mentioned in the above section, mitochondria are vital organelles to maintain cardiac homeostasis. CMs are dependent on ATP production via OXPHOS to fulfill the energy requirements of the organ. Therefore, the integrity of mitochondrial function is achieved through the implementation of an orchestrated MQC mechanism [144] (Figure 4). Disruptions in the MQC mechanisms have detrimental effects on the function and contractility of CMs, potentially leading to cellular apoptosis and tissue injury [145,146]. Indeed, mitochondrial dysfunction has a significant role in the pathogenesis of several CVs, such as hereditary cardiomyopathies, ischemia/reperfusion injury, hypertension, atherosclerosis, diabetes, and HF [147,148].
There are two levels of MQC mechanisms: molecular and organelle-level control. These orchestrated mechanisms are essential for the maintenance of mitochondrial function [149]. Mitochondrial chaperones, including heat shock proteins (HSP60, HSP70, and HSP90), are essential for the appropriate folding of proteins at the molecular level. These chaperones aid in the proper folding of newly synthesized proteins and the refolding of damaged ones before they are imported into the mitochondria. The ubiquitin–proteasome system (UPS) will then target proteins that are misfolded or damaged for degradation, trying to prevent the accumulation of dysfunctional proteins [150,151]. Organelle-level MQC mechanisms are activated when molecular-level mechanisms are insufficient to repair mitochondrial damage. These mechanisms address more severe mitochondrial dysfunction, assuring the maintenance of cellular health by addressing damaged mitochondria that cannot be repaired at the molecular level [152].

5.1. Mitochondrial Biogenesis

Mitochondrial biogenesis, the process by which cells increase mitochondrial mass, is essential for the preservation of cardiac homeostasis, particularly in the face of cellular injury [153,154]. By increasing mtDNA copy number and metabolic enzyme expression, mitochondria boost their metabolic capability. Mitobiogenesis involves a coordinated interaction between mitochondrial and nuclear genomes and is affected by several endogenous or exogenous stimuli such as cold, exercise, caloric restriction, cell division, and oxidative stress [155].
The transcriptional coregulator PPARγ coactivator 1α (PGC1α) is the main regulator of the mitobiogenesis process [156]. In response to external stimuli such as cold exposure, exercise, and fasting, PGC1α activates a variety of transcription factors such as Nuclear Respiratory Factors 1 and 2 (NRF1, 2) [157,158]. NRF1 or NRF2 induce mitochondrial transcription factor A (TFAM), which translocates to mitochondria where it transcribes mtDNA [159]. Moreover, mitochondria-resident microRNAs (miRs) assist in the regulation of the mitochondrial pool by targeting TFAM and other biogenesis factors, thereby increasing the mtDNA copy number and inducing the production of proteins that are essential for mitochondrial function [160,161,162].
The therapeutic potential of regulating the PGC1α signaling pathway is applicable to a variety of diseases, such as diabetic cardiomyopathy, obesity, and HF [163,164]. Indeed, PGC1α/β-deficient mice exhibit important mitochondrial structural dysregulation, including elongation and breakage, which results in fatal cardiomyopathy, small hearts, and reduced cardiac output [165,166]. In the context of HF, microarray analysis showed that cardiomyopathy-derived HF patients have reduced cardiac PGC1α expression compared to healthy individuals [167]. Similarly, in ischemic and non-ischemic animal models of cardiomyopathy (ICM and non-ICM), the PGC1α expression seems to be downregulated. The reduced PGC1α expression impacted mitochondrial content and mtDNA replication, particularly in the late phases of HF [168,169].
Nevertheless, the role of PGC-1α in HF needs to be fully addressed. Pereira et al. [170] showed in HF models that PGC1α overexpression did not improve myocardial contractility or mitochondrial function. Additionally, in transgenic mice with cardiac-specific overexpression of PGC-1α, excessive mitochondrial proliferation was observed, which was accompanied by structural disarray in the sarcoplasmic reticulum; this resulted in diminished myocardial contractility, cardiac enlargement, and non-ICM HF [171]. These studies agree with human studies that did not find changes in cardiac PGC1α expression in HF patients [172]. Specifically, the reduced PGC1α levels and complex IV activity were only observed in HFrEF [96]. This implies that the impact of PGC1α on HF may be contingent upon its level of expression, its interactions with other mitochondrial biogenesis effectors, and the broader intracellular environment.

5.2. Mitochondrial Dynamics

Mitochondria are very active organelles; through mitochondrial dynamics, they form a dynamic network driven by mitochondrial cytoskeleton movements, mitochondrial fusion and fission, and mitophagy [173,174,175]. Proteins located in the outer mitochondrial membrane (OMM), such as dynamin-related GTPases, also known as mitofusins (MFN1 and MFN2), and proteins located in the interior mitochondrial membrane (IMM), such as optic atrophy protein 1 (OPA1) and cardiolipin, are involved in mitochondrial fusion by which two mitochondria are fused sharing mtDNA, metabolites, and enzymes [176,177,178]. The fusion of the OMM initiates by homo- and hetero-dimerization of MFN1 and MFN2 while OPA1 drives the fusion of the IMM maintaining the cristae structure and normal membrane permeability, essential protection against oxidative damage [179].
The process of fission involves the separation of healthy mitochondria from a damaged mitochondrion [180,181]. The most important classes of proteins involved in the fission processes are dynamin-related protein 1 (DRP1), located in mitochondrial-associated membranes (MAMs), where mitochondrial fission primarily occurs. During the process, this GTPase is recruited to the OMM, where it interacts with receptors such as mitochondrial fission 1 protein (FIS1) and mitochondrial fission factor (MFF) to induce mitochondrial cleavage. Apart from acting as a receptor for DRP1, FIS1 can also inhibit fusion, then promote fission [182,183]. Furthermore, the endoplasmic reticulum (ER) can encircle mitochondria, which aids in the constriction of the membrane during fission [184]. Importantly, mitophagy is linked to fission, as it guarantees the elimination of dysfunctional mitochondria, thereby preserving the health of the mitochondrial population and cellular homeostasis [185,186].
Imbalances in mitochondrial fission and fusion seem to be related to cardiac hypertrophy and the development of HF. It has been shown that cardiac deletion of MFN1/MFN2 leads to an eccentric ventricular remodeling while DRP1 absence causes dilated cardiomyopathy in cardiac genetic-modified mice [187]. Moreover, MFN2 downregulation was observed in a rodent model of cardiac hypertrophy [188], and MFN2 upregulation could constrain myocardial hypertrophy [189,190]. A study also suggests a role for OPA1 in HF progression. In a HF mouse model of pressure overload with HFD feeding, OPA1 was found to facilitate energy metabolism and increase cardiac FA utilization, therefore reducing ROS production, mitochondrial fragmentation and cardiac dysfunction [191]. Interestingly, it has been recently proposed that the potential consequences of an imbalance between mitochondria fusion and fission may be more severe than the simultaneous cessation of both processes. This idea is supported by Song et al. (2017), which has demonstrated that MFN1/MFN2/DRP1 triple knockout mice exhibit a unique form of cardiac hypertrophy and longer survival time when compared to MFN1/MFN2 or DRP1 cardiac knockout mice [192].
Indeed, the dynamics of mitochondria, particularly the equilibrium between fission and fusion, seem to be dysregulated in HFpEF. While the number of studies on this subject is low, new evidence indicates a transition towards higher levels of mitochondrial fission in HFpEF. Specifically, increased DRP1 levels were found in LV tissues of HFpEF patients compared to HFrEF or healthy individuals, with limited changes in the fusion proteins OPA1 and MFN2 [96]. Moreover, abnormal mitochondria and reduced MFN2 levels were observed in the skeletal muscle of elderly HFpEF patients, which was related to a decreased peak VO2 and 6 min walk distance when compared to HFrEF or control individuals [193]. In line with this finding, Bode et al. (2020) reported increased mitochondrial fission in LA CMs derived from ZFS1 obese rats [194], and a multi-omics analysis revealed that the inflammatory response and mitochondrial fission were the main mechanisms responsible for myocyte stiffening in the DSSR model of HFpEF [195]. However, it remains to be elucidated whether the suggested fission abnormalities reflect the irregular muscle oxygen consumption, or if is a consequence of dysfunctional mitochondria triggered by mechanisms underlying HFpEF development, such as the increased mitochondrial oxidative stress.

5.3. Mitophagy

Mitophagy, the specific degradation of mitochondria by autophagy, is a fundamental process within the mechanisms of mitochondrial quality control. The turnover of dysfunctional mitochondria prevents the accumulation of impaired mitochondria, which is critical for cellular differentiation and for tissues with high energy demands, such as the cardiovascular system [148,196]. Several studies have already suggested that mitophagy inefficiency is linked to the progression of CVDs, like HF [197,198].
The mitophagy mechanisms and pathways have extensively been reviewed [199,200,201,202]. Briefly, the elimination of damaged mitochondria through mitophagy involves loading the damaged or dysfunctional mitochondria into autophagosomes via autophagy receptors like light chain-3 (LC3). The autophagosomes then fuse with lysosomes and the mitochondria are degraded by the activity of lysosomal enzymes [203]. Mitophagy can basically occur through ubiquitin-dependent or independent pathways. Ubiquitin-dependent pathways are normally triggered by the loss of mitochondrial membrane potential (MMP) and involve the PINK1/Parkin complex; ubiquitin-independent pathways, also called receptor-dependent pathways, involve BNIP3, FUNDC1, and lipid-dependent mechanisms [199,200].
Mitophagy is a central mechanism for mitochondrial quality and quantity control, through which dysfunctional mitochondria are degraded via autophagy to maintain homeostasis or, more specifically, mitohormesis [204,205,206,207]. It is well accepted that mitochondrial dysfunction impacts the progression and pathogenesis of HF; therefore, mitophagy is suggested to be cardioprotective [208]. In this sense, failing human hearts have reduced expression levels of autophagy-related proteins, such as Beclin1, LC3II, and PINK1 [209,210]. Mice with transverse aortic constriction (TAC)-induced HF show reduced PINK1 expression and Parkin-related mitophagy, as well as increased ROS production and CM apoptosis [211]. In addition, PINK1 ablation in mice compromises the mitochondrial translocation of Parkin and impairs mitochondrial homeostasis, thus facilitating CM apoptosis, myocardial fibrosis, and reduced capillary density [210]. These features reflect a HF phenotype, which is further evidenced by the onset of LVD and cardiac hypertrophy in young mice [210]. Nonetheless, these studies highlight that selective mild activation of PINK1-PRKN mitophagy in CMs could potentially prevent HF development and reduce cardiac injury.
Although recent studies suggest the role of autophagy in the development of HFpEF, the functional implications of mitophagy in HFpEF remain to be fully addressed. RNAseq analysis performed in HFpEF patients highlighted a potential link between endoplasmic reticulum stress, angiogenesis, and autophagy genes with the development of HFpEF [146]. In animal models, transcriptome analysis revealed that inflammation, extracellular matrix, endothelial function, sarcomere/cytoskeleton, apoptosis/autophagy, and those related to cell cycle and mitotic cell cycle processes were the most important upregulated pathways involved in HFpEF [212,213]. Moreover, it has been evidenced that mitophagy and autophagy are impaired in aged animal models [214,215]. Collaborating with these findings, spermidine, a natural polyamine, was found to enhance cardiac autophagy, mitophagy, and mitochondrial respiration; reduce inflammation; and prevent cardiac hypertrophy as well as diastolic dysfunction in HFpEF induced by DSSRs and HFD [216].
Taking all into account, it is increasingly clear that mitochondrial dynamics are crucial to health and the development of cardiac disease, with a belief that mitochondrial fusion is advantageous and fission is detrimental; however, this binary perspective may be overly simplistic. Fission is essential for the separation of damaged mitochondria for removal by mitophagy, and the exacerbation of myocardial injury may result from the excessive inhibition of this process. Therefore, it is imperative to achieve the appropriate equilibrium in the regulation of mitochondrial dynamics to facilitate potential clinical applications.

6. Mitochondrial Dysfunction and ROS Generation: Implications in HFpEF

6.1. Mitochondrial ROS Generation and Damage

Mitochondria complexes I and III are believed to be a major source of ROS. The superoxide radical (O2•−) is formed when a small percentage of electrons in the ETC prematurely reduce oxygen (O2) [217,218]. Under typical physiological conditions, ROS production comprises approximately 2% of the total oxygen consumed by mitochondria. Typically, an efficient antioxidant system situated within the mitochondrial matrix neutralizes the “excessive” mtROS [175,217] (Figure 5).
The first step in the neutralization process is conducted by a manganese-dependent superoxide dismutase (the mitochondrial isoform SOD2), which converts O2 to O2 and hydrogen peroxide (H2O2). The H2O2 is then reduced to water by glutathione peroxidase (GPX) and the mitochondrial isoforms of peroxiredoxin (PRX3 and PRX5) [219]. The regeneration of GPX and PRX is dependent on NADPH, whose reduction depends on TCA cycle activity. Therefore, to sustain the antioxidant systems and maintain ATP production, the TCA cycle integrity is needed. In this sense, mitochondrial dysfunction can disrupt the TCA cycle enzyme activity and OXPHOS, impairing ATP production and the redox balance, as characterized by variations in the NAD+/NADH, FAD/FADH2, and NADP+/NADPH ratios. The altered redox state then impacts the cell scavenger’s ability, leading to an increase in ROS release [147,220,221].
In mitochondria, there are numerous non-ETC sources of ROS. For instance, several mitochondrial flavoenzymes, such as α-ketoglutarate dehydrogenase, pyruvate dehydrogenase, glycerol phosphate, electron transfer flavoprotein (ETF), and ETF-oxidoreductase, are now recognized as significant ROS sources. In certain situations, they may produce a significantly greater quantity of O2 than ETC [221,222,223]. Moreover, during the FAO in the mitochondrial matrix, a significant amount of H2O2 is generated by flavoproteins situated upstream of complex III, such as short-, medium-, long-, and very long-chain acyl-CoA dehydrogenases (SCAD, MCAD, LCAD, and VLCAD, all belonging to essential enzymes for FAO) [224,225]. Furthermore, monoamine oxidase (MAO) in the OMM deactivates neurotransmitters such as serotonin and produces H2O2 as a secondary product [223,226]. Apart from mitochondria, there are other important cellular sources of ROS. The enzymes xanthine oxidase (XO), NADPH oxidase (Noxs), and uncoupled NOS also play a role in ROS production inside the cell [221,227].
It is widely known that ROS generation beyond the antioxidant system’s buffering capability contributes to oxidative stress, which can cause alterations in DNA, proteins, and lipids [223,228]. High concentrations of ROS cause non-specific damage to intracellular macromolecules, which typically generate more reactive intermediates and initiate a chain of damage multiplication [229]. Dysregulated ROS generation by FA results in both oxidative injury to the mitochondria and mitochondrial uncoupling, which in turn leads to impaired ATP production [230,231], suggesting a pathophysiological association between oxidative stress and mitochondrial dysfunction.
Mitochondria have their own circular, double-stranded DNA (mtDNA), which is 16,569 base pairs in length and encodes thirteen subunits of mammalian respiratory complexes, 22 tRNAs, and two rRNAs [232]. Compared to nuclear DNA, mtDNA is particularly vulnerable to oxidative stress. Numerous studies conducted on humans and animals have indicated that mtDNA may contain a substantially greater number of oxidized bases than nuclear DNA [232,233]. This is because mtDNA is in closer proximity to respiratory chains that produce ROS and lacks the protection of histone-like proteins observed with nuclear DNA. ROS (H2O2 and O2) generated in the mitochondrial inner membrane can easily react with proximal mtDNA, resulting in the accumulation of point mutations and/or deletions and a reduced DNA copy number. This process impairs mitochondrial functions [223,233,234].
Induced oxidative damage to mtDNA results in reduced amounts of mitochondrial transcripts and proteins, as well as impaired ETC performance and ATP generation. This, in turn, stimulates the creation of ROS in mitochondria and leads to mutations in mtDNA, so creating a self-perpetuating cycle of damage amplification [235,236]. Multiple investigations have shown that increased oxidative damage to mtDNA plays a role in the progression of cardiomyopathy and HF [237,238,239]. Heteroplasmy of mitochondria occurs when wild-type and ROS-mutated mtDNA coexist in the same mitochondrial structure. Once the proportion of aberrant mtDNA exceeds specific thresholds (60–70%), it disrupts mitochondrial homeostatic processes and reduces bioenergetic capability, mitophagy, and mitochondrial fusion and fission events. Cardiomyopathy develops when defective mitochondria are accumulated in CMs due to altered metabolic processes and communication pathways, reduced ATP generation, as well as cell shrinkage or death [240]. Since mitochondria can function as intracellular signaling centers that regulate nuclear gene expression at both the transcriptional and epigenetic levels, the rise in mitochondrial DNA heteroplasmy undeniably not only leads to malfunction in mitochondria but also triggers extensive changes in gene expression [241,242,243]. Thus far, more than 400 mitochondrial DNA abnormalities have been identified to be linked with human disorders such as HF [244].
Apart from mtDNA, lipids and proteins are also influenced by ROS. Lipid peroxidation may be elevated by the accumulation of lipid-derived metabolites and FA in the internal mitochondrial membrane. The development of metabolic cardiomyopathy is influenced by lipid peroxidation products, including malondialdehyde (MDA), 4-hydroxynonenal (4-HNE), and 4-hydroxyhexenal (4-HNE) [245]. There is evidence that lipid peroxidation products can damage mitochondrial membranes, suppress cyclooxygenase (COX), and activate uncoupling protein (UCP)-2, resulting in ETC dysfunction and an exacerbation of ROS production [245,246].
Cardiolipin, one of the primary phospholipids of the mitochondrial inner membranes, is also susceptible to modification by oxidants or ROS [247]. Oxidative modification of cardiolipin has significant effects on mitochondrial functions, such as opening mitochondrial permeability transition pores (mPTPs), reducing ETC activity, and impairing mitochondrial membrane fluidity [247]. Additionally, cardiac pathologies are associated with modifications to the mitochondrial proteome and associated mitochondrial dysfunctions [248]. It has been documented that the mitochondrial proteome is altered in response to cardiac oxidative stress [249,250]. In diabetic mice, TAC causes substantial cardiac mitochondrial proteome remodeling, which can be substantially reduced by the scavenging of mitochondrial ROS by mitochondrial catalase [250].

6.2. Mitochondrial ROS as Coordinators of Cellular Function and Mitohormesis

Historically, mtROS have been believed to be detrimental byproducts of cellular respiration, with a focus on aging and mitochondrial malfunction. However, recent studies have shown that mtROS are essential modulators of cellular signaling, adaptability, and stress responses rather than being hazardous byproducts [251]. Mitohormesis is the theory that low-to-moderate amounts of mtROS might activate protective pathways, which in turn improve cellular resilience and eventually sustain survival under hardship, potentially prolonging human life [204] (Figure 6). Mild stress caused by hypoxia, metabolic changes, or nutritional deprivation may induce the nucleus and cytoplasm to undergo metabolic and biochemical adaptations, which can assist cells in coping with more severe stressors in the future, rather than causing damage [204,217].
Studies in C. elegans have provided one of the most conspicuous examples of mitohormesis. Glucose deprivation raised mitochondrial respiration and ROS production, and the increased ROS triggered antioxidant defenses, promoting cellular adaptations, stress resistance, and increasing lifespan [252]. Pretreatment with antioxidants like N-acetylcysteine (NAC) prevented the hormetic response, highlighting ROS as active players in the adaptive processes that improve cellular function and longevity [252,253].
In line with this observation, mitohormesis has been associated with an extended lifespan in numerous organisms (Figure 6). For instance, the lifespan of Drosophila can be regulated by a redox-dependent mitohormetic response [254]. Similarly, in yeast, impaired TOR signaling increases mtROS production and extends the chronological lifespan [255]. Moreover, lifespan extension has been also observed in certain long-lived mutants, exemplified by C. elegans with impaired insulin/IGF-1 signaling, contingent upon mtROS [256]. In another instance, the activation of the hypoxia-inducible factor HIF-1 through the inhibition of mitochondrial respiration was triggered by oxidants [257]. A similar hormetic response, which involves the mtROS-dependent activation of HIF-1, was demonstrated to be crucial in the rescue of the AMPK-null mutant of C. elegans [258]. Therefore, the increased ROS, rather than being detrimental, was essential for the extension of lifespan, and mtROS seem to function as signaling molecules that induce adaptive responses, such as metabolic reprogramming and improved stress resistance.
Similarly, mtROS can function as second messengers that initiate cardioprotective signaling pathways under some specific circumstances. For example, the heart may be safeguarded from IR injury by mild oxidative stress that is induced by intermittent or moderate ROS production [259]. In CMs, modest concentrations of the mitochondrial redox cycler MitoParaquat (MitoPQ) were discovered to elevate mtROS levels without affecting mitochondrial function or calcium handling. In animal models, this moderate increase in ROS resulted in an increase in calcium accumulation in the sarcoplasmic reticulum, which in turn protected cells from anoxia/reoxygenation injury and reduced infarct size. However, excessive ROS induces loss of MMP, the opening of the mPTP, and altered calcium homeostasis [260]. This paradox exemplifies that at moderate levels, ROS can offer cellular defense; however, after reaching a threshold, increased ROS can impair mitochondrial and CM functions.
In addition to their function as protective messengers in CMs, mtROS have become essential regulators of cellular signaling [261]. ROS are essential messengers between mitochondria and other cellular organelles, particularly the nucleus. For instance, mtROS are involved in the regulation of macroautophagy [262], the modulation of hypoxia responses through HIF-1 [263], and the activation of cytosolic stress kinases [264]. MtROS can stimulate signaling pathways that promote cell survival and response to external stresses when the oxidative stress is moderate. Particularly, mtROS can activate the Nrf2 pathway, which enhances antioxidant defenses [265,266]. Experimental studies on C. elegans have shown that the initiation of defensive reactions, such as the AMPK pathway, relies on signaling that is dependent on mtROS [252,258].
Mitohormesis also underscores the reasons why antioxidant therapies, which are designed to decrease ROS levels, have been mainly unsuccessful in clinical trials [267,268,269]. Antioxidants may prevent the activation of protective hormetic pathways, impairing the stress management capacity. It is suggested that the increased mtROS production during physical exercise stimulates adaptative responses by the expression of antioxidant enzymes in the skeletal muscle [270]. Nevertheless, the advantageous effects of exercise are inhibited when individuals are administered antioxidant supplements (e.g., vitamin C and vitamin E) [270,271], suggesting that mtROS are required for the protective adaptative processes in response to exercise and other mild stressors.
Traditionally, mtROS have been considered hazardous agents in aging and disease. Additionally, the 1950 free radical theory of aging connected mtROS to cellular malfunction and to the development of several diseases such as cancer, diabetes, and dementia/neurodegeneration [272,273,274]. However, subsequent studies have contradicted this viewpoint, showing that mtROS are not only secondary consequences of mitochondrial malfunction but also essential regulators of cellular activity [222,251]. MtROS are now recognized as fundamental molecules for signaling biological pathways such as transcription, epigenetic, and stress responses [222]. Moreover, mtROS act as key mediators of cellular resilience and survival [275,276]. Recently, van Soest et al. have reported that mtROS released from mitochondria are unlikely to directly damage nuclear genomic DNA, even at extremely high levels [277]. These surprising results raise questions about the contribution of mtROS to chromosomal DNA mutations, and consequently to oncogenic transformation and aging.
Taking all into account, mtROS are more than just harmful respiratory byproducts. They act as regulatory signal molecules during metabolism, cell survival, and stress resistance. Mild-to-moderate ROS levels activate cellular defensive pathways, boosting cell resilience and longevity (Figure 6). Understanding the dual roles of mtROS in cellular signaling and damage is essential for developing targeted therapies that can mitigate disease progression while preserving necessary cellular functions.

6.3. Mitochondrial ROS and Heart Failure

As discussed in the previous section, mtROS are critical regulators of numerous intracellular responses through redox signaling pathways. Nevertheless, the function of mtROS in physiology or pathology is contingent upon their concentration, type, and the location of their formation [222]. The primary function of low levels of mtROS is to participate in physiological processes, including excitation–contraction coupling (ECC), cell proliferation, and differentiation. However, high levels of mtROS impair essential molecule structures and functions in CMs via oxidative stress-regulated signaling, which subsequently increases cardiac metabolism and malfunctions of mitochondria, ion channels, and transporters, as well as inflammation and CM death. Ultimately, these modifications lead to myocardial remodeling (hypertrophy, fibrosis, and apoptosis) and eventually HF [277,278] (Figure 6).
Substantial mitochondrial remodeling, encompassing energy production, Ca2+ regulation, and the generation of ROS or reactive nitrogen species (RNS), has been well characterized during HF progression [279,280,281,282]. The subpopulations of mitochondria were first observed in the cell periphery, followed by interfibrillar mitochondria in HF [283]. Moreover, studies showed increased ROS generation during ischemic cardiac reperfusion [284,285], confirming that ROS can also enhance the generation of mtROS in pathological circumstances, leading to tissue death and inflammation.
CM apoptosis has a substantial impact on hypertrophic remodeling and cardiac dysfunction [286]. Multiple oxidative stress downstream signaling pathways have been implicated in CM apoptosis [287]. Previous investigations have demonstrated that the JNK signaling pathway is activated when human cardiac progenitor cells (CPCs) are exposed to H2O2, resulting in the inhibition of cell death [288]. The generation of ROS enhanced palmitate-induced apoptosis in CPCs, while the production of ROS was reduced by pretreatment with the NAC in human CPCs [289]. The activation of proapoptotic signaling pathways such as JNK, p38, and ASK1; the suppression of antiapoptotic signaling pathways like the PI3K/AKT pathway and the ERK1/2 pathway; and the immediate effects of ROS on mitochondria that trigger the release of cytochrome c may contribute to ROS-induced mitochondria-dependent apoptosis in CMs [248].
Substantial evidence of elevated ROS production in HF, which contributes to disease progression, has been obtained through extensive research in both human and animal models [290]. ROS imbalance was found in human ventricles explanted from patients with dilated cardiomyopathy. Using electro–paramagnetic resonance (EPR), the authors reported augmented mtROS production and increased myocardial oxidative stress. Further, enzyme activity assays revealed decreased activity of MnSOD in the mitochondrial matrix [291]. Moreover, the significance of mtROS in HF is reinforced by the phenotype of MnSOD knockout mice. Specifically, mice lacking MnSOD exhibit fatal dilated cardiomyopathy and HF, severe oxidative damage to the mitochondria and DNA, as well as a reduction in energy production. Additionally, mutant mice are exceedingly susceptible to hyperoxia, and the addition of exogenous ROS scavengers can suppress mitochondrial defects and attenuate cardiac remodeling [292,293]. Furthermore, new studies confirmed the causal involvement of mtROS imbalance in the pathogenesis of HF and offered mechanistic insight into the potential beneficial effects of antioxidants on HF [292,293]. Additional supportive evidence was recently provided using a guinea pig model of nonischemic HF [294]. In this study, the authors used a mitochondria-targeted ROS sensor to demonstrate an increase of mtROS production in quiescent and contracting LV myocytes from failing hearts, and they proposed that mtROS imbalance disrupts the optimal redox signaling that promotes the expression of genes involved in mitochondrial functions and ion handling [294].
Direct oxidation of histone deacetylase 4 (HDAC4) by elevated ROS can also activate hypertrophic signaling [295], which is linked to the nuclear export and de-suppression of transcription factors involved in cardiac hypertrophy, such as the myocyte enhancer factor 2 (MEF2) and calcineurin-NFAT [295,296]. The cardiac hypertrophic program can also be activated by chronic exposure to various agonists, such as AngII and endothelin-1, through the activation of NADPH-mediated ROS production [297].
ETC has recently garnered attention as a modulator of lipid peroxidation and GSH levels, which are responsible for ferroptosis (iron-dependent cell death) [298,299]. Contrary to complex III, complex I generated mtROS in isolated cardiac mitochondria, which promoted mitochondrial lipid peroxidation, as shown by the presence of lipid peroxy radicals [298]. The inhibition of both complex I and complex III greatly intensified the ferroptosis of CMs triggered by RSL3 (ferroptosis inducer) [300]. At the same time, the synthesis of GSH can safeguard cells from oxidative stress-induced cell death, including ferroptosis [301]. Despite the fact that mitochondria are unable to synthesize GSH, they roughly transport 10–15% of the total cellular GSH into the matrix via dicarboxylate carrier (DIC) and oxoglutarate carrier (OGC) [300]. Inhibiting DIC and OGC caused increased mtROS, membrane depolarization, and GSH depletion, as well as aggravated ferroptosis in CMs. Additionally, the ferroptotic PEox species were accumulated in mitochondria during cardiac I/R injury [300]. These data imply a causal role for mtROS in cardiac ferroptosis and dysfunctions.
It is worth noting that during the progression of HF, mitochondrial dysfunction can be both the cause and consequence of increased mtROS [302,303]. Enhanced mtROS is correlated with increased ATP generation in mitochondria, which is a result of increased ETC activity. Recent findings suggest that the acetylation of enzymes in failing hearts suppresses ATP synthase activity, in contrast to the notion that elevated energy demand in persistently stressed hearts can stimulate ATP production. When ATP generation is obstructed, it impairs ETC and increases mtROS emissions [304]. Further mtROS generation and more severe mitochondrial damage are the result of the initial increase in oxidative stress caused by mitochondrial injury [305,306]. MtROS are also implicated in a diverse array of cellular functions during HF. The opening of mPTP and the subsequent cell death are facilitated by increased mtROS. Both mtDNA damage and impaired mitochondrial biogenesis can be attributed to severe oxidative stress. The enzymatic activities inhibited by the oxidative alteration of proteins and lipids ultimately limit the capacity of mitochondrial respiration [307,308,309]. Moreover, inflammation is first triggered by the release of mtDNA. It has been recently demonstrated that mtROS function as essential mediators of inflammation, hypoxic signaling, the MAP kinase pathway, and retrograde communication between the nucleus and mitochondria in the context of HF [310].
Collectively, in HF, ROS imbalance can influence ETC coupling and mitochondrial bioenergetics, resulting in a vicious cycle that encompasses the oxidative modification of redox-sensitive kinases, cytosolic sodium accumulation, and further mtROS emission. This, in turn, can induce CM apoptosis, cardiac ferroptosis, arrhythmias, and cardiac remodeling through the post-translational oxidative modification of regulatory proteins, lipids, mtDNA, and nuclear DNA [311,312].

6.4. Mitochondrial ROS and HFpEF: A Fraction of the Whole?

Although studies on HFpEF and oxidative stress are scarcer, it has been suggested that oxidative and nitrosative stress could drive the onset and/or progression of HFpEF [5,84,313,314]. Chronic HF patients (CHF-NYHA functional class III) have increased levels of mtROS in circulating peripheral blood mononuclear cells (PBMCs) and worse mitochondrial functions when compared to mild HF patients (class I–II). MtROS generation in PBMCs is favorably linked with urine 8-hydroxydeoxyguanosine, a systemic oxidative stress measure, in CHF patients. Importantly, mtROS generation in PBMCs is highly associated with plasma levels of B-type natriuretic peptide, a diagnostic for HF severity, and inversely correlated with peak oxygen uptake, a measure of exercise capability, in HFpEF patients [315], suggesting a pathological role for mtROS in HFpEF (Figure 7).
NO and NOS are oxidized in the presence of ROS, resulting in a decrease in the bioavailability of NO [316], which controls mitochondrial pore opening [317]. Vasodilator NO is strictly regulated by hemoglobin and myoglobin, as well as O2 gradients in the myocardium to ensure that blood flow is appropriately adjusted to the metabolic requirements of the tissue. Specifically, NO production is reduced when O2 concentrations are high, and it is increased when O2 concentrations are low [317]. Under hypoxic conditions, NO reversibly binds to the O2 binding site on cytochrome c oxidase, thereby modifying the mitochondrial O2 consumption in tissues. The dysregulation of this mechanism in HF is linked to the progression of compensated to decompensated HF and impaired myocardial O2 consumption (MVO2) [318].
One of the most common comorbidities in HFpEF patients is diabetes, presenting in 40–45% of HFpEF patients [319,320]. The concurrent accumulation of glycative post-translational modifications has frequently been associated with the detrimental effects of oxidative stress on the development and progression of numerous CVDs, such as diabetic cardiovascular complications, atherosclerosis, and ultimately HF [321]. Glycation is the process by which a sugar non-enzymatically reacts with an amine group of a protein to produce advanced glycation end products (AGEs). These irreversible compounds compromise the structure and function of intra- and extracellular proteins, as well as binding their signal-transducing receptor (receptor for advanced glycation end products, RAGE) and causing a series of deleterious downstream effects [322]. It has been demonstrated that ROS increases the generation of AGEs, which in turn further enhances ROS production, thereby creating a detrimental positive feedback cycle of malignant AGE and ROS accumulation known as glycoxidative stress [321,323].
The most recent hypotheses on the cause of HFpEF are rooted in the increase of systemic inflammation and metabolic disruption, both of which are induced by non-cardiac comorbidities associated with HFpEF [324,325]. The change in the favored oxidative substrate, the rise in ROS levels, and defective mitochondria result in energy depletion due to an imbalance between ATP synthesis and demand. Several downstream signaling pathways are simultaneously activated, causing disruption to the normal heart architecture and initiating systemic inflammation that can lead to endothelial dysfunction of the coronary arteries and other unwanted events in HFpEF. Collectively, these processes stimulate the heart to undergo remodeling and hypertrophy in HFpEF, whereas, in HFrEF, the predominant cause is the loss of CMs mostly caused by ischemia, resulting in a weak, thin, or floppy heart muscle [67,326]. Previous studies have proposed that aberrant mitochondrial energy activity and oxidative stress occur before structural myocardial remodeling [327], implying a causal link between mtROS and cardiac dysfunctions.
Moreover, studies reported that the NLRP3 inflammasome can be assembled with mtDNA and N-formyl peptides, which can elicit inflammatory responses in mitochondria [328,329]. The activation of the NLRP3 inflammasome is now widely recognized as a critical link between mitochondrial dysfunctions and the integrity of the innate immune system [330,331]. During the progression of HF, cardiac inflammation is the result of increased mitochondrial injury caused by increased oxidative stress and mitophagy. It has been shown that mtDNA is released into the circulation during TAC surgery, and increased mtDNA is also detected in myocardial infarction patients, which activates circulating immune cells [332]. Additionally, the activation of NLRP3 can be influenced by mitochondrial redox conditions and the availability of NAD+ [333].
The critical function of mtROS as both a signaling molecule and a direct inducer of inflammation in HF has been underscored by recent research. Proinflammatory cytokines and chemokines induced by mtROS can worsen the inflammatory response [334]. Additionally, mtROS activates the NF-kB pathway, activating inflammatory genes and linking mitochondrial dysfunction to the inflammatory processes observed in HF [335]. During “Mito-inflammation”, the mitochondria-related inflammatory response, mitochondria function as both inducers of mitochondrial damage-associated molecular patterns (mtDAMPs) and the downstream effects of intracellular signaling pathways that are initiated by exogenous pathogen-associated molecular patterns (PAMPs) [336]. During mito-inflammation, mitochondria release mtDNA, ATP, mtROS, cardiolipin, and mitochondrial Ca2+ into cytosol or extracellular milieu in order to stimulate the expression and release of a variety of inflammatory mediators. MtROS may either directly promote oxidative damage to intraorganelle molecules and mtDNA or travel freely across the OMM. Upon its release into the cytosol, mtROS may activate proinflammatory signaling pathways, including the NF-kB, HIF, and AP-1 pathways, thereby contributing to the secretion of proinflammatory cytokines, such as IL-1 and IL-8 [331,337]. As recently proposed, the chronic activation of these proinflammatory pathways is linked to increased systemic inflammation and oxidative stress, causing capillary rarefaction and mitochondrial dysfunction, as well as the impairment of endothelium-dependent vasodilation. This, in turn, could contribute to abnormalities in the heart, vasculature, skeletal muscle, and other organs in HFpEF through a variety of pathways [21,338].
Taken together, emerging evidence suggests a bidirectional crosstalk between mitochondria metabolic stress and inflammation in the pathogenesis of the HFpEF disorder. Excessive mtROS production has a detrimental impact on mitochondrial function, resulting in a vicious cycle of increased ROS production, mitochondrial damage, and inflammation. MtROS can facilitate the release of mtDNA and other mtDAMPs, which can activate proinflammatory signaling pathways and further amplify inflammation as well as mtROS generation (Figure 7). The complexity of HFpEF is further exacerbated by the well-established heterogeneity among HFpEF phenotypes.

7. Therapeutics for HFpEF: Targeting mtROS and Associated Signal Molecules

Although the management of HFpEF continues to be a challenge, multiple treatments encompassing the management of comorbidities, consideration of non-pharmacological options, and implementation of guideline-directed medical therapy have currently been entering clinical trials, with mixed success (Table 2). A multidisciplinary approach is the optimal method, which would facilitate coordinated treatment plans to enhance quality of life and decrease hospitalizations by enabling early diagnosis and comprehensive management of HFpEF patients.

7.1. Conventional Pharmacological Therapy

7.1.1. Diuretics

In the event of acute decompensation or to maintain euvolemia, diuretics should be administered to patients with HFpEF who have fluid retention to alleviate congestion and ameliorate symptoms [339]. Attention should be taken with obese patients that may have poor diuresis tolerance and a greater risk of renal dysfunction during decongestion [340].

7.1.2. Sodium-Glucose Co-Transporter 2 (SGLT2) Inhibitors

SGLT2 inhibitors including dapagliflozin, empagliflozin, and sotagliflozin have been historically used in the treatment of type 2 diabetes, but emerging clinical evidence also shows that they also provide significant cardiovascular benefits in patients with HF. Specifically, in the DELIVER and EMPEROR-Preserved trials, the risk of HF hospitalization was reduced among individuals with HF and an ejection fraction ≥40% using dapagliflozin and empagliflozin [127,341]. In a meta-analysis that combined both trials, the SGLT2 inhibitors demonstrated a consistent reduction in the composite of CV-related death or first HF hospitalization, with HR 0.88, 95% CI 0.77–1.00 for CV-related death and HR 0.74, 95% CI 0.67–0.83 for HF hospitalization [342]. Sotagliflozin and empagliflozin were associated with improved outcomes, including mortality and HF events, among individuals admitted with acute decompensated HF in the SOLOIST-WHF and EMPULSE trials [343,344].

7.1.3. Mineralocorticoid Receptor Antagonists (MARs)

There is evidence to suggest that MRAs may be beneficial in HFpEF [345]. In mice, the expression of proinflammatory factors in adipocytes was reduced with MRA treatment [346]. Nevertheless, the TOPCAT trial did not demonstrate a substantial advantage in a composite outcome of CV-related mortality, aborted cardiac arrest, or HF hospitalization among patients with HF and an ejection fraction of ≥45% [14]. However, a subsequent subgroup analysis of the TOPCAT trial revealed substantial regional variation in the treatment effect and outcomes, which was more favorable in patients with higher BMI [14,347]. Consequently, the MAR spironolactone is recommended for HFpEF at a grade of 2b according to the AHA/ACC/HFSA Guideline for the Management of HF [339].

7.1.4. Angiotensin Receptor–Neprilysin Inhibitors

In comparison to valsartan (an angiotensin II receptor blocker), sacubitril–valsartan (a molecule combining angiotensin receptor blocking agent and neprilysin inhibitor) did not achieve a substantial decrease in CV-related mortality or HF hospitalization in individuals with HF and an EF ≥ 45% who were enrolled in the PARAGON-HF trial [348]. Nevertheless, a pre-specified subgroup analysis revealed that sacubitril–valsartan provided a benefit in individuals with LVEF below the median of 57% and in women in comparison to men. This benefit was primarily due to a decrease in HF hospitalization [349]. The FDA has recently approved sacubitril–valsartan for all patients with HF, with the benefits being most apparent in patients with a LVEF below normal. Recent expert consensus recommendations recommend that women with HF be treated with sacubitril–valsartan and spironolactone across the LVEF spectrum, unless otherwise contraindicated, due to the fact that women typically have higher EF than men because of their smaller size of the LV cavity [350,351].

7.1.5. Angiotensin Receptor Blockers

The use of angiotensin receptor blockers alone can be interpreted as the establishment of a contraindication or other restriction on the use of angiotensin receptor–neprilysin inhibitors. In the CHARM-Preserved trial [352], treatment with an angiotensin receptor blocker candesartan was associated with a lower incidence of HF hospitalization compared to the placebo in patients with HF and an EF ≥ 40%. However, the I-PRESERVE trial did not demonstrate any observed benefit from the use of irbesartan, another angiotensin receptor blocker [13].

7.1.6. β-Blockers

While β-blockers are not recommended for HFpEF, they are often used for other purposes, such as hypertension control, heart rate control, or in the treatment of angina in CAD. Recent meta-analysis has shown that β-blockers reduce all-cause and CV-related death in HFrEF and HFmrEF, but not in HFpEF [353]. Moreover, there is no consistent benefit from β-blockers in AF patients [353]. Potentially, β-blockers may worsen chronotropic intolerance in HFpEF, resulting in reduced exercise capacity. Specifically, β-blocker discontinuation in HFpEF patients with chronotropic incompetence can enhance maximal functional capacity, especially in those with reduced end-systolic LV volume [354]. However, a systematic review and meta-analysis of HFpEF studies showed that β-blocker therapy reduces all-cause mortality by 19% but has no impact on HF hospitalization or its composite with mortality [355].

7.2. Mitochondria Oxidative Stress-Targeted Therapy

Mitochondrion-targeted medicines have pros and cons over standard drugs. These medicines target mitochondrial dysfunction in HF, allowing for more precise cellular intervention and better outcomes [305,356]. By focusing on mitochondrial pathways, such drugs may avoid systemic side effects of traditional pharmaceuticals. Directly improving mitochondrial activity increases ATP synthesis, reduces oxidative stress, and improves cellular energy metabolism, improving heart function and patient health. However, developing and delivering mitochondrion-targeted medicines can be more complicated than standard pharmaceuticals, making it difficult to reach mitochondria safely and effectively. Although most of these medicines are targeted, off-target consequences are possible, especially if they affect mitochondrial function in noncardiac organs. In general, mitochondrial therapy acts by changing metabolic substrate utilization, improving OXPHOS, minimizing oxidative damage, or optimizing mitochondrial dynamics [357,358,359].
Regarding HF, evidence shows that the selective activation of the mitochondrial antioxidant system could protect against oxidative damage and the progression of cardiac dysfunction [294,360]. Indeed, mtROS-scavenging compounds are now undergoing human testing and have demonstrated good tissue permeability and subcellular targeting capabilities (e.g., NCT03506633, NCT02966665, NCT01925937, and ClinicalTrials.gov). Here we discuss some promising therapeutics targeting mitochondrial oxidative stress in the context of HF.

7.2.1. Elamipretide

Elamipretide (Bendavia, MTP-131, SS31) is a mitochondria-targeted antioxidant. By preferentially binding to cardiolipin in the cytochrome c/cardiolipin complex, it optimizes mitochondrial electron transport, ATP generation and biogenesis [361,362]. Several animal models have demonstrated the efficacy of this water-soluble antioxidant, preventing doxorubicin-induced cardiotoxicity and mitochondrial damage in mice [363,364]. In failing hearts, the treatment with elamipretide was able to restore mitochondrial biogenesis and downregulate oxidation genes [365]. Moreover, post-MI rats treated with elamipretide showed improved cardiac functions, smaller infarct size, favorable pathological LV remodeling, and reduced ROS generation in non-infarcted areas [366]. In a canine model of microembolization-induced HF, elamipretide treatment at 0.5 mg/kg per day reduced LV dysfunctions, which was linked to a lower level of mtROS and better mitochondrial activity [367].
In addition to animal models, elamipretide has also been studied in patients. Elamipretide improves mitochondrial function in failing human hearts in vitro [368]. In HF patients, 4 weeks of elamipretide treatment was safe and well tolerated, but it did not improve LV end systolic volume [369]. Moreover, for MI patients, elamipretide was not able to restore heart function or improve cardiac functional measurements [370]. Therefore, the effects of elamipretide treatment in HF, particularly in HFpEF, need to be fully explored in more clinical trials.

7.2.2. CoQ10

Coenzyme Q10 (CoQ10), a lipid-soluble quinone found in all cells, is crucial for electron transport and ATP production [371]. The benefits of CoQ10 in HF treatment are now widely accepted. In a 3-month study of 2664 NYHA class II and III patients, CoQ10 adjunctive therapy reduced adverse effects and improved clinical symptoms [372]. In addition, 420 HF patients receiving CoQ10 treatment (300 mg per day) for two years had improved NYHA class and HF survival [373]. Interestingly, CoQ10 seems to have longer positive effects on the heart. In a 5-year trial, selenium plus CoQ10 supplementation reduced CV mortality from 12.6% to 5.9% [374], and after a follow-up of twelve years, healthy participants and subgroups of patients with diabetes, hypertension, or ischemic heart disease on CoQ10 supplementation had a lower CV mortality risk [375]. This study shows that CoQ10’s protective effect on the heart lasts beyond the intervention period, proving its long-term efficacy in the heart.

7.2.3. MitoQ

MitoQ (mitoquinone) is a CoQ (Coenzyme Q) derivative. MitoQ mimics natural CoQ and possesses high antioxidant action. MitoQ prevents superoxide-induced lipid peroxidation and mitochondrial damage [376]. Data from animal and human studies showed that MitoQ therapy prevents mitochondrial oxidative damage [377,378,379]. In a pressure overload animal model induced by ascending aortic constriction, MitoQ administration for 7 days reduced cardiac apoptosis, hypertrophic remodeling, and LV dysfunction via modulating mitochondrial-associated redox signaling [380]. Moreover, MitoQ was able to restore the cardiac mitochondrial network in HF animals [381] and reduce the hypoxia-induced cardiac dysfunctions in sheep and chicken [376]. Importantly, 6 weeks of MitoQ supplementation could improve vascular endothelial dysfunction and reduce aortic stiffness in older persons [382], suggesting a potential CV beneficial to aged people.

7.3. Non-Pharmacological Management

7.3.1. Dietary Interventions, Low Carbohydrate Diet

In obese HFpEF patients, exercise tolerance and aerobic capacity are enhanced by nutritional or behavioral interventions that involve calorie restriction or weight loss [383,384]. Additionally, weight loss in obese HFpEF patients induced by gastric bypass surgery was associated with a regression of LV mass and wall thickness, as well as improvements in LV relaxation and LA filling pressure [383]. Moreover, a decrease in plasma sphingolipidome levels, implicated as a lower level of lipotoxicity, and an improvement in quality of life were also associated with surgery-induced weight loss [383]. Indeed, it was found that obesity, diabetes, and impaired cardiorespiratory capacity were associated with sugar consumption in patients with HFpEF [384]. The toxic effects such as cardiac dysfunctions, primarily affecting diastole, of high sugar consumption or Western diet were confirmed by experimental models and preclinical studies [385]. Normalizing the postprandial hyperglycemic index and insulinemic peaks by the modification/restriction of carbohydrate intake has been demonstrated to enhance all metabolic syndrome (MetS) manifestations and reduce CV risk [386]. Similarly, the consumption of unsaturated FAs (UFA) such as monounsaturated FAs (MUFAs) and polyunsaturated FAs (PUFAs) was correlated with improved cardiorespiratory capacity and diastolic functions, as well as increased fat-free mass in obese HFpEF patients, regardless of calorie intake. Similarly, mice fed a high-fat diet that was high in UFA and low in carbohydrates exhibited reduced weight gain and preserved myocardial functions [387]. In fact, UFA has been shown to suppress inflammation by reducing NLRP3 inflammasome activation in macrophages and in a type II diabetes model [388]. In addition, a low carbohydrate diet has been linked to lower rates of severe CV events and overall mortality in individuals with high CV risk, regardless of their body weight or overall dietary caloric intake [389]. Furthermore, a low-sugar diet can limit the generation of AGEs and, consequently, the deleterious glycoxidative stress [322]. Altogether, these findings highlight the capacity of a low-sugar/carbohydrate diet to modulate overall health and cardiac functions. They also demonstrate that the quality of nutrients, rather than the quantity of calories, could prevent CV events and control MetS manifestations.

7.3.2. Exercise Training

In accordance with the concept of mitohormesis, exercise-induced oxidative stress leads to adaptive responses that enhance the endogenous antioxidant defense capacity and improve insulin sensitivity [270,271]. Exercise intolerance or dyspnoea upon exertion are important hallmarks of HFpEF. In this sense, exercise training is one of the few non-pharmacological interventions that has been successfully applied to HFpEF patients [390,391]. Exercise increases cardiorespiratory capacity measured by VO2peak and quality of life in HFrEF and HFpEF [390,391]. The gain in VO2peak and the improvement in physical capacity are linked to atrial reverse remodeling and improved LV diastolic functions in HFpEF patients [392]. The benefits of exercise in relation to VO2peak appear not to be different between high-intensity intervals vs. moderate continuous training [393].

8. Conclusions

Despite the significant progress in our understanding over the past decades, HFpEF continues to be an intriguing condition for clinicians and researchers because it encompasses more than a single pathology, thereby presenting as a multifactorial and intricate disease. Therefore, a variety of mechanisms are expected to be involved in this pathology; however, an integrated understanding of HFpEF has yet to be achieved. ROS generated by mitochondria play dual roles in cellular physiology and pathology. Understanding the regulatory mechanisms of mtROS in HFpEF will be critical for developing precision therapies to address mitochondrial oxidative stress and its downstream effects. Nevertheless, due to the complicated nature of HFpEF, the current knowledge represents just a fraction of the whole. Further novel insights into the mtROS-inflammatory mechanisms contributing to HFpEF pathophysiology have the potential to benefit millions of individuals with HFpEF worldwide.

Author Contributions

Conceptualization, C.S.M., A.Z. and Q.X.; methodology, formal analysis, and investigation, C.S.M. and A.Z.; Figure preparation, A.Z. and Q.X.; resources, Q.X.; writing—original draft preparation, C.S.M. and A.Z.; writing—review and editing, Q.X.; supervision, Q.X.; project administration, Q.X.; funding acquisition, Q.X. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by British Heart Foundation (PG/15/11/31279, PG/15/86/31723, PG/20/10458, and PG/23/11371 to Q.X). This work forms part of the research portfolio for the National Institute for Health Research Biomedical Research Center at Barts.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. McDonagh, T.A.; Metra, M.; Adamo, M.; Gardner, R.S.; Baumbach, A.; Böhm, M.; Burri, H.; Butler, J.; Čelutkienė, J.; Chioncel, O.; et al. 2021 ESC Guidelines for the diagnosis and treatment of acute and chronic heart failure. Eur. Heart J. 2021, 42, 3599–3726. [Google Scholar] [CrossRef] [PubMed]
  2. Virani, S.S.; Alonso, A.; Aparicio, H.J.; Benjamin, E.J.; Bittencourt, M.S.; Callaway, C.W.; Carson, A.P.; Chamberlain, A.M.; Cheng, S.; Delling, F.N.; et al. Heart Disease and Stroke Statistics—2021 Update. Circulation 2021, 143, e254–e743. [Google Scholar] [CrossRef] [PubMed]
  3. Paulus, W.J.; Tschöpe, C. A Novel Paradigm for Heart Failure with Preserved Ejection Fraction Comorbidities Drive Myocardial Dysfunction and Remodeling Through Coronary Microvascular Endothelial Inflammation. J. Am. Coll. Cardiol. 2013, 62, 263–271. [Google Scholar] [CrossRef] [PubMed]
  4. Mishra, S.; Kass, D.A. Cellular and molecular pathobiology of heart failure with preserved ejection fraction. Nat. Rev. Cardiol. 2021, 18, 400–423. [Google Scholar] [CrossRef] [PubMed]
  5. Schiattarella, G.G.; Altamirano, F.; Tong, D.; French, K.M.; Villalobos, E.; Kim, S.Y.; Luo, X.; Jiang, N.; May, H.I.; Wang, Z.V.; et al. Nitrosative stress drives heart failure with preserved ejection fraction. Nature 2019, 568, 351–356. [Google Scholar] [CrossRef]
  6. James, S.L.; Abate, D.; Abate, K.H.; Abay, S.M.; Abbafati, C.; Abbasi, N.; Abbastabar, H.; Abd-Allah, F.; Abdela, J.; Abdelalim, A.; et al. Global, regional, and national incidence, prevalence, and years lived with disability for 354 diseases and injuries for 195 countries and territories, 1990–2017: A systematic analysis for the Global Burden of Disease Study 2017. Lancet 2018, 392, 1789–1858. [Google Scholar] [CrossRef]
  7. Teramoto, K.; Teng, T.-H.K.; Chandramouli, C.; Tromp, J.; Sakata, Y.; Lam, C.S. Epidemiology and Clinical Features of Heart Failure with Preserved Ejection Fraction. Card. Fail. Rev. 2022, 8, e27. [Google Scholar] [CrossRef]
  8. Mentz, R.J.; Kelly, J.P.; von Lueder, T.G.; Voors, A.A.; Lam, C.S.; Cowie, M.R.; Kjeldsen, K.; Jankowska, E.A.; Atar, D.; Butler, J.; et al. Noncardiac Comorbidities in Heart Failure with Reduced Versus Preserved Ejection Fraction. J. Am. Coll. Cardiol. 2014, 64, 2281–2293. [Google Scholar] [CrossRef]
  9. Owan, T.E.; Hodge, D.O.; Herges, R.M.; Jacobsen, S.J.; Roger, V.L.; Redfield, M.M. Trends in Prevalence and Outcome of Heart Failure with Preserved Ejection Fraction. N. Engl. J. Med. 2006, 355, 251–259. [Google Scholar] [CrossRef]
  10. Savarese, G.; Lund, L.H. Global Public Health Burden of Heart Failure. Card. Fail. Rev. 2017, 3, 7–11. [Google Scholar] [CrossRef]
  11. Bhatia, R.S.; Tu, J.V.; Lee, D.S.; Austin, P.C.; Fang, J.; Haouzi, A.; Gong, Y.; Liu, P.P. Outcome of Heart Failure with Preserved Ejection Fraction in a Population-Based Study. N. Engl. J. Med. 2006, 355, 260–269. [Google Scholar] [CrossRef] [PubMed]
  12. Gerber, Y.; Weston, S.A.; Redfield, M.M.; Chamberlain, A.M.; Manemann, S.M.; Jiang, R.; Killian, J.M.; Roger, V.L. A Contemporary Appraisal of the Heart Failure Epidemic in Olmsted County, Minnesota, 2000 to 2010. JAMA Intern. Med. 2015, 175, 996. [Google Scholar] [CrossRef] [PubMed]
  13. Massie, B.M.; Carson, P.E.; McMurray, J.J.; Komajda, M.; McKelvie, R.; Zile, M.R.; Anderson, S.; Donovan, M.; Iverson, E.; Staiger, C.; et al. Irbesartan in Patients with Heart Failure and Preserved Ejection Fraction. N. Engl. J. Med. 2008, 359, 2456–2467. [Google Scholar] [CrossRef] [PubMed]
  14. Pitt, B.; Pfeffer, M.A.; Assmann, S.F.; Boineau, R.; Anand, I.S.; Claggett, B.; Clausell, N.; Desai, A.S.; Diaz, R.; Fleg, J.L.; et al. Spironolactone for Heart Failure with Preserved Ejection Fraction. N. Engl. J. Med. 2014, 370, 1383–1392. [Google Scholar] [CrossRef] [PubMed]
  15. Vaduganathan, M.; Patel, R.B.; Michel, A.; Shah, S.J.; Senni, M.; Gheorghiade, M.; Butler, J. Mode of Death in Heart Failure with Preserved Ejection Fraction. J. Am. Coll. Cardiol. 2017, 69, 556–569. [Google Scholar] [CrossRef] [PubMed]
  16. Loungani, R.S.; Teerlink, J.R.; Metra, M.; Allen, L.A.; Butler, J.; Carson, P.E.; Chen, C.-W.; Cotter, G.; Davison, B.A.; Eapen, Z.J.; et al. Cause of Death in Patients with Acute Heart Failure: Insights From RELAX-AHF-2. JACC Heart Fail. 2020, 8, 999–1008. [Google Scholar] [CrossRef]
  17. Lee, D.S.; Gona, P.; Albano, I.; Larson, M.G.; Benjamin, E.J.; Levy, D.; Kannel, W.B.; Vasan, R.S. A Systematic Assessment of Causes of Death After Heart Failure Onset in the Community. Circ. Heart Fail. 2011, 4, 36–43. [Google Scholar] [CrossRef]
  18. Shahim, A.; Hourqueig, M.; Lund, L.H.; Savarese, G.; Oger, E.; Venkateshvaran, A.; Benson, L.; Daubert, J.C.; Linde, C.; Donal, E.; et al. Long-term outcomes in heart failure with preserved ejection fraction: Predictors of cardiac and non-cardiac mortality. ESC Heart Fail. 2023, 10, 1835–1846. [Google Scholar] [CrossRef]
  19. Köseoğlu, F.D.; Özlek, B. Anemia and Iron Deficiency Predict All-Cause Mortality in Patients with Heart Failure and Preserved Ejection Fraction: 6-Year Follow-Up Study. Diagnostics 2024, 14, 209. [Google Scholar] [CrossRef]
  20. Omote, K.; Verbrugge, F.H.; Borlaug, B.A. Heart Failure with Preserved Ejection Fraction: Mechanisms and Treatment Strategies. Annu. Rev. Med. 2022, 73, 321–337. [Google Scholar] [CrossRef]
  21. Schiattarella, G.G.; Alcaide, P.; Condorelli, G.; Gillette, T.G.; Heymans, S.; Jones, E.A.V.; Kallikourdis, M.; Lichtman, A.; Marelli-Berg, F.; Shah, S.; et al. Immunometabolic Mechanisms of Heart Failure with Preserved Ejection Fraction. Nat. Cardiovasc. Res. 2022, 1, 211–222. [Google Scholar] [CrossRef] [PubMed]
  22. Amano, S.U.; Cohen, J.L.; Vangala, P.; Tencerova, M.; Nicoloro, S.M.; Yawe, J.C.; Shen, Y.; Czech, M.P.; Aouadi, M. Local Proliferation of Macrophages Contributes to Obesity-Associated Adipose Tissue Inflammation. Cell Metab. 2014, 19, 162–171. [Google Scholar] [CrossRef] [PubMed]
  23. Ouchi, N.; Parker, J.L.; Lugus, J.J.; Walsh, K. Adipokines in inflammation and metabolic disease. Nat. Rev. Immunol. 2011, 11, 85–97. [Google Scholar] [CrossRef] [PubMed]
  24. Dubrock, H.M.; Abouezzeddine, O.F.; Redfield, M.M. High-sensitivity C-reactive protein in heart failure with preserved ejection fraction. PLoS ONE 2018, 13, e0201836. [Google Scholar] [CrossRef] [PubMed]
  25. Pfeffer, M.A.; Shah, A.M.; Borlaug, B.A. Heart Failure with Preserved Ejection Fraction In Perspective. Circ. Res. 2019, 124, 1598–1617. [Google Scholar] [CrossRef] [PubMed]
  26. Zile, M.R.; Baicu, C.F.; Gaasch, W.H. Diastolic Heart Failure—Abnormalities in Active Relaxation and Passive Stiffness of the Left Ventricle. N. Engl. J. Med. 2004, 350, 1953–1959. [Google Scholar] [CrossRef]
  27. Borlaug, B.A.; Jaber, W.A.; Ommen, S.R.; Lam, C.S.P.; Redfield, M.M.; Nishimura, R.A. Diastolic relaxation and compliance reserve during dynamic exercise in heart failure with preserved ejection fraction. Heart 2011, 97, 964–969. [Google Scholar] [CrossRef]
  28. Methawasin, M.; Strom, J.G.; Slater, R.E.; Fernandez, V.; Saripalli, C.; Granzier, H. Experimentally Increasing the Compliance of Titin Through RNA Binding Motif-20 (RBM20) Inhibition Improves Diastolic Function In a Mouse Model of Heart Failure with Preserved Ejection Fraction. Circulation 2016, 134, 1085–1099. [Google Scholar] [CrossRef]
  29. Van Heerebeek, L.; Hamdani, N.; Falcão-Pires, I.; Leite-Moreira, A.F.; Begieneman, M.P.V.; Bronzwaer, J.G.F.; Van Der Velden, J.; Stienen, G.J.M.; Laarman, G.J.; Somsen, A.; et al. Low Myocardial Protein Kinase G Activity in Heart Failure with Preserved Ejection Fraction. Circulation 2012, 126, 830–839. [Google Scholar] [CrossRef]
  30. Zile, M.R.; Baicu, C.F.; Ikonomidis, J.S.; Stroud, R.E.; Nietert, P.J.; Bradshaw, A.D.; Slater, R.; Palmer, B.M.; Van Buren, P.; Meyer, M.; et al. Myocardial Stiffness in Patients with Heart Failure and a Preserved Ejection Fraction. Circulation 2015, 131, 1247–1259. [Google Scholar] [CrossRef]
  31. Phan, T.T.; Abozguia, K.; Shivu, G.N.; Mahadevan, G.; Ahmed, I.; Williams, L.; Dwivedi, G.; Patel, K.; Steendijk, P.; Ashrafian, H.; et al. Heart failure with preserved ejection fraction is characterized by dynamic impairment of active relaxation and contraction of the left ventricle on exercise and associated with myocardial energy deficiency. J. Am. Coll. Cardiol. 2009, 54, 402–409. [Google Scholar] [CrossRef] [PubMed]
  32. Sorimachi, H.; Burkhoff, D.; Verbrugge, F.H.; Omote, K.; Obokata, M.; Reddy, Y.N.V.; Takahashi, N.; Sunagawa, K.; Borlaug, B.A. Obesity, venous capacitance, and venous compliance in heart failure with preserved ejection fraction. Eur. J. Heart Fail. 2021, 23, 1648–1658. [Google Scholar] [CrossRef] [PubMed]
  33. Borlaug, B.A.; Lam, C.S.; Roger, V.L.; Rodeheffer, R.J.; Redfield, M.M. Contractility and ventricular systolic stiffening in hypertensive heart disease insights into the pathogenesis of heart failure with preserved ejection fraction. J. Am. Coll. Cardiol. 2009, 54, 410–418. [Google Scholar] [CrossRef] [PubMed]
  34. Tan, Y.T.; Wenzelburger, F.; Lee, E.; Heatlie, G.; Leyva, F.; Patel, K.; Frenneaux, M.; Sanderson, J.E. The Pathophysiology of Heart Failure with Normal Ejection Fraction. J. Am. Coll. Cardiol. 2009, 54, 36–46. [Google Scholar] [CrossRef] [PubMed]
  35. Backhaus, S.J.; Lange, T.; George, E.F.; Hellenkamp, K.; Gertz, R.J.; Billing, M.; Wachter, R.; Steinmetz, M.; Kutty, S.; Raaz, U.; et al. Exercise Stress Real-Time Cardiac Magnetic Resonance Imaging for Noninvasive Characterization of Heart Failure with Preserved Ejection Fraction. Circulation 2021, 143, 1484–1498. [Google Scholar] [CrossRef]
  36. Reddy, Y.N.V.; Obokata, M.; Egbe, A.; Yang, J.H.; Pislaru, S.; Lin, G.; Carter, R.; Borlaug, B.A. Left atrial strain and compliance in the diagnostic evaluation of heart failure with preserved ejection fraction. Eur. J. Heart Fail. 2019, 21, 891–900. [Google Scholar] [CrossRef]
  37. Reddy, Y.N.V.; Obokata, M.; Verbrugge, F.H.; Lin, G.; Borlaug, B.A. Atrial Dysfunction in Patients with Heart Failure with Preserved Ejection Fraction and Atrial Fibrillation. J. Am. Coll. Cardiol. 2020, 76, 1051–1064. [Google Scholar] [CrossRef]
  38. Borlaug, B.A.; Reddy, Y.N. The Role of the Pericardium in Heart Failure: Implications for Pathophysiology and Treatment. JACC Heart Fail. 2019, 7, 574–585. [Google Scholar] [CrossRef]
  39. Freed, B.H.; Daruwalla, V.; Cheng, J.Y.; Aguilar, F.G.; Beussink, L.; Choi, A.; Klein, D.A.; Dixon, D.; Baldridge, A.; Rasmussen-Torvik, L.J.; et al. Prognostic Utility and Clinical Significance of Cardiac Mechanics in Heart Failure with Preserved Ejection Fraction. Circ. Cardiovasc. Imaging 2016, 9, e003754. [Google Scholar] [CrossRef]
  40. Telles, F.; Nanayakkara, S.; Evans, S.; Patel, H.C.; Mariani, J.A.; Vizi, D.; William, J.; Marwick, T.H.; Kaye, D.M. Impaired left atrial strain predicts abnormal exercise haemodynamics in heart failure with preserved ejection fraction. Eur. J. Heart Fail. 2019, 21, 495–505. [Google Scholar] [CrossRef]
  41. Obokata, M.; Reddy, Y.N.V.; Melenovsky, V.; Pislaru, S.; Borlaug, B.A. Deterioration in right ventricular structure and function over time in patients with heart failure and preserved ejection fraction. Eur. Heart J. 2019, 40, 689–697. [Google Scholar] [CrossRef] [PubMed]
  42. Zakeri, R.; Chamberlain, A.M.; Roger, V.L.; Redfield, M.M. Temporal Relationship and Prognostic Significance of Atrial Fibrillation in Heart Failure Patients with Preserved Ejection Fraction. Circulation 2013, 128, 1085–1093. [Google Scholar] [CrossRef] [PubMed]
  43. Guazzi, M.; Naeije, R. Pulmonary Hypertension in Heart Failure. J. Am. Coll. Cardiol. 2017, 69, 1718–1734. [Google Scholar] [CrossRef] [PubMed]
  44. Gorter, T.M.; Obokata, M.; Reddy, Y.N.V.; Melenovsky, V.; Borlaug, B.A. Exercise unmasks distinct pathophysiologic features in heart failure with preserved ejection fraction and pulmonary vascular disease. Eur. Heart J. 2018, 39, 2825–2835. [Google Scholar] [CrossRef] [PubMed]
  45. Verbrugge, F.H.; Guazzi, M.; Testani, J.M.; Borlaug, B.A. Altered Hemodynamics and End-Organ Damage in Heart Failure. Circulation 2020, 142, 998–1012. [Google Scholar] [CrossRef]
  46. Gorter, T.M.; Hoendermis, E.S.; Van Veldhuisen, D.J.; Voors, A.A.; Lam, C.S.P.; Geelhoed, B.; Willems, T.P.; Van Melle, J.P. Right ventricular dysfunction in heart failure with preserved ejection fraction: A systematic review and meta-analysis. Eur. J. Heart Fail. 2016, 18, 1472–1487. [Google Scholar] [CrossRef]
  47. Gorter, T.M.; Van Veldhuisen, D.J.; Bauersachs, J.; Borlaug, B.A.; Celutkiene, J.; Coats, A.J.S.; Crespo-Leiro, M.G.; Guazzi, M.; Harjola, V.P.; Heymans, S.; et al. Right heart dysfunction and failure in heart failure with preserved ejection fraction: Mechanisms and management. Position statement on behalf of the Heart Failure Association of the European Society of Cardiology. Eur. J. Heart Fail. 2018, 20, 16–37. [Google Scholar] [CrossRef]
  48. Hoeper, M.M.; Meyer, K.; Rademacher, J.; Fuge, J.; Welte, T.; Olsson, K.M. Diffusion Capacity and Mortality in Patients with Pulmonary Hypertension Due to Heart Failure with Preserved Ejection Fraction. JACC Heart Fail. 2016, 4, 441–449. [Google Scholar] [CrossRef]
  49. Fermoyle, C.C.; Stewart, G.M.; Borlaug, B.A.; Johnson, B.D. Simultaneous Measurement of Lung Diffusing Capacity and Pulmonary Hemodynamics Reveals Exertional Alveolar-Capillary Dysfunction in Heart Failure with Preserved Ejection Fraction. J. Am. Heart Assoc. 2021, 10, e019950. [Google Scholar] [CrossRef]
  50. Chirinos, J.A.; Bhattacharya, P.; Kumar, A.; Proto, E.; Konda, P.; Segers, P.; Akers, S.R.; Townsend, R.R.; Zamani, P. Impact of Diabetes Mellitus on Ventricular Structure, Arterial Stiffness, and Pulsatile Hemodynamics in Heart Failure with Preserved Ejection Fraction. J. Am. Heart Assoc. 2019, 8, e011457. [Google Scholar] [CrossRef]
  51. Reddy, Y.N.V.; Andersen, M.J.; Obokata, M.; Koepp, K.E.; Kane, G.C.; Melenovsky, V.; Olson, T.P.; Borlaug, B.A. Arterial Stiffening with Exercise in Patients with Heart Failure and Preserved Ejection Fraction. J. Am. Coll. Cardiol. 2017, 70, 136–148. [Google Scholar] [CrossRef] [PubMed]
  52. Schwartzenberg, S.; Redfield, M.M.; From, A.M.; Sorajja, P.; Nishimura, R.A.; Borlaug, B.A. Effects of vasodilation in heart failure with preserved or reduced ejection fraction: Implications of distinct pathophysiologies on response to therapy. J. Am. Coll. Cardiol. 2012, 59, 442–451. [Google Scholar] [CrossRef] [PubMed]
  53. Akiyama, E.; Sugiyama, S.; Matsuzawa, Y.; Konishi, M.; Suzuki, H.; Nozaki, T.; Ohba, K.; Matsubara, J.; Maeda, H.; Horibata, Y.; et al. Incremental prognostic significance of peripheral endothelial dysfunction in patients with heart failure with normal left ventricular ejection fraction. J. Am. Coll. Cardiol. 2012, 60, 1778–1786. [Google Scholar] [CrossRef] [PubMed]
  54. Sanders-Van Wijk, S.; Tromp, J.; Beussink-Nelson, L.; Hage, C.; Svedlund, S.; Saraste, A.; Swat, S.A.; Sanchez, C.; Njoroge, J.; Tan, R.-S.; et al. Proteomic Evaluation of the Comorbidity-Inflammation Paradigm in Heart Failure with Preserved Ejection Fraction. Circulation 2020, 142, 2029–2044. [Google Scholar] [CrossRef]
  55. Ahmad, A.; Corban, M.T.; Toya, T.; Verbrugge, F.H.; Sara, J.D.; Lerman, L.O.; Borlaug, B.A.; Lerman, A. Coronary microvascular dysfunction is associated with exertional haemodynamic abnormalities in patients with heart failure with preserved ejection fraction. Eur. J. Heart Fail. 2021, 23, 765–772. [Google Scholar] [CrossRef]
  56. Obokata, M.; Reddy, Y.N.V.; Melenovsky, V.; Kane, G.C.; Olson, T.P.; Jarolim, P.; Borlaug, B.A. Myocardial Injury and Cardiac Reserve in Patients with Heart Failure and Preserved Ejection Fraction. J. Am. Coll. Cardiol. 2018, 72, 29–40. [Google Scholar] [CrossRef]
  57. Shah, S.J.; Lam, C.S.P.; Svedlund, S.; Saraste, A.; Hage, C.; Tan, R.-S.; Beussink-Nelson, L.; Ljung Faxén, U.; Fermer, M.L.; Broberg, M.A.; et al. Prevalence and correlates of coronary microvascular dysfunction in heart failure with preserved ejection fraction: PROMIS-HFpEF. Eur. Heart J. 2018, 39, 3439–3450. [Google Scholar] [CrossRef]
  58. Yang, J.H.; Obokata, M.; Reddy, Y.N.; Redfield, M.M.; Lerman, A.; Borlaug, B.A. Endothelium-dependent and independent coronary microvascular dysfunction in patients with heart failure with preserved ejection fraction. Eur. J. Heart Fail. 2020, 22, 432–441. [Google Scholar] [CrossRef]
  59. Abudiab, M.M.; Redfield, M.M.; Melenovsky, V.; Olson, T.P.; Kass, D.A.; Johnson, B.D.; Borlaug, B.A. Cardiac output response to exercise in relation to metabolic demand in heart failure with preserved ejection fraction. Eur. J. Heart Fail. 2013, 15, 776–785. [Google Scholar] [CrossRef]
  60. Borlaug, B.A.; Melenovsky, V.; Russell, S.D.; Kessler, K.; Pacak, K.; Becker, L.C.; Kass, D.A. Impaired Chronotropic and Vasodilator Reserves Limit Exercise Capacity in Patients with Heart Failure and a Preserved Ejection Fraction. Circulation 2006, 114, 2138–2147. [Google Scholar] [CrossRef]
  61. Sarma, S.; Stoller, D.; Hendrix, J.; Howden, E.; Lawley, J.; Livingston, S.; Adams-Huet, B.; Holmes, C.; Goldstein, D.S.; Levine, B.D. Mechanisms of Chronotropic Incompetence in Heart Failure with Preserved Ejection Fraction. Circ. Heart Fail. 2020, 13, e006331. [Google Scholar] [CrossRef] [PubMed]
  62. Sarma, S.; Howden, E.; Lawley, J.; Samels, M.; Levine, B.D. Central Command and the Regulation of Exercise Heart Rate Response in Heart Failure with Preserved Ejection Fraction. Circulation 2021, 143, 783–789. [Google Scholar] [CrossRef] [PubMed]
  63. Kouta Funakoshi, M.; Kazuya Hosokawa, P.; Takuya Kishi, P.; Tomomi Ide, P.; Kenji Sunagawa, P. Striking volume intolerance is induced by mimicking arterial baroreflex failure in normal left ventricular function. Basic. Sci. Exp. Study 2013, 20, 53–59. [Google Scholar] [CrossRef] [PubMed]
  64. Bhella, P.S.; Prasad, A.; Heinicke, K.; Hastings, J.L.; Arbab-Zadeh, A.; Adams-Huet, B.; Pacini, E.L.; Shibata, S.; Palmer, M.D.; Newcomer, B.R.; et al. Abnormal haemodynamic response to exercise in heart failure with preserved ejection fraction. Eur. J. Heart Fail. 2011, 13, 1296–1304. [Google Scholar] [CrossRef] [PubMed]
  65. Houstis, N.E.; Eisman, A.S.; Pappagianopoulos, P.P.; Wooster, L.; Bailey, C.S.; Wagner, P.D.; Lewis, G.D. Exercise Intolerance in Heart Failure with Preserved Ejection Fraction. Circulation 2018, 137, 148–161. [Google Scholar] [CrossRef]
  66. Kitzman, D.W.; Nicklas, B.; Kraus, W.E.; Lyles, M.F.; Eggebeen, J.; Morgan, T.M.; Haykowsky, M. Skeletal muscle abnormalities and exercise intolerance in older patients with heart failure and preserved ejection fraction. Am. J. Physiol. Heart Circ. Physiol. 2014, 306, H1364–H1370. [Google Scholar] [CrossRef]
  67. Kumar, A.A.; Kelly, D.P.; Chirinos, J.A. Mitochondrial Dysfunction in Heart Failure with Preserved Ejection Fraction. Circulation 2019, 139, 1435–1450. [Google Scholar] [CrossRef]
  68. Okayama, H.; Hamada, M.; Kawakami, H.; Ikeda, S.; Hashida, H.; Shigematsu, Y.; Hiwada, K. Alterations in expression of sarcoplasmic reticulum gene in Dahl rats during the transition from compensatory myocardial hypertrophy to heart failure. J. Hypertens. 1997, 15, 1767–1774. [Google Scholar] [CrossRef]
  69. Qu, P.; Hamada, M.; Ikeda, S.; Hiasa, G.; Shigematsu, Y.; Hiwada, K. Time-Course Changes in Left Ventricular Geometry and Function during the Development of Hypertension in Dahl Salt-Sensitive Rats. Hypertens. Res. 2000, 23, 613–623. [Google Scholar] [CrossRef]
  70. Chen-Izu, Y.; Chen, L.; Bányász, T.; McCulle, S.L.; Norton, B.; Scharf, S.M.; Agarwal, A.; Patwardhan, A.; Izu, L.T.; Balke, C.W. Hypertension-induced remodeling of cardiac excitation-contraction coupling in ventricular myocytes occurs prior to hypertrophy development. Am. J. Physiol.-Heart Circ. Physiol. 2007, 293, H3301–H3310. [Google Scholar] [CrossRef]
  71. Weber, K.T.; Janicki, J.S.; Pick, R.; Capasso, J.; Anversa, P. Myocardial fibrosis and pathologic hypertrophy in the rat with renovascular hypertension. Am. J. Cardiol. 1990, 65, 1g–7g. [Google Scholar] [CrossRef] [PubMed]
  72. Nishio, M.; Sakata, Y.; Mano, T.; Yoshida, J.; Ohtani, T.; Takeda, Y.; Miwa, T.; Masuyama, T.; Yamamoto, K.; Hori, M. Therapeutic effects of angiotensin II type 1 receptor blocker at an advanced stage of hypertensive diastolic heart failure. J. Hypertens. 2007, 25, 455–461. [Google Scholar] [CrossRef] [PubMed]
  73. Yamamoto, K.; Masuyama, T.; Sakata, Y.; Mano, T.; Nishikawa, N.; Kondo, H.; Akehi, N.; Kuzuya, T.; Miwa, T.; Hori, M. Roles of renin-angiotensin and endothelin systems in development of diastolic heart failure in hypertensive hearts. Cardiovasc. Res. 2000, 47, 274–283. [Google Scholar] [CrossRef]
  74. Lupón, J.; Gavidia-Bovadilla, G.; Ferrer, E.; de Antonio, M.; Perera-Lluna, A.; López-Ayerbe, J.; Domingo, M.; Núñez, J.; Zamora, E.; Moliner, P.; et al. Heart Failure with Preserved Ejection Fraction Infrequently Evolves Toward a Reduced Phenotype in Long-Term Survivors. Circ. Heart Fail. 2019, 12, e005652. [Google Scholar] [CrossRef] [PubMed]
  75. Nagata, K.; Liao, R.; Eberli, F.R.; Satoh, N.; Chevalier, B.; Apstein, C.S.; Suter, T.M. Early changes in excitation-contraction coupling: Transition from compensated hypertrophy to failure in Dahl salt-sensitive rat myocytes. Cardiovasc. Res. 1998, 37, 467–477. [Google Scholar] [CrossRef] [PubMed]
  76. Capasso, J.M.; Palackal, T.; Olivetti, G.; Anversa, P. Left ventricular failure induced by long-term hypertension in rats. Circ. Res. 1990, 66, 1400–1412. [Google Scholar] [CrossRef]
  77. Wallner, M.; Eaton, D.M.; Berretta, R.M.; Liesinger, L.; Schittmayer, M.; Gindlhuber, J.; Wu, J.; Jeong, M.Y.; Lin, Y.H.; Borghetti, G.; et al. HDAC inhibition improves cardiopulmonary function in a feline model of diastolic dysfunction. Sci. Transl. Med. 2020, 12, eaay7205. [Google Scholar] [CrossRef]
  78. Gallet, R.; de Couto, G.; Simsolo, E.; Valle, J.; Sun, B.; Liu, W.; Tseliou, E.; Zile, M.R.; Marbán, E. Cardiosphere-derived cells reverse heart failure with preserved ejection fraction (HFpEF) in rats by decreasing fibrosis and inflammation. JACC Basic. Transl. Sci. 2016, 1, 14–28. [Google Scholar] [CrossRef]
  79. Tofovic, S.P.; Kusaka, H.; Kost, C.K., Jr.; Bastacky, S. Renal function and structure in diabetic, hypertensive, obese ZDFxSHHF-hybrid rats. Ren. Fail. 2000, 22, 387–406. [Google Scholar] [CrossRef]
  80. Leite, S.; Cerqueira, R.J.; Ibarrola, J.; Fontoura, D.; Fernández-Celis, A.; Zannad, F.; Falcão-Pires, I.; Paulus, W.J.; Leite-Moreira, A.F.; Rossignol, P.; et al. Arterial Remodeling and Dysfunction in the ZSF1 Rat Model of Heart Failure with Preserved Ejection Fraction. Circ. Heart Fail. 2019, 12, e005596. [Google Scholar] [CrossRef]
  81. Lai, Y.C.; Tabima, D.M.; Dube, J.J.; Hughan, K.S.; Vanderpool, R.R.; Goncharov, D.A.; St Croix, C.M.; Garcia-Ocaña, A.; Goncharova, E.A.; Tofovic, S.P.; et al. SIRT3-AMP-Activated Protein Kinase Activation by Nitrite and Metformin Improves Hyperglycemia and Normalizes Pulmonary Hypertension Associated with Heart Failure with Preserved Ejection Fraction. Circulation 2016, 133, 717–731. [Google Scholar] [CrossRef] [PubMed]
  82. Hopf, A.E.; Andresen, C.; Kötter, S.; Isić, M.; Ulrich, K.; Sahin, S.; Bongardt, S.; Röll, W.; Drove, F.; Scheerer, N.; et al. Diabetes-Induced Cardiomyocyte Passive Stiffening Is Caused by Impaired Insulin-Dependent Titin Modification and Can Be Modulated by Neuregulin-1. Circ. Res. 2018, 123, 342–355. [Google Scholar] [CrossRef] [PubMed]
  83. Salah, E.M.; Bastacky, S.I.; Jackson, E.K.; Tofovic, S.P. Captopril Attenuates Cardiovascular and Renal Disease in a Rat Model of Heart Failure with Preserved Ejection Fraction. J. Cardiovasc. Pharmacol. 2018, 71, 205–214. [Google Scholar] [CrossRef] [PubMed]
  84. Tong, D.; Schiattarella, G.G.; Jiang, N.; May, H.I.; Lavandero, S.; Gillette, T.G.; Hill, J.A. Female Sex Is Protective in a Preclinical Model of Heart Failure with Preserved Ejection Fraction. Circulation 2019, 140, 1769–1771. [Google Scholar] [CrossRef]
  85. Munagala, V.K.; Hart, C.Y.; Burnett, J.C., Jr.; Meyer, D.M.; Redfield, M.M. Ventricular structure and function in aged dogs with renal hypertension: A model of experimental diastolic heart failure. Circulation 2005, 111, 1128–1135. [Google Scholar] [CrossRef]
  86. Kawaguchi, M.; Hay, I.; Fetics, B.; Kass, D.A. Combined ventricular systolic and arterial stiffening in patients with heart failure and preserved ejection fraction: Implications for systolic and diastolic reserve limitations. Circulation 2003, 107, 714–720. [Google Scholar] [CrossRef]
  87. Reiter, U.; Reiter, G.; Manninger, M.; Adelsmayr, G.; Schipke, J.; Alogna, A.; Rajces, A.; Stalder, A.F.; Greiser, A.; Mühlfeld, C.; et al. Early-stage heart failure with preserved ejection fraction in the pig: A cardiovascular magnetic resonance study. J. Cardiovasc. Magn. Reson. 2016, 18, 63. [Google Scholar] [CrossRef]
  88. Sorop, O.; Heinonen, I.; van Kranenburg, M.; van de Wouw, J.; de Beer, V.J.; Nguyen, I.T.N.; Octavia, Y.; van Duin, R.W.B.; Stam, K.; van Geuns, R.J.; et al. Multiple common comorbidities produce left ventricular diastolic dysfunction associated with coronary microvascular dysfunction, oxidative stress, and myocardial stiffening. Cardiovasc. Res. 2018, 114, 954–964. [Google Scholar] [CrossRef]
  89. Schwarzl, M.; Hamdani, N.; Seiler, S.; Alogna, A.; Manninger, M.; Reilly, S.; Zirngast, B.; Kirsch, A.; Steendijk, P.; Verderber, J.; et al. A porcine model of hypertensive cardiomyopathy: Implications for heart failure with preserved ejection fraction. Am. J. Physiol. Heart Circ. Physiol. 2015, 309, H1407–H1418. [Google Scholar] [CrossRef]
  90. Ferrari, R.; Censi, S.; Mastrorilli, F.; Boraso, A. Prognostic benefits of heart rate reduction in cardiovascular disease. Eur. Heart J. Suppl. 2003, 5, G10–G14. [Google Scholar] [CrossRef]
  91. Nguyen, B.Y.; Ruiz-Velasco, A.; Bui, T.; Collins, L.; Wang, X.; Liu, W. Mitochondrial function in the heart: The insight into mechanisms and therapeutic potentials. Br. J. Pharmacol. 2019, 176, 4302–4318. [Google Scholar] [CrossRef] [PubMed]
  92. Barth, E.; Stämmler, G.; Speiser, B.; Schaper, J. Ultrastructural quantitation of mitochondria and myofilaments in cardiac muscle from 10 different animal species including man. J. Mol. Cell. Cardiol. 1992, 24, 669–681. [Google Scholar] [CrossRef] [PubMed]
  93. Schaper, J.; Meiser, E.; Stämmler, G. Ultrastructural morphometric analysis of myocardium from dogs, rats, hamsters, mice, and from human hearts. Circ. Res. 1985, 56, 377–391. [Google Scholar] [CrossRef] [PubMed]
  94. Lopaschuk, G.D.; Karwi, Q.G.; Tian, R.; Wende, A.R.; Abel, E.D. Cardiac Energy Metabolism in Heart Failure. Circ. Res. 2021, 128, 1487–1513. [Google Scholar] [CrossRef]
  95. Chaanine, A.H.; Joyce, L.D.; Stulak, J.M.; Maltais, S.; Joyce, D.L.; Dearani, J.A.; Klaus, K.; Nair, K.S.; Hajjar, R.J.; Redfield, M.M. Mitochondrial Morphology, Dynamics, and Function in Human Pressure Overload or Ischemic Heart Disease with Preserved or Reduced Ejection Fraction. Circ. Heart Fail. 2019, 12, e005131. [Google Scholar] [CrossRef]
  96. Stoldt, S.; Wenzel, D.; Kehrein, K.; Riedel, D.; Ott, M.; Jakobs, S. Spatial orchestration of mitochondrial translation and OXPHOS complex assembly. Nat. Cell Biol. 2018, 20, 528–534. [Google Scholar] [CrossRef]
  97. Vercellino, I.; Sazanov, L.A. The assembly, regulation and function of the mitochondrial respiratory chain. Nat. Rev. Mol. Cell Biol. 2022, 23, 141–161. [Google Scholar] [CrossRef]
  98. Kretzschmar, T.; Wu, J.M.F.; Schulze, P.C. Mitochondrial Homeostasis Mediates Lipotoxicity in the Failing Myocardium. Int. J. Mol. Sci. 2021, 22, 1498. [Google Scholar] [CrossRef]
  99. Arnold, P.K.; Finley, L.W.S. Regulation and function of the mammalian tricarboxylic acid cycle. J. Biol. Chem. 2023, 299, 102838. [Google Scholar] [CrossRef]
  100. Martínez-Reyes, I.; Chandel, N.S. Mitochondrial TCA cycle metabolites control physiology and disease. Nat. Commun. 2020, 11, 102. [Google Scholar] [CrossRef]
  101. Spinelli, J.B.; Haigis, M.C. The multifaceted contributions of mitochondria to cellular metabolism. Nat. Cell Biol. 2018, 20, 745–754. [Google Scholar] [CrossRef] [PubMed]
  102. Wisneski, J.A.; Stanley, W.C.; Neese, R.A.; Gertz, E.W. Effects of acute hyperglycemia on myocardial glycolytic activity in humans. J. Clin. Investig. 1990, 85, 1648–1656. [Google Scholar] [CrossRef] [PubMed]
  103. Aubert, G.; Vega, R.B.; Kelly, D.P. Perturbations in the gene regulatory pathways controlling mitochondrial energy production in the failing heart. Biochim. Biophys. Acta 2013, 1833, 840–847. [Google Scholar] [CrossRef] [PubMed]
  104. Kolwicz, S.C., Jr.; Purohit, S.; Tian, R. Cardiac metabolism and its interactions with contraction, growth, and survival of cardiomyocytes. Circ. Res. 2013, 113, 603–616. [Google Scholar] [CrossRef] [PubMed]
  105. Da Dalt, L.; Cabodevilla, A.G.; Goldberg, I.J.; Norata, G.D. Cardiac lipid metabolism, mitochondrial function, and heart failure. Cardiovasc. Res. 2023, 119, 1905–1914. [Google Scholar] [CrossRef] [PubMed]
  106. van Hall, G. Lactate kinetics in human tissues at rest and during exercise. Acta Physiol. 2010, 199, 499–508. [Google Scholar] [CrossRef]
  107. Chen, Y.J.; Mahieu, N.G.; Huang, X.; Singh, M.; Crawford, P.A.; Johnson, S.L.; Gross, R.W.; Schaefer, J.; Patti, G.J. Lactate metabolism is associated with mammalian mitochondria. Nat. Chem. Biol. 2016, 12, 937–943. [Google Scholar] [CrossRef]
  108. Matsuura, T.R.; Puchalska, P.; Crawford, P.A.; Kelly, D.P. Ketones and the Heart: Metabolic Principles and Therapeutic Implications. Circ. Res. 2023, 132, 882–898. [Google Scholar] [CrossRef]
  109. Evans, M.; Cogan, K.E.; Egan, B. Metabolism of ketone bodies during exercise and training: Physiological basis for exogenous supplementation. J. Physiol. 2017, 595, 2857–2871. [Google Scholar] [CrossRef]
  110. Ho, K.L.; Karwi, Q.G.; Wagg, C.; Zhang, L.; Vo, K.; Altamimi, T.; Uddin, G.M.; Ussher, J.R.; Lopaschuk, G.D. Ketones can become the major fuel source for the heart but do not increase cardiac efficiency. Cardiovasc. Res. 2021, 117, 1178–1187. [Google Scholar] [CrossRef]
  111. McGarrah, R.W.; White, P.J. Branched-chain amino acids in cardiovascular disease. Nat. Rev. Cardiol. 2023, 20, 77–89. [Google Scholar] [CrossRef] [PubMed]
  112. Osellame, L.D.; Blacker, T.S.; Duchen, M.R. Cellular and molecular mechanisms of mitochondrial function. Best Pract. Res. Clin. Endocrinol. Metab. 2012, 26, 711–723. [Google Scholar] [CrossRef]
  113. Sivanand, S.; Viney, I.; Wellen, K.E. Spatiotemporal Control of Acetyl-CoA Metabolism in Chromatin Regulation. Trends Biochem. Sci. 2018, 43, 61–74. [Google Scholar] [CrossRef]
  114. Jaakkola, P.; Mole, D.R.; Tian, Y.M.; Wilson, M.I.; Gielbert, J.; Gaskell, S.J.; von Kriegsheim, A.; Hebestreit, H.F.; Mukherji, M.; Schofield, C.J.; et al. Targeting of HIF-alpha to the von Hippel-Lindau ubiquitylation complex by O2-regulated prolyl hydroxylation. Science 2001, 292, 468–472. [Google Scholar] [CrossRef]
  115. Epstein, A.C.; Gleadle, J.M.; McNeill, L.A.; Hewitson, K.S.; O’Rourke, J.; Mole, D.R.; Mukherji, M.; Metzen, E.; Wilson, M.I.; Dhanda, A.; et al. C. elegans EGL-9 and mammalian homologs define a family of dioxygenases that regulate HIF by prolyl hydroxylation. Cell 2001, 107, 43–54. [Google Scholar] [CrossRef] [PubMed]
  116. Vujic, A.; Koo, A.N.M.; Prag, H.A.; Krieg, T. Mitochondrial redox and TCA cycle metabolite signaling in the heart. Free Radic. Biol. Med. 2021, 166, 287–296. [Google Scholar] [CrossRef] [PubMed]
  117. Jin, Z.; Wei, W.; Yang, M.; Du, Y.; Wan, Y. Mitochondrial complex I activity suppresses inflammation and enhances bone resorption by shifting macrophage-osteoclast polarization. Cell Metab. 2014, 20, 483–498. [Google Scholar] [CrossRef] [PubMed]
  118. Neubauer, S. The failing heart--an engine out of fuel. N. Engl. J. Med. 2007, 356, 1140–1151. [Google Scholar] [CrossRef]
  119. Lai, L.; Leone, T.C.; Keller, M.P.; Martin, O.J.; Broman, A.T.; Nigro, J.; Kapoor, K.; Koves, T.R.; Stevens, R.; Ilkayeva, O.R.; et al. Energy metabolic reprogramming in the hypertrophied and early stage failing heart: A multisystems approach. Circ. Heart Fail. 2014, 7, 1022–1031. [Google Scholar] [CrossRef]
  120. Allard, M.F.; Schönekess, B.O.; Henning, S.L.; English, D.R.; Lopaschuk, G.D. Contribution of oxidative metabolism and glycolysis to ATP production in hypertrophied hearts. Am. J. Physiol. 1994, 267, H742–H750. [Google Scholar] [CrossRef]
  121. Pound, K.M.; Sorokina, N.; Ballal, K.; Berkich, D.A.; Fasano, M.; Lanoue, K.F.; Taegtmeyer, H.; O’Donnell, J.M.; Lewandowski, E.D. Substrate-enzyme competition attenuates upregulated anaplerotic flux through malic enzyme in hypertrophied rat heart and restores triacylglyceride content: Attenuating upregulated anaplerosis in hypertrophy. Circ. Res. 2009, 104, 805–812. [Google Scholar] [CrossRef] [PubMed]
  122. Hahn, V.S.; Petucci, C.; Kim, M.S.; Bedi, K.C., Jr.; Wang, H.; Mishra, S.; Koleini, N.; Yoo, E.J.; Margulies, K.B.; Arany, Z.; et al. Myocardial Metabolomics of Human Heart Failure with Preserved Ejection Fraction. Circulation 2023, 147, 1147–1161. [Google Scholar] [CrossRef]
  123. Tong, D.; Schiattarella, G.G.; Jiang, N.; Altamirano, F.; Szweda, P.A.; Elnwasany, A.; Lee, D.I.; Yoo, H.; Kass, D.A.; Szweda, L.I.; et al. NAD(+) Repletion Reverses Heart Failure with Preserved Ejection Fraction. Circ. Res. 2021, 128, 1629–1641. [Google Scholar] [CrossRef] [PubMed]
  124. O’Sullivan, J.F.; Li, M.; Koay, Y.C.; Wang, X.S.; Guglielmi, G.; Marques, F.Z.; Nanayakkara, S.; Mariani, J.; Slaughter, E.; Kaye, D.M. Cardiac Substrate Utilization and Relationship to Invasive Exercise Hemodynamic Parameters in HFpEF. JACC Basic Transl. Sci. 2024, 9, 281–299. [Google Scholar] [CrossRef] [PubMed]
  125. Sun, Q.; Güven, B.; Wagg, C.S.; Almeida de Oliveira, A.; Silver, H.; Zhang, L.; Chen, B.; Wei, K.; Ketema, E.B.; Karwi, Q.G.; et al. Mitochondrial fatty acid oxidation is the major source of cardiac adenosine triphosphate production in heart failure with preserved ejection fraction. Cardiovasc. Res. 2024, 120, 360–371. [Google Scholar] [CrossRef] [PubMed]
  126. Zordoky, B.N.; Sung, M.M.; Ezekowitz, J.; Mandal, R.; Han, B.; Bjorndahl, T.C.; Bouatra, S.; Anderson, T.; Oudit, G.Y.; Wishart, D.S.; et al. Metabolomic fingerprint of heart failure with preserved ejection fraction. PLoS ONE 2015, 10, e0124844. [Google Scholar] [CrossRef]
  127. Anker, S.D.; Butler, J.; Filippatos, G.; Ferreira, J.P.; Bocchi, E.; Böhm, M.; Brunner-La Rocca, H.P.; Choi, D.J.; Chopra, V.; Chuquiure-Valenzuela, E.; et al. Empagliflozin in Heart Failure with a Preserved Ejection Fraction. N. Engl. J. Med. 2021, 385, 1451–1461. [Google Scholar] [CrossRef]
  128. Nassif, M.E.; Windsor, S.L.; Borlaug, B.A.; Kitzman, D.W.; Shah, S.J.; Tang, F.; Khariton, Y.; Malik, A.O.; Khumri, T.; Umpierrez, G.; et al. The SGLT2 inhibitor dapagliflozin in heart failure with preserved ejection fraction: A multicenter randomized trial. Nat. Med. 2021, 27, 1954–1960. [Google Scholar] [CrossRef]
  129. Koleini, N.; Meddeb, M.; Zhao, L.; Keykhaei, M.; Kwon, S.; Farshidfar, F.; Hahn, V.S.; Pearce, E.L.; Sharma, K.; Kass, D.A. Landscape of glycolytic metabolites and their regulating proteins in myocardium from human heart failure with preserved ejection fraction. Eur. J. Heart Fail. 2024. online ahead of print. [Google Scholar] [CrossRef]
  130. Liu, X.; Zhang, Y.; Deng, Y.; Yang, L.; Ou, W.; Xie, M.; Ding, L.; Jiang, C.; Yu, H.; Li, Q.; et al. Mitochondrial protein hyperacetylation underpins heart failure with preserved ejection fraction in mice. J. Mol. Cell. Cardiol. 2022, 165, 76–85. [Google Scholar] [CrossRef]
  131. Méndez-Fernández, A.; Fernández-Mora, Á.; Bernal-Ramírez, J.; Alves-Figueiredo, H.; Nieblas, B.; Salazar-Ramírez, F.; Maldonado-Ruiz, R.; Zazueta, C.; García, N.; Lozano, O.; et al. Distinguishing pathophysiological features of heart failure with reduced and preserved ejection fraction: A comparative analysis of two mouse models. J. Physiol. 2024. online ahead of print. [Google Scholar] [CrossRef] [PubMed]
  132. Hirschey, M.D.; Shimazu, T.; Goetzman, E.; Jing, E.; Schwer, B.; Lombard, D.B.; Grueter, C.A.; Harris, C.; Biddinger, S.; Ilkayeva, O.R.; et al. SIRT3 regulates mitochondrial fatty-acid oxidation by reversible enzyme deacetylation. Nature 2010, 464, 121–125. [Google Scholar] [CrossRef] [PubMed]
  133. Deng, Y.; Xie, M.; Li, Q.; Xu, X.; Ou, W.; Zhang, Y.; Xiao, H.; Yu, H.; Zheng, Y.; Liang, Y.; et al. Targeting Mitochondria-Inflammation Circuit by β-Hydroxybutyrate Mitigates HFpEF. Circ. Res. 2021, 128, 232–245. [Google Scholar] [CrossRef] [PubMed]
  134. Levitas, A.; Muhammad, E.; Harel, G.; Saada, A.; Caspi, V.C.; Manor, E.; Beck, J.C.; Sheffield, V.; Parvari, R. Familial neonatal isolated cardiomyopathy caused by a mutation in the flavoprotein subunit of succinate dehydrogenase. Eur. J. Hum. Genet. 2010, 18, 1160–1165. [Google Scholar] [CrossRef]
  135. Van Vranken, J.G.; Bricker, D.K.; Dephoure, N.; Gygi, S.P.; Cox, J.E.; Thummel, C.S.; Rutter, J. SDHAF4 promotes mitochondrial succinate dehydrogenase activity and prevents neurodegeneration. Cell Metab. 2014, 20, 241–252. [Google Scholar] [CrossRef]
  136. Wang, X.; Zhang, X.; Cao, K.; Zeng, M.; Fu, X.; Zheng, A.; Zhang, F.; Gao, F.; Zou, X.; Li, H.; et al. Cardiac disruption of SDHAF4-mediated mitochondrial complex II assembly promotes dilated cardiomyopathy. Nat. Commun. 2022, 13, 3947. [Google Scholar] [CrossRef]
  137. Gibb, A.A.; Murray, E.K.; Eaton, D.M.; Huynh, A.T.; Tomar, D.; Garbincius, J.F.; Kolmetzky, D.W.; Berretta, R.M.; Wallner, M.; Houser, S.R.; et al. Molecular Signature of HFpEF: Systems Biology in a Cardiac-Centric Large Animal Model. JACC Basic Transl. Sci. 2021, 6, 650–672. [Google Scholar] [CrossRef]
  138. Miranda-Silva, D.; Wüst, R.C.I.; Conceição, G.; Gonçalves-Rodrigues, P.; Gonçalves, N.; Gonçalves, A.; Kuster, D.W.D.; Leite-Moreira, A.F.; van der Velden, J.; de Sousa Beleza, J.M.; et al. Disturbed cardiac mitochondrial and cytosolic calcium handling in a metabolic risk-related rat model of heart failure with preserved ejection fraction. Acta Physiol (Oxf) 2020, 228, e13378. [Google Scholar] [CrossRef]
  139. Schwartz, B.; Gjini, P.; Gopal, D.M.; Fetterman, J.L. Inefficient Batteries in Heart Failure: Metabolic Bottlenecks Disrupting the Mitochondrial Ecosystem. JACC Basic Transl. Sci. 2022, 7, 1161–1179. [Google Scholar] [CrossRef]
  140. Karamanlidis, G.; Lee, C.F.; Garcia-Menendez, L.; Kolwicz, S.C., Jr.; Suthammarak, W.; Gong, G.; Sedensky, M.M.; Morgan, P.G.; Wang, W.; Tian, R. Mitochondrial complex I deficiency increases protein acetylation and accelerates heart failure. Cell Metab. 2013, 18, 239–250. [Google Scholar] [CrossRef]
  141. Hahn, V.S.; Knutsdottir, H.; Luo, X.; Bedi, K.; Margulies, K.B.; Haldar, S.M.; Stolina, M.; Yin, J.; Khakoo, A.Y.; Vaishnav, J.; et al. Myocardial Gene Expression Signatures in Human Heart Failure with Preserved Ejection Fraction. Circulation 2021, 143, 120–134. [Google Scholar] [CrossRef] [PubMed]
  142. Norata, G.D.; Tavori, H.; Pirillo, A.; Fazio, S.; Catapano, A.L. Biology of proprotein convertase subtilisin kexin 9: Beyond low-density lipoprotein cholesterol lowering. Cardiovasc. Res. 2016, 112, 429–442. [Google Scholar] [CrossRef] [PubMed]
  143. Da Dalt, L.; Castiglioni, L.; Baragetti, A.; Audano, M.; Svecla, M.; Bonacina, F.; Pedretti, S.; Uboldi, P.; Benzoni, P.; Giannetti, F.; et al. PCSK9 deficiency rewires heart metabolism and drives heart failure with preserved ejection fraction. Eur. Heart J. 2021, 42, 3078–3090. [Google Scholar] [CrossRef] [PubMed]
  144. Ni, H.M.; Williams, J.A.; Ding, W.X. Mitochondrial dynamics and mitochondrial quality control. Redox Biol. 2015, 4, 6–13. [Google Scholar] [CrossRef]
  145. Hu, Y.F.; Chen, Y.J.; Lin, Y.J.; Chen, S.A. Inflammation and the pathogenesis of atrial fibrillation. Nat. Rev. Cardiol. 2015, 12, 230–243. [Google Scholar] [CrossRef]
  146. Gisterå, A.; Hansson, G.K. The immunology of atherosclerosis. Nat. Rev. Nephrol. 2017, 13, 368–380. [Google Scholar] [CrossRef]
  147. Suomalainen, A.; Battersby, B.J. Mitochondrial diseases: The contribution of organelle stress responses to pathology. Nat. Rev. Mol. Cell Biol. 2018, 19, 77–92. [Google Scholar] [CrossRef]
  148. Bravo-San Pedro, J.M.; Kroemer, G.; Galluzzi, L. Autophagy and Mitophagy in Cardiovascular Disease. Circ. Res. 2017, 120, 1812–1824. [Google Scholar] [CrossRef]
  149. Ng, M.Y.W.; Wai, T.; Simonsen, A. Quality control of the mitochondrion. Dev. Cell 2021, 56, 881–905. [Google Scholar] [CrossRef]
  150. Bahr, T.; Katuri, J.; Liang, T.; Bai, Y. Mitochondrial chaperones in human health and disease. Free Radic. Biol. Med. 2022, 179, 363–374. [Google Scholar] [CrossRef]
  151. Becker, J.; Craig, E.A. Heat-shock proteins as molecular chaperones. Eur. J. Biochem. 1994, 219, 11–23. [Google Scholar] [CrossRef] [PubMed]
  152. Kwon, Y.T.; Ciechanover, A. The Ubiquitin Code in the Ubiquitin-Proteasome System and Autophagy. Trends Biochem. Sci. 2017, 42, 873–886. [Google Scholar] [CrossRef] [PubMed]
  153. Jornayvaz, F.R.; Shulman, G.I. Regulation of mitochondrial biogenesis. Essays Biochem. 2010, 47, 69–84. [Google Scholar] [CrossRef] [PubMed]
  154. Del Re, D.P.; Amgalan, D.; Linkermann, A.; Liu, Q.; Kitsis, R.N. Fundamental Mechanisms of Regulated Cell Death and Implications for Heart Disease. Physiol. Rev. 2019, 99, 1765–1817. [Google Scholar] [CrossRef] [PubMed]
  155. Gupta, S.; Kass, G.E.; Szegezdi, E.; Joseph, B. The mitochondrial death pathway: A promising therapeutic target in diseases. J. Cell. Mol. Med. 2009, 13, 1004–1033. [Google Scholar] [CrossRef] [PubMed]
  156. Puigserver, P.; Wu, Z.; Park, C.W.; Graves, R.; Wright, M.; Spiegelman, B.M. A cold-inducible coactivator of nuclear receptors linked to adaptive thermogenesis. Cell 1998, 92, 829–839. [Google Scholar] [CrossRef] [PubMed]
  157. Wu, Z.; Puigserver, P.; Andersson, U.; Zhang, C.; Adelmant, G.; Mootha, V.; Troy, A.; Cinti, S.; Lowell, B.; Scarpulla, R.C.; et al. Mechanisms controlling mitochondrial biogenesis and respiration through the thermogenic coactivator PGC-1. Cell 1999, 98, 115–124. [Google Scholar] [CrossRef]
  158. Zong, H.; Ren, J.M.; Young, L.H.; Pypaert, M.; Mu, J.; Birnbaum, M.J.; Shulman, G.I. AMP kinase is required for mitochondrial biogenesis in skeletal muscle in response to chronic energy deprivation. Proc. Natl. Acad. Sci. USA 2002, 99, 15983–15987. [Google Scholar] [CrossRef]
  159. Virbasius, J.V.; Scarpulla, R.C. Activation of the human mitochondrial transcription factor A gene by nuclear respiratory factors: A potential regulatory link between nuclear and mitochondrial gene expression in organelle biogenesis. Proc. Natl. Acad. Sci. USA 1994, 91, 1309–1313. [Google Scholar] [CrossRef]
  160. Das, S.; Bedja, D.; Campbell, N.; Dunkerly, B.; Chenna, V.; Maitra, A.; Steenbergen, C. miR-181c regulates the mitochondrial genome, bioenergetics, and propensity for heart failure in vivo. PLoS ONE 2014, 9, e96820. [Google Scholar] [CrossRef]
  161. Das, S.; Ferlito, M.; Kent, O.A.; Fox-Talbot, K.; Wang, R.; Liu, D.; Raghavachari, N.; Yang, Y.; Wheelan, S.J.; Murphy, E.; et al. Nuclear miRNA regulates the mitochondrial genome in the heart. Circ. Res. 2012, 110, 1596–1603. [Google Scholar] [CrossRef] [PubMed]
  162. Dogan, A.E.; Hamid, S.M.; Yildirim, A.D.; Yildirim, Z.; Sen, G.; Riera, C.E.; Gottlieb, R.A.; Erbay, E. PACT establishes a posttranscriptional brake on mitochondrial biogenesis by promoting the maturation of miR-181c. J. Biol. Chem. 2022, 298, 102050. [Google Scholar] [CrossRef] [PubMed]
  163. Fang, W.J.; Wang, C.J.; He, Y.; Zhou, Y.L.; Peng, X.D.; Liu, S.K. Resveratrol alleviates diabetic cardiomyopathy in rats by improving mitochondrial function through PGC-1α deacetylation. Acta Pharmacol. Sin. 2018, 39, 59–73. [Google Scholar] [CrossRef] [PubMed]
  164. Haemmerle, G.; Moustafa, T.; Woelkart, G.; Büttner, S.; Schmidt, A.; van de Weijer, T.; Hesselink, M.; Jaeger, D.; Kienesberger, P.C.; Zierler, K.; et al. ATGL-mediated fat catabolism regulates cardiac mitochondrial function via PPAR-α and PGC-1. Nat. Med. 2011, 17, 1076–1085. [Google Scholar] [CrossRef]
  165. Martin, O.J.; Lai, L.; Soundarapandian, M.M.; Leone, T.C.; Zorzano, A.; Keller, M.P.; Attie, A.D.; Muoio, D.M.; Kelly, D.P. A role for peroxisome proliferator-activated receptor γ coactivator-1 in the control of mitochondrial dynamics during postnatal cardiac growth. Circ. Res. 2014, 114, 626–636. [Google Scholar] [CrossRef]
  166. Lai, L.; Leone, T.C.; Zechner, C.; Schaeffer, P.J.; Kelly, S.M.; Flanagan, D.P.; Medeiros, D.M.; Kovacs, A.; Kelly, D.P. Transcriptional coactivators PGC-1alpha and PGC-lbeta control overlapping programs required for perinatal maturation of the heart. Genes. Dev. 2008, 22, 1948–1961. [Google Scholar] [CrossRef]
  167. Sihag, S.; Cresci, S.; Li, A.Y.; Sucharov, C.C.; Lehman, J.J. PGC-1α and ERRα target gene downregulation is a signature of the failing human heart. J. Mol. Cell. Cardiol. 2009, 46, 201–212. [Google Scholar] [CrossRef]
  168. Sun, C.-K.; Chang, L.-T.; Sheu, J.-J.; Wang, C.-Y.; Youssef, A.A.; Wu, C.-J.; Chua, S.; Yip, H.-K. Losartan Preserves Integrity of Cardiac Gap Junctions and PGC-1 alpha Gene Expression and Prevents Cellular Apoptosis in Remote Area of Left Ventricular Myocardium Following Acute Myocardial Infarction. Int. Heart J. 2007, 48, 533–546. [Google Scholar] [CrossRef]
  169. Watson, P.A.; Reusch, J.E.B.; McCune, S.A.; Leinwand, L.A.; Luckey, S.W.; Konhilas, J.P.; Brown, D.A.; Chicco, A.J.; Sparagna, G.C.; Long, C.S.; et al. Restoration of CREB function is linked to completion and stabilization of adaptive cardiac hypertrophy in response to exercise. Am. J. Physiol.-Heart Circ. Physiol. 2007, 293, H246–H259. [Google Scholar] [CrossRef]
  170. Pereira, R.O.; Wende, A.R.; Crum, A.; Hunter, D.; Olsen, C.D.; Rawlings, T.; Riehle, C.; Ward, W.F.; Abel, E.D. Maintaining PGC-1α expression following pressure overload-induced cardiac hypertrophy preserves angiogenesis but not contractile or mitochondrial function. FASEB J. 2014, 28, 3691–3702. [Google Scholar] [CrossRef]
  171. Russell, L.K.; Mansfield, C.M.; Lehman, J.J.; Kovacs, A.; Courtois, M.; Saffitz, J.E.; Medeiros, D.M.; Valencik, M.L.; McDonald, J.A.; Kelly, D.P. Cardiac-specific induction of the transcriptional coactivator peroxisome proliferator-activated receptor gamma coactivator-1alpha promotes mitochondrial biogenesis and reversible cardiomyopathy in a developmental stage-dependent manner. Circ. Res. 2004, 94, 525–533. [Google Scholar] [CrossRef] [PubMed]
  172. Hu, X.; Xu, X.; Lu, Z.; Zhang, P.; Fassett, J.; Zhang, Y.; Xin, Y.; Hall, J.L.; Viollet, B.; Bache, R.J.; et al. AMP Activated Protein Kinase-α2 Regulates Expression of Estrogen-Related Receptor-α, a Metabolic Transcription Factor Related to Heart Failure Development. Hypertension 2011, 58, 696–703. [Google Scholar] [CrossRef] [PubMed]
  173. Chan, D.C. Mitochondrial Dynamics and Its Involvement in Disease. Annu. Rev. Pathol. 2020, 15, 235–259. [Google Scholar] [CrossRef] [PubMed]
  174. Parra, V.; Verdejo, H.; del Campo, A.; Pennanen, C.; Kuzmicic, J.; Iglewski, M.; Hill, J.A.; Rothermel, B.A.; Lavandero, S. The complex interplay between mitochondrial dynamics and cardiac metabolism. J. Bioenerg. Biomembr. 2011, 43, 47–51. [Google Scholar] [CrossRef] [PubMed]
  175. Atici, A.E.; Crother, T.R.; Noval Rivas, M. Mitochondrial quality control in health and cardiovascular diseases. Front. Cell. Dev. Biol. 2023, 11, 1290046. [Google Scholar] [CrossRef]
  176. Paradies, G.; Paradies, V.; Ruggiero, F.M.; Petrosillo, G. Role of Cardiolipin in Mitochondrial Function and Dynamics in Health and Disease: Molecular and Pharmacological Aspects. Cells 2019, 8, 728. [Google Scholar] [CrossRef]
  177. Olichon, A.; ElAchouri, G.; Baricault, L.; Delettre, C.; Belenguer, P.; Lenaers, G. OPA1 alternate splicing uncouples an evolutionary conserved function in mitochondrial fusion from a vertebrate restricted function in apoptosis. Cell Death Differ. 2007, 14, 682–692. [Google Scholar] [CrossRef]
  178. Chen, H.; Detmer, S.A.; Ewald, A.J.; Griffin, E.E.; Fraser, S.E.; Chan, D.C. Mitofusins Mfn1 and Mfn2 coordinately regulate mitochondrial fusion and are essential for embryonic development. J. Cell Biol. 2003, 160, 189–200. [Google Scholar] [CrossRef]
  179. Song, Z.; Chen, H.; Fiket, M.; Alexander, C.; Chan, D.C. OPA1 processing controls mitochondrial fusion and is regulated by mRNA splicing, membrane potential, and Yme1L. J. Cell Biol. 2007, 178, 749–755. [Google Scholar] [CrossRef]
  180. Elgass, K.; Pakay, J.; Ryan, M.T.; Palmer, C.S. Recent advances into the understanding of mitochondrial fission. Biochim. Biophys. Acta (BBA)—Mol. Cell Res. 2013, 1833, 150–161. [Google Scholar] [CrossRef]
  181. Ong, S.-B.; Hausenloy, D.J. Mitochondrial morphology and cardiovascular disease. Cardiovasc. Res. 2010, 88, 16–29. [Google Scholar] [CrossRef] [PubMed]
  182. Atkins, K.; Dasgupta, A.; Chen, K.H.; Mewburn, J.; Archer, S.L. The role of Drp1 adaptor proteins MiD49 and MiD51 in mitochondrial fission: Implications for human disease. Clin. Sci. 2016, 130, 1861–1874. [Google Scholar] [CrossRef] [PubMed]
  183. Otera, H.; Ishihara, N.; Mihara, K. New insights into the function and regulation of mitochondrial fission. Biochim. Biophys. Acta (BBA)—Mol. Cell Res. 2013, 1833, 1256–1268. [Google Scholar] [CrossRef] [PubMed]
  184. Friedman, J.R.; Lackner, L.L.; West, M.; DiBenedetto, J.R.; Nunnari, J.; Voeltz, G.K. ER Tubules Mark Sites of Mitochondrial Division. Science 2011, 334, 358–362. [Google Scholar] [CrossRef]
  185. Quiles, J.M.; Gustafsson, Å.B. The role of mitochondrial fission in cardiovascular health and disease. Nat. Rev. Cardiol. 2022, 19, 723–736. [Google Scholar] [CrossRef]
  186. Zaffagnini, G.; Martens, S. Mechanisms of Selective Autophagy. J. Mol. Biol. 2016, 428, 1714–1724. [Google Scholar] [CrossRef]
  187. Song, M.; Mihara, K.; Chen, Y.; Scorrano, L.; Dorn, G.W., II. Mitochondrial Fission and Fusion Factors Reciprocally Orchestrate Mitophagic Culling in Mouse Hearts and Cultured Fibroblasts. Cell Metab. 2015, 21, 273–286. [Google Scholar] [CrossRef]
  188. Fang, L.; Moore, X.-L.; Gao, X.-M.; Dart, A.M.; Lim, Y.L.; Du, X.-J. Down-regulation of mitofusin-2 expression in cardiac hypertrophy in vitro and in vivo. Life Sci. 2007, 80, 2154–2160. [Google Scholar] [CrossRef]
  189. Yu, H.; Guo, Y.; Mi, L.; Wang, X.; Li, L.; Gao, W. Mitofusin 2 Inhibits Angiotensin II-Induced Myocardial Hypertrophy. J. Cardiovasc. Pharmacol. Ther. 2011, 16, 205–211. [Google Scholar] [CrossRef]
  190. Sun, D.; Li, C.; Liu, J.; Wang, Z.; Liu, Y.; Luo, C.; Chen, Y.; Wen, S. Expression Profile of microRNAs in Hypertrophic Cardiomyopathy and Effects of microRNA-20 in Inducing Cardiomyocyte Hypertrophy Through Regulating Gene MFN2. DNA Cell Biol. 2019, 38, 796–807. [Google Scholar] [CrossRef]
  191. Guo, Y.; Wang, Z.; Qin, X.; Xu, J.; Hou, Z.; Yang, H.; Mao, X.; Xing, W.; Li, X.; Zhang, X.; et al. Enhancing fatty acid utilization ameliorates mitochondrial fragmentation and cardiac dysfunction via rebalancing optic atrophy 1 processing in the failing heart. Cardiovasc. Res. 2018, 114, 979–991. [Google Scholar] [CrossRef] [PubMed]
  192. Song, M.; Franco, A.; Fleischer, J.A.; Zhang, L.; Dorn, G.W., II. Abrogating Mitochondrial Dynamics in Mouse Hearts Accelerates Mitochondrial Senescence. Cell Metab. 2017, 26, 872–883. [Google Scholar] [CrossRef] [PubMed]
  193. Molina, A.J.A.; Bharadwaj, M.S.; Van Horn, C.; Nicklas, B.J.; Lyles, M.F.; Eggebeen, J.; Haykowsky, M.J.; Brubaker, P.H.; Kitzman, D.W. Skeletal Muscle Mitochondrial Content, Oxidative Capacity, and Mfn2 Expression Are Reduced in Older Patients with Heart Failure and Preserved Ejection Fraction and Are Related to Exercise Intolerance. JACC Heart Fail. 2016, 4, 636–645. [Google Scholar] [CrossRef] [PubMed]
  194. Bode, D.; Wen, Y.; Hegemann, N.; Primessnig, U.; Parwani, A.; Boldt, L.H.; MPieske, B.; RHeinzel, F.; Hohendanner, F. Oxidative Stress and Inflammatory Modulation of Ca2+ Handling in Metabolic HFpEF-Related Left Atrial Cardiomyopathy. Antioxidants 2020, 9, 860. [Google Scholar] [CrossRef] [PubMed]
  195. Zhang, W.; Zhang, H.; Yao, W.; Li, L.; Niu, P.; Huo, Y.; Tan, W. Morphometric, Hemodynamic, and Multi-Omics Analyses in Heart Failure Rats with Preserved Ejection Fraction. Int. J. Mol. Sci. 2020, 21, 3362. [Google Scholar] [CrossRef]
  196. McWilliams, T.G.; Prescott, A.R.; Montava-Garriga, L.; Ball, G.; Singh, F.; Barini, E.; Muqit, M.M.K.; Brooks, S.P.; Ganley, I.G. Basal Mitophagy Occurs Independently of PINK1 in Mouse Tissues of High Metabolic Demand. Cell Metab. 2018, 27, 439–449. [Google Scholar] [CrossRef]
  197. Tan, Y.; Li, M.; Wu, G.; Lou, J.; Feng, M.; Xu, J.; Zhou, J.; Zhang, P.; Yang, H.; Dong, L.; et al. Short-term but not long-term high fat diet feeding protects against pressure overload-induced heart failure through activation of mitophagy. Life Sci. 2021, 272, 119242. [Google Scholar] [CrossRef]
  198. Tong, M.; Saito, T.; Zhai, P.; Oka, S.-I.; Mizushima, W.; Nakamura, M.; Ikeda, S.; Shirakabe, A.; Sadoshima, J. Mitophagy Is Essential for Maintaining Cardiac Function During High Fat Diet-Induced Diabetic Cardiomyopathy. Circ. Res. 2019, 124, 1360–1371. [Google Scholar] [CrossRef]
  199. Liu, Y.; Wang, Y.; Bi, Y.; Zhao, Z.; Wang, S.; Lin, S.; Yang, Z.; Wang, X.; Mao, J. Emerging role of mitophagy in heart failure: From molecular mechanism to targeted therapy. Cell Cycle 2023, 22, 906–918. [Google Scholar] [CrossRef]
  200. Ajoolabady, A.; Chiong, M.; Lavandero, S.; Klionsky, D.J.; Ren, J. Mitophagy in cardiovascular diseases: Molecular mechanisms, pathogenesis, and treatment. Trends Mol. Med. 2022, 28, 836–849. [Google Scholar] [CrossRef]
  201. Sanhueza-Olivares, F.; Troncoso, M.F.; Pino-De La Fuente, F.; Martinez-Bilbao, J.; Riquelme, J.A.; Norambuena-Soto, I.; Villa, M.; Lavandero, S.; Castro, P.F.; Chiong, M. A potential role of autophagy-mediated vascular senescence in the pathophysiology of HFpEF. Front. Endocrinol. 2022, 13, 1057349. [Google Scholar] [CrossRef] [PubMed]
  202. Titus, A.S.; Sung, E.A.; Zablocki, D.; Sadoshima, J. Mitophagy for cardioprotection. Basic Res. Cardiol. 2023, 118, 42. [Google Scholar] [CrossRef] [PubMed]
  203. Rüb, C.; Wilkening, A.; Voos, W. Mitochondrial quality control by the Pink1/Parkin system. Cell Tissue Res. 2017, 367, 111–123. [Google Scholar] [CrossRef] [PubMed]
  204. Yun, J.; Finkel, T. Mitohormesis. Cell Metab. 2014, 19, 757–766. [Google Scholar] [CrossRef]
  205. Twig, G.; Shirihai, O.S. The Interplay Between Mitochondrial Dynamics and Mitophagy. Antioxid. Redox Signal. 2010, 14, 1939–1951. [Google Scholar] [CrossRef]
  206. Kubli, D.A.; Gustafsson, Å.B. Mitochondria and Mitophagy. Circ. Res. 2012, 111, 1208–1221. [Google Scholar] [CrossRef]
  207. Morales, P.E.; Arias-Durán, C.; Ávalos-Guajardo, Y.; Aedo, G.; Verdejo, H.E.; Parra, V.; Lavandero, S. Emerging role of mitophagy in cardiovascular physiology and pathology. Mol. Asp. Med. 2020, 71, 100822. [Google Scholar] [CrossRef]
  208. Mozaffarian, D.; Benjamin, E.J.; Go, A.S.; Arnett, D.K.; Blaha, M.J.; Cushman, M.; Das, S.R.; de Ferranti, S.; Després, J.P.; Fullerton, H.J.; et al. Heart Disease and Stroke Statistics-2016 Update: A Report From the American Heart Association. Circulation 2016, 133, e38–e360. [Google Scholar] [CrossRef]
  209. Kassiotis, C.; Ballal, K.; Wellnitz, K.; Vela, D.; Gong, M.; Salazar, R.; Frazier, O.H.; Taegtmeyer, H. Markers of autophagy are downregulated in failing human heart after mechanical unloading. Circulation 2009, 120, S191–S197. [Google Scholar] [CrossRef]
  210. Billia, F.; Hauck, L.; Konecny, F.; Rao, V.; Shen, J.; Mak, T.W. PTEN-inducible kinase 1 (PINK1)/Park6 is indispensable for normal heart function. Proc. Natl. Acad. Sci. USA 2011, 108, 9572–9577. [Google Scholar] [CrossRef]
  211. Wang, B.; Nie, J.; Wu, L.; Hu, Y.; Wen, Z.; Dong, L.; Zou, M.H.; Chen, C.; Wang, D.W. AMPKα2 Protects Against the Development of Heart Failure by Enhancing Mitophagy via PINK1 Phosphorylation. Circ. Res. 2018, 122, 712–729. [Google Scholar] [CrossRef] [PubMed]
  212. Summer, G.; Kuhn, A.R.; Munts, C.; Miranda-Silva, D.; Leite-Moreira, A.F.; Lourenço, A.P.; Heymans, S.; Falcão-Pires, I.; van Bilsen, M. A directed network analysis of the cardiome identifies molecular pathways contributing to the development of HFpEF. J. Mol. Cell. Cardiol. 2020, 144, 66–75. [Google Scholar] [CrossRef] [PubMed]
  213. Roh, J.D.; Houstis, N.; Yu, A.; Chang, B.; Yeri, A.; Li, H.; Hobson, R.; Lerchenmüller, C.; Vujic, A.; Chaudhari, V.; et al. Exercise training reverses cardiac aging phenotypes associated with heart failure with preserved ejection fraction in male mice. Aging Cell 2020, 19, e13159. [Google Scholar] [CrossRef]
  214. Shinmura, K.; Tamaki, K.; Sano, M.; Murata, M.; Yamakawa, H.; Ishida, H.; Fukuda, K. Impact of long-term caloric restriction on cardiac senescence: Caloric restriction ameliorates cardiac diastolic dysfunction associated with aging. J. Mol. Cell. Cardiol. 2011, 50, 117–127. [Google Scholar] [CrossRef] [PubMed]
  215. Hoshino, A.; Mita, Y.; Okawa, Y.; Ariyoshi, M.; Iwai-Kanai, E.; Ueyama, T.; Ikeda, K.; Ogata, T.; Matoba, S. Cytosolic p53 inhibits Parkin-mediated mitophagy and promotes mitochondrial dysfunction in the mouse heart. Nat. Commun. 2013, 4, 2308. [Google Scholar] [CrossRef] [PubMed]
  216. Eisenberg, T.; Abdellatif, M.; Schroeder, S.; Primessnig, U.; Stekovic, S.; Pendl, T.; Harger, A.; Schipke, J.; Zimmermann, A.; Schmidt, A.; et al. Cardioprotection and lifespan extension by the natural polyamine spermidine. Nat. Med. 2016, 22, 1428–1438. [Google Scholar] [CrossRef]
  217. Bou-Teen, D.; Kaludercic, N.; Weissman, D.; Turan, B.; Maack, C.; Di Lisa, F.; Ruiz-Meana, M. Mitochondrial ROS and mitochondria-targeted antioxidants in the aged heart. Free Radic. Biol. Med. 2021, 167, 109–124. [Google Scholar] [CrossRef]
  218. Ali, M.A.; Gioscia-Ryan, R.; Yang, D.; Sutton, N.R.; Tyrrell, D.J. Cardiovascular aging: Spotlight on mitochondria. Am. J. Physiol. Heart Circ. Physiol. 2024, 326, H317–H333. [Google Scholar] [CrossRef]
  219. Aon, M.A.; Cortassa, S.; O’Rourke, B. Redox-optimized ROS balance: A unifying hypothesis. Biochim. Biophys. Acta (BBA)—Bioenerg. 2010, 1797, 865–877. [Google Scholar] [CrossRef]
  220. Müller, M.; Donhauser, E.; Maske, T.; Bischof, C.; Dumitrescu, D.; Rudolph, V.; Klinke, A. Mitochondrial Integrity Is Critical in Right Heart Failure Development. Int. J. Mol. Sci. 2023, 24, 11108. [Google Scholar] [CrossRef]
  221. Kowaltowski, A.J.; de Souza-Pinto, N.C.; Castilho, R.F.; Vercesi, A.E. Mitochondria and reactive oxygen species. Free Radic. Biol. Med. 2009, 47, 333–343. [Google Scholar] [CrossRef] [PubMed]
  222. Figueira, T.R.; Barros, M.H.; Camargo, A.A.; Castilho, R.F.; Ferreira, J.C.B.; Kowaltowski, A.J.; Sluse, F.E.; Souza-Pinto, N.C.; Vercesi, A.E. Mitochondria as a Source of Reactive Oxygen and Nitrogen Species: From Molecular Mechanisms to Human Health. Antioxid. Redox Signal. 2012, 18, 2029–2074. [Google Scholar] [CrossRef] [PubMed]
  223. Zorov, D.B.; Juhaszova, M.; Sollott, S.J. Mitochondrial Reactive Oxygen Species (ROS) and ROS-Induced ROS Release. Physiol. Rev. 2014, 94, 909–950. [Google Scholar] [CrossRef] [PubMed]
  224. Zhang, Y.; Bharathi, S.S.; Beck, M.E.; Goetzman, E.S. The fatty acid oxidation enzyme long-chain acyl-CoA dehydrogenase can be a source of mitochondrial hydrogen peroxide. Redox Biol. 2019, 26, 101253. [Google Scholar] [CrossRef] [PubMed]
  225. Cardoso, A.R.; Kakimoto, P.A.H.B.; Kowaltowski, A.J. Diet-Sensitive Sources of Reactive Oxygen Species in Liver Mitochondria: Role of Very Long Chain Acyl-CoA Dehydrogenases. PLoS ONE 2013, 8, e77088. [Google Scholar] [CrossRef]
  226. Kaludercic, N.; Carpi, A.; Menabò, R.; Di Lisa, F.; Paolocci, N. Monoamine oxidases (MAO) in the pathogenesis of heart failure and ischemia/reperfusion injury. Biochim. Biophys. Acta (BBA)—Mol. Cell Res. 2011, 1813, 1323–1332. [Google Scholar] [CrossRef]
  227. Bedard, K.; Krause, K.H. The NOX family of ROS-generating NADPH oxidases: Physiology and pathophysiology. Physiol. Rev. 2007, 87, 245–313. [Google Scholar] [CrossRef]
  228. Chen, Z.; Jin, Z.X.; Cai, J.; Li, R.; Deng, K.Q.; Ji, Y.X.; Lei, F.; Li, H.P.; Lu, Z.; Li, H. Energy substrate metabolism and oxidative stress in metabolic cardiomyopathy. J. Mol. Med. 2022, 100, 1721–1739. [Google Scholar] [CrossRef]
  229. Campbell, E.L.; Colgan, S.P. Control and dysregulation of redox signalling in the gastrointestinal tract. Nat. Rev. Gastroenterol. Hepatol. 2019, 16, 106–120. [Google Scholar] [CrossRef]
  230. Boudina, S.; Sena, S.; O’Neill, B.T.; Tathireddy, P.; Young, M.E.; Abel, E.D. Reduced Mitochondrial Oxidative Capacity and Increased Mitochondrial Uncoupling Impair Myocardial Energetics in Obesity. Circulation 2005, 112, 2686–2695. [Google Scholar] [CrossRef]
  231. Boudina, S.; Bugger, H.; Sena, S.; O’Neill, B.T.; Zaha, V.G.; Ilkun, O.; Wright, J.J.; Mazumder, P.K.; Palfreyman, E.; Tidwell, T.J.; et al. Contribution of Impaired Myocardial Insulin Signaling to Mitochondrial Dysfunction and Oxidative Stress in the Heart. Circulation 2009, 119, 1272–1283. [Google Scholar] [CrossRef] [PubMed]
  232. Simões, I.C.M.; Fontes, A.; Pinton, P.; Zischka, H.; Wieckowski, M.R. Mitochondria in non-alcoholic fatty liver disease. Int. J. Biochem. Cell Biol. 2018, 95, 93–99. [Google Scholar] [CrossRef] [PubMed]
  233. Muftuoglu, M.; Mori, M.P.; Souza-Pinto, N.C.d. Formation and repair of oxidative damage in the mitochondrial DNA. Mitochondrion 2014, 17, 164–181. [Google Scholar] [CrossRef] [PubMed]
  234. Kawahara, H.; Fukura, M.; Tsuchishima, M.; Takase, S. Mutation of Mitochondrial DNA in Livers From Patients with Alcoholic Hepatitis and Nonalcoholic Steatohepatitis. Alcohol. Clin. Exp. Res. 2007, 31, S54–S60. [Google Scholar] [CrossRef] [PubMed]
  235. Pohjoismäki, J.L.O.; Goffart, S.; Tyynismaa, H.; Willcox, S.; Ide, T.; Kang, D.; Suomalainen, A.; Karhunen, P.J.; Griffith, J.D.; Holt, I.J.; et al. Human Heart Mitochondrial DNA Is Organized in Complex Catenated Networks Containing Abundant Four-way Junctions and Replication Forks. J. Biol. Chem. 2009, 284, 21446–21457. [Google Scholar] [CrossRef]
  236. Frahm, T.; Mohamed, S.A.; Bruse, P.; Gemünd, C.; Oehmichen, M.; Meissner, C. Lack of age-related increase of mitochondrial DNA amount in brain, skeletal muscle and human heart. Mech. Ageing Dev. 2005, 126, 1192–1200. [Google Scholar] [CrossRef]
  237. Wang, J.; Wilhelmsson, H.; Graff, C.; Li, H.; Oldfors, A.; Rustin, P.; Brüning, J.C.; Kahn, C.R.; Clayton, D.A.; Barsh, G.S.; et al. Dilated cardiomyopathy and atrioventricular conduction blocks induced by heart-specific inactivation of mitochondrial DNA gene expression. Nat. Genet. 1999, 21, 133–137. [Google Scholar] [CrossRef]
  238. Li, H.; Wang, J.; Wilhelmsson, H.; Hansson, A.; Thorén, P.; Duffy, J.; Rustin, P.; Larsson, N.-G. Genetic modification of survival in tissue-specific knockout mice with mitochondrial cardiomyopathy. Proc. Natl. Acad. Sci. USA 2000, 97, 3467–3472. [Google Scholar] [CrossRef]
  239. Marin-Garcia, J.; Goldenthal, M.J.; Moe, G.W. Mitochondrial pathology in cardiac failure. Cardiovasc. Res. 2001, 49, 17–26. [Google Scholar] [CrossRef]
  240. Elorza, A.A.; Soffia, J.P. mtDNA Heteroplasmy at the Core of Aging-Associated Heart Failure. An Integrative View of OXPHOS and Mitochondrial Life Cycle in Cardiac Mitochondrial Physiology. Front. Cell. Dev. Biol. 2021, 9, 625020. [Google Scholar] [CrossRef]
  241. Lakshmanan, L.N.; Yee, Z.; Ng, L.F.; Gunawan, R.; Halliwell, B.; Gruber, J. Clonal expansion of mitochondrial DNA deletions is a private mechanism of aging in long-lived animals. Aging Cell 2018, 17, e12814. [Google Scholar] [CrossRef] [PubMed]
  242. Tuppen, H.A.L.; Blakely, E.L.; Turnbull, D.M.; Taylor, R.W. Mitochondrial DNA mutations and human disease. Biochim. Biophys. Acta (BBA)—Bioenerg. 2010, 1797, 113–128. [Google Scholar] [CrossRef] [PubMed]
  243. Li, H.; Slone, J.; Fei, L.; Huang, T. Mitochondrial DNA Variants and Common Diseases: A Mathematical Model for the Diversity of Age-Related mtDNA Mutations. Cells 2019, 8, 608. [Google Scholar] [CrossRef] [PubMed]
  244. Li, M.; Schönberg, A.; Schaefer, M.; Schroeder, R.; Nasidze, I.; Stoneking, M. Detecting Heteroplasmy from High-Throughput Sequencing of Complete Human Mitochondrial DNA Genomes. Am. J. Hum. Genet. 2010, 87, 237–249. [Google Scholar] [CrossRef] [PubMed]
  245. Martinelli, I.; Tomassoni, D.; Moruzzi, M.; Roy, P.; Cifani, C.; Amenta, F.; Tayebati, S.K. Cardiovascular Changes Related to Metabolic Syndrome: Evidence in Obese Zucker Rats. Int. J. Mol. Sci. 2020, 21, 2035. [Google Scholar] [CrossRef]
  246. Chen, Z.; Tian, R.; She, Z.; Cai, J.; Li, H. Role of oxidative stress in the pathogenesis of nonalcoholic fatty liver disease. Free Radic. Biol. Med. 2020, 152, 116–141. [Google Scholar] [CrossRef]
  247. Li, J.; Romestaing, C.; Han, X.; Li, Y.; Hao, X.; Wu, Y.; Sun, C.; Liu, X.; Jefferson, L.S.; Xiong, J.; et al. Cardiolipin Remodeling by ALCAT1 Links Oxidative Stress and Mitochondrial Dysfunction to Obesity. Cell Metab. 2010, 12, 154–165. [Google Scholar] [CrossRef]
  248. 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]
  249. Tappia, S.P.; Adameova, A.; Dhalla, S.N. Attenuation of Diabetes-induced Cardiac and Subcellular Defects by Sulphur-containing Amino Acids. Curr. Med. Chem. 2018, 25, 336–345. [Google Scholar] [CrossRef]
  250. Wilson, A.J.; Gill, E.K.; Abudalo, R.A.; Edgar, K.S.; Watson, C.J.; Grieve, D.J. Reactive oxygen species signalling in the diabetic heart: Emerging prospect for therapeutic targeting. Heart 2018, 104, 293–299. [Google Scholar] [CrossRef]
  251. Palma, F.R.; Gantner, B.N.; Sakiyama, M.J.; Kayzuka, C.; Shukla, S.; Lacchini, R.; Cunniff, B.; Bonini, M.G. ROS production by mitochondria: Function or dysfunction? Oncogene 2024, 43, 295–303. [Google Scholar] [CrossRef]
  252. Schulz, T.J.; Zarse, K.; Voigt, A.; Urban, N.; Birringer, M.; Ristow, M. Glucose restriction extends Caenorhabditis elegans life span by inducing mitochondrial respiration and increasing oxidative stress. Cell Metab. 2007, 6, 280–293. [Google Scholar] [CrossRef]
  253. Yang, W.; Hekimi, S. A mitochondrial superoxide signal triggers increased longevity in Caenorhabditis elegans. PLoS Biol. 2010, 8, e1000556. [Google Scholar] [CrossRef]
  254. Owusu-Ansah, E.; Song, W.; Perrimon, N. Muscle mitohormesis promotes longevity via systemic repression of insulin signaling. Cell 2013, 155, 699–712. [Google Scholar] [CrossRef]
  255. Pan, Y.; Schroeder, E.A.; Ocampo, A.; Barrientos, A.; Shadel, G.S. Regulation of yeast chronological life span by TORC1 via adaptive mitochondrial ROS signaling. Cell Metab. 2011, 13, 668–678. [Google Scholar] [CrossRef]
  256. Zarse, K.; Schmeisser, S.; Groth, M.; Priebe, S.; Beuster, G.; Kuhlow, D.; Guthke, R.; Platzer, M.; Kahn, C.R.; Ristow, M. Impaired insulin/IGF1 signaling extends life span by promoting mitochondrial L-proline catabolism to induce a transient ROS signal. Cell Metab. 2012, 15, 451–465. [Google Scholar] [CrossRef]
  257. Lee, S.J.; Hwang, A.B.; Kenyon, C. Inhibition of respiration extends C. elegans life span via reactive oxygen species that increase HIF-1 activity. Curr. Biol. 2010, 20, 2131–2136. [Google Scholar] [CrossRef]
  258. Xie, M.; Roy, R. Increased levels of hydrogen peroxide induce a HIF-1-dependent modification of lipid metabolism in AMPK compromised C. elegans dauer larvae. Cell Metab. 2012, 16, 322–335. [Google Scholar] [CrossRef]
  259. Otani, H. Reactive Oxygen Species as Mediators of Signal Transduction in Ischemic Preconditioning. Antioxid. Redox Signal. 2004, 6, 449–469. [Google Scholar] [CrossRef]
  260. Antonucci, S.; Mulvey, J.F.; Burger, N.; Di Sante, M.; Hall, A.R.; Hinchy, E.C.; Caldwell, S.T.; Gruszczyk, A.V.; Deshwal, S.; Hartley, R.C.; et al. Selective mitochondrial superoxide generation in vivo is cardioprotective through hormesis. Free Radic. Biol. Med. 2019, 134, 678–687. [Google Scholar] [CrossRef]
  261. Finkel, T. Signal transduction by mitochondrial oxidants. J. Biol. Chem. 2012, 287, 4434–4440. [Google Scholar] [CrossRef]
  262. Scherz-Shouval, R.; Shvets, E.; Fass, E.; Shorer, H.; Gil, L.; Elazar, Z. Reactive oxygen species are essential for autophagy and specifically regulate the activity of Atg4. EMBO J. 2019, 38, e101812. [Google Scholar] [CrossRef]
  263. Chandel, N.S.; Maltepe, E.; Goldwasser, E.; Mathieu, C.E.; Simon, M.C.; Schumacker, P.T. Mitochondrial reactive oxygen species trigger hypoxia-induced transcription. Proc. Natl. Acad. Sci. USA 1998, 95, 11715–11720. [Google Scholar] [CrossRef]
  264. Kamata, H.; Honda, S.; Maeda, S.; Chang, L.; Hirata, H.; Karin, M. Reactive oxygen species promote TNFalpha-induced death and sustained JNK activation by inhibiting MAP kinase phosphatases. Cell 2005, 120, 649–661. [Google Scholar] [CrossRef]
  265. Ma, Q. Role of nrf2 in oxidative stress and toxicity. Annu. Rev. Pharmacol. Toxicol. 2013, 53, 401–426. [Google Scholar] [CrossRef]
  266. Paek, J.; Lo, J.Y.; Narasimhan, S.D.; Nguyen, T.N.; Glover-Cutter, K.; Robida-Stubbs, S.; Suzuki, T.; Yamamoto, M.; Blackwell, T.K.; Curran, S.P. Mitochondrial SKN-1/Nrf mediates a conserved starvation response. Cell Metab. 2012, 16, 526–537. [Google Scholar] [CrossRef]
  267. Bjelakovic, G.; Nikolova, D.; Gluud, L.L.; Simonetti, R.G.; Gluud, C. Mortality in randomized trials of antioxidant supplements for primary and secondary prevention: Systematic review and meta-analysis. JAMA 2007, 297, 842–857. [Google Scholar] [CrossRef]
  268. Ramamoorthy, V.; Rubens, M.; Saxena, A.; Shehadeh, N. Selenium and vitamin E for prostate cancer--justifications for the SELECT study. Asian Pac. J. Cancer Prev. 2015, 16, 2619–2627. [Google Scholar] [CrossRef]
  269. Lonn, E.; Bosch, J.; Yusuf, S.; Sheridan, P.; Pogue, J.; Arnold, J.M.; Ross, C.; Arnold, A.; Sleight, P.; Probstfield, J.; et al. Effects of long-term vitamin E supplementation on cardiovascular events and cancer: A randomized controlled trial. JAMA 2005, 293, 1338–1347. [Google Scholar] [CrossRef]
  270. Gomez-Cabrera, M.C.; Domenech, E.; Romagnoli, M.; Arduini, A.; Borras, C.; Pallardo, F.V.; Sastre, J.; Viña, J. Oral administration of vitamin C decreases muscle mitochondrial biogenesis and hampers training-induced adaptations in endurance performance. Am. J. Clin. Nutr. 2008, 87, 142–149. [Google Scholar] [CrossRef]
  271. Ristow, M.; Zarse, K.; Oberbach, A.; Klöting, N.; Birringer, M.; Kiehntopf, M.; Stumvoll, M.; Kahn, C.R.; Blüher, M. Antioxidants prevent health-promoting effects of physical exercise in humans. Proc. Natl. Acad. Sci. USA 2009, 106, 8665–8670. [Google Scholar] [CrossRef]
  272. Rass, U.; Ahel, I.; West, S.C. Defective DNA repair and neurodegenerative disease. Cell 2007, 130, 991–1004. [Google Scholar] [CrossRef]
  273. Horwitz, E.; Krogvold, L.; Zhitomirsky, S.; Swisa, A.; Fischman, M.; Lax, T.; Dahan, T.; Hurvitz, N.; Weinberg-Corem, N.; Klochendler, A.; et al. β-Cell DNA Damage Response Promotes Islet Inflammation in Type 1 Diabetes. Diabetes 2018, 67, 2305–2318. [Google Scholar] [CrossRef]
  274. Schroyer, A.L.; Stimes, N.W.; Abi Saab, W.F.; Chadee, D.N. MLK3 phosphorylation by ERK1/2 is required for oxidative stress-induced invasion of colorectal cancer cells. Oncogene 2018, 37, 1031–1040. [Google Scholar] [CrossRef]
  275. Zhang, J.; Simpson, C.M.; Berner, J.; Chong, H.B.; Fang, J.; Ordulu, Z.; Weiss-Sadan, T.; Possemato, A.P.; Harry, S.; Takahashi, M.; et al. Systematic identification of anticancer drug targets reveals a nucleus-to-mitochondria ROS-sensing pathway. Cell 2023, 186, 2361–2379. [Google Scholar] [CrossRef]
  276. van Soest, D.M.K.; Polderman, P.E.; den Toom, W.T.F.; Keijer, J.P.; van Roosmalen, M.J.; Leyten, T.M.F.; Lehmann, J.; Zwakenberg, S.; De Henau, S.; van Boxtel, R.; et al. Mitochondrial H(2)O(2) release does not directly cause damage to chromosomal DNA. Nat. Commun. 2024, 15, 2725. [Google Scholar] [CrossRef]
  277. D’Oria, R.; Schipani, R.; Leonardini, A.; Natalicchio, A.; Perrini, S.; Cignarelli, A.; Laviola, L.; Giorgino, F. The Role of Oxidative Stress in Cardiac Disease: From Physiological Response to Injury Factor. Oxidative Med. Cell. Longev. 2020, 2020, 5732956. [Google Scholar] [CrossRef]
  278. Ni, Y.; Deng, J.; Liu, X.; Li, Q.; Zhang, J.; Bai, H.; Zhang, J. Echinacoside reverses myocardial remodeling and improves heart function via regulating SIRT1/FOXO3a/MnSOD axis in HF rats induced by isoproterenol. J. Cell. Mol. Med. 2021, 25, 203–216. [Google Scholar] [CrossRef]
  279. Murphy, E.; Ardehali, H.; Balaban, R.S.; DiLisa, F.; Dorn, G.W.; Kitsis, R.N.; Otsu, K.; Ping, P.; Rizzuto, R.; Sack, M.N.; et al. Mitochondrial Function, Biology, and Role in Disease. Circ. Res. 2016, 118, 1960–1991. [Google Scholar] [CrossRef]
  280. Santulli, G.; Xie, W.; Reiken, S.R.; Marks, A.R. Mitochondrial calcium overload is a key determinant in heart failure. Proc. Natl. Acad. Sci. USA 2015, 112, 11389–11394. [Google Scholar] [CrossRef]
  281. Gong, G.; Liu, X.; Wang, W. Regulation of metabolism in individual mitochondria during excitation-contraction coupling. J. Mol. Cell. Cardiol. 2014, 76, 235–246. [Google Scholar] [CrossRef]
  282. Doenst, T.; Pytel, G.; Schrepper, A.; Amorim, P.; Färber, G.; Shingu, Y.; Mohr, F.W.; Schwarzer, M. Decreased rates of substrate oxidation ex vivo predict the onset of heart failure and contractile dysfunction in rats with pressure overload. Cardiovasc. Res. 2009, 86, 461–470. [Google Scholar] [CrossRef]
  283. Hollander, J.M.; Thapa, D.; Shepherd, D.L. Physiological and structural differences in spatially distinct subpopulations of cardiac mitochondria: Influence of cardiac pathologies. Am. J. Physiol.-Heart Circ. Physiol. 2014, 307, H1–H14. [Google Scholar] [CrossRef]
  284. Hausenloy, D.J.; Yellon, D.M. Myocardial ischemia-reperfusion injury: A neglected therapeutic target. J. Clin. Investig. 2013, 123, 92–100. [Google Scholar] [CrossRef]
  285. Cadenas, S.; Aragonés, J.; Landázuri, M.O. Mitochondrial reprogramming through cardiac oxygen sensors in ischaemic heart disease. Cardiovasc. Res. 2010, 88, 219–228. [Google Scholar] [CrossRef]
  286. Cesselli, D.; Jakoniuk, I.; Barlucchi, L.; Beltrami, A.P.; Hintze, T.H.; Nadal-Ginard, B.; Kajstura, J.; Leri, A.; Anversa, P. Oxidative Stress–Mediated Cardiac Cell Death Is a Major Determinant of Ventricular Dysfunction and Failure in Dog Dilated Cardiomyopathy. Circ. Res. 2001, 89, 279–286. [Google Scholar] [CrossRef]
  287. Barlaka, E.; Görbe, A.; Gáspár, R.; Pálóczi, J.; Ferdinandy, P.; Lazou, A. Activation of PPARβ/δ protects cardiac myocytes from oxidative stress-induced apoptosis by suppressing generation of reactive oxygen/nitrogen species and expression of matrix metalloproteinases. Pharmacol. Res. 2015, 95–96, 102–110. [Google Scholar] [CrossRef]
  288. Laviola, L.; Leonardini, A.; Melchiorre, M.; Orlando, M.R.; Peschechera, A.; Bortone, A.; Paparella, D.; Natalicchio, A.; Perrini, S.; Giorgino, F. Glucagon-Like Peptide-1 Counteracts Oxidative Stress-Dependent Apoptosis of Human Cardiac Progenitor Cells by Inhibiting the Activation of the c-Jun N-terminal Protein Kinase Signaling Pathway. Endocrinology 2012, 153, 5770–5781. [Google Scholar] [CrossRef]
  289. Leonardini, A.; D’Oria, R.; Incalza, M.A.; Caccioppoli, C.; Andrulli Buccheri, V.; Cignarelli, A.; Paparella, D.; Margari, V.; Natalicchio, A.; Perrini, S.; et al. GLP-1 Receptor Activation Inhibits Palmitate-Induced Apoptosis via Ceramide in Human Cardiac Progenitor Cells. J. Clin. Endocrinol. Metab. 2017, 102, 4136–4147. [Google Scholar] [CrossRef]
  290. Sawyer, D.B. Oxidative Stress in Heart Failure: What Are We Missing? Am. J. Med. Sci. 2011, 342, 120–124. [Google Scholar] [CrossRef]
  291. Sam, F.; Kerstetter, D.L.; Pimental, D.R.; Mulukutla, S.; Tabaee, A.; Bristow, M.R.; Colucci, W.S.; Sawyer, D.B. Increased Reactive Oxygen Species Production and Functional Alterations in Antioxidant Enzymes in Human Failing Myocardium. J. Card. Fail. 2005, 11, 473–480. [Google Scholar] [CrossRef]
  292. Melov, S.; Coskun, P.; Patel, M.; Tuinstra, R.; Cottrell, B.; Jun, A.S.; Zastawny, T.H.; Dizdaroglu, M.; Goodman, S.I.; Huang, T.-T.; et al. Mitochondrial disease in superoxide dismutase 2 mutant mice. Proc. Natl. Acad. Sci. USA 1999, 96, 846–851. [Google Scholar] [CrossRef]
  293. Melov, S.; Doctrow, S.R.; Schneider, J.A.; Haberson, J.; Patel, M.; Coskun, P.E.; Huffman, K.; Wallace, D.C.; Malfroy, B. Lifespan extension and rescue of spongiform encephalopathy in superoxide dismutase 2 nullizygous mice treated with superoxide dismutase-catalase mimetics. J. Neurosci. 2001, 21, 8348–8353. [Google Scholar] [CrossRef]
  294. Dey, S.; DeMazumder, D.; Sidor, A.; Foster, D.B.; O’Rourke, B. Mitochondrial ROS Drive Sudden Cardiac Death and Chronic Proteome Remodeling in Heart Failure. Circ. Res. 2018, 123, 356–371. [Google Scholar] [CrossRef]
  295. Ago, T.; Liu, T.; Zhai, P.; Chen, W.; Li, H.; Molkentin, J.D.; Vatner, S.F.; Sadoshima, J. A Redox-Dependent Pathway for Regulating Class II HDACs and Cardiac Hypertrophy. Cell 2008, 133, 978–993. [Google Scholar] [CrossRef]
  296. Wei, J.; Joshi, S.; Speransky, S.; Crowley, C.; Jayathilaka, N.; Lei, X.; Wu, Y.; Gai, D.; Jain, S.; Hoosien, M.; et al. Reversal of pathological cardiac hypertrophy via the MEF2-coregulator interface. JCI Insight 2017, 2, e91068. [Google Scholar] [CrossRef]
  297. Oudot, A.; Vergely, C.; Ecarnot-Laubriet, A.; Rochette, L. Angiotensin II activates NADPH oxidase in isolated rat hearts subjected to ischaemia–reperfusion. Eur. J. Pharmacol. 2003, 462, 145–154. [Google Scholar] [CrossRef]
  298. Lyamzaev, K.G.; Panteleeva, A.A.; Simonyan, R.A.; Avetisyan, A.V.; Chernyak, B.V. Mitochondrial Lipid Peroxidation Is Responsible for Ferroptosis. Cells 2023, 12, 611. [Google Scholar] [CrossRef]
  299. Dixon, S.J.; Lemberg, K.M.; Lamprecht, M.R.; Skouta, R.; Zaitsev, E.M.; Gleason, C.E.; Patel, D.N.; Bauer, A.J.; Cantley, A.M.; Yang, W.S.; et al. Ferroptosis: An Iron-Dependent Form of Nonapoptotic Cell Death. Cell 2012, 149, 1060–1072. [Google Scholar] [CrossRef]
  300. Jang, S.; Chapa-Dubocq, X.R.; Tyurina, Y.Y.; St Croix, C.M.; Kapralov, A.A.; Tyurin, V.A.; Bayır, H.; Kagan, V.E.; Javadov, S. Elucidating the contribution of mitochondrial glutathione to ferroptosis in cardiomyocytes. Redox Biol. 2021, 45, 102021. [Google Scholar] [CrossRef]
  301. Jiang, X.; Stockwell, B.R.; Conrad, M. Ferroptosis: Mechanisms, biology and role in disease. Nat. Rev. Mol. Cell Biol. 2021, 22, 266–282. [Google Scholar] [CrossRef]
  302. Jiang, X.; Zhang, K.; Gao, C.; Ma, W.; Liu, M.; Guo, X.; Bao, G.; Han, B.; Hu, H.; Zhao, Z. Activation of FMS-like tyrosine kinase 3 protects against isoprenaline-induced cardiac hypertrophy by improving autophagy and mitochondrial dynamics. FASEB J. 2022, 36, e22672. [Google Scholar] [CrossRef]
  303. Mercer, J.R.; Cheng, K.K.; Figg, N.; Gorenne, I.; Mahmoudi, M.; Griffin, J.; Vidal-Puig, A.; Logan, A.; Murphy, M.P.; Bennett, M. DNA damage links mitochondrial dysfunction to atherosclerosis and the metabolic syndrome. Circ. Res. 2010, 107, 1021–1031. [Google Scholar] [CrossRef]
  304. Xu, H.-x.; Cui, S.-m.; Zhang, Y.-m.; Ren, J. Mitochondrial Ca2+ regulation in the etiology of heart failure: Physiological and pathophysiological implications. Acta Pharmacol. Sin. 2020, 41, 1301–1309. [Google Scholar] [CrossRef]
  305. Luan, Y.; Ren, K.D.; Luan, Y.; Chen, X.; Yang, Y. Mitochondrial Dynamics: Pathogenesis and Therapeutic Targets of Vascular Diseases. Front. Cardiovasc. Med. 2021, 8, 770574. [Google Scholar] [CrossRef]
  306. Luan, Y.; Luan, Y.; Feng, Q.; Chen, X.; Ren, K.D.; Yang, Y. Emerging Role of Mitophagy in the Heart: Therapeutic Potentials to Modulate Mitophagy in Cardiac Diseases. Oxidative Med. Cell. Longev. 2021, 2021, 3259963. [Google Scholar] [CrossRef]
  307. Fang, X.; Cai, Z.; Wang, H.; Han, D.; Cheng, Q.; Zhang, P.; Gao, F.; Yu, Y.; Song, Z.; Wu, Q.; et al. Loss of Cardiac Ferritin H Facilitates Cardiomyopathy via Slc7a11-Mediated Ferroptosis. Circ. Res. 2020, 127, 486–501. [Google Scholar] [CrossRef]
  308. Zhou, L.; Ma, B.; Han, X. The role of autophagy in angiotensin II-induced pathological cardiac hypertrophy. J. Mol. Endocrinol. 2016, 57, R143–R152. [Google Scholar] [CrossRef]
  309. Miao, W.; Chen, M.; Chen, M.; Cui, C.; Zhu, Y.; Luo, X.; Wu, B. Nr2f2 Overexpression Aggravates Ferroptosis and Mitochondrial Dysfunction by Regulating the PGC-1 α Signaling in Diabetes-Induced Heart Failure Mice. Mediat. Inflamm. 2022, 2022, 8373389. [Google Scholar] [CrossRef]
  310. Zuo, B.; Fan, X.; Xu, D.; Zhao, L.; Zhang, B.; Li, X. Deciphering the mitochondria-inflammation axis: Insights and therapeutic strategies for heart failure. Int. Immunopharmacol. 2024, 139, 112697. [Google Scholar] [CrossRef]
  311. Kohli, S.; Ahuja, S.; Rani, V. Transcription Factors in Heart: Promising Therapeutic Targets in Cardiac Hypertrophy. Curr. Cardiol. Rev. 2011, 7, 262–271. [Google Scholar] [CrossRef]
  312. Pu, W.T.; Ma, Q.; Izumo, S. NFAT Transcription Factors Are Critical Survival Factors That Inhibit Cardiomyocyte Apoptosis During Phenylephrine Stimulation In Vitro. Circ. Res. 2003, 92, 725–731. [Google Scholar] [CrossRef]
  313. Qiu, Z.; Fan, Y.; Wang, Z.; Huang, F.; Li, Z.; Sun, Z.; Hua, S.; Jin, W.; Chen, Y. Catestatin Protects Against Diastolic Dysfunction by Attenuating Mitochondrial Reactive Oxygen Species Generation. J. Am. Heart Assoc. 2023, 12, e029470. [Google Scholar] [CrossRef]
  314. Schiattarella, G.G.; Altamirano, F.; Kim, S.Y.; Tong, D.; Ferdous, A.; Piristine, H.; Dasgupta, S.; Wang, X.; French, K.M.; Villalobos, E.; et al. Xbp1s-FoxO1 axis governs lipid accumulation and contractile performance in heart failure with preserved ejection fraction. Nat. Commun. 2021, 12, 1684. [Google Scholar] [CrossRef]
  315. Shirakawa, R.; Yokota, T.; Nakajima, T.; Takada, S.; Yamane, M.; Furihata, T.; Maekawa, S.; Nambu, H.; Katayama, T.; Fukushima, A.; et al. Mitochondrial reactive oxygen species generation in blood cells is associated with disease severity and exercise intolerance in heart failure patients. Sci. Rep. 2019, 9, 14709. [Google Scholar] [CrossRef]
  316. Negi, S.I.; Jeong, E.-M.; Shukrullah, I.; Veleder, E.; Jones, D.P.; Fan, T.-H.M.; Varadarajan, S.; Danilov, S.M.; Fukai, T.; Dudley, S.C., Jr. Renin-Angiotensin Activation and Oxidative Stress in Early Heart Failure with Preserved Ejection Fraction. BioMed Res. Int. 2015, 2015, 825027. [Google Scholar] [CrossRef]
  317. Brown, G.C.; Borutaite, V. Nitric oxide and mitochondrial respiration in the heart. Cardiovasc. Res. 2007, 75, 283–290. [Google Scholar] [CrossRef]
  318. Mital, S.; Loke, K.E.; Addonizio, L.J.; Oz, M.C.; Hintze, T.H. Left ventricular assist device implantation augments nitric oxide dependent control of mitochondrial respiration in failing human hearts. J. Am. Coll. Cardiol. 2000, 36, 1897–1902. [Google Scholar] [CrossRef]
  319. Echouffo-Tcheugui, J.B.; Xu, H.; DeVore, A.D.; Schulte, P.J.; Butler, J.; Yancy, C.W.; Bhatt, D.L.; Hernandez, A.F.; Heidenreich, P.A.; Fonarow, G.C. Temporal trends and factors associated with diabetes mellitus among patients hospitalized with heart failure: Findings from Get with The Guidelines-Heart Failure registry. Am. Heart J. 2016, 182, 9–20. [Google Scholar] [CrossRef]
  320. Yap, J.; Tay, W.T.; Teng, T.K.; Anand, I.; Richards, A.M.; Ling, L.H.; MacDonald, M.R.; Chandramouli, C.; Tromp, J.; Siswanto, B.B.; et al. Association of Diabetes Mellitus on Cardiac Remodeling, Quality of Life, and Clinical Outcomes in Heart Failure with Reduced and Preserved Ejection Fraction. J. Am. Heart Assoc. 2019, 8, e013114. [Google Scholar] [CrossRef]
  321. Schleicher, E.; Friess, U. Oxidative stress, AGE, and atherosclerosis. Kidney Int. 2007, 72, S17–S26. [Google Scholar] [CrossRef]
  322. Chaudhuri, J.; Bains, Y.; Guha, S.; Kahn, A.; Hall, D.; Bose, N.; Gugliucci, A.; Kapahi, P. The Role of Advanced Glycation End Products in Aging and Metabolic Diseases: Bridging Association and Causality. Cell Metab. 2018, 28, 337–352. [Google Scholar] [CrossRef]
  323. Zhang, M.; Kho, A.L.; Anilkumar, N.; Chibber, R.; Pagano, P.J.; Shah, A.M.; Cave, A.C. Glycated Proteins Stimulate Reactive Oxygen Species Production in Cardiac Myocytes. Circulation 2006, 113, 1235–1243. [Google Scholar] [CrossRef]
  324. Maack, C.; Lehrke, M.; Backs, J.; Heinzel, F.R.; Hulot, J.-S.; Marx, N.; Paulus, W.J.; Rossignol, P.; Taegtmeyer, H.; Bauersachs, J.; et al. Heart failure and diabetes: Metabolic alterations and therapeutic interventions: A state-of-the-art review from the Translational Research Committee of the Heart Failure Association–European Society of Cardiology. Eur. Heart J. 2018, 39, 4243–4254. [Google Scholar] [CrossRef]
  325. Kasner, M.; Aleksandrov, A.S.; Westermann, D.; Lassner, D.; Gross, M.; von Haehling, S.; Anker, S.D.; Schultheiss, H.-P.; Tschöpe, C. Functional iron deficiency and diastolic function in heart failure with preserved ejection fraction. Int. J. Cardiol. 2013, 168, 4652–4657. [Google Scholar] [CrossRef]
  326. Bates, A.; Shen, Q.; Hiebert, J.B.; Thimmesch, A.; Pierce, J.D. Myocardial energetics and ubiquinol in diastolic heart failure. Nurs. Health Sci. 2014, 16, 428–433. [Google Scholar] [CrossRef]
  327. Luptak, I.; Qin, F.; Sverdlov, A.L.; Pimentel, D.R.; Panagia, M.; Croteau, D.; Siwik, D.A.; Bachschmid, M.M.; He, H.; Balschi, J.A.; et al. Energetic Dysfunction Is Mediated by Mitochondrial Reactive Oxygen Species and Precedes Structural Remodeling in Metabolic Heart Disease. Antioxid. Redox Signal. 2019, 31, 539–549. [Google Scholar] [CrossRef]
  328. Hu, Q.; Wood, C.R.; Cimen, S.; Venkatachalam, A.B.; Alwayn, I.P.J. Mitochondrial Damage-Associated Molecular Patterns (MTDs) Are Released during Hepatic Ischemia Reperfusion and Induce Inflammatory Responses. PLoS ONE 2015, 10, e0140105. [Google Scholar] [CrossRef]
  329. Zhao, C.; Itagaki, K.; Gupta, A.; Odom, S.; Sandler, N.; Hauser, C.J. Mitochondrial damage-associated molecular patterns released by abdominal trauma suppress pulmonary immune responses. J. Trauma Acute Care Surg. 2014, 76, 1222–1227. [Google Scholar] [CrossRef]
  330. Pollara, J.; Edwards, R.W.; Lin, L.; Bendersky, V.A.; Brennan, T.V. Circulating mitochondria in deceased organ donors are associated with immune activation and early allograft dysfunction. JCI Insight 2018, 3, e121622. [Google Scholar] [CrossRef]
  331. Gong, T.; Liu, L.; Jiang, W.; Zhou, R. DAMP-sensing receptors in sterile inflammation and inflammatory diseases. Nat. Rev. Immunol. 2020, 20, 95–112. [Google Scholar] [CrossRef]
  332. Unterreiner, A.; Rubert, J.; Kauffmann, M.; Fruhauf, A.; Heiser, D.; Erbel, P.; Schlapbach, A.; Eder, J.; Bodendorf, U.; Boettcher, A.; et al. Pharmacological inhibition of IKKβ dampens NLRP3 inflammasome activation after priming in the human myeloid cell line THP-1. Biochem. Biophys. Res. Commun. 2021, 545, 177–182. [Google Scholar] [CrossRef]
  333. Hulina, A.; Grdić Rajković, M.; Jakšić Despot, D.; Jelić, D.; Dojder, A.; Čepelak, I.; Rumora, L. Extracellular Hsp70 induces inflammation and modulates LPS/LTA-stimulated inflammatory response in THP-1 cells. Cell Stress Chaperones 2018, 23, 373–384. [Google Scholar] [CrossRef]
  334. Li, H.; Leung, J.C.K.; Yiu, W.H.; Chan, L.Y.Y.; Li, B.; Lok, S.W.Y.; Xue, R.; Zou, Y.; Lai, K.N.; Tang, S.C.W. Tubular β-catenin alleviates mitochondrial dysfunction and cell death in acute kidney injury. Cell Death Dis. 2022, 13, 1061. [Google Scholar] [CrossRef]
  335. Ummarino, D. Mitochondrial calcium efflux essential for heart function. Nat. Rev. Cardiol. 2017, 14, 317. [Google Scholar] [CrossRef]
  336. Luan, Y.; Luan, Y.; Jiao, Y.; Liu, H.; Huang, Z.; Feng, Q.; Pei, J.; Yang, Y.; Ren, K. Broadening Horizons: Exploring mtDAMPs as a Mechanism and Potential Intervention Target in Cardiovascular Diseases. Aging Dis. 2023. [Google Scholar] [CrossRef]
  337. Silvis, M.J.M.; Kaffka Genaamd Dengler, S.E.; Odille, C.A.; Mishra, M.; van der Kaaij, N.P.; Doevendans, P.A.; Sluijter, J.P.G.; de Kleijn, D.P.V.; de Jager, S.C.A.; Bosch, L.; et al. Damage-Associated Molecular Patterns in Myocardial Infarction and Heart Transplantation: The Road to Translational Success. Front. Immunol. 2020, 11, 599511. [Google Scholar] [CrossRef]
  338. Pandey, A.; Shah, S.J.; Butler, J.; Kellogg, D.L., Jr.; Lewis, G.D.; Forman, D.E.; Mentz, R.J.; Borlaug, B.A.; Simon, M.A.; Chirinos, J.A.; et al. Exercise Intolerance in Older Adults with Heart Failure with Preserved Ejection Fraction: JACC State-of-the-Art Review. J. Am. Coll. Cardiol. 2021, 78, 1166–1187. [Google Scholar] [CrossRef]
  339. Heidenreich, P.A.; Bozkurt, B.; Aguilar, D.; Allen, L.A.; Byun, J.J.; Colvin, M.M.; Deswal, A.; Drazner, M.H.; Dunlay, S.M.; Evers, L.R.; et al. 2022 AHA/ACC/HFSA Guideline for the Management of Heart Failure: A Report of the American College of Cardiology/American Heart Association Joint Committee on Clinical Practice Guidelines. Circulation 2022, 145, e895–e1032. [Google Scholar] [CrossRef]
  340. Reddy, Y.N.V.; Obokata, M.; Testani, J.M.; Felker, G.M.; Tang, W.H.W.; Abou-Ezzeddine, O.F.; Sun, J.L.; Chakrabothy, H.; McNulty, S.; Shah, S.J.; et al. Adverse Renal Response to Decongestion in the Obese Phenotype of Heart Failure with Preserved Ejection Fraction. J. Card. Fail. 2020, 26, 101–107. [Google Scholar] [CrossRef]
  341. Solomon, S.D.; McMurray, J.J.V.; Claggett, B.; de Boer, R.A.; DeMets, D.; Hernandez, A.F.; Inzucchi, S.E.; Kosiborod, M.N.; Lam, C.S.P.; Martinez, F.; et al. Dapagliflozin in Heart Failure with Mildly Reduced or Preserved Ejection Fraction. N. Engl. J. Med. 2022, 387, 1089–1098. [Google Scholar] [CrossRef]
  342. Vaduganathan, M.; Docherty, K.F.; Claggett, B.L.; Jhund, P.S.; de Boer, R.A.; Hernandez, A.F.; Inzucchi, S.E.; Kosiborod, M.N.; Lam, C.S.P.; Martinez, F.; et al. SGLT-2 inhibitors in patients with heart failure: A comprehensive meta-analysis of five randomised controlled trials. Lancet 2022, 400, 757–767. [Google Scholar] [CrossRef]
  343. Bhatt, D.L.; Szarek, M.; Steg, P.G.; Cannon, C.P.; Leiter, L.A.; McGuire, D.K.; Lewis, J.B.; Riddle, M.C.; Voors, A.A.; Metra, M.; et al. Sotagliflozin in Patients with Diabetes and Recent Worsening Heart Failure. N. Engl. J. Med. 2021, 384, 117–128. [Google Scholar] [CrossRef]
  344. Voors, A.A.; Angermann, C.E.; Teerlink, J.R.; Collins, S.P.; Kosiborod, M.; Biegus, J.; Ferreira, J.P.; Nassif, M.E.; Psotka, M.A.; Tromp, J.; et al. The SGLT2 inhibitor empagliflozin in patients hospitalized for acute heart failure: A multinational randomized trial. Nat. Med. 2022, 28, 568–574. [Google Scholar] [CrossRef]
  345. Packer, M. Obesity-Associated Heart Failure as a Theoretical Target for Treatment with Mineralocorticoid Receptor Antagonists. JAMA Cardiol. 2018, 3, 883–887. [Google Scholar] [CrossRef]
  346. Guo, C.; Ricchiuti, V.; Lian, B.Q.; Yao, T.M.; Coutinho, P.; Romero, J.R.; Li, J.; Williams, G.H.; Adler, G.K. Mineralocorticoid receptor blockade reverses obesity-related changes in expression of adiponectin, peroxisome proliferator-activated receptor-gamma, and proinflammatory adipokines. Circulation 2008, 117, 2253–2261. [Google Scholar] [CrossRef]
  347. Pfeffer, M.A.; Claggett, B.; Assmann, S.F.; Boineau, R.; Anand, I.S.; Clausell, N.; Desai, A.S.; Diaz, R.; Fleg, J.L.; Gordeev, I.; et al. Regional variation in patients and outcomes in the Treatment of Preserved Cardiac Function Heart Failure with an Aldosterone Antagonist (TOPCAT) trial. Circulation 2015, 131, 34–42. [Google Scholar] [CrossRef]
  348. Solomon, S.D.; McMurray, J.J.V.; Anand, I.S.; Ge, J.; Lam, C.S.P.; Maggioni, A.P.; Martinez, F.; Packer, M.; Pfeffer, M.A.; Pieske, B.; et al. Angiotensin–Neprilysin Inhibition in Heart Failure with Preserved Ejection Fraction. N. Engl. J. Med. 2019, 381, 1609–1620. [Google Scholar] [CrossRef]
  349. McMurray, J.J.V.; Jackson, A.M.; Lam, C.S.P.; Redfield, M.M.; Anand, I.S.; Ge, J.; Lefkowitz, M.P.; Maggioni, A.P.; Martinez, F.; Packer, M.; et al. Effects of Sacubitril-Valsartan Versus Valsartan in Women Compared with Men with Heart Failure and Preserved Ejection Fraction: Insights From PARAGON-HF. Circulation 2020, 141, 338–351. [Google Scholar] [CrossRef]
  350. Chyou, J.Y.; Qin, H.; Butler, J.; Voors, A.A.; Lam, C.S.P. Sex-related similarities and differences in responses to heart failure therapies. Nat. Rev. Cardiol. 2024, 21, 498–516. [Google Scholar] [CrossRef]
  351. Kittleson, M.M.; Panjrath, G.S.; Amancherla, K.; Davis, L.L.; Deswal, A.; Dixon, D.L.; Januzzi, J.L., Jr.; Yancy, C.W. 2023 ACC Expert Consensus Decision Pathway on Management of Heart Failure with Preserved Ejection Fraction: A Report of the American College of Cardiology Solution Set Oversight Committee. J Am Coll Cardiol 2023, 81, 1835–1878. [Google Scholar] [CrossRef]
  352. Yusuf, S.; Pfeffer, M.A.; Swedberg, K.; Granger, C.B.; Held, P.; McMurray, J.J.; Michelson, E.L.; Olofsson, B.; Ostergren, J. Effects of candesartan in patients with chronic heart failure and preserved left-ventricular ejection fraction: The CHARM-Preserved Trial. Lancet 2003, 362, 777–781. [Google Scholar] [CrossRef]
  353. Cleland, J.G.F.; Bunting, K.V.; Flather, M.D.; Altman, D.G.; Holmes, J.; Coats, A.J.S.; Manzano, L.; McMurray, J.J.V.; Ruschitzka, F.; van Veldhuisen, D.J.; et al. Beta-blockers for heart failure with reduced, mid-range, and preserved ejection fraction: An individual patient-level analysis of double-blind randomized trials. Eur. Heart J. 2018, 39, 26–35. [Google Scholar] [CrossRef]
  354. Palau, P.; de la Espriella, R.; Seller, J.; Santas, E.; Domínguez, E.; Bodí, V.; Sanchis, J.; Núñez, E.; Bayés-Genís, A.; Bertomeu-González, V.; et al. β-Blocker withdrawal and Functional Capacity Improvement in Patients with Heart Failure with Preserved Ejection Fraction. JAMA Cardiol. 2024, 9, 392–396. [Google Scholar] [CrossRef]
  355. Kaddoura, R.; Madurasinghe, V.; Chapra, A.; Abushanab, D.; Al-Badriyeh, D.; Patel, A. Beta-blocker therapy in heart failure with preserved ejection fraction (B-HFpEF): A systematic review and meta-analysis. Curr. Probl. Cardiol. 2024, 49, 102376. [Google Scholar] [CrossRef]
  356. Luan, Y.; Jin, Y.; Zhang, P.; Li, H.; Yang, Y. Mitochondria-associated endoplasmic reticulum membranes and cardiac hypertrophy: Molecular mechanisms and therapeutic targets. Front. Cardiovasc. Med. 2022, 9, 1015722. [Google Scholar] [CrossRef]
  357. Dabravolski, S.A.; Nikiforov, N.G.; Starodubova, A.V.; Popkova, T.V.; Orekhov, A.N. The Role of Mitochondria-Derived Peptides in Cardiovascular Diseases and Their Potential as Therapeutic Targets. Int. J. Mol. Sci. 2021, 22, 8770. [Google Scholar] [CrossRef]
  358. Ramachandra, C.J.A.; Hernandez-Resendiz, S.; Crespo-Avilan, G.E.; Lin, Y.H.; Hausenloy, D.J. Mitochondria in acute myocardial infarction and cardioprotection. EBioMedicine 2020, 57, 102884. [Google Scholar] [CrossRef]
  359. Bhatti, J.S.; Bhatti, G.K.; Reddy, P.H. Mitochondrial dysfunction and oxidative stress in metabolic disorders—A step towards mitochondria based therapeutic strategies. Biochim. Biophys. Acta Mol. Basis Dis. 2017, 1863, 1066–1077. [Google Scholar] [CrossRef]
  360. 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]
  361. Birk, A.V.; Chao, W.M.; Bracken, C.; Warren, J.D.; Szeto, H.H. Targeting mitochondrial cardiolipin and the cytochrome c/cardiolipin complex to promote electron transport and optimize mitochondrial ATP synthesis. Br. J. Pharmacol. 2014, 171, 2017–2028. [Google Scholar] [CrossRef]
  362. Eirin, A.; Ebrahimi, B.; Kwon, S.H.; Fiala, J.A.; Williams, B.J.; Woollard, J.R.; He, Q.; Gupta, R.C.; Sabbah, H.N.; Prakash, Y.S.; et al. Restoration of Mitochondrial Cardiolipin Attenuates Cardiac Damage in Swine Renovascular Hypertension. J. Am. Heart Assoc. 2016, 5, e003118. [Google Scholar] [CrossRef]
  363. Dai, D.-F.; Chen, T.; Szeto, H.; Nieves-Cintrón, M.; Kutyavin, V.; Santana, L.F.; Rabinovitch, P.S. Mitochondrial Targeted Antioxidant Peptide Ameliorates Hypertensive Cardiomyopathy. J. Am. Coll. Cardiol. 2011, 58, 73–82. [Google Scholar] [CrossRef]
  364. Zhang, L.; Feng, M.; Wang, X.; Zhang, H.; Ding, J.; Cheng, Z.; Qian, L. Peptide Szeto-Schiller 31 ameliorates doxorubicin-induced cardiotoxicity by inhibiting the activation of the p38 MAPK signaling pathway. Int. J. Mol. Med. 2021, 47, 63. [Google Scholar] [CrossRef]
  365. Shi, J.; Dai, W.; Hale, S.L.; Brown, D.A.; Wang, M.; Han, X.; Kloner, R.A. Bendavia restores mitochondrial energy metabolism gene expression and suppresses cardiac fibrosis in the border zone of the infarcted heart. Life Sci. 2015, 141, 170–178. [Google Scholar] [CrossRef]
  366. Dai, W.; Shi, J.; Gupta, R.C.; Sabbah, H.N.; Hale, S.L.; Kloner, R.A. Bendavia, a mitochondria-targeting peptide, improves postinfarction cardiac function, prevents adverse left ventricular remodeling, and restores mitochondria-related gene expression in rats. J. Cardiovasc. Pharmacol. 2014, 64, 543–553. [Google Scholar] [CrossRef]
  367. Sabbah, H.N.; Gupta, R.C.; Kohli, S.; Wang, M.; Hachem, S.; Zhang, K. Chronic therapy with elamipretide (MTP-131), a novel mitochondria-targeting peptide, improves left ventricular and mitochondrial function in dogs with advanced heart failure. Circ. Heart Fail. 2016, 9, e002206. [Google Scholar] [CrossRef]
  368. Chatfield, K.C.; Sparagna, G.C.; Chau, S.; Phillips, E.K.; Ambardekar, A.V.; Aftab, M.; Mitchell, M.B.; Sucharov, C.C.; Miyamoto, S.D.; Stauffer, B.L. Elamipretide Improves Mitochondrial Function in the Failing Human Heart. JACC Basic. Transl. Sci. 2019, 4, 147–157. [Google Scholar] [CrossRef]
  369. Butler, J.; Khan, M.S.; Anker, S.D.; Fonarow, G.C.; Kim, R.J.; Nodari, S.; O’Connor, C.M.; Pieske, B.; Pieske-Kraigher, E.; Sabbah, H.N.; et al. Effects of Elamipretide on Left Ventricular Function in Patients with Heart Failure with Reduced Ejection Fraction: The PROGRESS-HF Phase 2 Trial. J. Card. Fail. 2020, 26, 429–437. [Google Scholar] [CrossRef]
  370. Gibson, C.M.; Giugliano, R.P.; Kloner, R.A.; Bode, C.; Tendera, M.; Jánosi, A.; Merkely, B.; Godlewski, J.; Halaby, R.; Korjian, S.; et al. EMBRACE STEMI study: A Phase 2a trial to evaluate the safety, tolerability, and efficacy of intravenous MTP-131 on reperfusion injury in patients undergoing primary percutaneous coronary intervention. Eur. Heart J. 2016, 37, 1296–1303. [Google Scholar] [CrossRef]
  371. Crane, F.L.; Hatefi, Y.; Lester, R.L.; Widmer, C. Isolation of a quinone from beef heart mitochondria. Biochim. Biophys. Acta 1957, 25, 220–221. [Google Scholar] [CrossRef]
  372. Baggio, E.; Gandini, R.; Plancher, A.C.; Passeri, M.; Carmosino, G. Italian multicenter study on the safety and efficacy of coenzyme Q10 as adjunctive therapy in heart failure. Mol. Asp. Med. 1994, 15, s287–s294. [Google Scholar] [CrossRef]
  373. Mortensen, S.A.; Rosenfeldt, F.; Kumar, A.; Dolliner, P.; Filipiak, K.J.; Pella, D.; Alehagen, U.; Steurer, G.; Littarru, G.P. The Effect of Coenzyme Q10 on Morbidity and Mortality in Chronic Heart Failure: Results From Q-SYMBIO: A Randomized Double-Blind Trial. JACC Heart Fail. 2014, 2, 641–649. [Google Scholar] [CrossRef]
  374. Alehagen, U.; Johansson, P.; Björnstedt, M.; Rosén, A.; Dahlström, U. Cardiovascular mortality and N-terminal-proBNP reduced after combined selenium and coenzyme Q10 supplementation: A 5-year prospective randomized double-blind placebo-controlled trial among elderly Swedish citizens. Int. J. Cardiol. 2013, 167, 1860–1866. [Google Scholar] [CrossRef]
  375. Alehagen, U.; Aaseth, J.; Alexander, J.; Johansson, P. Still reduced cardiovascular mortality 12 years after supplementation with selenium and coenzyme Q10 for four years: A validation of previous 10-year follow-up results of a prospective randomized double-blind placebo-controlled trial in elderly. PLoS ONE 2018, 13, e0193120. [Google Scholar] [CrossRef]
  376. Botting, K.J.; Skeffington, K.L.; Niu, Y.; Allison, B.J.; Brain, K.L.; Itani, N.; Beck, C.; Logan, A.; Murray, A.J.; Murphy, M.P.; et al. Translatable mitochondria-targeted protection against programmed cardiovascular dysfunction. Sci. Adv. 2020, 6, eabb1929. [Google Scholar] [CrossRef]
  377. Xiao, L.; Xu, X.; Zhang, F.; Wang, M.; Xu, Y.; Tang, D.; Wang, J.; Qin, Y.; Liu, Y.; Tang, C.; et al. The mitochondria-targeted antioxidant MitoQ ameliorated tubular injury mediated by mitophagy in diabetic kidney disease via Nrf2/PINK1. Redox Biol. 2017, 11, 297–311. [Google Scholar] [CrossRef]
  378. Pak, O.; Scheibe, S.; Esfandiary, A.; Gierhardt, M.; Sydykov, A.; Logan, A.; Fysikopoulos, A.; Veit, F.; Hecker, M.; Kroschel, F.; et al. Impact of the mitochondria-targeted antioxidant MitoQ on hypoxia-induced pulmonary hypertension. Eur. Respir. J. 2018, 51, 1701024. [Google Scholar] [CrossRef]
  379. Hao, L.; Sun, Q.; Zhong, W.; Zhang, W.; Sun, X.; Zhou, Z. Mitochondria-targeted ubiquinone (MitoQ) enhances acetaldehyde clearance by reversing alcohol-induced posttranslational modification of aldehyde dehydrogenase 2: A molecular mechanism of protection against alcoholic liver disease. Redox Biol. 2018, 14, 626–636. [Google Scholar] [CrossRef]
  380. Goh, K.Y.; He, L.; Song, J.; Jinno, M.; Rogers, A.J.; Sethu, P.; Halade, G.V.; Rajasekaran, N.S.; Liu, X.; Prabhu, S.D.; et al. Mitoquinone ameliorates pressure overload-induced cardiac fibrosis and left ventricular dysfunction in mice. Redox Biol. 2019, 21, 101100. [Google Scholar] [CrossRef]
  381. Kim, S.; Song, J.; Ernst, P.; Latimer, M.N.; Ha, C.-M.; Goh, K.Y.; Ma, W.; Rajasekaran, N.-S.; Zhang, J.; Liu, X.; et al. MitoQ regulates redox-related noncoding RNAs to preserve mitochondrial network integrity in pressure-overload heart failure. Am. J. Physiol.-Heart Circ. Physiol. 2020, 318, H682–H695. [Google Scholar] [CrossRef] [PubMed]
  382. Rossman, M.J.; Santos-Parker, J.R.; Steward, C.A.C.; Bispham, N.Z.; Cuevas, L.M.; Rosenberg, H.L.; Woodward, K.A.; Chonchol, M.; Gioscia-Ryan, R.A.; Murphy, M.P.; et al. Chronic Supplementation with a Mitochondrial Antioxidant (MitoQ) Improves Vascular Function in Healthy Older Adults. Hypertension 2018, 71, 1056–1063. [Google Scholar] [CrossRef] [PubMed]
  383. Mikhalkova, D.; Holman, S.R.; Jiang, H.; Saghir, M.; Novak, E.; Coggan, A.R.; O’Connor, R.; Bashir, A.; Jamal, A.; Ory, D.S.; et al. Bariatric Surgery-Induced Cardiac and Lipidomic Changes in Obesity-Related Heart Failure with Preserved Ejection Fraction. Obesity 2018, 26, 284–290. [Google Scholar] [CrossRef] [PubMed]
  384. Kitzman, D.W.; Brubaker, P.; Morgan, T.; Haykowsky, M.; Hundley, G.; Kraus, W.E.; Eggebeen, J.; Nicklas, B.J. Effect of Caloric Restriction or Aerobic Exercise Training on Peak Oxygen Consumption and Quality of Life in Obese Older Patients with Heart Failure with Preserved Ejection Fraction: A Randomized Clinical Trial. JAMA 2016, 315, 36–46. [Google Scholar] [CrossRef] [PubMed]
  385. Carbone, S.; Mauro, A.G.; Mezzaroma, E.; Kraskauskas, D.; Marchetti, C.; Buzzetti, R.; Van Tassell, B.W.; Abbate, A.; Toldo, S. A high-sugar and high-fat diet impairs cardiac systolic and diastolic function in mice. Int. J. Cardiol. 2015, 198, 66–69. [Google Scholar] [CrossRef]
  386. von Bibra, H.; Ströhle, A.; St. John Sutton, M.; Worm, N. Dietary therapy in heart failure with preserved ejection fraction and/or left ventricular diastolic dysfunction in patients with metabolic syndrome. Int. J. Cardiol. 2017, 234, 7–15. [Google Scholar] [CrossRef]
  387. Carbone, S.; Canada, J.M.; Buckley, L.F.; Trankle, C.R.; Billingsley, H.E.; Dixon, D.L.; Mauro, A.G.; Dessie, S.; Kadariya, D.; Mezzaroma, E.; et al. Dietary Fat, Sugar Consumption, and Cardiorespiratory Fitness in Patients with Heart Failure with Preserved Ejection Fraction. JACC Basic. Transl. Sci. 2017, 2, 513–525. [Google Scholar] [CrossRef]
  388. Yan, Y.; Jiang, W.; Spinetti, T.; Tardivel, A.; Castillo, R.; Bourquin, C.; Guarda, G.; Tian, Z.; Tschopp, J.; Zhou, R. Omega-3 Fatty Acids Prevent Inflammation and Metabolic Disorder through Inhibition of NLRP3 Inflammasome Activation. Immunity 2013, 38, 1154–1163. [Google Scholar] [CrossRef]
  389. Wang, D.D.; Li, Y.; Chiuve, S.E.; Stampfer, M.J.; Manson, J.E.; Rimm, E.B.; Willett, W.C.; Hu, F.B. Association of Specific Dietary Fats with Total and Cause-Specific Mortality. JAMA Intern. Med. 2016, 176, 1134–1145. [Google Scholar] [CrossRef]
  390. Tucker, W.J.; Lijauco, C.C.; Hearon, C.M., Jr.; Angadi, S.S.; Nelson, M.D.; Sarma, S.; Nanayakkara, S.; La Gerche, A.; Haykowsky, M.J. Mechanisms of the Improvement in Peak VO(2) with Exercise Training in Heart Failure with Reduced or Preserved Ejection Fraction. Heart Lung Circ. 2018, 27, 9–21. [Google Scholar] [CrossRef]
  391. Pandey, A.; Parashar, A.; Kumbhani, D.; Agarwal, S.; Garg, J.; Kitzman, D.; Levine, B.; Drazner, M.; Berry, J. Exercise training in patients with heart failure and preserved ejection fraction: Meta-analysis of randomized control trials. Circ. Heart Fail. 2015, 8, 33–40. [Google Scholar] [CrossRef] [PubMed]
  392. Edelmann, F.; Gelbrich, G.; Düngen, H.D.; Fröhling, S.; Wachter, R.; Stahrenberg, R.; Binder, L.; Töpper, A.; Lashki, D.J.; Schwarz, S.; et al. Exercise training improves exercise capacity and diastolic function in patients with heart failure with preserved ejection fraction: Results of the Ex-DHF (Exercise training in Diastolic Heart Failure) pilot study. J. Am. Coll. Cardiol. 2011, 58, 1780–1791. [Google Scholar] [CrossRef] [PubMed]
  393. Mueller, S.; Winzer, E.B.; Duvinage, A.; Gevaert, A.B.; Edelmann, F.; Haller, B.; Pieske-Kraigher, E.; Beckers, P.; Bobenko, A.; Hommel, J.; et al. Effect of High-Intensity Interval Training, Moderate Continuous Training, or Guideline-Based Physical Activity Advice on Peak Oxygen Consumption in Patients with Heart Failure with Preserved Ejection Fraction: A Randomized Clinical Trial. JAMA 2021, 325, 542–551. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Pathophysiology of HFpEF. Multiple cardiovascular (CV) and non-CV comorbidities sequentially cause chronic and low-grade systemic inflammation, mitochondrial dysfunction, and ROS generation. These pathological alterations consequently trigger pathological remodeling and functional changes in the heart, which in turn increase LV stiffness, eventually causing HFpEF. HFpEF, heart failure with preserved ejection fraction; LV, left ventricular; NO, nitric oxide; ROS, reactive oxygen species. Red upwards and downwards arrows indicate increase and decrease, respectively. Diagram was created with URL (BioRender.com) (accessed on 27 October 2024).
Figure 1. Pathophysiology of HFpEF. Multiple cardiovascular (CV) and non-CV comorbidities sequentially cause chronic and low-grade systemic inflammation, mitochondrial dysfunction, and ROS generation. These pathological alterations consequently trigger pathological remodeling and functional changes in the heart, which in turn increase LV stiffness, eventually causing HFpEF. HFpEF, heart failure with preserved ejection fraction; LV, left ventricular; NO, nitric oxide; ROS, reactive oxygen species. Red upwards and downwards arrows indicate increase and decrease, respectively. Diagram was created with URL (BioRender.com) (accessed on 27 October 2024).
Antioxidants 13 01330 g001
Figure 2. Mitochondrial metabolism in cardiac hemostasis. During cardiac hemostasis and under physiological conditions, the mitochondria in the heart produce 95% ATP by metabolizing a variety of fuels including fatty acids, glucose, lactate, ketones, pyruvate, and amino acids, primarily by mitochondrial oxidative phosphorylation (OXPHOS) through electron transport chain (ETC, or mitochondrial complex I–V). Tricarboxylic acid (TCA) cycle produces multiple metabolites (particularly NADH and FADH2), which enter ETC and function as electron carriers for ATP production. Additionally, 5% ATP is generated by anaerobic glycolysis. Diagram was created with URL (BioRender.com (accessed on 30 October 2024)).
Figure 2. Mitochondrial metabolism in cardiac hemostasis. During cardiac hemostasis and under physiological conditions, the mitochondria in the heart produce 95% ATP by metabolizing a variety of fuels including fatty acids, glucose, lactate, ketones, pyruvate, and amino acids, primarily by mitochondrial oxidative phosphorylation (OXPHOS) through electron transport chain (ETC, or mitochondrial complex I–V). Tricarboxylic acid (TCA) cycle produces multiple metabolites (particularly NADH and FADH2), which enter ETC and function as electron carriers for ATP production. Additionally, 5% ATP is generated by anaerobic glycolysis. Diagram was created with URL (BioRender.com (accessed on 30 October 2024)).
Antioxidants 13 01330 g002
Figure 3. Mitochondrial metabolism in HFrEF and HFpEF. A healthy heart uses fatty acids (FAs) as its primary energy fuel to produce ATP through fatty acid oxidation (FAO), while glucose becomes the main energy fuel for ATP generation in heart failure via glucose oxidation. Although impaired mitochondrial oxidative phosphorylation (OXPHOS) is a common feature for both HFrEF and HFpEF, evidence suggests that the failing heart in HFpEF attempts to compensate for the impaired OXPHOS by increasing BCAA (branched chain amino acid) rather than glucose oxidation. The latter is predominately observed in HFrEF. Multiple molecules (such as Sirt3, NAD+, ketone body β-hydroxybutyrate, SGLT2 inhibitor) have been reported as the key regulators for mitochondrial metabolism, which could be the novel therapeutics for HFpEF. Interestingly, PCSK9 (proprotein convertase subtilisin/kexin type 9) genetic deficiency may cause HFpEF by impairing FAO, raising caution for the clinical applications of PCSK9 inhibitors. Red upwards and downwards arrows indicate increase and decrease, respectively. Diagram was created with URL (BioRender.com) (accessed on 27 October 2024).
Figure 3. Mitochondrial metabolism in HFrEF and HFpEF. A healthy heart uses fatty acids (FAs) as its primary energy fuel to produce ATP through fatty acid oxidation (FAO), while glucose becomes the main energy fuel for ATP generation in heart failure via glucose oxidation. Although impaired mitochondrial oxidative phosphorylation (OXPHOS) is a common feature for both HFrEF and HFpEF, evidence suggests that the failing heart in HFpEF attempts to compensate for the impaired OXPHOS by increasing BCAA (branched chain amino acid) rather than glucose oxidation. The latter is predominately observed in HFrEF. Multiple molecules (such as Sirt3, NAD+, ketone body β-hydroxybutyrate, SGLT2 inhibitor) have been reported as the key regulators for mitochondrial metabolism, which could be the novel therapeutics for HFpEF. Interestingly, PCSK9 (proprotein convertase subtilisin/kexin type 9) genetic deficiency may cause HFpEF by impairing FAO, raising caution for the clinical applications of PCSK9 inhibitors. Red upwards and downwards arrows indicate increase and decrease, respectively. Diagram was created with URL (BioRender.com) (accessed on 27 October 2024).
Antioxidants 13 01330 g003
Figure 4. Mitochondrial quality control (MQC). MQC involved in mitochondrial biogenesis (new mitochondria generation), dynamics (fission and fusion), and mitophagy (degradation of the damaged or dysfunctional mitochondria) is a critical mechanism for protecting mitochondria and cellular functions. Decreased mitochondrial biogenesis, increased mitochondrial fission, and impaired mitophagy underscore HFpEF pathogenesis. Red upwards and downwards arrows indicate increase and decrease, respectively. Diagram was created with URL (BioRender.com) (accessed on 27 October 2024).
Figure 4. Mitochondrial quality control (MQC). MQC involved in mitochondrial biogenesis (new mitochondria generation), dynamics (fission and fusion), and mitophagy (degradation of the damaged or dysfunctional mitochondria) is a critical mechanism for protecting mitochondria and cellular functions. Decreased mitochondrial biogenesis, increased mitochondrial fission, and impaired mitophagy underscore HFpEF pathogenesis. Red upwards and downwards arrows indicate increase and decrease, respectively. Diagram was created with URL (BioRender.com) (accessed on 27 October 2024).
Antioxidants 13 01330 g004
Figure 5. Mitochondrial ROS generation and potential detrimental effects. Although ROS can be generated by other cellular oxidative systems such as xanthine oxidase (XO), NADPH oxidase (Noxs), and uncoupled NOS, mitochondrial ROS (mtROS) generated from complexes I and III are the major source for cellular ROS. Cells including cardiomyocytes have an efficient antioxidant system situated within the mitochondrial matrix, neutralizing the “excessive” mtROS to fine-tuning cellular redox signaling. Imbalanced redox signaling generates excessive mtROS, which in turn oxidize DNAs, proteins, lipids, and mtDNAs and alter their corresponding functions, thereby causing a variety of cellular dysfunctions, injuries, and even death. Green upwards and downwards arrows indicate increase and decrease, respectively. Diagram was created with URL (BioRender.com) (accessed on 27 October 2024).
Figure 5. Mitochondrial ROS generation and potential detrimental effects. Although ROS can be generated by other cellular oxidative systems such as xanthine oxidase (XO), NADPH oxidase (Noxs), and uncoupled NOS, mitochondrial ROS (mtROS) generated from complexes I and III are the major source for cellular ROS. Cells including cardiomyocytes have an efficient antioxidant system situated within the mitochondrial matrix, neutralizing the “excessive” mtROS to fine-tuning cellular redox signaling. Imbalanced redox signaling generates excessive mtROS, which in turn oxidize DNAs, proteins, lipids, and mtDNAs and alter their corresponding functions, thereby causing a variety of cellular dysfunctions, injuries, and even death. Green upwards and downwards arrows indicate increase and decrease, respectively. Diagram was created with URL (BioRender.com) (accessed on 27 October 2024).
Antioxidants 13 01330 g005
Figure 6. Mitochondrial ROS in mitohormesis and mitochondrial dysfunction in cardiomyocytes (CMs). Fine-tuning regulation of mitochondrial ROS (mtROS) is critical for cellular functions. While mitohormesis with low-to-mild levels of mtROS could activate protective pathways, improve cellular resilience against stress, and increase lifespan, excessive mtROS have detrimental effects on CMs such as hypertrophy, fibrosis, and apoptosis, eventually causing HFpEF. Green upwards and downwards arrows indicate increase and decrease, respectively. Diagram was created with URL (BioRender.com) (accessed on 27 October 2024).
Figure 6. Mitochondrial ROS in mitohormesis and mitochondrial dysfunction in cardiomyocytes (CMs). Fine-tuning regulation of mitochondrial ROS (mtROS) is critical for cellular functions. While mitohormesis with low-to-mild levels of mtROS could activate protective pathways, improve cellular resilience against stress, and increase lifespan, excessive mtROS have detrimental effects on CMs such as hypertrophy, fibrosis, and apoptosis, eventually causing HFpEF. Green upwards and downwards arrows indicate increase and decrease, respectively. Diagram was created with URL (BioRender.com) (accessed on 27 October 2024).
Antioxidants 13 01330 g006
Figure 7. Mitochondrial ROS in HFpEF. In HFpEF, there is a bidirectional crosstalk between mitochondria metabolic stress and inflammation, namely “mito-inflammation”. Excessive levels of mtROS initiate a vicious cycle of increased mtROS production, mitochondrial damage, and inflammation. Increased mtROS promotes the release of mtDNA and other mtDAMPs, which in turn activate proinflammatory signaling pathways, amplifying inflammation and further increasing mtROS generation. In HFpEF with diabetes, high levels of glucose and increased mtROS trigger glycoxidative stress, a detrimental positive feedback cycle of malignant AGE and ROS accumulation. The resulting mito-inflammation causes endothelial dysfunction, vasodilation impairment, and capillary rarefaction, all contributing to HFpEF pathogenesis. Moreover, the changed oxidative substrates, increased ROS, and defective mitochondria due to metabolic disruption and system inflammation also cause energy depletion, underscoring their role in endothelial dysfunction and HFpEF. Green upwards and downwards arrows indicate increase and decrease, respectively. Diagram was created with URL (BioRender.com) (accessed on 27 October 2024).
Figure 7. Mitochondrial ROS in HFpEF. In HFpEF, there is a bidirectional crosstalk between mitochondria metabolic stress and inflammation, namely “mito-inflammation”. Excessive levels of mtROS initiate a vicious cycle of increased mtROS production, mitochondrial damage, and inflammation. Increased mtROS promotes the release of mtDNA and other mtDAMPs, which in turn activate proinflammatory signaling pathways, amplifying inflammation and further increasing mtROS generation. In HFpEF with diabetes, high levels of glucose and increased mtROS trigger glycoxidative stress, a detrimental positive feedback cycle of malignant AGE and ROS accumulation. The resulting mito-inflammation causes endothelial dysfunction, vasodilation impairment, and capillary rarefaction, all contributing to HFpEF pathogenesis. Moreover, the changed oxidative substrates, increased ROS, and defective mitochondria due to metabolic disruption and system inflammation also cause energy depletion, underscoring their role in endothelial dysfunction and HFpEF. Green upwards and downwards arrows indicate increase and decrease, respectively. Diagram was created with URL (BioRender.com) (accessed on 27 October 2024).
Antioxidants 13 01330 g007
Table 1. Animal models of HFpEF.
Table 1. Animal models of HFpEF.
Animal Models
Hypertensive PhenotypeCardio–Metabolic Phenotype
FeaturesRodentLarge AnimalRodentLarge Animal
ModelDSSRsSHRsAortic ConstrictionAldosterone InfusionAged Dogs + PerinephritisAortic-Banded CatsZSF1 Ageing l-NAME + HFDDOCA + Salt-Loaded Pigs + HFD
HypertensionYYNYYYYNYY
Pulmonary CongestionNNMildMild-YY-Y-
Diastolic DysfunctionYYMildYYYYYYY
LVHYYYYYYYMildYY
Exercise IntoleranceYY----YYYY
ObesityNNNNNNYNYY
Preserved LVEFY *Y *Y *Y *YY *YYYY
DOCA, deoxycorticosterone acetate; DSSRs, Dahl salt-sensitive; HFD, high-fat diet; l-NAME, N-nitro-l-arginine methyl ester; SHRs, spontaneously hypertensive rats; LVH, left ventricular hypertrophy; LVEF, left ventricular ejection fraction. * During the initial phase, the LVEF is preserved.
Table 2. Pharmacological and non-pharmacological management of HFpEF.
Table 2. Pharmacological and non-pharmacological management of HFpEF.
TherapeuticsTarget(s)RationaleMajor Findings/OutcomeMain Adverse Effects
Non-pharmacological management
Dietary interventions; low carbohydrate dietAdipose tissue (especially visceral fat), liver, and pancreasImprove insulin sensitivity; reduce inflammation and glycoxidative stressImprove MetS manifestations and cardiovascular riskRisk of hypoglycemia in individuals on medication
Exercise trainingCardiac and skeletal muscles Improve LV diastolic function and cardiac output Improvement in VO2peak peak
and physical capacity
Muscle soreness, fatigue, and injury from over-exertion
Conventional/repurposing drugs
DiureticsKidneys
Cardiovascular system
Alleviate symptoms of fluid overloadReduce congestion;
control of blood pressure and pulmonary congestion
Electrolyte imbalance;
hypotension
SGLT2 inhibitorsKidneysReduce glucose and sodium reabsorption, leading to natriuresis and osmotic diuresisReduce fluid overload, inflammation, and oxidative stress;
improve cardiac energy efficiency;
reduce hospitalization and cardiovascular death
Volume depletion;
electrolyte imbalance
Mineralocorticoid receptor antagonistsMineralocorticoid receptors in the kidneys, heart, and blood vesselsReduce sodium retention, lower blood pressure, and alleviate fluid overload;
reduce myocardial fibrosis
Reduce hospitalizations;
reduce myocardial remodeling
Hyperkalemia;
gynecomastia;
renal impairment
Angiotensin receptor–neprilysin inhibitorsAngiotensin II type 1 receptor and neprilysinReduce vascular stiffness and blood pressure; increase natriuresisReduce hospitalizations;
reverse cardiac remodeling
Hypotension;
hyperkalemia;
angioedema
Angiotensin receptor blockersAngiotensin II type 1 receptor/Renin–angiotensin–aldosterone systemReduce the heart’s workload by lowering blood pressureReduce hospitalizationsHypotension
Hyperkalemia
Renal dysfunction
β-Blockersβ-adrenergic receptors (heart and blood vessels)Reduce heart rate and improve diastolic filling timeSlight reduction in hospitalizations;
can improve exercise tolerance
Bradycardia;
hypotension
Mitochondria oxidative stress-targeted therapy
Elamipretide–BendaviaMitochondriaImprove mitochondrial bioenergetics;
improve heart’s energy capacity
Improve mitochondrial function and cardiac remodelingInjection site reactions
Coenzyme Q10MitochondriaImprove energy production;
reduce oxidative stress
Improve LVEF and exercise intolerance;
may reduce hospitalization
Gastrointestinal issues
MitoQ: mitoquinoneMitochondriaNeutralize oxidative stress;
restore mitochondrial function
Enhance mitochondrial respiration;
may improve exercise tolerance and reduce fatigue
Gastrointestinal issues
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Martinez, C.S.; Zheng, A.; Xiao, Q. Mitochondrial Reactive Oxygen Species Dysregulation in Heart Failure with Preserved Ejection Fraction: A Fraction of the Whole. Antioxidants 2024, 13, 1330. https://doi.org/10.3390/antiox13111330

AMA Style

Martinez CS, Zheng A, Xiao Q. Mitochondrial Reactive Oxygen Species Dysregulation in Heart Failure with Preserved Ejection Fraction: A Fraction of the Whole. Antioxidants. 2024; 13(11):1330. https://doi.org/10.3390/antiox13111330

Chicago/Turabian Style

Martinez, Caroline Silveira, Ancheng Zheng, and Qingzhong Xiao. 2024. "Mitochondrial Reactive Oxygen Species Dysregulation in Heart Failure with Preserved Ejection Fraction: A Fraction of the Whole" Antioxidants 13, no. 11: 1330. https://doi.org/10.3390/antiox13111330

APA Style

Martinez, C. S., Zheng, A., & Xiao, Q. (2024). Mitochondrial Reactive Oxygen Species Dysregulation in Heart Failure with Preserved Ejection Fraction: A Fraction of the Whole. Antioxidants, 13(11), 1330. https://doi.org/10.3390/antiox13111330

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