Next Article in Journal
Deep Learning Approaches for lncRNA-Mediated Mechanisms: A Comprehensive Review of Recent Developments
Previous Article in Journal
Hyaluronic Acid: A Powerful Biomolecule with Wide-Ranging Applications—A Comprehensive Review
Previous Article in Special Issue
Impact of Prenatal Exposure to Maternal Diabetes and High-Fat Diet on Postnatal Myocardial Ketone Body Metabolism in Rats
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Unravelling the Interplay between Cardiac Metabolism and Heart Regeneration

1
National Heart Research Institute Singapore, National Heart Centre Singapore, Singapore 169609, Singapore
2
Cardiovascular & Metabolic Disorders Program, Duke-National University of Singapore Medical School, Singapore 169857, Singapore
3
Yong Loo Lin School of Medicine, National University of Singapore, Singapore 119077, Singapore
4
The Hatter Cardiovascular Institute, University College London, London WC1E 6HX, UK
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(12), 10300; https://doi.org/10.3390/ijms241210300
Submission received: 30 May 2023 / Revised: 14 June 2023 / Accepted: 16 June 2023 / Published: 18 June 2023
(This article belongs to the Special Issue Cardiac Metabolism in Heart Failure)

Abstract

:
Ischemic heart disease (IHD) is the leading cause of heart failure (HF) and is a significant cause of morbidity and mortality globally. An ischemic event induces cardiomyocyte death, and the ability for the adult heart to repair itself is challenged by the limited proliferative capacity of resident cardiomyocytes. Intriguingly, changes in metabolic substrate utilisation at birth coincide with the terminal differentiation and reduced proliferation of cardiomyocytes, which argues for a role of cardiac metabolism in heart regeneration. As such, strategies aimed at modulating this metabolism-proliferation axis could, in theory, promote heart regeneration in the setting of IHD. However, the lack of mechanistic understanding of these cellular processes has made it challenging to develop therapeutic modalities that can effectively promote regeneration. Here, we review the role of metabolic substrates and mitochondria in heart regeneration, and discuss potential targets aimed at promoting cardiomyocyte cell cycle re-entry. While advances in cardiovascular therapies have reduced IHD-related deaths, this has resulted in a substantial increase in HF cases. A comprehensive understanding of the interplay between cardiac metabolism and heart regeneration could facilitate the discovery of novel therapeutic targets to repair the damaged heart and reduce risk of HF in patients with IHD.

1. Introduction

Ischemic heart disease (IHD) is the most prevalent cardiovascular disease and remains the leading cause of death globally. Although IHD-related mortality rates have decreased over time due to primary prevention, and improved diagnosis and treatment, the rise in absolute numbers of IHD cases is a serious cause for concern. Crucially, IHD can increase the risk of heart failure (HF) by 8-fold and, as expected, is the most frequent underlying cause of HF [1,2]. HF itself affects ~64 million individuals and is a leading cause of hospitalisation [3]. Currently, HF affects 1–1.3% of the global population and is expected to rise to 3% and affect more than 70% of those aged above 65 by 2030 [4]. While the ageing of the population will undoubtedly contribute to this foreseeable increase in HF prevalence, improved survival outcomes in IHD patients with acute myocardial infarction (AMI) may also contribute to an increase in HF cases [4,5]. As such, there is an unmet need to identify novel therapeutic strategies to prevent the onset and progression of HF in patients with IHD.
Myocardial ischemia instigates profound derangement in cellular energetics and metabolism in the heart which induces injury and eventual death of cardiomyocytes [6,7]. Cardiomyocytes undergo apoptosis at a rate of 0.002% in normal human hearts, 0.12–0.7% in failing hearts from patients with NYHA class III and IV [8,9,10], and a staggering 17% in ischemic hearts [10,11]. The adult human heart is one of the least regenerative organs, but it is widely accepted that the heart’s regenerative ability is preserved in mammals during the early neonatal stage, and throughout the entire lifespan in certain lower vertebrate species [12,13,14]. In support, cardiomyocyte proliferation has been observed in mice subjected to apical resection and MI within the first week of birth [14,15], but this regenerative window was restricted to 2 days postpartum in larger mammals [16,17]. Heart regeneration has also been observed in a human newborn with severe MI where functional cardiac recovery had occurred within weeks after the initial extensive myocardial damage, translating into long-term normal cardiac function [18]. While the ability to promote heart regeneration in the setting of AMI is appealing, very little is known about the molecular pathways underlying cardiomyocyte proliferation. Alternatively, several efforts have been made to transplant various cell types into the infarcted heart. However, suboptimal delivery, homing, engraftment, and survival of these transplanted cells remain a major issue for it to be considered as a viable therapeutic modality [19]. Interestingly, marked changes in cardiac metabolism have been found to occur during heart development as evidenced by the predominant utilisation of glycolysis in foetal hearts and oxidative phosphorylation in adult hearts, the latter coinciding with terminal differentiation and reduced proliferation of cardiomyocytes [20,21]. These findings lend support to the assumption that the transition from glycolytic to oxidative metabolism, and the resultant increase in reactive oxygen species (ROS) production is a key driver of DNA damage and cell cycle arrest in cardiomyocytes [22].
With the aim of understanding the complexities associated with cardiac metabolism and heart regeneration, we review the role of metabolic substrates and mitochondria in heart regeneration, and discuss potential targets aimed at promoting cardiomyocyte cell cycle re-entry.

2. The Role of Metabolic Substrates in Heart Regeneration

The heart beats around 100,000 times each day and consumes ~8% of total ATP produced by the body [11,23]. Despite being one of the most energy-consuming organs, the heart stores limited amounts of ATP, which are sufficient to maintain its function for only a few seconds in the presence of nutrient shortage [24]. In order to meet the high energy demands, the heart can utilise a variety of energy substrates (e.g., fatty acids, glucose, ketones, and amino acids), albeit at different proportions. Importantly, the heart can readily switch to a specific type of fuel during cardiac development, and in response to physiological and pathological stress [25]. In this section, we focus on the major metabolic substrates utilised by the heart and discuss their potential roles in heart regeneration.

2.1. Glucose Metabolism

Cells with high proliferative capacity display elevated glycolic rates for energy production, despite the presence of adequate oxygen [26]. Glycolysis is an inefficient way to generate energy as only 2 ATPs are generated per molecule of glucose (in contrast to 36 ATPs generated by oxidative phosphorylation); however, this metabolic pathway is a major contributor for the biosynthesis of cellular components, such as lipids, amino acids, and nucleotides [26]. Studies in zebrafish, and in neonatal and adult mice have shown a pro-glycolytic metabolic profile is favourable for heart regeneration. In support, analysis of the transcriptome and proteome profile in zebrafish at 7 days post-cryoinjury revealed an alteration in cardiomyocyte metabolism from mitochondrial oxidative phosphorylation to glycolysis at the border and remote zones [27]. In other studies, lactate, an end product of glycolysis, promoted cell cycle progression in neonatal mouse cardiomyocytes and in human induced pluripotent stem-cell-derived cardiomyocytes (iPSC-CMs) by regulating the expression of genes involved in cell fate and proliferation [28].
In neonatal and healthy adult hearts, glycolysis contributes to ~44% and ~5% of total ATP produced, respectively [29,30]. The cardiomyocyte plasma membrane is impermeable to glucose; as such, glucose uptake is mediated by glucose transporters (GLUTs) of which 14 members have been identified in various tissues to date. GLUT1 and GLUT4 are the most abundantly expressed isoforms in the heart [31], and notably, both transporters display an expression profile coinciding with different stages of cardiac development, with GLUT1 and GLUT4 being predominantly expressed in foetal and adult hearts, respectively [32]. The upregulation of GLUT1 has been found to play important roles in cardiac development and in response to cardiac stress [33,34]. Consistently, cardiac-specific overexpression of GLUT1 increased the percentage of proliferative cardiomyocytes and reduced fibrosis in cryoinjured neonatal mice by promoting nucleotide biosynthesis [35] (Figure 1). While these findings support the upregulation of GLUT1 as a potential mediator of heart regeneration, it is important to note that glycolysis is a fundamental metabolic pathway in active inflammatory cells (e.g., neutrophils, proinflammatory macrophages) [36], and inhibition of GLUT1 has been considered a potential therapeutic intervention for attenuating pro-inflammatory responses following AMI [37].
Besides their role in metabolism, certain glycolytic enzymes have been found to be interlinked with cell cycle regulatory pathways. Pyruvate kinase (PK) is an important enzyme which regulates the conversion of phosphoenolpyruvate and ADP to pyruvate and ATP [38]. These enzymes also show developmental-stage-specific expression profiles, with the PK muscle isoform 2 (PKM2) being predominantly expressed during embryonic and postnatal development, and PKM1 being the dominant isoform in adult cardiomyocytes [39]. Interestingly, cardiomyocyte-specific overexpression of PKM2 in post-MI adult mice has been shown to increase cardiomyocyte proliferation, and improved cardiac function and long-term survival by elevating targets downstream of β-catenin signalling [39]. Conversely, PKM2 expression has been found to be upregulated in stressed hearts where it has contributed to HF [40], while inhibition of the PKM2/β-catenin axis in post-MI mice reduced infarct size, increased the percentage of proliferative cardiomyocytes, improved mitochondrial function, and enhanced angiogenesis, which was attributed to the activation of target genes associated with cell proliferation [41]. Pyruvate dehydrogenase kinases (PDKs) are another family of enzymes with important roles in regulating pyruvate metabolism and metabolic flexibility [42]. Transcriptomic analyses in zebrafish revealed PDK enzymes, PDK2b, PDK3b, and PDK4b are up-regulated at the border zone in cryoinjured hearts, and overexpression of PDK3 increased the number of proliferating cardiomyocytes, albeit with no reduction in scared regions [43] (Figure 1).
Other signalling pathways not directly involved in glycolytic metabolism have also been implicated in cardiomyocyte proliferation. For instance, the Nrg1/ErbB pathway is indispensable for cardiac development as it regulates cardiomyocyte growth and survival and mediates terminal differentiation of cardiomyocytes from iPSCs [44,45]. The Nrg1/ErbB2 axis has also been shown to mediate profound changes in metabolism by up-regulating glycolytic genes at the border zone of cryoinjured zebrafish hearts [46] (Figure 1). Furthermore, cardiac-specific overexpression of ErbB2 after 3 weeks of permanent left anterior descending (LAD) coronary artery ligation in mice promoted de-differentiation and proliferation of cardiomyocytes, resulting in improved cardiac function [47]. While it is unclear how this Nrg1/ErbB2 axis mediates metabolic remodelling, it could be through the modulation of fatty acid metabolism as transient overexpression of ErbB2 in a breast cancer cell line promoted glycolysis and cell migration by upregulating fatty acid synthase involved in neoplastic lipogenesis [48,49].
Collectively, these findings suggest a switch from glucose oxidation to glycolysis is sufficient to promote cell cycle re-entry in adult cardiomyocytes (Figure 1). However, considering the ubiquitous expression of GLUTs in multiple cardiac cell types and that certain cell cycle activators are associated with cardiac hypertrophy and cancer [34,50,51], future strategies aimed at targeting these pathways should be investigated in a cell-type-specific manner to circumvent potential off-target effects.

2.2. Fatty Acid Metabolism

The foetal heart barely utilises fatty acids for efficient ATP production, which could be attributed to its low mitochondrial content and the limited availability of fatty acids in the placenta [52]. However, immediately after birth, the infant heart switches rapidly to fatty acid β-oxidation (FAO) [52,53], and this metabolic pathway generates the majority of ATP (~40–60% of total ATP) throughout adulthood [30]. Compared to other substrates, FAO generates the majority of ATP; however, as this process consumes the largest amount of oxygen, fatty acids are considered the least efficient energy substrate [54].
Long-chain fatty acid (LCFA) uptake into the cytosol is mediated by the fatty acid transport protein, CD36/FAT [54], and although the role of CD36 in cardiomyocyte proliferation is unclear, CD36 knockdown in mouse endothelial cells (ECs) has been found to prevent angiogenesis and vascular repair in response to hindlimb ischemia [55]. ECs play important roles in establishing an intact vasculature and in guiding cardiomyocyte organization in response to injury [56], while angiogenesis is essential for heart regeneration, as it restores blood flow to damaged myocardial tissue [57,58,59]. EC-specific inhibition of Notch signalling has been shown to impair fatty acid transport, resulting in abnormalities of the cardiac metabolome and vascular densities in adult mouse hearts [60]. In this study, genetic ablation of RBP-Jκ (a core component of the Notch pathway) reduced LCFA transport to the heart, resulting in a switch to glucose metabolism as the main source of energy, which, in turn, promoted cardiac hypertrophy and HF. Indeed, several studies support Notch signalling as a critical mediator of angiogenesis and cardiomyocyte proliferation [61,62], but it is unclear whether this regenerative ability is regulated by cardiac metabolism.
In the cytosol, LCFAs are esterified to fatty acyl CoAs by fatty acyl coA synthase and the resulting acyl groups are transferred into the mitochondria for FAO [54]. Acyl coA synthetase long-chain family member 1 (ACSL1) is a key rate-limiting enzyme which regulates LCFA uptake rates by increasing esterification to form fatty acyl CoA [63,64] (Figure 1). Interestingly, the expression of ACSL1 has been found to increase with age while the expression of glucose metabolism-related enzymes (glucose-6-phosphate 1-dehydrogenase, hexokinase 1, hexokinase 3, PKM) decrease with age [65]. These findings allow for the speculation that increased FAO during ageing may be an important suppressor of heart regeneration. In support, cardiac-specific knockdown of ACSL1 in neonatal mice promoted cardiomyocyte proliferation even at 60 days of age as evidenced by increased expression of cell proliferation markers [65]. Similarly, knockdown of ACSL1 in post-MI adult mice improved cardiac function by inducing cardiomyocyte proliferation. Mechanistic studies have revealed that inhibition of ACSL1 mediates cell proliferation via an Akt-FoxO1 axis, as suppression of this pathway decreased the expression of positive cell cycle regulators, cyclin B1, cyclin D2 and cyclin-dependent kinase 1 [65]. While these findings suggest a decrease in FAO is necessary for cardiomyocyte proliferation, other studies have shown that increased FAO (mediated by ACSL1) improves re-endothelisation after vessel injury. Furthermore, cyclic mechanical stretching of endothelial progenitor cells (EPCs) enhances vascular adhesion and endothelial differentiation by activating ACSL1 to increase FAO, while transplantation of ACSL1-overexpressing EPCs in rats with carotid injury results in improved vascular homing and repair [66].
Peroxisome-proliferator-activated receptors (PPARs) are key regulators of FAO [67]. PPARα is a master ligand-activated transcriptional factor that coordinates the expression of lipid metabolism genes, such as CD36 and mitochondrial FAO enzymes (e.g., carnitine palmitoyltransferases, acyl-CoA dehydrogenases) [68] (Figure 1). Experimental evidence suggests that PPARα plays a biphasic role in cell proliferation, as administration of the PPARα agonist, GW7647, was initially found to stimulate cardiomyocyte proliferation in infant mice at P4, but later promoted cardiomyocyte hypertrophy and binucleation, and reduced cardiomyocyte proliferation rates at P5 [69]. These findings suggest PPARα may have a more prominent role in the terminal differentiation of cardiomyocytes. Indeed, increased expression of PPARα has been observed during the differentiation of mouse embryonic stem cells into cardiomyocytes, while pharmacological inhibition of PPARα prevented this process, as evidenced by a decreased expression of cardiac specific genes [70].
Collectively, these findings support a role for FAO in promoting metabolic maturation in the adult heart with concurrent reduction in regenerative capacity [71,72]. It remains controversial whether inhibition of FAO is sufficient to extend the regenerative window as this may also delay cardiomyocyte maturation. Hence, the interplay between metabolic transitions and cardiomyocyte maturation needs to be carefully considered as inhibition of FAO at inopportune settings could result in adverse outcomes. In support, analysis of endomyocardial biopsies from failing hearts has revealed a decrease in PPARA mRNA, implying a decrease in FAO in the setting of HF [73]. Finally, recent findings suggest that neither GW7647 nor the FAO inhibitor, etomoxir, could promote cardiomyocyte proliferation in post-MI mouse hearts [74].

2.3. Ketone Body Metabolism

Ketone bodies such as acetoacetate, β-hydroxybutyrate, and acetone accumulate in the systemic circulation under conditions of prolonged fasting, insulin deprivation, and extreme exercise [75,76,77], with recent evidence supporting an increased utilisation of ketones in failing hearts [78]. The role of ketones in the developing heart is less clear. Comparative proteomics analysis of cells from three stages of cardiac differentiation (iPSCs, cardiac progenitor cells, and cardiomyocytes) has revealed ketogenic substrates are upregulated as result of increased expression of ketogenic enzymes, 3-hydroxymethyl-3-methylglutaryl-CoAlyase, 3-hydroxy-3-methylglutaryl-CoA synthase 2 (HMGCS2), and 3-hydroxybutyrate dehydrogenase 1 [79]. In another multi-omics study, increased ketogenesis was found to occur in neonatal mouse hearts at P7 when compared to hearts at E18.5 [80]. Interestingly, the expression of HMGCS2 (a rate-limiting enzyme of ketogenesis) gradually decreased after weaning and its levels reached normal physiological levels by postnatal day 56. Since the occurrence of early ketogenesis coincided with the regenerative window in neonatal mouse hearts, it could be speculated that ketone body utilisation plays a role in cardiomyocyte proliferation. In support of this idea, overexpression of HMGCS2 improved cardiac function in post-MI mice by increasing the percentage of phospho-histone H3 (PHH3)+ (a cell proliferation marker) cardiomyocytes [81]. Furthermore, ketone body oxidation was shown to enhance EC proliferation and angiogenesis in vitro and in mice subjected to pressure overload [82]. Whether this enhanced angiogenic capacity of ECs is sufficient to improve cardiac function remains to be validated as left ventricular (LV) ejection fraction was not investigated in this study, while the LV anterior wall thickness at end-diastole remained thickened even after ketogenic diet in the setting of pressure overload. As the primary concern for promoting ketone body oxidation is the risk of acidosis [83], future studies should investigate the safety range of circulating ketone bodies to maximise their potential benefits in heart regeneration, whilst preventing adverse effects.

2.4. Amino Acid Metabolism

Although amino acids are one of the smallest contributors of ATP production (~2% of total ATP) [84], they are intricately involved in several pathological conditions [30]. High circulating levels of the branched-chain amino acids (BCAAs), leucine, valine, and isoleucine, are associated with insulin resistance, type 2 diabetes, coronary artery diseases, and HF [85,86]. Interestingly, BCAAs were found to modulate liver regeneration and function in patients who had undergone hepatectomy [87]. BCAAs and their metabolites mediate cell growth, proliferation, and tumour progression by activating the mammalian target of rapamycin complex 1 (mTORC1) pathway [88] (Figure 1). Leucine, in particular, is a key stimulator of mTORC1-mediated cell growth through blockage of the mTORC1 inhibitor, sestrin 2 [89]. Given that mTORC1 is a mediator of cardiac hypertrophy [90], BCAAs may have an undefined role in cardiomyocyte proliferation and growth.
Multi-omics analysis has revealed that glutamine is enriched during heart regeneration in zebrafish and in neonatal mice but reduced in adult mice which have lost regenerative capacity [91]. This dynamic change in glutamine expression was found to correlate with the regulation of mTORC1, which plays an important role in cell growth and proliferation [92,93]. Interestingly, activation of the Wnt/β-catenin pathway in regenerating zebrafish hearts rescued the negative effects that mTORC1 inhibition exerted on cardiomyocyte proliferation [91] (Figure 1). Given that the Wnt/β-catenin pathway is a master regulator of cardiac development and cardiomyocyte terminal differentiation [94], further studies are needed to tease out the interplay between Wnt/β-catenin signalling and glutamine in cardiomyocyte proliferation [95].
In summary, the adult heart is incapable of repairing damaged tissue, which is, in part, attributed to the switch in metabolic substrate utilisation after birth. Accumulating evidence supports the possibility that multiple metabolic components are involved in the regulation of cardiomyocyte proliferation. Indeed, conditions of high glycolysis and reduced FAO have been proposed to promote heart regeneration, but the roles of ketone body and amino acid metabolism in cardiomyocyte proliferation is controversial and unclear. As such, future studies are needed to elucidate the dynamic interplay between these metabolic pathways and the cell proliferation pathways which they regulate.

3. The Role of Mitochondria in Heart Regeneration

The heart, a perpetually active organ with substantial energy needs relies on a dense and highly functional mitochondrial network to maintain energy requirements and overall performance [23,96]. However, the role of mitochondria in the heart extends beyond energy production, as these dynamic organelles also orchestrate a range of cellular functions, including signal transduction, calcium regulation, oxidative stress management, and apoptosis [97,98,99]. Recent findings have begun to shed light on the intricate link between mitochondria metabolism and cardiovascular diseases [100,101,102]. This includes understanding how early defects in mitochondrial oxidative phosphorylation manifest in HF, and the pivotal role of mitochondrial metabolic impairment in MI-induced cardiac damage [103,104,105].
The role of mitochondria has extended into the areas of cell fate determination and development [106]. Crucial functions of mitochondria in stem cells have been highlighted in several reports [107,108], indicating mitochondrial characteristics, such as morphology, localization, abundance, and function, could serve as markers of pluripotency, underscoring the multifaceted roles these organelles play in cardiac health and disease [106,109].

3.1. Oxidative Phosphorylation and Reactive Oxygen Species

Mitochondria are the primary sites of oxidative phosphorylation, a process that generates ATP with superior efficiency compared to glycolysis [110]. However, this increased efficiency is linked with the endogenous generation of ROS as a by-product of ATP production due to electron leakage [111,112]. Overproduction of ROS can lead to oxidative DNA damage in cardiomyocytes, triggering a cell cycle checkpoint and arresting the cell cycle [22,113]. Notably, mitochondrial dysfunction, such as when induced by mitochondrial transcription factor A inactivation, has been shown to elevate ROS production, activate the DNA damage response, and induce cardiomyocyte cell cycle arrest, eventually leading to lethal cardiomyopathy [114]. Meanwhile, elevated ROS levels have also been linked to disturbances in mitochondrial and antioxidant proteins, leading to cardiac hypertrophy [115]. On the contrary, when using microRNA or CRISPR/Cas9 technology to silence key genes involved in the mitochondrial electron transport chain or tricarboxylic acid cycle, a reduction in mitochondrial number was observed followed by a decrease in ROS and an increase in cardiomyocyte proliferation [116,117]. These findings support the causal relationship between disturbed mitochondrial function and ROS production, which eventually interferes with cardiomyocyte proliferation (Figure 2).
In heart diseases such as AMI, redox (reduction–oxidation) changes in the injured heart may affect the proliferation and differentiation of cycling cardiomyocytes or progenitor cells [118,119]. This indicates that ROS levels generally correlate with stem cell differentiation, and increased oxidative stress post-cardiac injury could potentially induce the terminal differentiation of glycolytic cardiac progenitor cells (CPCs) [120,121]. Furthermore, cardiac hypoxia, a common consequence of cardiac injury, may play a crucial role in the recruitment of cycling cells or their progeny to the injured site [122]. Ineffectively managed ROS levels are also reported to be associated with cardiac ageing, cardiomyopathy, and a decline in the CPC population, resulting in reduced cardiac cell turnover [123,124].
Given the critical roles of mitochondrial function and ROS production in cardiac health, targeting these mitochondrial injuries with an emphasis on reducing oxidative damage could offer a promising strategy to delay the progression of HF.

3.2. Hypoxia Conditioning

The heart was initially recognised as the least regenerative organ, but this notion has been significantly developed over the past few decades [125,126]. Accumulating evidence suggests the existence of progenitor cells, such as c-kit+ cells, stem cell antigen-1+ cells, side population (SP) cells, and cardio-sphere-derived cells, within the adult heart, albeit in limited numbers [22,127,128,129]. These resident CPCs have demonstrated remarkable capabilities for self-renewal and differentiation into multiple cardiovascular lineages, including endothelial, smooth muscle, and myocardial cells, and this phenomenon is observable both in vitro and in vivo [19,130].
Studies have shown that a hypoxic niche environment may regulate signalling pathways to sustain the dedifferentiation and survival of foetal cardiovascular progenitor cells [131,132], whereas high oxygen concentrations coincide with the stagnation of cardiomyocyte proliferation [133]. Moderate hypoxia (SaO2 75–85%) can also bolster cell cycle activities in postnatal human cardiomyocytes [134]. Along these lines, the preserved self-renewal and decreased mitochondrial ROS levels were further observed in murine CPCs residing in hypoxic niches [135]. Moreover, intracellular ROS production was found to be maintained at low levels in several resident CPCs in the adult heart to preserve their quiescence and/or multipotency [119].
Long-term systemic hypoxemia could potentially reduce mitochondrial respiration and consequent ROS production in adult cardiomyocytes, which may promote cardiomyocyte re-entry into the cell cycle, thereby stimulating the proliferation of terminally differentiated cardiomyocytes [119,136] (Figure 2). Interestingly, exercise such as treadmill running have been shown to effectively reduce cardiomyocyte ROS accumulation and induce mitochondrial uncoupling, which coincided with heart regeneration [133]. Similarly, gradual exposure to severe systemic hypoxemia in mice resulted in the inhibition of oxidative metabolism, decreased ROS production and oxidative DNA damage, and reactivation of cardiomyocyte mitosis [137]. Besides ROS, other potential mechanisms through which hypoxia attenuates cardiac stem cell apoptosis have been identified. For instance, low oxygen tension has been shown to stabilize and activate several critical transcription factors and signalling pathways, such as hypoxia inducible factor 1α (HIF-1α) and Yes-associated protein (YAP), which, in turn, induced metabolic remodelling towards glycolysis to facilitate cardiomyocyte proliferation [138,139]. Importantly, HIF-1α and YAP control glycolysis by regulating the expression of glycolytic enzymes, and deletion or downregulation of these players has been shown to impair glucose uptake and glycolysis, which decreased cardiomyocyte proliferation [138].
Hypoxia-induced proliferation has been shown to be sufficient for promoting heart regeneration following MI [113]. Hypoxemia exposure following MI appears to induce robust regeneration, reduce myocardial fibrosis, and improve LV systolic function [137]. Additionally, a moderate level of hypoxia in combination with a mitochondrial ROS scavenger reversed the hypertrophic growth of LV cardiomyocytes by inducing cell cycle re-entry in terminally differentiated cardiomyocytes, thereby resulting in a substantial recovery of cardiac function [112].
The therapeutic potential of hypoxia/antioxidant treatment extends beyond resident CPCs. Hypoxia conditioning and antioxidant treatment can also benefit transplanted cardiac stem/progenitor cells aimed at repairing infarcted hearts, as this approach is challenged by low survival rates of donor cells [140,141]. For instance, hypoxic preconditioning of c-kit+ CPCs improves their survival and homing after engraftment into an infarcted heart [142]. Similarly, overexpressing sulfiredoxin 1, reduces ROS generation, and mitochondrial membrane potential, while enhancing the primary antioxidant systems and increasing the migration, proliferation, and cardiac differentiation of CPCs [143]. Another antioxidant, MHY-1684, has been shown to enhance the angiogenic capacity of CPCs in the setting of ROS-related diabetic cardiomyopathy by decreasing hyperglycemia-induced mitochondrial ROS generation and attenuating mitochondrial fragmentation. These findings indicate a potential therapeutic role for antioxidants in modulating mitochondrial dynamics and function in response to oxidative stress [142].

3.3. Mitochondrial Quality Control

Mitochondrial dynamics, encompassing biogenesis, fusion, fission, and mitophagy are crucial for regulating mitochondrial morphology and maintaining mitochondrial health, especially for cellular responses to metabolic cues or environmental stresses [144,145]. Abnormal mitochondrial dynamics such as excessive mitochondrial fragmentation and impaired fusion have been implicated as drivers of HF and cardiac ischemia/reperfusion (IRI) injury [146,147], while maintenance of normal mitochondrial structure and function could facilitate heart regeneration [148]. There is growing evidence suggesting that mitochondrial quality control in cardiomyocytes can enhance cardiac function, save cardiomyocytes from death, and prevent worsening of cardiovascular diseases under external environmental stress [144,149,150] (Figure 2).
Studies have shown that the differentiation process of c-kit+ progenitor cells involves mitochondrial fragmentation, which is controlled by the calcineurin-Drp1 pathway [151], while an environmental stressor such as high glucose milieu can alter mitochondrial dynamics and increase the expression of fission-related proteins such as Fis1 and Drp1, leading to a significant decrease in the tube-forming ability of CPCs [152]. Interestingly, pharmacological inhibition of mitochondrial fragmentation can help to maintain the undifferentiated state of c-kit+ progenitor cells. Mitochondrial division inhibitor 1 (mdivi-1), which inhibits Drp1-dependent mitochondrial fission, has shown promise in enhancing the survival of human W8B2+ cardiac stem cells, but surprisingly, the cytoprotective effects of mdivi-1 in simulated IRI models does not appear to be governed by changes in mitochondrial morphology, membrane potential, or ROS production [153].
The balance of mitophagy and mitochondrial biogenesis is of great importance in cardiomyocyte proliferation and differentiation. Mitophagy is induced during differentiation of adult CPCs and is mediated by mitophagy receptors [154]. Disrupting BNIP3L- and FUNDC1-mediated mitophagy during differentiation leads to sustained mitochondrial fission and the formation of dysfunctional mitochondria, resulting in increased susceptibility to cell death and failure to survive in an infarcted heart [155]. In a hypoxia/reperfusion injury cellular model, transfection of miR-494-3p mimic (inhibitor of PGC1α) improved cardiomyocyte proliferation activity by inhibiting mitochondrial biogenesis, thereby preventing the occurrence of cardiomyocyte apoptosis and autophagy [156].
In summary, the intricate role of mitochondria in cardiac regeneration is woven into multiple facets of cellular processes. Mitochondrial oxidative phosphorylation and ROS play pivotal roles in powering the energy-demanding process of cardiomyocyte proliferation, whereas mitochondrial dynamics and mitophagy are crucial in upholding mitochondrial quality control, thereby, mutually facilitating the regeneration of robust cardiac cells. Enhancing our comprehension of the mechanisms underlying mitochondrial dysfunction and identifying innovative ways to bolster mitochondrial function and foster heart regeneration could pave the way for novel therapeutic strategies.

4. Potential Strategies to Promote Heart Regeneration through Metabolic Modulation

In the past decade, studies have focused on cell-based therapies to promote heart regeneration by administrating stem cells into injured hearts [157]. Several clinical trials have been conducted which evaluated the direct intracoronary or intramyocardial delivery of multiple sources of adult stem cells (e.g., bone marrow cells and adipose tissue derived cells) into the heart to regenerate the damaged myocardium [158,159]. However, there is no approved phase III clinical trial for AMI patients, and the results from smaller trials are inconsistent in terms of the efficacy and potential risk of administration routes [158,159]. As such, there is an unmet need to discover effective and safe therapies to promote cardiomyocyte proliferation and heart regeneration. In this section, we focus on potential targets that can modulate cardiac metabolism to promote heart regeneration (summarised in Table 1).

4.1. Long Non-Coding RNAs

Long non-coding RNAs (lncRNAs) are non-coding transcripts of >200 nucleotides in length and which are abundantly expressed in the cardiovascular system where they regulate cardiac development and disease [168,169]. LncRNAs are emerging as regulators of glucose and fatty acid metabolism, but only a few cardiac-specific lncRNAs have been investigated to date [170,171]. For instance, the cardiomyocyte-enriched lncRNA, LncHrt, has been shown to preserve cardiac metabolism and improve cardiac function in post-MI adult mice by activating the LKB1–AMPK pathway via sirtuin 2 (Sirt2) [172]. Importantly, lncRNAs have been found to regulate cardiomyocyte proliferation, angiogenesis, and heart regeneration by mediating transcriptional and epigenetic remodelling in the heart [173]. Bioinformatic assessment of neonatal mouse hearts has revealed differentially expressed lncRNAs at P1 and P7 that were associated with cardiac metabolism and cell proliferation [174]. Moreover, overexpression of Sirt1 antisense lncRNA in adult mouse hearts led to the stabilisation of Sirt1 mRNA which promoted an increase in ki67+ and PHH3+ cardiomyocytes [175]. Accumulating evidence support Sirt1 activity in the positive regulation of PPARα and PGC1α which promotes fatty acid metabolism in phenylephrine-induced cardiomyocytes hypertrophy in neonatal rats [176]. These findings further highlight the controversies regarding the inhibition of FAO to promote cardiomyocyte proliferation [65,66,71].

4.2. Hormones

Glucocorticoids are steroid hormones released by adrenal glands in response to stress and play important roles in cardiovascular diseases and in the maturation of foetal cardiac function [177,178,179]. In neonatal mouse hearts, physiological exposure to glucocorticoids was found to reduce cardiomyocyte proliferation, whereas ablation of glucocorticoid receptors extended the time window of cell cycle exit, which coincided with increased expression of glycolytic genes [160]. Consistently, dexamethasone (a potent glucocorticoid) inhibited the proliferation of isolated cardiomyocytes from newborn P2 rats, as evidenced by an increased number of binucleated cardiomyocytes, consisting of reduced cyclin D2 expression [162]. It could be speculated that the mechanism by which glucocorticoids suppress cardiomyocyte proliferation is by modulating FAO, as dexamethasone was shown to induce expression of FAO-related genes in mouse foetal cardiomyocytes [161]. Collectively, these findings suggest that inhibition of glucocorticoid signalling may promote cardiomyocyte proliferation in neonatal mice, but whether this approach can exert similar effects in adult mammalian hearts warrants further investigation; however, caution is advised when attempting to suppress glucocorticoid signalling, as this pathway has been shown to protect cardiomyocytes from cell death [180].
In higher vertebrates, thyroid hormones are secreted by the thyroid gland and exist as two active forms: triiodothyronine (T3) and thyroxine (T4) [181]. The heart is a target organ of thyroid hormones and recent studies have shown this hormone regulates cardiac metabolism through multiple mechanisms [182]. An increase in circulating levels of thyroid hormones have been observed after birth where it triggers cardiomyocyte cell cycle arrest and suppresses regenerative potential [167]. This has led to the speculation that changes in the levels of thyroid hormones could mediate the interplay between metabolism and cell cycle arrest [183]. Consistent with an anti-proliferative role, T3 was found to decrease BrdU uptake in ovine foetal cardiomyocytes by reducing cyclin D1 expression [163]. Similarly, a novel interaction between adrenergic-thyroid hormone in postnatal mice led to an increase in metabolic rates which may impair cardiomyocyte division and limit proliferation [184]. While these findings suggest that reduced thyroid levels promote cardiomyocyte proliferation, it will be challenging to translate this approach to clinical settings as decreased levels of thyroid hormones in the adult heart has been associated with an increased risk of cardiovascular diseases [181]. Conflicting findings have also been reported, whereby T3 administration in neonatal mouse hearts stimulated cardiomyocyte proliferation by inducing mitochondrial ROS production, which, in turn, activated JNK2α2-mediated, IGF1-dependent Erk1/2 proliferative pathways [164]. This finding conflicts with the ROS paradigm where increased ROS production is considered a key suppressor of cardiomyocyte proliferation [22,112,114,143]. In other studies, inhibition of DUSP5 (the nuclear phospho-Erk1/2-specific phosphatase) increased T3-activited phospho-Erk1/2 levels, resulting in ventricular cardiomyocyte proliferation of ~15% in young adult mice [165]. Though T3 has been shown to induce cardiomyocyte proliferation in both neonatal and young adult mice, caution is advised when considering this approach as a therapeutic modality to repair the damaged heart, as high-dose T4 administration has been found to stimulate hypertrophy of existing cardiomyocyte, rather than promote hyperplastic growth [185]. Collectively, these findings suggest that modulation of thyroid hormones can mediate cardiomyocyte proliferation, but it is important to identify the time windows when this hormone should be suppressed or elevated to promote effective regeneration.
To summarise, lncRNAs and hormones are promising targets that could potentially promote cardiomyocyte proliferation via direct and/or indirect modulation of cardiac metabolism. A comprehensive understanding of their mechanistic properties in modulating epigenetics and endogenous influences is key to their further development as therapeutic modalities for repairing the damaged heart in the setting of AMI.

5. Conclusions and Future Directions

HF is, in most cases, a progressive condition with a poor prognosis that imposes a global economic burden of ~$108 billion per year. IHD is the most frequent underlying cause of HF, as an ischemic event, can induce substantial cardiomyocyte death which precipitates adverse remodelling and cardiac dysfunction [1,2]. For decades, a vast number of studies have explored the possibility of transplanting cardiac and non-cardiac cells into the damaged myocardium with the aim of restoring cardiac function, either through direct action of the donor cells or via paracrine mechanisms [186]. However, the outcomes of these studies have been controversial, with large clinical studies reporting no obvious improvements in cardiac function following cell transplantation [187]. It would be ideal if we had sufficient knowledge of how to re-activate endogenous pathways that regulate the cell cycle in adult cardiomyocytes, given that the heart is one of the least regenerative organs in the human body [12,13]. Findings from rodents do support the potential for heart regeneration during the early stages after birth [15]; however, the extremely narrow regenerative window in larger mammals [16,17] calls into question whether cardiomyocyte proliferation can be considered a realistic therapeutic option, whilst underscoring the differences in regenerative capacity between species. Future studies in mammals that are more closely related to humans could help to provide novel insight on re-activating the cell cycle in the adult heart.
Changes in cardiac metabolism underlie the pathophysiology of several cardiac diseases [20,30], yet its role in heart regeneration has been barely explored, which is surprising, given that shifts in substrate utilisation before and after birth coincide with the terminal differentiation and reduced proliferation of cardiomyocytes [20,21]. Experimental studies do suggest that reduced oxidative metabolism favours cardiomyocyte re-entry into the cell cycle [137], but this raises the question of why heart regeneration is not initiated in the setting of HF, given that oxidative metabolism is already dampened [188]. Most likely, there are fundamental differences between the metabolism-proliferation axis in foetal and failing hearts, despite both adapting (or maladapting) to a reduced oxidative metabolism profile, and future studies should focus on these differences to elucidate metabolic pathways that can be potentially modulated without predisposing a pathological outcome.
Activation of the renin–angiotensin system (RAS) and increased production of the main effector, angiotensin II (Ang II), are instrumental in cardiac remodelling by promoting myofibroblast proliferation and matrix synthesis. Increased expression of angiotensin-converting enzyme has been observed in cardiomyocytes adjacent to the infarct scar and in nonmyocytes within the scarred tissue in MI patients [189], suggesting that RAS exerts pleiotropic effects on several cell types. Moreover, an increased expression of Ang II receptor type 1 (AT1) has been observed in infarcted rat hearts with increased Ang II binding affinity in the endothelium and myofibers [190]. In cardiomyocytes, Ang II is a potent inducer of hypertrophy, with studies also supporting a pathogenic role in metabolic perturbations. For instance, acute exposure of rat neonatal cardiomyocytes to Ang II was found to result in increased glucose uptake [191], while prolonged exposure of adult rat cardiomyocytes elicited downregulation of FAO pathways [192]. Importantly, these metabolic changes were associated with cardiomyocyte hypertrophy, rather than hyperplasia. RAS playing a role in cardiomyocyte proliferation is unlikely given that its inhibition was not found to mediate cardiomyocyte proliferation post-MI, but did increase vascular densities in the border zone [193]. Consistently, upregulation of AT1 was found to decrease microvessel densities in the setting of MI, while its inhibition promoted angiogenesis [194,195]. Conversely, Ang II can trigger VEGF synthesis in mesenchymal stem cells (MSCs) and injection of Ang II-treated MSCs into the border zone of infarcted hearts led to substantial improvements in cardiac function and reductions in infarct size and fibrosis [196].
Finally, if regeneration were to occur, would the increase in cardiomyocyte (and or endothelial cell) numbers be sufficient to provide meaningful improvements in cardiac function? A typical MI can cause the loss of ~1 billion cardiomyocytes in the adult heart, so the regeneration of this many cardiomyocytes will require a comprehensive understanding of its cell cycle arrest and re-entry pathways. To conclude, the targeting of cardiac metabolism to promote heart regeneration is an attractive concept, but one that comes with many future challenges. Nevertheless, the use of multi-omics platforms coupled with human samples and preclinical models could help to guide this area of research towards identifying suitable targets which can be adopted into clinical practice to improve health outcomes in patients with IHD.

Funding

Chrishan Ramachandra is supported by the Goh Cardiovascular Research Award (Duke-NUS-GCR/2022/0027). Derek Hausenloy is supported by the Duke-NUS Signature Research Programme funded by the Ministry of Health, Singapore Ministry of Health’s National Medical Research Council under its Singapore Translational Research Investigator Award (MOH-STaR21jun-0003), Centre Grant scheme (NMRC CG21APR1006), and Collaborative Centre Grant scheme (NMRC/CG21APRC006). En Ping Yap is supported by the Soo Jia Sien Heart Research Fund.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Cleland, J.G.; McGowan, J. Heart failure due to ischaemic heart disease: Epidemiology, pathophysiology and progression. J. Cardiovasc. Pharmacol. 1999, 33 (Suppl. 3), S17–S29. [Google Scholar] [CrossRef]
  2. Vedin, O.; Lam, C.S.P.; Koh, A.S.; Benson, L.; Teng, T.H.K.; Tay, W.T.; Braun, O.O.; Savarese, G.; Dahlstrom, U.; Lund, L.H. Significance of ischemic heart disease in patients with heart failure and preserved, midrange, and reduced ejection fraction: A nationwide cohort study. Circ. Heart Fail. 2017, 10, e003875. [Google Scholar] [CrossRef] [PubMed]
  3. Disease, G.B.D.; Injury, I.; Prevalence, C. 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] [Green Version]
  4. Groenewegen, A.; Rutten, F.H.; Mosterd, A.; Hoes, A.W. Epidemiology of heart failure. Eur. J. Heart Fail. 2020, 22, 1342–1356. [Google Scholar] [CrossRef]
  5. Triposkiadis, F.; Xanthopoulos, A.; Butler, J. Cardiovascular aging and heart failure: JACC review topic of the week. J. Am. Coll. Cardiol. 2019, 74, 804–813. [Google Scholar] [CrossRef] [PubMed]
  6. Orogo, A.M.; Gustafsson, A.B. Cell death in the myocardium: My heart won’t go on. IUBMB Life 2013, 65, 651–656. [Google Scholar] [CrossRef]
  7. Yehualashet, A.S.; Belachew, T.F.; Kifle, Z.D.; Abebe, A.M. Targeting cardiac metabolic pathways: A role in ischemic management. Vasc. Health Risk Manag. 2020, 16, 353–365. [Google Scholar] [CrossRef]
  8. van Empel, V.P.; Bertrand, A.T.; Hofstra, L.; Crijns, H.J.; Doevendans, P.A.; De Windt, L.J. Myocyte apoptosis in heart failure. Cardiovasc. Res. 2005, 67, 21–29. [Google Scholar] [CrossRef]
  9. Saraste, A.; Pulkki, K.; Kallajoki, M.; Heikkila, P.; Laine, P.; Mattila, S.; Nieminen, M.S.; Parvinen, M.; Voipio-Pulkki, L.M. Cardiomyocyte apoptosis and progression of heart failure to transplantation. Eur. J. Clin. Investig. 1999, 29, 380–386. [Google Scholar] [CrossRef]
  10. Crow, M.T.; Mani, K.; Nam, Y.J.; Kitsis, R.N. The mitochondrial death pathway and cardiac myocyte apoptosis. Circ. Res. 2004, 95, 957–970. [Google Scholar] [CrossRef] [Green Version]
  11. Narula, J.; Haider, N.; Virmani, R.; DiSalvo, T.G.; Kolodgie, F.D.; Hajjar, R.J.; Schmidt, U.; Semigran, M.J.; Dec, G.W.; Khaw, B.A. Apoptosis in myocytes in end-stage heart failure. N. Engl. J. Med. 1996, 335, 1182–1189. [Google Scholar] [CrossRef] [PubMed]
  12. Laflamme, M.A.; Murry, C.E. Heart regeneration. Nature 2011, 473, 326–335. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Chiong, M.; Wang, Z.V.; Pedrozo, Z.; Cao, D.J.; Troncoso, R.; Ibacache, M.; Criollo, A.; Nemchenko, A.; Hill, J.A.; Lavandero, S. Cardiomyocyte death: Mechanisms and translational implications. Cell Death Dis. 2011, 2, e244. [Google Scholar] [CrossRef] [Green Version]
  14. Neubauer, S. The failing heart--an engine out of fuel. N. Engl. J. Med. 2007, 356, 1140–1151. [Google Scholar] [CrossRef] [Green Version]
  15. Lam, N.T.; Sadek, H.A. Neonatal heart regeneration: Comprehensive literature review. Circulation 2018, 138, 412–423. [Google Scholar] [CrossRef]
  16. Zhu, W.; Zhang, E.; Zhao, M.; Chong, Z.; Fan, C.; Tang, Y.; Hunter, J.D.; Borovjagin, A.V.; Walcott, G.P.; Chen, J.Y.; et al. Regenerative potential of neonatal porcine hearts. Circulation 2018, 138, 2809–2816. [Google Scholar] [CrossRef]
  17. Ye, L.; D’Agostino, G.; Loo, S.J.; Wang, C.X.; Su, L.P.; Tan, S.H.; Tee, G.Z.; Pua, C.J.; Pena, E.M.; Cheng, R.B.; et al. Early regenerative capacity in the porcine heart. Circulation 2018, 138, 2798–2808. [Google Scholar] [CrossRef]
  18. Haubner, B.J.; Schneider, J.; Schweigmann, U.; Schuetz, T.; Dichtl, W.; Velik-Salchner, C.; Stein, J.I.; Penninger, J.M. Functional recovery of a human neonatal heart after severe myocardial infarction. Circ. Res. 2016, 118, 216–221. [Google Scholar] [CrossRef] [Green Version]
  19. Yap, E.P.; Chan, X.; Yu, F.; Cong, S.; Ramachandra, C.J.A. Mending a broken heart: Can gene modulation bolster therapeutic performance of adult stem cells? Cond. Med. 2021, 4, 71–87. [Google Scholar]
  20. Lopaschuk, G.D.; Jaswal, J.S. Energy metabolic phenotype of the cardiomyocyte during development, differentiation, and postnatal maturation. J. Cardiovasc. Pharmacol. 2010, 56, 130–140. [Google Scholar] [CrossRef] [PubMed]
  21. Lopaschuk, G.D.; Collins-Nakai, R.L.; Itoi, T. Developmental changes in energy substrate use by the heart. Cardiovasc. Res. 1992, 26, 1172–1180. [Google Scholar] [CrossRef]
  22. Puente, B.N.; Kimura, W.; Muralidhar, S.A.; Moon, J.; Amatruda, J.F.; Phelps, K.L.; Grinsfelder, D.; Rothermel, B.A.; Chen, R.; Garcia, J.A.; et al. The oxygen-rich postnatal environment induces cardiomyocyte cell-cycle arrest through DNA damage response. Cell 2014, 157, 565–579. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Brown, D.A.; Perry, J.B.; Allen, M.E.; Sabbah, H.N.; Stauffer, B.L.; Shaikh, S.R.; Cleland, J.G.; Colucci, W.S.; Butler, J.; Voors, A.A.; et al. Expert consensus document: Mitochondrial function as a therapeutic target in heart failure. Nat. Rev. Cardiol. 2017, 14, 238–250. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Doenst, T.; Nguyen, T.D.; Abel, E.D. Cardiac metabolism in heart failure: Implications beyond ATP production. Circ. Res. 2013, 113, 709–724. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Karwi, Q.G.; Uddin, G.M.; Ho, K.L.; Lopaschuk, G.D. Loss of metabolic flexibility in the failing heart. Front. Cardiovasc. Med. 2018, 5, 68. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Vander Heiden, M.G.; Cantley, L.C.; Thompson, C.B. Understanding the Warburg effect: The metabolic requirements of cell proliferation. Science 2009, 324, 1029–1033. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Spelat, R.; Ferro, F.; Contessotto, P.; Aljaabary, A.; Martin-Saldana, S.; Jin, C.; Karlsson, N.G.; Grealy, M.; Hilscher, M.M.; Magni, F.; et al. Metabolic reprogramming and membrane glycan remodeling as potential drivers of zebrafish heart regeneration. Commun. Biol. 2022, 5, 1365. [Google Scholar] [CrossRef]
  28. Ordono, J.; Perez-Amodio, S.; Ball, K.; Aguirre, A.; Engel, E. The generation of a lactate-rich environment stimulates cell cycle progression and modulates gene expression on neonatal and hiPSC-derived cardiomyocytes. Biomater. Adv. 2022, 139, 213035. [Google Scholar] [CrossRef]
  29. Lopaschuk, G.D.; Spafford, M.A.; Marsh, D.R. Glycolysis is predominant source of myocardial ATP production immediately after birth. Am. J. Physiol. 1991, 261, H1698–H1705. [Google Scholar] [CrossRef]
  30. 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]
  31. Wood, I.S.; Trayhurn, P. Glucose transporters (GLUT and SGLT): Expanded families of sugar transport proteins. Br. J. Nutr. 2003, 89, 3–9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Abel, E.D. Glucose transport in the heart. Front. Biosci. 2004, 9, 201–215. [Google Scholar] [CrossRef] [PubMed]
  33. Smoak, I.W.; Branch, S. Glut-1 expression and its response to hypoglycemia in the embryonic mouse heart. Anat. Embryol. 2000, 201, 327–333. [Google Scholar] [CrossRef] [PubMed]
  34. Morissette, M.R.; Howes, A.L.; Zhang, T.; Heller Brown, J. Upregulation of GLUT1 expression is necessary for hypertrophy and survival of neonatal rat cardiomyocytes. J. Mol. Cell Cardiol. 2003, 35, 1217–1227. [Google Scholar] [CrossRef]
  35. Fajardo, V.M.; Feng, I.; Chen, B.Y.; Perez-Ramirez, C.A.; Shi, B.; Clark, P.; Tian, R.; Lien, C.L.; Pellegrini, M.; Christofk, H.; et al. GLUT1 overexpression enhances glucose metabolism and promotes neonatal heart regeneration. Sci. Rep. 2021, 11, 8669. [Google Scholar] [CrossRef]
  36. Soto-Heredero, G.; Gomez de Las Heras, M.M.; Gabande-Rodriguez, E.; Oller, J.; Mittelbrunn, M. Glycolysis—A key player in the inflammatory response. FEBS J. 2020, 287, 3350–3369. [Google Scholar] [CrossRef] [Green Version]
  37. Chen, Z.; Dudek, J.; Maack, C.; Hofmann, U. Pharmacological inhibition of GLUT1 as a new immunotherapeutic approach after myocardial infarction. Biochem. Pharmacol. 2021, 190, 114597. [Google Scholar] [CrossRef]
  38. Israelsen, W.J.; Vander Heiden, M.G. Pyruvate kinase: Function, regulation and role in cancer. Semin. Cell Dev. Biol. 2015, 43, 43–51. [Google Scholar] [CrossRef] [Green Version]
  39. Magadum, A.; Singh, N.; Kurian, A.A.; Munir, I.; Mehmood, T.; Brown, K.; Sharkar, M.T.K.; Chepurko, E.; Sassi, Y.; Oh, J.G.; et al. Pkm2 regulates cardiomyocyte cell cycle and promotes cardiac regeneration. Circulation 2020, 141, 1249–1265. [Google Scholar] [CrossRef]
  40. Rees, M.L.; Subramaniam, J.; Li, Y.; Hamilton, D.J.; Frazier, O.H.; Taegtmeyer, H. A PKM2 signature in the failing heart. Biochem. Biophys. Res. Commun. 2015, 459, 430–436. [Google Scholar] [CrossRef] [Green Version]
  41. Hauck, L.; Dadson, K.; Chauhan, S.; Grothe, D.; Billia, F. Inhibiting the Pkm2/b-catenin axis drives in vivo replication of adult cardiomyocytes following experimental MI. Cell Death Differ. 2021, 28, 1398–1417. [Google Scholar] [CrossRef] [PubMed]
  42. Zhang, S.; Hulver, M.W.; McMillan, R.P.; Cline, M.A.; Gilbert, E.R. The pivotal role of pyruvate dehydrogenase kinases in metabolic flexibility. Nutr. Metab. 2014, 11, 10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Fukuda, R.; Marin-Juez, R.; El-Sammak, H.; Beisaw, A.; Ramadass, R.; Kuenne, C.; Guenther, S.; Konzer, A.; Bhagwat, A.M.; Graumann, J.; et al. Stimulation of glycolysis promotes cardiomyocyte proliferation after injury in adult zebrafish. EMBO Rep. 2020, 21, e49752. [Google Scholar] [CrossRef]
  44. Ramachandra, C.J.; Mehta, A.; Wong, P.; Shim, W. ErbB4 Activated p38gamma MAPK isoform mediates early cardiogenesis through NKx2.5 in human pluripotent stem cells. Stem Cells 2016, 34, 288–298. [Google Scholar] [CrossRef] [Green Version]
  45. Ramachandra, C.J.; Mehta, A.; Lua, C.H.; Chitre, A.; Ja, K.P.; Shim, W. ErbB receptor tyrosine kinase: A molecular switch between cardiac and neuroectoderm specification in human pluripotent stem cells. Stem Cells 2016, 34, 2461–2470. [Google Scholar] [CrossRef] [Green Version]
  46. Honkoop, H.; de Bakker, D.E.; Aharonov, A.; Kruse, F.; Shakked, A.; Nguyen, P.D.; de Heus, C.; Garric, L.; Muraro, M.J.; Shoffner, A.; et al. Single-cell analysis uncovers that metabolic reprogramming by ErbB2 signaling is essential for cardiomyocyte proliferation in the regenerating heart. eLife 2019, 8, e50163. [Google Scholar] [CrossRef]
  47. Aharonov, A.; Shakked, A.; Umansky, K.B.; Savidor, A.; Genzelinakh, A.; Kain, D.; Lendengolts, D.; Revach, O.Y.; Morikawa, Y.; Dong, J.; et al. ERBB2 drives YAP activation and EMT-like processes during cardiac regeneration. Nat. Cell Biol. 2020, 22, 1346–1356. [Google Scholar] [CrossRef]
  48. Zhou, L.; Jiang, S.; Fu, Q.; Smith, K.; Tu, K.; Li, H.; Zhao, Y. FASN, ErbB2-mediated glycolysis is required for breast cancer cell migration. Oncol. Rep. 2016, 35, 2715–2722. [Google Scholar] [CrossRef] [Green Version]
  49. Li, J.; Dong, L.; Wei, D.; Wang, X.; Zhang, S.; Li, H. Fatty acid synthase mediates the epithelial-mesenchymal transition of breast cancer cells. Int. J. Biol. Sci. 2014, 10, 171–180. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Kashihara, T.; Mukai, R.; Oka, S.I.; Zhai, P.; Nakada, Y.; Yang, Z.; Mizushima, W.; Nakahara, T.; Warren, J.S.; Abdellatif, M.; et al. YAP mediates compensatory cardiac hypertrophy through aerobic glycolysis in response to pressure overload. J. Clin. Investig. 2022, 132, e150595. [Google Scholar] [CrossRef]
  51. Ancey, P.B.; Contat, C.; Boivin, G.; Sabatino, S.; Pascual, J.; Zangger, N.; Perentes, J.Y.; Peters, S.; Abel, E.D.; Kirsch, D.G.; et al. GLUT1 expression in tumor-associated neutrophils promotes lung cancer growth and resistance to radiotherapy. Cancer Res. 2021, 81, 2345–2357. [Google Scholar] [CrossRef] [PubMed]
  52. Onay-Besikci, A. Regulation of cardiac energy metabolism in newborn. Mol. Cell Biochem. 2006, 287, 1–11. [Google Scholar] [CrossRef] [PubMed]
  53. Iruretagoyena, J.I.; Davis, W.; Bird, C.; Olsen, J.; Radue, R.; Teo Broman, A.; Kendziorski, C.; Splinter BonDurant, S.; Golos, T.; Bird, I.; et al. Metabolic gene profile in early human fetal heart development. Mol. Hum. Reprod. 2014, 20, 690–700. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Lopaschuk, G.D.; Ussher, J.R.; Folmes, C.D.; Jaswal, J.S.; Stanley, W.C. Myocardial fatty acid metabolism in health and disease. Physiol. Rev. 2010, 90, 207–258. [Google Scholar] [CrossRef]
  55. Bou Khzam, L.; Son, N.H.; Mullick, A.E.; Abumrad, N.A.; Goldberg, I.J. Endothelial cell CD36 deficiency prevents normal angiogenesis and vascular repair. Am. J. Transl. Res. 2020, 12, 7737–7761. [Google Scholar]
  56. Mathison, M.; Rosengart, T.K. Heart regeneration: The endothelial cell comes first. J. Thorac. Cardiovasc. Surg. 2018, 155, 1128–1129. [Google Scholar] [CrossRef] [Green Version]
  57. Singh, S.; Prakash, S.; Gupta, S.K. Angiogenesis: A critical determinant for cardiac regeneration. Mol. Ther. Nucleic Acids 2022, 29, 88–89. [Google Scholar] [CrossRef]
  58. Ja, K.P.; Miao, Q.; Zhen Tee, N.G.; Lim, S.Y.; Nandihalli, M.; Ramachandra, C.J.A.; Mehta, A.; Shim, W. iPSC-derived human cardiac progenitor cells improve ventricular remodelling via angiogenesis and interstitial networking of infarcted myocardium. J. Cell Mol. Med. 2016, 20, 323–332. [Google Scholar] [CrossRef] [Green Version]
  59. Myu Mai Ja, K.P.; Lim, K.P.; Chen, A.; Ting, S.; Li, S.Q.; Tee, N.; Ramachandra, C.; Mehta, A.; Wong, P.; Oh, S.; et al. Construction of a vascularized hydrogel for cardiac tissue formation in a porcine model. J. Tissue Eng. Regen. Med. 2018, 12, e2029–e2038. [Google Scholar] [CrossRef]
  60. Jabs, M.; Rose, A.J.; Lehmann, L.H.; Taylor, J.; Moll, I.; Sijmonsma, T.P.; Herberich, S.E.; Sauer, S.W.; Poschet, G.; Federico, G.; et al. Inhibition of endothelial notch signaling impairs fatty acid transport and leads to metabolic and vascular remodeling of the adult heart. Circulation 2018, 137, 2592–2608. [Google Scholar] [CrossRef] [Green Version]
  61. Zhao, L.; Ben-Yair, R.; Burns, C.E.; Burns, C.G. Endocardial Notch signaling promotes cardiomyocyte proliferation in the regenerating zebrafish heart through Wnt pathway antagonism. Cell Rep. 2019, 26, 546–554.e545. [Google Scholar] [CrossRef] [Green Version]
  62. Wang, W.; Hu, Y.F.; Pang, M.; Chang, N.; Yu, C.; Li, Q.; Xiong, J.W.; Peng, Y.; Zhang, R. BMP and Notch signaling pathways differentially regulate cardiomyocyte proliferation during ventricle regeneration. Int. J. Biol. Sci. 2021, 17, 2157–2166. [Google Scholar] [CrossRef] [PubMed]
  63. Schneider, H.; Staudacher, S.; Poppelreuther, M.; Stremmel, W.; Ehehalt, R.; Fullekrug, J. Protein mediated fatty acid uptake: Synergy between CD36/FAT-facilitated transport and acyl-CoA synthetase-driven metabolism. Arch. Biochem. Biophys. 2014, 546, 8–18. [Google Scholar] [CrossRef]
  64. Goldenberg, J.R.; Wang, X.; Lewandowski, E.D. Acyl CoA synthetase-1 links facilitated long chain fatty acid uptake to intracellular metabolic trafficking differently in hearts of male versus female mice. J. Mol. Cell Cardiol. 2016, 94, 1–9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Li, Y.; Yang, M.; Tan, J.; Shen, C.; Deng, S.; Fu, X.; Gao, S.; Li, H.; Zhang, X.; Cai, W. Targeting ACSL1 promotes cardiomyocyte proliferation and cardiac regeneration. Life Sci. 2022, 294, 120371. [Google Scholar] [CrossRef] [PubMed]
  66. Han, Y.; Yan, J.; Li, Z.Y.; Fan, Y.J.; Jiang, Z.L.; Shyy, J.Y.; Chien, S. Cyclic stretch promotes vascular homing of endothelial progenitor cells via Acsl1 regulation of mitochondrial fatty acid oxidation. Proc. Natl. Acad. Sci. USA 2023, 120, e2219630120. [Google Scholar] [CrossRef] [PubMed]
  67. Gilde, A.J.; van der Lee, K.A.; Willemsen, P.H.; Chinetti, G.; van der Leij, F.R.; van der Vusse, G.J.; Staels, B.; van Bilsen, M. Peroxisome proliferator-activated receptor (PPAR) alpha and PPARbeta/delta, but not PPARgamma, modulate the expression of genes involved in cardiac lipid metabolism. Circ. Res. 2003, 92, 518–524. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Djouadi, F.; Brandt, J.M.; Weinheimer, C.J.; Leone, T.C.; Gonzalez, F.J.; Kelly, D.P. The role of the peroxisome proliferator-activated receptor alpha (PPAR alpha) in the control of cardiac lipid metabolism. Prostaglandins Leukot. Essent. Fatty Acids 1999, 60, 339–343. [Google Scholar] [CrossRef]
  69. Cao, T.; Liccardo, D.; LaCanna, R.; Zhang, X.; Lu, R.; Finck, B.N.; Leigh, T.; Chen, X.; Drosatos, K.; Tian, Y. Fatty acid oxidation promotes cardiomyocyte proliferation rate but does not change cardiomyocyte number in infant mice. Front. Cell Dev. Biol. 2019, 7, 42. [Google Scholar] [CrossRef]
  70. Ding, L.; Liang, X.; Zhu, D.; Lou, Y. Peroxisome proliferator-activated receptor alpha is involved in cardiomyocyte differentiation of murine embryonic stem cells in vitro. Cell Biol. Int. 2007, 31, 1002–1009. [Google Scholar] [CrossRef]
  71. Lim, G.B. Inhibiting fatty acid oxidation promotes cardiomyocyte proliferation. Nat. Rev. Cardiol. 2020, 17, 266–267. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Ramachandra, C.J.A.; Mehta, A.; Wong, P.; Ja, K.; Fritsche-Danielson, R.; Bhat, R.V.; Hausenloy, D.J.; Kovalik, J.P.; Shim, W. Fatty acid metabolism driven mitochondrial bioenergetics promotes advanced developmental phenotypes in human induced pluripotent stem cell derived cardiomyocytes. Int. J. Cardiol. 2018, 272, 288–297. [Google Scholar] [CrossRef] [PubMed]
  73. Czarnowska, E.; Domal-Kwiatkowska, D.; Reichman-Warmusz, E.; Bierla, J.B.; Sowinska, A.; Ratajska, A.; Goral-Radziszewska, K.; Wojnicz, R. The correlation of PPARalpha activity and cardiomyocyte metabolism and structure in idiopathic dilated cardiomyopathy during heart failure progression. PPAR Res. 2016, 2016, 7508026. [Google Scholar] [CrossRef] [Green Version]
  74. Roy, R.L.T.; Gao, E.; Zhang, X.Y.; Tian, Y. Activation or inhibition of PPARα-mediated fatty acid β-oxidation does not active cardiomyocyte proliferation in normal or infarcted adult mice. bioRxiv 2019. preprint. [Google Scholar] [CrossRef]
  75. Balasse, E.O.; Fery, F. Ketone body production and disposal: Effects of fasting, diabetes, and exercise. Diabetes Metab. Rev. 1989, 5, 247–270. [Google Scholar] [CrossRef] [PubMed]
  76. Wentz, A.E.; d’Avignon, D.A.; Weber, M.L.; Cotter, D.G.; Doherty, J.M.; Kerns, R.; Nagarajan, R.; Reddy, N.; Sambandam, N.; Crawford, P.A. Adaptation of myocardial substrate metabolism to a ketogenic nutrient environment. J. Biol. Chem. 2010, 285, 24447–24456. [Google Scholar] [CrossRef] [Green Version]
  77. Puchalska, P.; Crawford, P.A. Multi-dimensional roles of ketone bodies in fuel metabolism, signaling, and therapeutics. Cell Metab. 2017, 25, 262–284. [Google Scholar] [CrossRef] [Green Version]
  78. Yurista, S.R.; Nguyen, C.T.; Rosenzweig, A.; de Boer, R.A.; Westenbrink, B.D. Ketone bodies for the failing heart: Fuels that can fix the engine? Trends Endocrinol. Metab. 2021, 32, 814–826. [Google Scholar] [CrossRef]
  79. Kim, S.; Jeon, J.M.; Kwon, O.K.; Choe, M.S.; Yeo, H.C.; Peng, X.; Cheng, Z.; Lee, M.Y.; Lee, S. Comparative proteomic analysis reveals the upregulation of ketogenesis in cardiomyocytes differentiated from induced pluripotent stem cells. Proteomics 2019, 19, e1800284. [Google Scholar] [CrossRef]
  80. Chong, D.; Gu, Y.; Zhang, T.; Xu, Y.; Bu, D.; Chen, Z.; Xu, N.; Li, L.; Zhu, X.; Wang, H.; et al. Neonatal ketone body elevation regulates postnatal heart development by promoting cardiomyocyte mitochondrial maturation and metabolic reprogramming. Cell Discov. 2022, 8, 106. [Google Scholar] [CrossRef]
  81. Cheng, Y.Y.; Gregorich, Z.; Prajnamitra, R.P.; Lundy, D.J.; Ma, T.Y.; Huang, Y.H.; Lee, Y.C.; Ruan, S.C.; Lin, J.H.; Lin, P.J.; et al. Metabolic changes associated with cardiomyocyte dedifferentiation enable adult mammalian cardiac regeneration. Circulation 2022, 146, 1950–1967. [Google Scholar] [CrossRef]
  82. Weis, E.M.; Puchalska, P.; Nelson, A.B.; Taylor, J.; Moll, I.; Hasan, S.S.; Dewenter, M.; Hagenmuller, M.; Fleming, T.; Poschet, G.; et al. Ketone body oxidation increases cardiac endothelial cell proliferation. EMBO Mol. Med. 2022, 14, e14753. [Google Scholar] [CrossRef]
  83. Cong, S.C., X.; Yap, E.P.; Ramachandra, C.J.A.; Hausenloy, D.J. Insights into the potential cardioprotective mechanisms of SGLT2 inhibitors. Cond. Med. 2022, 5, 1–10. [Google Scholar]
  84. Murashige, D.; Jang, C.; Neinast, M.; Edwards, J.J.; Cowan, A.; Hyman, M.C.; Rabinowitz, J.D.; Frankel, D.S.; Arany, Z. Comprehensive quantification of fuel use by the failing and nonfailing human heart. Science 2020, 370, 364–368. [Google Scholar] [CrossRef] [PubMed]
  85. McGarrah, R.W.; White, P.J. Branched-chain amino acids in cardiovascular disease. Nat. Rev. Cardiol. 2023, 20, 77–89. [Google Scholar] [CrossRef] [PubMed]
  86. Cuomo, P.; Capparelli, R.; Iannelli, A.; Iannelli, D. Role of branched-chain amino acid metabolism in type 2 diabetes, obesity, cardiovascular disease and non-alcoholic fatty liver disease. Int. J. Mol. Sci. 2022, 23, 4325. [Google Scholar] [CrossRef]
  87. Beppu, T.; Nitta, H.; Hayashi, H.; Imai, K.; Okabe, H.; Nakagawa, S.; Hashimoto, D.; Chikamoto, A.; Ishiko, T.; Yoshida, M.; et al. Effect of branched-chain amino acid supplementation on functional liver regeneration in patients undergoing portal vein embolization and sequential hepatectomy: A randomized controlled trial. J. Gastroenterol. 2015, 50, 1197–1205. [Google Scholar] [CrossRef]
  88. Wang, J.; Wang, W.; Zhu, F.; Duan, Q. The role of branched chain amino acids metabolic disorders in tumorigenesis and progression. Biomed. Pharmacother. 2022, 153, 113390. [Google Scholar] [CrossRef]
  89. Vellai, T. How the amino acid leucine activates the key cell-growth regulator mTOR. Nature 2021, 596, 192–194. [Google Scholar] [CrossRef]
  90. Davogustto, G.E.; Salazar, R.L.; Vasquez, H.G.; Karlstaedt, A.; Dillon, W.P.; Guthrie, P.H.; Martin, J.R.; Vitrac, H.; De La Guardia, G.; Vela, D.; et al. Metabolic remodeling precedes mTORC1-mediated cardiac hypertrophy. J. Mol. Cell Cardiol. 2021, 158, 115–127. [Google Scholar] [CrossRef]
  91. Miklas, J.W.; Levy, S.; Hofsteen, P.; Mex, D.I.; Clark, E.; Muster, J.; Robitaille, A.M.; Sivaram, G.; Abell, L.; Goodson, J.M.; et al. Amino acid primed mTOR activity is essential for heart regeneration. iScience 2022, 25, 103574. [Google Scholar] [CrossRef]
  92. Dowling, R.J.; Topisirovic, I.; Alain, T.; Bidinosti, M.; Fonseca, B.D.; Petroulakis, E.; Wang, X.; Larsson, O.; Selvaraj, A.; Liu, Y.; et al. mTORC1-mediated cell proliferation, but not cell growth, controlled by the 4E-BPs. Science 2010, 328, 1172–1176. [Google Scholar] [CrossRef] [Green Version]
  93. Csibi, A.; Fendt, S.M.; Li, C.; Poulogiannis, G.; Choo, A.Y.; Chapski, D.J.; Jeong, S.M.; Dempsey, J.M.; Parkhitko, A.; Morrison, T.; et al. The mTORC1 pathway stimulates glutamine metabolism and cell proliferation by repressing SIRT4. Cell 2013, 153, 840–854. [Google Scholar] [CrossRef] [Green Version]
  94. Mehta, A.; Ramachandra, C.J.; Sequiera, G.L.; Sudibyo, Y.; Nandihalli, M.; Yong, P.J.; Koh, C.H.; Shim, W. Phasic modulation of Wnt signaling enhances cardiac differentiation in human pluripotent stem cells by recapitulating developmental ontogeny. Biochim. Biophys. Acta 2014, 1843, 2394–2402. [Google Scholar] [CrossRef] [Green Version]
  95. Balatskyi, V.V.; Sowka, A.; Dobrzyn, P.; Piven, O.O. WNT/beta-catenin pathway is a key regulator of cardiac function and energetic metabolism. Acta Physiol. 2023, 237, e13912. [Google Scholar] [CrossRef] [PubMed]
  96. 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] [Green Version]
  97. Picard, M.; Shirihai, O.S. Mitochondrial signal transduction. Cell Metab. 2022, 34, 1620–1653. [Google Scholar] [CrossRef] [PubMed]
  98. Rossi, A.; Pizzo, P.; Filadi, R. Calcium, mitochondria and cell metabolism: A functional triangle in bioenergetics. Biochim. Biophys. Acta Mol. Cell Res. 2019, 1866, 1068–1078. [Google Scholar] [CrossRef]
  99. Bauer, T.M.; Murphy, E. Role of mitochondrial calcium and the permeability transition pore in regulating cell death. Circ. Res. 2020, 126, 280–293. [Google Scholar] [CrossRef]
  100. Ramachandra, C.J.A.; Chua, J.; Cong, S.; Kp, M.M.J.; Shim, W.; Wu, J.C.; Hausenloy, D.J. Human-induced pluripotent stem cells for modelling metabolic perturbations and impaired bioenergetics underlying cardiomyopathies. Cardiovasc. Res. 2021, 117, 694–711. [Google Scholar] [CrossRef]
  101. Lazou, A.; Ramachandra, C.J. Protecting the mitochondria in cardiac disease. Int. J. Mol. Sci. 2022, 23, 8115. [Google Scholar] [CrossRef] [PubMed]
  102. Cong, S.; Ramachandra, C.J.A.; Mai Ja, K.M.; Yap, J.; Shim, W.; Wei, L.; Hausenloy, D.J. Mechanisms underlying diabetic cardiomyopathy: From pathophysiology to novel therapeutic targets. Cond. Med. 2020, 3, 82–97. [Google Scholar]
  103. Lemieux, H.; Semsroth, S.; Antretter, H.; Hofer, D.; Gnaiger, E. Mitochondrial respiratory control and early defects of oxidative phosphorylation in the failing human heart. Int. J. Biochem. Cell Biol. 2011, 43, 1729–1738. [Google Scholar] [CrossRef] [PubMed]
  104. Carley, A.N.; Taegtmeyer, H.; Lewandowski, E.D. Matrix revisited: Mechanisms linking energy substrate metabolism to the function of the heart. Circ. Res. 2014, 114, 717–729. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. 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] [PubMed]
  106. Xu, X.; Duan, S.; Yi, F.; Ocampo, A.; Liu, G.H.; Izpisua Belmonte, J.C. Mitochondrial regulation in pluripotent stem cells. Cell Metab. 2013, 18, 325–332. [Google Scholar] [CrossRef] [Green Version]
  107. Khacho, M.; Harris, R.; Slack, R.S. Mitochondria as central regulators of neural stem cell fate and cognitive function. Nat. Rev. Neurosci. 2019, 20, 34–48. [Google Scholar] [CrossRef]
  108. Chakrabarty, R.P.; Chandel, N.S. Mitochondria as signaling organelles control mammalian stem cell fate. Cell Stem Cell 2021, 28, 394–408. [Google Scholar] [CrossRef]
  109. Ramachandra, C.J.A.; Mai Ja, K.P.M.; Lin, Y.H.; Shim, W.; Boisvert, W.A.; Hausenloy, D.J. Induced pluripotent stem cells for modelling energetic alterations in hypertrophic cardiomyopathy. Cond. Med. 2019, 2, 142–151. [Google Scholar]
  110. 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]
  111. Chen, Y.R.; Zweier, J.L. Cardiac mitochondria and reactive oxygen species generation. Circ. Res. 2014, 114, 524–537. [Google Scholar] [CrossRef] [Green Version]
  112. Sakaguchi, A.; Nishiyama, C.; Kimura, W. Cardiac regeneration as an environmental adaptation. Biochim. Biophys. Acta Mol. Cell Res. 2020, 1867, 118623. [Google Scholar] [CrossRef]
  113. Sakaguchi, A.; Kimura, W. Metabolic regulation of cardiac regeneration: Roles of hypoxia, energy homeostasis, and mitochondrial dynamics. Curr. Opin. Genet. Dev. 2021, 70, 54–60. [Google Scholar] [CrossRef]
  114. Zhang, D.; Li, Y.; Heims-Waldron, D.; Bezzerides, V.; Guatimosim, S.; Guo, Y.; Gu, F.; Zhou, P.; Lin, Z.; Ma, Q.; et al. Mitochondrial cardiomyopathy caused by elevated reactive oxygen species and impaired cardiomyocyte proliferation. Circ. Res. 2018, 122, 74–87. [Google Scholar] [CrossRef]
  115. Ramachandra, C.J.A.; Cong, S.; Chan, X.; Yap, E.P.; Yu, F.; Hausenloy, D.J. Oxidative stress in cardiac hypertrophy: From molecular mechanisms to novel therapeutic targets. Free Radic. Biol. Med. 2021, 166, 297–312. [Google Scholar] [CrossRef] [PubMed]
  116. Chen, Y.; Wu, G.; Li, M.; Hesse, M.; Ma, Y.; Chen, W.; Huang, H.; Liu, Y.; Xu, W.; Tang, Y.; et al. LDHA-mediated metabolic reprogramming promoted cardiomyocyte proliferation by alleviating ROS and inducing M2 macrophage polarization. Redox Biol. 2022, 56, 102446. [Google Scholar] [CrossRef] [PubMed]
  117. Pandey, R.; Velasquez, S.; Durrani, S.; Jiang, M.; Neiman, M.; Crocker, J.S.; Benoit, J.B.; Rubinstein, J.; Paul, A.; Ahmed, R.P. MicroRNA-1825 induces proliferation of adult cardiomyocytes and promotes cardiac regeneration post ischemic injury. Am. J. Transl. Res. 2017, 9, 3120–3137. [Google Scholar] [PubMed]
  118. Dambrova, M.; Zuurbier, C.J.; Borutaite, V.; Liepinsh, E.; Makrecka-Kuka, M. Energy substrate metabolism and mitochondrial oxidative stress in cardiac ischemia/reperfusion injury. Free Radic. Biol. Med. 2021, 165, 24–37. [Google Scholar] [CrossRef]
  119. Kimura, W.; Muralidhar, S.; Canseco, D.C.; Puente, B.; Zhang, C.C.; Xiao, F.; Abderrahman, Y.H.; Sadek, H.A. Redox signaling in cardiac renewal. Antioxid. Redox Signal. 2014, 21, 1660–1673. [Google Scholar] [CrossRef]
  120. Baharlooie, M.; Peymani, M.; Nasr-Esfahani, M.H.; Ghaedi, K. Pioglitazone mediates cardiac progenitor formation through increasing ROS levels. BioMed Res. Int. 2022, 2022, 1480345. [Google Scholar] [CrossRef]
  121. Liang, J.; Wu, M.; Chen, C.; Mai, M.; Huang, J.; Zhu, P. Roles of reactive oxygen species in cardiac differentiation, reprogramming, and regenerative therapies. Oxid. Med. Cell Longev. 2020, 2020, 2102841. [Google Scholar] [CrossRef]
  122. Ghadge, S.K.; Muhlstedt, S.; Ozcelik, C.; Bader, M. SDF-1alpha as a therapeutic stem cell homing factor in myocardial infarction. Pharmacol. Ther. 2011, 129, 97–108. [Google Scholar] [CrossRef] [PubMed]
  123. Herrero, D.; Albericio, G.; Higuera, M.; Herranz-Lopez, M.; Garcia-Brenes, M.A.; Cordero, A.; Roche, E.; Sepulveda, P.; Mora, C.; Bernad, A. The vascular niche for adult cardiac progenitor cells. Antioxidants 2022, 11, 882. [Google Scholar] [CrossRef] [PubMed]
  124. 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] [PubMed]
  125. Lazar, E.; Sadek, H.A.; Bergmann, O. Cardiomyocyte renewal in the human heart: Insights from the fall-out. Eur. Heart J. 2017, 38, 2333–2342. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Senyo, S.E.; Steinhauser, M.L.; Pizzimenti, C.L.; Yang, V.K.; Cai, L.; Wang, M.; Wu, T.D.; Guerquin-Kern, J.L.; Lechene, C.P.; Lee, R.T. Mammalian heart renewal by pre-existing cardiomyocytes. Nature 2013, 493, 433–436. [Google Scholar] [CrossRef] [Green Version]
  127. Neidig, L.E.; Weinberger, F.; Palpant, N.J.; Mignone, J.; Martinson, A.M.; Sorensen, D.W.; Bender, I.; Nemoto, N.; Reinecke, H.; Pabon, L.; et al. Evidence for minimal cardiogenic potential of stem cell antigen 1-positive cells in the adult mouse heart. Circulation 2018, 138, 2960–2962. [Google Scholar] [CrossRef]
  128. Ferreira-Martins, J.; Ogorek, B.; Cappetta, D.; Matsuda, A.; Signore, S.; D’Amario, D.; Kostyla, J.; Steadman, E.; Ide-Iwata, N.; Sanada, F.; et al. Cardiomyogenesis in the developing heart is regulated by c-kit-positive cardiac stem cells. Circ. Res. 2012, 110, 701–715. [Google Scholar] [CrossRef] [Green Version]
  129. Sandstedt, J.; Jonsson, M.; Kajic, K.; Sandstedt, M.; Lindahl, A.; Dellgren, G.; Jeppsson, A.; Asp, J. Left atrium of the human adult heart contains a population of side population cells. Basic. Res. Cardiol. 2012, 107, 255. [Google Scholar] [CrossRef] [Green Version]
  130. Hesse, M.; Welz, A.; Fleischmann, B.K. Heart regeneration and the cardiomyocyte cell cycle. Pflugers Arch. 2018, 470, 241–248. [Google Scholar] [CrossRef] [Green Version]
  131. Rigaud, V.O.; Zarka, C.; Kurian, J.; Harlamova, D.; Elia, A.; Kasatkin, N.; Johnson, J.; Behanan, M.; Kraus, L.; Pepper, H.; et al. UCP2 modulates cardiomyocyte cell cycle activity, acetyl-CoA, and histone acetylation in response to moderate hypoxia. JCI Insight 2022, 7, e155475. [Google Scholar] [CrossRef]
  132. Knox, C.; Camberos, V.; Ceja, L.; Monteon, A.; Hughes, L.; Longo, L.; Kearns-Jonker, M. Long-term hypoxia maintains a state of dedifferentiation and enhanced stemness in fetal cardiovascular progenitor cells. Int. J. Mol. Sci. 2021, 22, 9382. [Google Scholar] [CrossRef] [PubMed]
  133. Bo, B.; Li, S.; Zhou, K.; Wei, J. The regulatory role of oxygen metabolism in exercise-induced cardiomyocyte regeneration. Front. Cell Dev. Biol. 2021, 9, 664527. [Google Scholar] [CrossRef] [PubMed]
  134. Ye, L.; Qiu, L.; Feng, B.; Jiang, C.; Huang, Y.; Zhang, H.; Zhang, H.; Hong, H.; Liu, J. Role of blood oxygen saturation during post-natal human cardiomyocyte cell cycle activities. JACC Basic. Transl. Sci. 2020, 5, 447–460. [Google Scholar] [CrossRef]
  135. Korski, K.I.; Kubli, D.A.; Wang, B.J.; Khalafalla, F.G.; Monsanto, M.M.; Firouzi, F.; Echeagaray, O.H.; Kim, T.; Adamson, R.M.; Dembitsky, W.P.; et al. Hypoxia prevents mitochondrial dysfunction and senescence in human c-Kit(+) cardiac progenitor cells. Stem Cells 2019, 37, 555–567. [Google Scholar] [CrossRef]
  136. Payan, S.M.; Hubert, F.; Rochais, F. Cardiomyocyte proliferation, a target for cardiac regeneration. Biochim. Biophys. Acta Mol. Cell Res. 2020, 1867, 118461. [Google Scholar] [CrossRef]
  137. Nakada, Y.; Canseco, D.C.; Thet, S.; Abdisalaam, S.; Asaithamby, A.; Santos, C.X.; Shah, A.M.; Zhang, H.; Faber, J.E.; Kinter, M.T.; et al. Hypoxia induces heart regeneration in adult mice. Nature 2017, 541, 222–227. [Google Scholar] [CrossRef]
  138. Mohamed, T.M.A.; Abouleisa, R.; Hill, B.G. Metabolic determinants of cardiomyocyte proliferation. Stem Cells 2022, 40, 458–467. [Google Scholar] [CrossRef] [PubMed]
  139. Zheng, J.; Zhang, M.; Weng, H. Induction of the mitochondrial NDUFA4L2 protein by HIF-1a regulates heart regeneration by promoting the survival of cardiac stem cell. Biochem. Biophys. Res. Commun. 2018, 503, 2226–2233. [Google Scholar] [CrossRef]
  140. Feyen, D.A.M.; Gaetani, R.; Doevendans, P.A.; Sluijter, J.P.G. Stem cell-based therapy: Improving myocardial cell delivery. Adv. Drug. Deliv. Rev. 2016, 106, 104–115. [Google Scholar] [CrossRef]
  141. Barbash, I.M.; Chouraqui, P.; Baron, J.; Feinberg, M.S.; Etzion, S.; Tessone, A.; Miller, L.; Guetta, E.; Zipori, D.; Kedes, L.H.; et al. Systemic delivery of bone marrow-derived mesenchymal stem cells to the infarcted myocardium: Feasibility, cell migration, and body distribution. Circulation 2003, 108, 863–868. [Google Scholar] [CrossRef]
  142. Jang, W.B.; Park, J.H.; Ji, S.T.; Lee, N.K.; Kim, D.Y.; Kim, Y.J.; Jung, S.Y.; Kang, S.; Lamichane, S.; Lamichane, B.D.; et al. Cytoprotective roles of a novel compound, MHY-1684, against hyperglycemia-induced oxidative stress and mitochondrial dysfunction in human cardiac progenitor cells. Oxid. Med. Cell Longev. 2018, 2018, 4528184. [Google Scholar] [CrossRef] [PubMed]
  143. Li, X.; He, P.; Wang, X.L.; Zhang, S.; Devejian, N.; Bennett, E.; Cai, C. Sulfiredoxin-1 enhances cardiac progenitor cell survival against oxidative stress via the upregulation of the ERK/NRF2 signal pathway. Free Radic. Biol. Med. 2018, 123, 8–19. [Google Scholar] [CrossRef]
  144. Fan, H.; He, Z.; Huang, H.; Zhuang, H.; Liu, H.; Liu, X.; Yang, S.; He, P.; Yang, H.; Feng, D. Mitochondrial quality control in cardiomyocytes: A critical role in the progression of cardiovascular diseases. Front. Physiol. 2020, 11, 252. [Google Scholar] [CrossRef] [Green Version]
  145. 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] [PubMed]
  146. Ong, S.B.; Kwek, X.Y.; Katwadi, K.; Hernandez-Resendiz, S.; Crespo-Avilan, G.E.; Ismail, N.I.; Lin, Y.H.; Yap, E.P.; Lim, S.Y.; Ja, K.; et al. Targeting mitochondrial fission using mdivi-1 in a clinically relevant large animal model of acute myocardial infarction: A pilot study. Int. J. Mol. Sci. 2019, 20, 3972. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Dorn, G.W., 2nd; Vega, R.B.; Kelly, D.P. Mitochondrial biogenesis and dynamics in the developing and diseased heart. Genes. Dev. 2015, 29, 1981–1991. [Google Scholar] [CrossRef] [Green Version]
  148. Li, L.; Tao, G.; Hill, M.C.; Zhang, M.; Morikawa, Y.; Martin, J.F. Pitx2 maintains mitochondrial function during regeneration to prevent myocardial fat deposition. Development 2018, 145, dev168609. [Google Scholar] [CrossRef] [Green Version]
  149. Picca, A.; Mankowski, R.T.; Burman, J.L.; Donisi, L.; Kim, J.S.; Marzetti, E.; Leeuwenburgh, C. Mitochondrial quality control mechanisms as molecular targets in cardiac ageing. Nat. Rev. Cardiol. 2018, 15, 543–554. [Google Scholar] [CrossRef]
  150. Kalkhoran, S.B.; Hernandez-Resendiz, S.; Ong, S.G.; Ramachandra, C.J.A.; Hausenloy, D.J. Mitochondrial shaping proteins as novel treatment targets for cardiomyopathies. Cond. Med. 2020, 3, 216–226. [Google Scholar]
  151. Rahman, A.; Li, Y.; Ismail, N.I.; Chan, T.K.; Li, Y.; Xu, D.; Zhou, H.; Ong, S.B. The calcineurin-Drp1-mediated mitochondrial fragmentation is aligned with the differentiation of c-Kit cardiac progenitor cells. Int. J. Stem Cells 2023, 16, 123–134. [Google Scholar] [CrossRef] [PubMed]
  152. Choi, H.Y.; Park, J.H.; Jang, W.B.; Ji, S.T.; Jung, S.Y.; Kim da, Y.; Kang, S.; Kim, Y.J.; Yun, J.; Kim, J.H.; et al. High glucose causes human cardiac progenitor cell dysfunction by promoting mitochondrial fission: Role of a GLUT1 blocker. Biomol. Ther. 2016, 24, 363–370. [Google Scholar] [CrossRef] [Green Version]
  153. Rosdah, A.A.; Bond, S.T.; Sivakumaran, P.; Hoque, A.; Oakhill, J.S.; Drew, B.G.; Delbridge, L.M.D.; Lim, S.Y. Mdivi-1 protects human W8B2(+) cardiac stem cells from oxidative stress and simulated ischemia-reperfusion injury. Stem Cells Dev. 2017, 26, 1771–1780. [Google Scholar] [CrossRef] [PubMed]
  154. Lampert, M.A.; Gustafsson, A.B. Mitochondria and autophagy in adult stem cells: Proliferate or differentiate. J. Muscle Res. Cell Motil. 2020, 41, 355–362. [Google Scholar] [CrossRef]
  155. Lampert, M.A.; Orogo, A.M.; Najor, R.H.; Hammerling, B.C.; Leon, L.J.; Wang, B.J.; Kim, T.; Sussman, M.A.; Gustafsson, A.B. BNIP3L/NIX and FUNDC1-mediated mitophagy is required for mitochondrial network remodeling during cardiac progenitor cell differentiation. Autophagy 2019, 15, 1182–1198. [Google Scholar] [CrossRef] [PubMed]
  156. Mu, N.; Zhang, T.; Zhu, Y.; Lu, B.; Zheng, Q.; Duan, J. The mechanism by which miR-494-3p regulates PGC1-alpha-mediated inhibition of mitophagy in cardiomyocytes and alleviation of myocardial ischemia-reperfusion injury. BMC Cardiovasc. Disord. 2023, 23, 204. [Google Scholar] [CrossRef]
  157. Guo, Q.Y.; Yang, J.Q.; Feng, X.X.; Zhou, Y.J. Regeneration of the heart: From molecular mechanisms to clinical therapeutics. Mil. Med. Res. 2023, 10, 18. [Google Scholar] [CrossRef]
  158. Doppler, S.A.; Deutsch, M.A.; Lange, R.; Krane, M. Cardiac regeneration: Current therapies-future concepts. J. Thorac. Dis. 2013, 5, 683–697. [Google Scholar]
  159. Bolli, R.; Solankhi, M.; Tang, X.L.; Kahlon, A. Cell therapy in patients with heart failure: A comprehensive review and emerging concepts. Cardiovasc. Res. 2022, 118, 951–976. [Google Scholar] [CrossRef]
  160. Pianca, N.; Sacchi, F.; Umansky, K.B.; Chirivì, M.; Iommarini, L.; al, e. Glucocorticoid receptor antagonization propels endogenous cardiomyocyte proliferation and cardiac regeneration. Nat. Cardiovasc. Res. 2022, 1, 617–633. [Google Scholar] [CrossRef]
  161. Ivy, J.R.; Carter, R.N.; Zhao, J.F.; Buckley, C.; Urquijo, H.; Rog-Zielinska, E.A.; Panting, E.; Hrabalkova, L.; Nicholson, C.; Agnew, E.J.; et al. Glucocorticoids regulate mitochondrial fatty acid oxidation in fetal cardiomyocytes. J. Physiol. 2021, 599, 4901–4924. [Google Scholar] [CrossRef]
  162. Gay, M.S.; Dasgupta, C.; Li, Y.; Kanna, A.; Zhang, L. Dexamethasone induces cardiomyocyte terminal differentiation via epigenetic repression of cyclin D2 gene. J. Pharmacol. Exp. Ther. 2016, 358, 190–198. [Google Scholar] [CrossRef] [Green Version]
  163. Chattergoon, N.N.; Giraud, G.D.; Thornburg, K.L. Thyroid hormone inhibits proliferation of fetal cardiac myocytes in vitro. J. Endocrinol. 2007, 192, R1–R8. [Google Scholar] [CrossRef] [Green Version]
  164. Tan, L.; Bogush, N.; Naib, H.; Perry, J.; Calvert, J.W.; Martin, D.I.K.; Graham, R.M.; Naqvi, N.; Husain, A. Redox activation of JNK2alpha2 mediates thyroid hormone-stimulated proliferation of neonatal murine cardiomyocytes. Sci. Rep. 2019, 9, 17731. [Google Scholar] [CrossRef] [Green Version]
  165. Bogush, N.; Tan, L.; Naib, H.; Faizullabhoy, E.; Calvert, J.W.; Iismaa, S.E.; Gupta, A.; Ramchandran, R.; Martin, D.I.K.; Graham, R.M.; et al. DUSP5 expression in left ventricular cardiomyocytes of young hearts regulates thyroid hormone (T3)-induced proliferative ERK1/2 signaling. Sci. Rep. 2020, 10, 21918. [Google Scholar] [CrossRef] [PubMed]
  166. McClure, T.D.; Young, M.E.; Taegtmeyer, H.; Ning, X.H.; Buroker, N.E.; Lopez-Guisa, J.; Portman, M.A. Thyroid hormone interacts with PPARalpha and PGC-1 during mitochondrial maturation in sheep heart. Am. J. Physiol. Heart Circ. Physiol. 2005, 289, H2258–H2264. [Google Scholar] [CrossRef] [PubMed]
  167. Hirose, K.; Payumo, A.Y.; Cutie, S.; Hoang, A.; Zhang, H.; Guyot, R.; Lunn, D.; Bigley, R.B.; Yu, H.; Wang, J.; et al. Evidence for hormonal control of heart regenerative capacity during endothermy acquisition. Science 2019, 364, 184–188. [Google Scholar] [CrossRef]
  168. Anderson, K.M.; Anderson, D.M. LncRNAs at the heart of development and disease. Mamm. Genome 2022, 33, 354–365. [Google Scholar] [CrossRef] [PubMed]
  169. Mattick, J.S.; Amaral, P.P.; Carninci, P.; Carpenter, S.; Chang, H.Y.; Chen, L.L.; Chen, R.; Dean, C.; Dinger, M.E.; Fitzgerald, K.A.; et al. Long non-coding RNAs: Definitions, functions, challenges and recommendations. Nat. Rev. Mol. Cell Biol. 2023, 24, 430–447. [Google Scholar] [CrossRef]
  170. Mongelli, A.; Martelli, F.; Farsetti, A.; Gaetano, C. The dark that matters: Long non-coding RNAs as master regulators of cellular metabolism in non-communicable diseases. Front. Physiol. 2019, 10, 369. [Google Scholar] [CrossRef] [Green Version]
  171. Frank, S.; Aguirre, A.; Hescheler, J.; Kurian, L. A lncRNA perspective into (re)building the heart. Front. Cell Dev. Biol. 2016, 4, 128. [Google Scholar] [CrossRef] [Green Version]
  172. Liu, N.; Kataoka, M.; Wang, Y.; Pu, L.; Dong, X.; Fu, X.; Zhang, F.; Gao, F.; Liang, T.; Pei, J.; et al. LncRNA LncHrt preserves cardiac metabolic homeostasis and heart function by modulating the LKB1-AMPK signaling pathway. Basic. Res. Cardiol. 2021, 116, 48. [Google Scholar] [CrossRef]
  173. Yuan, T.; Krishnan, J. Non-coding RNAs in cardiac regeneration. Front. Physiol. 2021, 12, 650566. [Google Scholar] [CrossRef] [PubMed]
  174. Chen, Y.M.; Li, H.; Fan, Y.; Zhang, Q.J.; Li, X.; Wu, L.J.; Chen, Z.J.; Zhu, C.; Qian, L.M. Identification of differentially expressed lncRNAs involved in transient regeneration of the neonatal C57BL/6J mouse heart by next-generation high-throughput RNA sequencing. Oncotarget 2017, 8, 28052–28062. [Google Scholar] [CrossRef] [Green Version]
  175. Li, B.; Hu, Y.; Li, X.; Jin, G.; Chen, X.; Chen, G.; Chen, Y.; Huang, S.; Liao, W.; Liao, Y.; et al. Sirt1 antisense long noncoding RNA promotes cardiomyocyte proliferation by enhancing the stability of Sirt1. J. Am. Heart Assoc. 2018, 7, e009700. [Google Scholar] [CrossRef] [Green Version]
  176. Planavila, A.; Iglesias, R.; Giralt, M.; Villarroya, F. Sirt1 acts in association with PPARalpha to protect the heart from hypertrophy, metabolic dysregulation, and inflammation. Cardiovasc. Res. 2011, 90, 276–284. [Google Scholar] [CrossRef] [Green Version]
  177. Walker, B.R. Glucocorticoids and cardiovascular disease. Eur. J. Endocrinol. 2007, 157, 545–559. [Google Scholar] [CrossRef] [Green Version]
  178. Oakley, R.H.; Cidlowski, J.A. The biology of the glucocorticoid receptor: New signaling mechanisms in health and disease. J. Allergy Clin. Immunol. 2013, 132, 1033–1044. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Jellyman, J.K.; Fletcher, A.J.W.; Fowden, A.L.; Giussani, D.A. Glucocorticoid maturation of fetal cardiovascular function. Trends Mol. Med. 2020, 26, 170–184. [Google Scholar] [CrossRef] [PubMed]
  180. Ren, R.; Oakley, R.H.; Cruz-Topete, D.; Cidlowski, J.A. Dual role for glucocorticoids in cardiomyocyte hypertrophy and apoptosis. Endocrinology 2012, 153, 5346–5360. [Google Scholar] [CrossRef] [Green Version]
  181. Grais, I.M.; Sowers, J.R. Thyroid and the heart. Am. J. Med. 2014, 127, 691–698. [Google Scholar] [CrossRef] [Green Version]
  182. Janssen, R.; Muller, A.; Simonides, W.S. Cardiac thyroid hormone metabolism and heart failure. Eur. Thyroid. J. 2017, 6, 130–137. [Google Scholar] [CrossRef]
  183. Ross, I.; Omengan, D.B.; Huang, G.N.; Payumo, A.Y. Thyroid hormone-dependent regulation of metabolism and heart regeneration. J. Endocrinol. 2022, 252, R71–R82. [Google Scholar] [CrossRef]
  184. Payumo, A.Y.; Chen, X.; Hirose, K.; Chen, X.; Hoang, A.; Khyeam, S.; Yu, H.; Wang, J.; Chen, Q.; Powers, N.; et al. Adrenergic-thyroid hormone interactions drive postnatal thermogenesis and loss of mammalian heart regenerative capacity. Circulation 2021, 144, 1000–1003. [Google Scholar] [CrossRef]
  185. Gerdes, A.M.; Kriseman, J.; Bishop, S.P. Changes in myocardial cell size and number during the development and reversal of hyperthyroidism in neonatal rats. Lab. Investig. 1983, 48, 598–602. [Google Scholar] [PubMed]
  186. Parizadeh, S.M.; Jafarzadeh-Esfehani, R.; Ghandehari, M.; Parizadeh, M.R.; Ferns, G.A.; Avan, A.; Hassanian, S.M. Stem cell therapy: A novel approach for myocardial infarction. J. Cell Physiol. 2019, 234, 16904–16912. [Google Scholar] [CrossRef] [PubMed]
  187. Nguyen, P.K.; Rhee, J.W.; Wu, J.C. Adult stem cell therapy and heart failure, 2000 to 2016: A systematic review. JAMA Cardiol. 2016, 1, 831–841. [Google Scholar] [CrossRef] [Green Version]
  188. Zhou, B.; Tian, R. Mitochondrial dysfunction in pathophysiology of heart failure. J. Clin. Investig. 2018, 128, 3716–3726. [Google Scholar] [CrossRef] [Green Version]
  189. Hokimoto, S.; Yasue, H.; Fujimoto, K.; Yamamoto, H.; Nakao, K.; Kaikita, K.; Sakata, R.; Miyamoto, E. Expression of angiotensin-converting enzyme in remaining viable myocytes of human ventricles after myocardial infarction. Circulation 1996, 94, 1513–1518. [Google Scholar] [CrossRef] [PubMed]
  190. El-Sabban, M.E.; Hassan, K.A.; Birbari, A.E.; Bitar, K.M.; Bikhazi, A.B. Angiotensin II binding and extracellular matrix remodelling in a rat model of myocardial infarction. J. Renin Angiotensin Aldosterone Syst. 2000, 1, 369–378. [Google Scholar] [CrossRef] [Green Version]
  191. Stuck, B.J.; Lenski, M.; Bohm, M.; Laufs, U. Metabolic switch and hypertrophy of cardiomyocytes following treatment with angiotensin II are prevented by AMP-activated protein kinase. J. Biol. Chem. 2008, 283, 32562–32569. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  192. Pellieux, C.; Montessuit, C.; Papageorgiou, I.; Lerch, R. Angiotensin II downregulates the fatty acid oxidation pathway in adult rat cardiomyocytes via release of tumour necrosis factor-alpha. Cardiovasc. Res. 2009, 82, 341–350. [Google Scholar] [CrossRef] [Green Version]
  193. Segersvard, H.; Lakkisto, P.; Forsten, H.; Immonen, K.; Kosonen, R.; Palojoki, E.; Kankuri, E.; Harjula, A.; Laine, M.; Tikkanen, I. Effects of angiotensin II blockade on cardiomyocyte regeneration after myocardial infarction in rats. J. Renin Angiotensin Aldosterone Syst. 2015, 16, 92–102. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. de Boer, R.A.; Pinto, Y.M.; Suurmeijer, A.J.; Pokharel, S.; Scholtens, E.; Humler, M.; Saavedra, J.M.; Boomsma, F.; van Gilst, W.H.; van Veldhuisen, D.J. Increased expression of cardiac angiotensin II type 1 (AT1) receptors decreases myocardial microvessel density after experimental myocardial infarction. Cardiovasc. Res. 2003, 57, 434–442. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Sladek, T.; Sladkova, J.; Kolar, F.; Papousek, F.; Cicutti, N.; Korecky, B.; Rakusan, K. The effect of AT1 receptor antagonist on chronic cardiac response to coronary artery ligation in rats. Cardiovasc. Res. 1996, 31, 568–576. [Google Scholar] [CrossRef] [Green Version]
  196. Liu, C.; Fan, Y.; Zhou, L.; Zhu, H.Y.; Song, Y.C.; Hu, L.; Wang, Y.; Li, Q.P. Pretreatment of mesenchymal stem cells with angiotensin II enhances paracrine effects, angiogenesis, gap junction formation and therapeutic efficacy for myocardial infarction. Int. J. Cardiol. 2015, 188, 22–32. [Google Scholar] [CrossRef]
Figure 1. Schematic illustrating the interplay between cardiac metabolism and cardiomyocyte proliferation in neonatal and adult mammalian hearts. Neonatal heart metabolism (red), adult heart metabolisms (blue). Abbreviations: Glut1—glucose transporter type 1; Glut4—glucose transporter type 4; PK—pyruvate kinase; PDK—pyruvate dehydrogenase kinase; FACS—fatty acyl coA synthase; PPARα—peroxisome-proliferator-activated receptor alpha; BCAAs—branched-chain amino acids; BCKA—branch-chain alpha-keto acids; mTORC1—mammalian target of rapamycin complex 1; TCA—tricarboxylic acid cycle.
Figure 1. Schematic illustrating the interplay between cardiac metabolism and cardiomyocyte proliferation in neonatal and adult mammalian hearts. Neonatal heart metabolism (red), adult heart metabolisms (blue). Abbreviations: Glut1—glucose transporter type 1; Glut4—glucose transporter type 4; PK—pyruvate kinase; PDK—pyruvate dehydrogenase kinase; FACS—fatty acyl coA synthase; PPARα—peroxisome-proliferator-activated receptor alpha; BCAAs—branched-chain amino acids; BCKA—branch-chain alpha-keto acids; mTORC1—mammalian target of rapamycin complex 1; TCA—tricarboxylic acid cycle.
Ijms 24 10300 g001
Figure 2. Schematic illustrating the role of mitochondria in cardiac injury and regeneration. (Left) environmental stress, such as that caused by ischemia/reperfusion injury can lead to mitochondrial dysfunction, resulting in ROS overproduction and dysregulation of mitochondrial dynamics, thereby instigating DNA damage and suppressing cardiomyocyte proliferation. (Right) under hypoxic conditions or during antioxidant treatments, cardiomyocytes can switch from oxidative metabolism to glycolysis, potentially via the HIF-1α pathway, and undergo restoration of aberrant mitophagy and mitochondria biogenesis to promote cell cycle re-entry. Abbreviations: ROS—reactive oxygen species; HIF-1α—hypoxia inducible factor 1α.
Figure 2. Schematic illustrating the role of mitochondria in cardiac injury and regeneration. (Left) environmental stress, such as that caused by ischemia/reperfusion injury can lead to mitochondrial dysfunction, resulting in ROS overproduction and dysregulation of mitochondrial dynamics, thereby instigating DNA damage and suppressing cardiomyocyte proliferation. (Right) under hypoxic conditions or during antioxidant treatments, cardiomyocytes can switch from oxidative metabolism to glycolysis, potentially via the HIF-1α pathway, and undergo restoration of aberrant mitophagy and mitochondria biogenesis to promote cell cycle re-entry. Abbreviations: ROS—reactive oxygen species; HIF-1α—hypoxia inducible factor 1α.
Ijms 24 10300 g002
Table 1. List of studies investigating the effects of hormones as mediators of cardiomyocyte proliferation and cardiac metabolism.
Table 1. List of studies investigating the effects of hormones as mediators of cardiomyocyte proliferation and cardiac metabolism.
HormoneTreatmentAnimal/Cellular ModelEffects on Cardiomyocyte ProliferationProliferation Markers EvaluatedEffects on Cardiac MetabolismReferences
GlucocorticoidCardiomyocyte-specific knockout of glucocorticoid receptorsNeonatal C57BL/6 mice (P1)PromotedKi67; BrdU; Aurora B; NucleationDecreased fatty acid oxidation; increased glycolysis[160]
DexamethasonePregnant C57BL/6J mice (E13.5 or E16.5) Failed to induce fatty acid oxidation[161]
Pregnant C57BL/6J mice (E17.5) Decreased fatty acid oxidation
Neonatal Sprague Dawley rats (P2)InhibitedKi67; Nucleation [162]
Thyroid hormonesT3Ovine (Ovis aries) foetal cardiomyocytes (~135 days gestation)InhibitedBrdU [163]
Neonatal cardiomyocytes from C57BL/6 mice (P2-P4)IncreasedPHH3; EdUIncreased ROS production[164]
Neonatal C57BL/6 mice (P6)IncreasedPHH3; EdU [165]
ThyroidectomyNeonatal sheep (P30) Reduced mitochondrial maturation and biogenesis[166]
Myh6-Cre;ThraDN/+Neonatal C57BL/6 mice (P14)IncreasedKi67; EdU; Aurora B; NucleationDownregulation of mitochondrial genes[167]
Myh6-Cre;ThraDN/+Adult C57BL/6 mice post-IRI injuryIncreasedKi67; EdU; Aurora B; Nucleation
Abbreviation: Myh6-Cre;ThraDN/+—cardiomyocyte-specific overexpression of dominant negative thyroid hormone receptor alpha; IRI—ischemia/reperfusion injury; PHH3—phospho-histone H3.
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

Yu, F.; Cong, S.; Yap, E.P.; Hausenloy, D.J.; Ramachandra, C.J. Unravelling the Interplay between Cardiac Metabolism and Heart Regeneration. Int. J. Mol. Sci. 2023, 24, 10300. https://doi.org/10.3390/ijms241210300

AMA Style

Yu F, Cong S, Yap EP, Hausenloy DJ, Ramachandra CJ. Unravelling the Interplay between Cardiac Metabolism and Heart Regeneration. International Journal of Molecular Sciences. 2023; 24(12):10300. https://doi.org/10.3390/ijms241210300

Chicago/Turabian Style

Yu, Fan, Shuo Cong, En Ping Yap, Derek J. Hausenloy, and Chrishan J. Ramachandra. 2023. "Unravelling the Interplay between Cardiac Metabolism and Heart Regeneration" International Journal of Molecular Sciences 24, no. 12: 10300. https://doi.org/10.3390/ijms241210300

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