Skip to main content

Mitochondrial dysfunction at the crossroad of cardiovascular diseases and cancer

Abstract

A large body of evidence indicates the existence of a complex pathophysiological relationship between cardiovascular diseases and cancer. Mitochondria are crucial organelles whose optimal activity is determined by quality control systems, which regulate critical cellular events, ranging from intermediary metabolism and calcium signaling to mitochondrial dynamics, cell death and mitophagy. Emerging data indicate that impaired mitochondrial quality control drives myocardial dysfunction occurring in several heart diseases, including cardiac hypertrophy, myocardial infarction, ischaemia/reperfusion damage and metabolic cardiomyopathies. On the other hand, diverse human cancers also dysregulate mitochondrial quality control to promote their initiation and progression, suggesting that modulating mitochondrial homeostasis may represent a promising therapeutic strategy both in cardiology and oncology. In this review, first we briefly introduce the physiological mechanisms underlying the mitochondrial quality control system, and then summarize the current understanding about the impact of dysregulated mitochondrial functions in cardiovascular diseases and cancer. We also discuss key mitochondrial mechanisms underlying the increased risk of cardiovascular complications secondary to the main current anticancer strategies, highlighting the potential of strategies aimed at alleviating mitochondrial impairment-related cardiac dysfunction and tumorigenesis. It is hoped that this summary can provide novel insights into precision medicine approaches to reduce cardiovascular and cancer morbidities and mortalities.

Introduction

Cardiovascular disorders (CVDs) and cancer represent the two major causes of mortality and morbidity worldwide and, therefore, pose an unprecedented need to identify novel therapeutic treatments to effectively save the patients from death [1]. CVDs and cancer share several common environmental risk factors, such as alcohol abuse, obesity, sedentary life, and tobacco, and improving the lifestyle could be of benefit for patients suffering from these diseases [2]. Furthermore, the same genetic and cellular mechanisms could play a pathogenic role in both CVDs and cancer, including inflammatory signalling, reactive oxygen species (ROS), Wnt signalling pathway, and hypoxia-inducible factor-1α (HIF-1α) [1, 3,4,5]. In addition, the incidence of CVDs is significantly higher in cancer patients, and cancer survivors undergo a higher risk to develop CVDs [6], such as chronic heart failure [7]. Therefore, new diagnoses of malignancy in subjects with pre-existing HF can often occur, requiring a complex and early integrated cardio-oncological approach for their clinical management. The concept that cancer represents a risk factor for CVDs, such as heart disease, atherosclerosis, cerebrovascular disease, hypertension, and aortic aneurysm/dissection, is increasingly being recognized. Accordingly, cancer may directly affect cardiac function as evinced by the increased medium-term to long-term risk for one or more CVDs showed in survivors of most site-specific cancers compared with that for the general population [8].The site of cancer also represents an important factor in determining the risk of CVDs in cancer survivors. In this regard, a recent Surveillance, Epidemiology, and End Results program-based observational study in US population, which represents the largest study on the deaths from CVDs among patients with 28 cancer types with over 40 years of follow-up, reported that cancer patients exhibit an on average 2–6 times higher CVDs mortality risk than the general population, with the majority of deaths from CVDs occurring in patients presenting with breast, prostate, or bladder cancer. Conversely, patients with endometrial cancer showed a very high risk of dying from CVDs at the first year after diagnosis, which remains elevated compared to other cancer sites [9]. Cardiotoxicity, in turn, is a recognized side effect of anticancer treatments that can ultimately lead to therapy discontinuation in some oncological patients [10,11,12].

On this evidence, cardio-oncology has emerged as a common framework for cardiologists, oncologists, and hematologists to identify and target the most adverse cardiovascular events that prevent cancer patients to benefit from a wide variety of therapeutic approaches [13, 14].

Mitochondria are dynamic organelles that not only constitute the cell powerhouse by producing most of the chemical energy (adenosine-5-trisphosphate, ATP) that supports cellular activities through oxidative phosphorylation (OXPHOS) and fatty acid oxidation, but also serve as a signalling hub by regulating a plethora of cellular processes, including ROS generation and maintenance, phospholipid biosynthesis, intracellular Ca2+ regulation, angiogenesis, innate immune signalling, and iron-sulphur cluster biosynthesis [15,16,17,18,19,20]. Mitochondria are no longer regarded as static and passive sub-cellular structures [21]; they are now considered dynamic organelles that can change shape, structure, number, and intracellular positioning through a finely tuned regulatory network, known as quality control system, which fulfills the primary function to meet ATP supply with the high-energy demand of vital organs, such as heart and brain [15, 22,23,24,25]. Defects in mitochondrial function, as well as mitochondrial dynamics and quality control, have been reported in both CVDs and cancer, which present deregulated bioenergetics and metabolism [11, 26,27,28] and show a dramatic deregulation of mitochondrial-derived intracellular signals, such as ROS and Ca2+ [16, 27, 29]. In addition, mitochondria sense and respond to stressors signals, such as DNA damage and hypoxia, which is a hallmark of multiple disorders, including CVDs and cancer [4, 11, 30]. Finally, several anticancer therapies cause life-threatening cardiovascular complications by inducing mitochondrial stress or dysfunction, which may also initiate many secondary, systemic signalling pathways that further exacerbate cardiovascular damage [11]. On this basis, herein we first survey the molecular mechanisms which regulate mitochondrial dynamics and quality control and illustrate how they can be altered both in CVDs and cancer. In parallel, we delineate the emerging therapeutic avenues that aim at reducing the burden of mitochondrial damage in those regarded as the most life-threatening disorders worldwide. Finally, we highlight the cardiovascular complications deriving from anti-cancer therapies, and illustrate how mitochondria-targeted strategies could improve the therapeutic outcome of cancer patients without causing cardiotoxicity.

Mitochondrial physiology: focus on mitochondrial dynamics and quality control systems

Mitochondria are highly dynamic organelles that can adapt to different conditions by reversibly cycling through two distinct morphological states—a short rod-like shape and an interconnected pattern—by dividing and fusing the inner and outer mitochondrial membranes (IMM and OMM, respectively). Mitochondrial dynamics refers to the dynamic behaviour of mitochondria, including the fusion and fission of mitochondrial membranes [31]. Such fission–fusion cycle is intricately regulated by a highly conserved mechanism involving a group of proteins that are located on the IMM and OMM and belong to the dynamin family of GTPases. Mitofusin 1 and 2 (MFN1/2) and optic atrophy 1 (OPA1) are responsible for fusion, while dynamin-related protein 1 (DRP1) and dynamin-2 (DNM2) are the primary fission proteins [32,33,34].

Mitochondrial fusion

Mitochondrial fusion is the process by which two mitochondria combine to form one mitochondrion. A typical mitochondrial fusion reaction involves two mitochondria colliding end-to-end; the membrane fusion event occurs at the site of the collision. Fusion reactions can also occur end-to-side or even within a single mitochondrion to form structures that resemble rings. Because mitochondria have two sets of membranes, the fusion process begins with the outer membrane and continues with the inner membrane. These two membrane fusion events occur almost simultaneously, although a temporal distinction can sometimes be identified. OPA1 and MFN1/2 are responsible for triggering the IMM and OMM fusion processes during mitochondrial fusion [31]. Although the primary function of OPA1 is to control mitochondrial fusion, it also has unrelated activities, such as improving mitochondrial respiratory efficiency, stabilizing the structure of mitochondrial cristae, controlling the release of cytochrome c, and triggering the assembly of electron transport supercomplexes [35]. MFN2 has activities unrelated to nuclear fusion. Specifically, MFN2 strengthens mitochondrial calcium buffering by connecting mitochondria and the endoplasmic reticulum (ER) [36]. Phosphorylated MFN2 acts as a mitochondrial receptor for the protein Parkin and mediates mitophagy in cardiomyocytes [37]. The function of OPA1 is controlled not only by the cleavage of OPA1 by proteases but also by post-translational modifications. In cardiac mitochondria, there are two forms of OPA1: the long form of OPA1 (L-OPA1), which binds to the IMM, and the soluble short form of OPA1 (S-OPA1), which is located in the intermembrane space representing a product of L-OPA1 proteolytic cleavage. The presence of both L-OPA1 and S-OPA1 in mitochondria facilitates the process of mitochondrial fusion [38]. In humans, L-OPA1 has two distinct cleavage sites. The S1 site, found in all forms of L-OPA1, is cleaved by the protease OMA1, and the S2 site, only found in some forms of L-OPA1, is cleaved by YME1L. These two cleavage points are distinct from one another [39]. In most cases, YME1L acts as a mediator for the proteolysis of L-OPA1 at the S2 site. This process produces S-OPA1 and unsevered L-OPA1, both of which promote mitochondrial fusion and maintain cristae morphology. When cells are stressed, however, OMA1 activity increases, and OPA1 undergoes proteolysis at the S1 site. This causes all forms of L-OPA1 to be cleaved into S-OPA1, which causes mitochondrial fragmentation, cristae morphology disruption, and apoptosis [40]. OMA1 activation and L-OPA1 processing may be linked to the proapoptotic signalling pathways, and this activation may be dependent on the oligomerization of BAK and BAX [41]. The idea that elevated S-OPA1 levels and OMA1 activity impair mitochondrial and cardiac function was disproven by the work of Lee et al., who demonstrated that increasing S-OPA1 levels improved cell survival under oxidative stress by preserving mitochondrial cristae and energetics. The suppression of OMA1 expression leads to more oxidative stress-induced cell death, suggesting that OMA1-induced S-OPA1 production represents a survival mechanism in stressed cells [42]. Alongside proteolysis, post-translational modifications are another mechanism for controlling OPA1 function. Sirtuin 3, a mitochondrial deacetylase, protects cardiomyocytes from doxorubicin-induced cell death by increasing OPA1’s GTPase activity through deacetylation of OPA1 at Lys 926 and Lys 931 [43]. In addition, OPA1 was found to be O-GlcNAcylated in the presence of hyperglycemia, which reduces its GTPase activity and ultimately leads to cardiomyocyte death [44]. Like OPA1, MFN1/2 can be regulated through post-translational modifications. Cells are particularly susceptible to apoptotic stimuli when MFN1 is phosphorylated by ERK at T562, which reduces its efficiency in oligomerization and mitochondrial tethering but increases its binding to BAK [45]. In addition to upregulating the activity of protein kinase C beta II (PKC beta II), phosphorylation of MFN1 at Ser86 reduces MFN1 GTPase activity and leads to an increase in the accumulation of fragmented and dysfunctional mitochondria, both of which are associated with HF [46]. The phosphorylation of MFN1 appears to reduce its GTPase activity, which in turn reduces mitochondrial fusion and impairs mitochondrial function. Increased mitophagy and decreased accumulation of mitochondria with abnormal structure and function are achieved through PTEN-induced putative kinase 1 (PINK1)-mediated phosphorylation of MFN2, which in turn promotes Parkin translocation to mitochondria [47, 48]. Previous studies indicated that the protective effect of Notch1 on cardiomyocytes under myocardial reperfusion injury is associated with a decrease in MFN2 phosphorylation [49]. OPA1 has an important role in maintaining cardiac health and function. The homozygous mutation of OPA1 is lethal in mice, and even heterozygous OPA1+/- mice show significant cardiac dysfunction by the age of 12 months [50]. The overexpression of OPA1 encourages the assembly of respiratory chain supercomplexes (RCS) and the formation of tight cristae, in contrast to its conditional ablation, which compromises cristae integrity and slows respiratory function and growth [51].

Mitochondrial fission

Mitochondrial fission is the process by which a mitochondrion fragmentates into smaller mitochondria. DRP1 is a GTP-hydrolyzing enzyme that plays a crucial role in mediating mitochondrial fission [52, 53]. DRP1 and its receptors—mitochondrial fission protein 1 (FIS1), mitochondrial dynamics proteins of 49 and 51 kilodaltons (Mid49/51), and mitochondrial fission factor (MFF)—regulate mitochondrial fission [54]. The DRP1 protein is divided into four structural domains: the GTPase domain, which generates mechanical force; the self-assembly domain; the GTPase effector domain, which regulates the amount of GTPase activity; and the variable domain (B-insert), which mediates interactions between proteins. Phosphorylation, SUMOylation, and O-GlcNAcylation represent some examples of the plethora of post-translational modifications that have been linked to the B-insert gene [55]. Stress-induced alterations in the post-translational structure of DRP1, its translocation to the OMM, and its binding to receptors at INF2-marked precontraction sites, all contribute to the cellular response known as mitochondrial contraction. During this process, DRP1 oligomerizes and forms ring-like multimeric structures that resemble those of dynamin rings. In a GTP-dependent process, DRP1 allows the mitochondrion to divide into two separate mitochondria. DRP1 activation and mitochondrial fission are both influenced by posttranscriptional modifications, particularly phosphorylation and dephosphorylation, as well as SUMOylation and deSUMOylation. Ser616, Ser637, and Ser656 phosphorylation sites in DRP1 protein have been extensively studied and characterized [53, 56].

Ubiquitin proteasome system (UPS) and mitochondrial unfolded protein response (UPR.mt)

Typically, mislocalized or damaged proteins are degraded by the ubiquitin–proteasome system (UPS) in the cytosol [57, 58]. Proteases, such as CLPP and LONP1, are responsible for the degradation of damaged or misfolded proteins inside the mitochondria, while chaperones, such as HSP10, HSP60, and DNAJ, are responsible for ensuring that newly imported proteins fold properly. In response to mitochondrial stresses, such as oxidative stress, ischemia/reperfusion (I/R) injury (IRI), impairment of mitochondrial DNA or metabolic function, and/or perturbation of mitochondrial proteostasis, the mitochondrial unfolded protein response (UPRmt) is activated [59,60,61]. Particularly, UPRmt represents an adaptive program that controls mitochondrial homeostasis and reduces unfolded and misfolded protein amounts, thanks to its adaptive ability to respond to stressful conditions by promoting the transcription of genes encoding mitochondrial chaperones and proteases. This intricate quality control mechanism, also involved in sensing defects in mitochondrial translation and OXPHOS biogenesis, contributes to combat mitochondrial stress and dysfunction, exerting pleiotropic actions for maintaining cellular homeostasis [62,63,64,65]. The UPRmt is widely accepted as the first line of defense against mitochondrial damage. When stress levels rise above a certain threshold, mitophagy is triggered and removes dysfunctional mitochondria. The expression of UPRmt genes is regulated in mammals by the activating transcription factor 5 (ATF5), which contains a basic leucine zipper (bZIP) that binds to mitochondrial targeting sequences (MTSs) and nuclear localization sequences (NLSs). Under physiological conditions, ATF5 is degraded by the protease LON, which specifically targets ATF5 in the mitochondrial matrix [66,67,68]. When mitochondrial proteostasis is disrupted, ATF5 activates genes encoding for UPRmt-associated proteins [63]. In addition to ATF5, ATF4 and CHOP also play a role in UPRmt activation [63, 69]. How ATF4, CHOP, and ATF5 all contribute to UPRmt activation is currently not completely deciphered. However, the integrated stress response (ISR), which involves the regulation of protein biosynthesis in response to mitochondrial damage, is connected to these three transcription factors [63, 70]. By pharmacologically stimulating the ISR while blood flow through coronary arteries is restored, IRI can be prevented [71]. The primary regulator of the ISR, i.e., eIF2, is activated by the mitochondrial stress transmitter axis OMA1/DELE1/HRI [72]. Increased cardiomyocyte survival, decreased ATP demand, and reduced ROS production after IRI, are all results of eIF2 activation [73, 74]. After eIF2 activation, expression of the transcription factors ATF4, ATF5, and CHOP is enhanced, leading to increased expression of chaperones and proteases [75, 76]. In mammals, UPRmt activation correlates with the induction of the ISR and eIF2. However, additional investigation aimed at improving the knowledge regarding the role of mitochondrial dysfunction, the UPRmt, and the ISR in cardiac IRI is required.

Mitophagy

Unnecessary or broken cell components must be removed from all cells. The accumulation of toxic waste, the accommodation of new elements, and the reuse of old building blocks are all unfeasible without the complex system of autophagy [77, 78], subclassified in macroautophagy, microautophagy, and chaperone-mediated autophagy (CMA). To engulf and transport cargo to the lysosome, macroautophagy requires autophagosome formation. The direct translocation of unfolded proteins across the lysosomal membrane is facilitated by CMA. In microautophagy, smaller waste products are directly engulfed in the lysosomal matrix through invaginations of the lysosomal membrane [78, 79]. The best-characterized pathway for mitochondrial turnover, i.e., macroautophagy—henceforth referred to as autophagy—involves the degradation of large cellular components or even organelles, including mitochondria (mitophagy).

Mitophagy is responsible for removing damaged or redundant mitochondria. Mitophagy is crucial to the quality control system of the mitochondria being an integral component of the mitochondrial response to stressor inputs [66, 80]. As a consequence, when mitophagy is impaired, mitochondria do not work or there is an accumulation of mitochondria in the cytoplasm, which can alter cell homeostasis and lead to diseases. Two main mechanisms of mitophagy have been identified: ubiquitin-mediated mitophagy and receptor-mediated mitophagy, which may, respectively, occur through the PINK1-Parkin-mediated ubiquitin pathway and the FUN14 domain-containing protein 1 (FUNDC1) receptor-mediated pathway [81, 82]. Some essential mitochondrial proteins are ubiquitylated in PINK1-Parkin mediated regulation of mitophagy. The kinase PINK1 is first synthesized in the cytoplasm and subsequently translocates into the IMM, in which it is targeted to the ubiquitin-mediated proteasome pathway upon cleavage by PARL protease [83]. Selective accumulation of the active PINK1 in OMM occurs when PINK1 degradation is inhibited because of mitochondrial impairment or loss of mitochondrial membrane potential (ΔΨm) [84]. OMM-localized PINK1 homodimerizes, auto-phosphorylates, and phosphorylates ubiquitin (Ub) at Ser65. The phospho-Ser65-Ub, in turn, recruits cytosolic Parkin to the OMM, where it is phosphorylated by PINK1 at Ser65 and can thereby initiate its E3 ubiquitin ligase activity [85]. Parkin is thus able to ubiquitinate multiple targets that are located on the outer mitochondrial surface, including voltage-dependent anion channel (VDAC), MIRO1, and MFN1/2 [86]. Notably, MFN2 phosphorylation by Parkin is instrumental to prevent the fusion between dysfunctional and healthy mitochondria [37]. Autophagy receptors, such as Sequestosome 1 (p62/SQSTM1), optineurin (OPTN), neighbor of BRCA1 gene 1 (NBR1), and coiled-coil domain-containing protein 2 (NDP52), recognize the ubiquitinated mitochondrial proteins and link them to LC3-II for autophagic degradation. LC3-II binds the autophagosomal membrane to autophagy proteins (ATG), and together drive the OMM to fuse with lysosomes, thereby forming the autolysosomes. Finally, cathepsins and lipases degrade the autophagosome content removing damage mitochondria [87, 88].

Mitophagy may also occur in a ubiquitin-independent manner, which does not involve Parkin, but is mediated by mitophagic receptors that are expressed on the mitochondrial outer membrane [89, 90]. These receptors include FUNDC1 [91], as well as the Bcl-2 interacting protein 3 (BNIP3) and its analogue NIX [92]. The crucial component of these receptors is a conserved LC3-interacting region (LIR) that binds to and connects LC3 to the mitochondrial outer surface. FUNDC1 phosphorylation at Tyr18 by receptor-interacting serine/threonine kinase 3 (RIPK3) and SRC prevents the binding of the LIR domain to LC3-II. However, hypoxia, which represents a common hallmark of both CVDs and cancer, prevents FUNDC1 phosphorylation and thus enables LC3-II recognition and binding to FUNDC1, thereby targeting the mitochondria for mitophagy [81, 82]. Similarly, BNIP3 and NIX expression is increased by hypoxia and regulated by HIF-1α. However, they also belong to the pro-apoptotic BH3-only proteins and can also participate in OMM permeabilization, mitochondrial permeability transition pore (mPTP) opening and cytochrome C release. Therefore, their role in mitophagy needs to be further investigated [81, 82].

Mitochondrial dysfunction in cardiovascular diseases (CVDs)

The progressive decline of mitochondrial function occurring during CVDs is associated with alterations in the respiratory chain and ATP synthesis, excessive ROS production, and structural abnormality of mitochondria. These processes lead to cell damage and cardiomyocyte death occurring via apoptosis, triggered by cytochrome c release, or necrosis, induced by mPTP opening [93, 94]. Accordingly, it is widely accepted that inhibiting mPTP blunts the loss of cardiac myocytes driving numerous heart diseases, including myocardial IRI, diverse cardiomyopathies, and the cardiotoxic actions of anti-cancer agents [95]. Although there are no large, randomized studies assessing the therapeutic potential of strategies aimed at preventing/relieving mitochondrial dysfunction in CVDs, numerous pre-clinical researches have been conducted in this field highlighting mitochondrial function as a promising target for the treatment of these pathologic states, that represent a major health problem worldwide, as well as the main cause of death in the Western world [96, 97]. Indeed, it is widely accepted that maintaining mitochondrial function and integrity is a major factor for the physiological functions of cells and their survival, particularly for non-dividing cells endowed with a high-energy demand and producing high levels of oxidative stress, such as cardiomyocytes [96, 97]. Therefore, it is imperative to ensure and coordinate important quality control mechanisms in cardiomyocytes, taking part to mitochondrial biogenesis, mitochondrial dynamics, and mitophagy, to prevent mitochondrial dysfunction and the progression of CVDs [98].

Imbalance of mitochondrial dynamics and mitophagy

An emerging body of pre-clinical evidence indicates that alterations of intrinsic quality control mechanisms, including fusion, fission, and mitochondrial autophagy, are implicated in the progression of numerous CVDs (Fig. 1). As a matter of fact, a central link exists between the dysregulated mitochondrial dynamics and mitophagy, the accumulation of dysfunctional mitochondria and impaired ATP production, leading to mitochondrial dysfunction and cardiac disease. Recent reports also indicate that key metabolic changes drive pathological cardiac remodelling, and mitochondrial dynamic appears fundamental for the correct balance between energy demand and nutrient supply, suggesting that changes in mitochondrial morphology may act as a mechanism for bioenergetic adaptation during HF [99].

Fig. 1
figure 1

General representation illustrating the consequences of the imbalance of mitochondrial dynamics, culminating in heart disease. See text for details. Drp1 dynamin-related protein 1, Mfn1 mitofusin 1, Mfn2 mitofusin 2, OPA1 optic atrophy 1, PINK1 PTEN-induced putative kinase 1, ΔΨm mitochondrial membrane potential

One of the reason behind the association between alteration of mitochondrial dynamics and numerous CVDs lies in the fact that mitochondrial dynamics often represents the first response against mitochondrial stress, such as changes in ΔΨm, and this is particularly true for cardiomyocytes, since mitochondria are especially abundant in cardiac tissue, representing approximately 30% of the total cell volume, and since cardiomyocytes require high levels of ATP to support cardiac function [100].

Given the fundamental role of mitophagy, as an autophagic response targeting damaged, dysfunctional, and cytotoxic cardiac mitochondria, it can be assumed that disturbed mitophagy can cause mitochondrial damage and ultimately cardiomyocyte death. Direct evidence comes from genetic and pharmacological studies in rodent models of CVDs, indicating that altering autophagy or mitophagy, for example by limiting autophagic or mitophagic flux at the whole-body level or in specific compartments of the cardiovascular system, can increase the propensity to spontaneously develop cardiodegenerative disorders or influence disease severity in several cardiovascular conditions, including ischemic heart disease, cardiomyopathies, and atherosclerosis. In this regard, Bravo-San Pedro et al. provided important mechanistic details, highlighting the cardiovascular consequences of deleting autophagy- and mitophagy-relevant genes in the heart, including but not limited to Atg5 (autophagy related 5), Bnip3l (BCL2 interacting protein 3 like), Dnml1 (dynamin 1 like), Lamp2 (lysosomal-associated membrane protein 2), Mfn1, Mfn2, the proautophagic transcription factor FOXO3 (forkhead box O3), Park2 [Parkinson disease (autosomal recessive, juvenile) 2], parkin, and Pink1 (PTEN-induced putative kinase 1) [101]. Although the detailed molecular mechanisms by which disrupted mitophagy leads to mitochondrial damage and cardiomyocyte death, driving the progression of CVDs, are largely unknown, a direct link has been established between mitochondrial damage secondary to cardiovascular stress and the exacerbation of the imbalance of autophagy or mitophagy (excessive inhibition or promotion). This indicates that preserving autophagic/mitophagic responses is crucial to safeguard the cardiovascular homeostasis and prevent premature degenerative changes in cardiovascular tissues. For instance, in coronary heart disease/acute MI or IRI, where increasing studies are delineating the role of FUNDC1-, PINK/Parkin-, and BNIP3-dependent mitophagy, these pathways represent potential important targets to improve myocardial IRI (see the recent review by Liu and Mu for further details) [102]. Other studies reported the protective role of mitophagy during IRI and other numerous CVDs, including diabetic cardiomyopathy, HF, hypertension, arrhythmia, stroke, valvular dysfunction, and aging (see the recent review by Yang et al. for further details) [103].

Collectively, if many studies indicated mitophagy as a key protective program for mitochondrial function and cardiomyocyte homeostasis, other experimental findings reported that mitophagy may also play a negative role in heart disease, such as IRI and HF. Indeed, ischemic/hypoxic environment and/or the excess of ROS (i.e., during myocardial reperfusion) may activate mitophagy through BNIP3, Beclin-1 and ATG5, impacting ATP production, and exacerbating myocardial injury. A “mitophagy paradox” has also been hypothesized in IRI, since mitophagy seems to be protective (adaptive) during acute ischemic phase, where it contributes in removing dysfunctional mitochondria and preserving basal metabolic requirements, while the excessive mitochondrial fission, fragmentation and mitophagy (maladaptive) may induce unnecessary mitochondrial clearance and impair mitochondrial integrity and function during the reperfusion phase [104].

Noteworthy, the role of mitophagy during IRI appears even more complex if we consider that mitophagy also acts as “mediator” during inflammatory damage of cardiomyocytes or vascular endothelial cells to remove inflammation-induced damaged mitochondria under normal conditions [102]. It is known that inflammation plays a central role in the early stages of coronary heart disease, driving the formation of atherosclerotic plaque or myocardial fiber plaque [105]. Impaired mitophagy also associates with inflammation, particularly through the activation of NLRP3 inflammasome, which in turn can be responsible for endothelial damage and increased accumulation of cholesterol in macrophages [106]. Additionally, a defective mitophagy in mice lacking Fundc1, a newly characterized mitophagy receptor, is associated with metabolic disorders via MAPK signaling and inflammatory responses [107]. Conversely, a very recent study reported that tumor necrosis factor receptor-associated factor-2 (TRAF2), an innate immunity effector, is essential for physiological mitophagy in cardiomyocytes, since its myocardial loss (in inducible cardiac-myocyte specific TRAF2 knockout mice) impairs physiological mitophagy in the heart inducing cardiac inflammation, thus supporting a cardioprotective role of TRAF2-stimulated mitophagy [108].

On the other hand, an uncontrolled activation of mitophagy can lead to the accumulation of damaged mitochondria and inflammatory response activation, reduction of myocytes, and contractile impairment, which ultimately contribute to cardiac aging and HF [99]. Other evidence indicates that excessive mitophagy leads to mitochondrial population decline, impaired oxidative phosphorylation, and decreased ATP production, culminating in HF [109, 110]. According to these reports, in the failing heart, the stress-induced mitophagy acts as a maladaptive response to hemodynamic parameters such as pressure overload inducing a negative remodeling of the myocardium, mainly mediated by Beclin-1 [111, 112]. Overall, given the complex and multifaceted involvment of mitophagy in the regulation of myocardial injury, it should be considered that the protective or damaging role in activating mitophagy needs to be contextualized to the degree and duration of cell stress, the different stages of myocardial injury, as well as the degree of mitophagy that, physiologically can be beneficial, while at excessive or inadequate levels can be deleterious.

Implication of the fission protein DRP1 in CVDs

Direct evidence attesting the crucial implication of mitochondrial dynamics in CVDs derives from studies reporting how alterations in the role of endogenous factors, that physiologically regulate mitochondrial fission and fusion, can dramatically participate to the onset and progression of cardiac dysfunction. For instance, mitochondrial fission abnormalities generated by Drp1 disruption may induce mitochondrial elongation and inhibit mitochondrial autophagy, thereby resulting in mitochondrial dysfunction and promoting cardiac dysfunction and increased susceptibility to IRI in mice [113]. Similarly, downregulation of mitochondrial autophagy can induce mitochondrial dysfunction and HF, while its restoration mitigated the progression of HF in a mouse model of pressure overload, where endogenous Drp1 was crucial for mediating mitochondrial autophagy and maintaining both mitochondrial and cardiac function [114]. Another study indicated that Drp1 inhibition significantly contributed to the accumulation of altered cardiac mitochondria by repressing BNIP3-induced mitophagy, suggesting that that Drp1-mediated mitochondrial fission is a prerequisite for mitophagy during cardiac hypertrophy [56, 115].

The essential role of endogenous Drp1 was further demonstrated in adult mouse cardiomyocytes with ablated Drp1, that caused dramatic alterations in mitochondrial fission and promoted mitophagic mitochondrial depletion, contributing to the lethal cardiomyopathy [116]. In support of this, the homozygous deletion of Drp1 in mice is responsible for embryonic lethality, which is due to elongated mitochondria, altered apoptosis and reduced cell proliferation, while postnatal cardiac specific knock out of Drp1 leads to dilated cardiomyopathy and rapid lethality in mice [117, 118]. Based on these knowledge, the role of Drp1 in the pathogenesis of several CVDs, including pulmonary arterial hypertension, HF, cardiac hypertrophy, IRI and myocardial infarction (MI), has been extensively addressed in the last years, highlighting the potentiality of this factor to effectively represent a strategy to treat CVDs [56].

However, Drp1 upregulation has also been reported in several CVDs, where the excessive fission of mitochondria impairs the cardiac function (Fig. 2). For instance, a dominant-negative mutant form of Drp1 decreased mitochondrial fission, mPTP sensitivity, and cell death in cardiac-derived HL-1 cells subjected to IRI; a similar cardioprotective effect was obtained by inhibiting Drp1 with the Mitochondrial division inhibitor-1 (Mdivi-1) in HL-1 and mice exposed to coronary artery occlusion and reperfusion [119]. Additional reports confirmed the cardioprotective action of Drp1 inhibition against IRI in rat and mouse models by preserving mitochondrial function and reducing cell death [120,121,122,123]. Drp1 inhibition was also effective in relieving myocardial dysfunction in in vitro and in vivo models of insulin resistance and metabolic cardiomyopathy, as well as cardiac hypertrophy and HF [56, 100].

Fig. 2
figure 2

Proposed mechanisms underlying the detrimental effects deriving from an imbalanced mitochondrial dynamic in the heart, with particular regard to the implication of Drp1, a protein playing a fundamental role in heart physiology and cardiomyocyte survival during normal conditions, but that it is upregulated during cardiac pathological conditions, resulting in excessive fission. See text for details. CVDs cardiovascular diseases, Drp1 dynamin-related protein 1, VDAC voltage-dependent anion channel

Focusing on the very recent studies, Hu et al. provided novel evidence about the mechanism by which Drp1 can contribute to the lipid overload-induced heart hypertrophy and dysfunction through modulation of the intracellular redox environment [124]. In particular, by using high-fat diet (HFD)-induced metabolic dysfunction in mouse and monkey models, and in vitro palmitate model, the authors found that decreased NAD+ levels and increased of Drp1 acetylation at Lys642 occur. The consequent excessive Drp1 activation through phosphorylation, mitochondrial translocation, and oligomerization, showed higher GTPase activity, bound with voltage-dependent anion channel 1 (VDAC1) on mitochondria, to induce mitochondrial fission and cardiomyocyte death; this suggests that acetylation represents a novel post-translational Drp1 modification that regulates its activity, contributing to metabolic cardiomyopathy (Fig. 2).

Drp1 activation and mitochondrial fission are influenced by additional specific posttranscriptional modifications. In particular, Ser616 phosphorylation stimulates Drp1 oligomerization and mitochondrial fission by targeting Drp1 to the OMM [125]; conversely, Ser637 and Ser656 phosphorylation blocks Drp1 oligomerization and prevents mitochondrial fission [126, 127]. Ser616 phosphorylation is upregulated in response to IRI via either the activation of CDK1, PKC, PGAM5, ROCK, and GSK-3 or the inhibition of PKC [128]. IRI injury also promotes Ser637 dephosphorylation, which activates Drp1 and promotes mitochondrial fission [129]. Increased cardiac PIM-1 expression prevents cell death in response to IRI by increasing Drp1 phosphorylation at Ser637 [121]. PKA phosphorylates Ser656, while calcineurin dephosphorylates it [127, 130]. Nitrite-mediated cardiomyocyte protection against IRI has been linked to PKA activation, which inhibits Drp1 activity [130]. MiR-199 prevents calcineurin from phosphorylating Drp1 during MI, which prevents apoptosis [127]. SUMOylation by SUMO-1 promotes Drp1 localization to mitochondria, while SUMOylation by SUMO2/3 inhibits it [131]. SUMOylation can be degraded in response to the activity of SENP families [132]. A decrease in SUMO-2/3-modified Drp1 and an increase in mitochondrial fission are caused by increased SENP5 expression following cardiac IRI [133]. SENP3 exacerbates cardiac IRI by facilitating Drp1 translocation to mitochondria during reperfusion [134]. MFF undergoes posttranscriptional modifications in the same way that Drp1 does. Succinate induces MFF phosphorylation in a GPR91-dependent fashion during myocardial ischemic injury. In turn, MFF activation promotes mitochondrial fission after ERK1/2 activation [135]. Furthermore, NR4A1 expression is upregulated after cardiac microvascular reperfusion injury; this upregulation promotes MFF phosphorylation by activating CK2a, leading to fatal mitochondrial fission [136]. Myocardial reperfusion damage is characterized by an increase in MFF phosphorylation via activation of the JNK pathway and a heightened rate of mitochondrial fission [137]. It is unclear, however, how other posttranscriptional modifications of receptors may influence mitochondrial fission in myocardial reperfusion damage. The production of numerous daughter mitochondria, which in turn supply cardiomyocytes with ATP, is the result of efficient mitochondrial fission. In addition, damaged mitochondria cannot be removed during mitophagy without the help of a process called mitochondrial supply [138]. Particularly, the increased phosphorylation of the kinase JNK during IRI in mice activates MFF and Bnip3, that contribute to the fatal mitochondrial fission and mitophagy, respectively via the caspase 9-related apoptosis and mitochondrial metabolism alterations. The role of MFF during microcirculatory IRI has been deciphered using homozygous Mff-deficient (Mffgt) mice that, compared with the wild type counterpart, showed smaller infarct size, restored cardiac function, improved blood flow, and reduced microcirculatory perfusion alterations [139]. This study also provides evidence on the critical contribution of MFF-dependent mitochondrial fission in mediating the microcirculatory IRI through VDAC1/hexokinase 2-mediated mPTP opening, mitochondrial ROS overproduction, and cardiolipin oxidation-involved in cytochrome c leakage into the cytoplasm. Other studies reported an important implication of MFF-dependent mitochondrial division in the development of HF. In this regard, Chen et al. demonstrated that Mff mutant mice exhibit, in addition to increased mitophagy, reduced mitochondrial density and respiratory chain activity, and died at 13 weeks due to a severe dilated cardiomyopathy leading to HF [140].

In line with these results, a recent study contributed in deciphering the implications of pathological fission and mitochondrial failure in sepsis-induced cardiomyopathy, focusing on the role of Drp1/FIS1 (mitochondrial adaptor fission 1 interaction of Drp1) interaction. P110, an adaptor-specific peptide inhibitor of Drp1/FIS1 interaction, was found to inhibit lipopolysaccharide (LPS)-induced oxidative stress and mitochondrial fragmentation in H9c2 cells and Balb/c mice, which presented an improved cardiac function and reduced mortality. These data indicate that Drp1/FIS1-mediated mitochondrial fission represents an important mechanism of cardiac dysfunction during sepsis [141]. On the other hand, Sun et al. further elucidated the role of Drp1 during IRI in human and mouse models by indicating the beneficial contribution of the TBC domain family member 15 (TBC1D15) in the regulation of mitochondrial homeostasis through the interaction with Drp1 at the mitochondria–lysosome contacts [142]. This report highlights that TBC1D15 may also serve as a potential regulator of mitochondrial homeostasis in the heart and during IRI, suggesting its combination with Drp1 to support asymmetrical mitochondrial fission and clearance for preserving mitochondrial integrity.

Involvement of the mitochondrial fusion mitofusin 2 (Mfn2) and optic atrophy 1 (OPA1) in CVDs

The specific cardiac deletion of Mfn2 leads to cardiac hypertrophy with systolic dysfunction [37, 143]. On the other hand, it has been demonstrated that Mfn2 prevents the excessive accumulation of autophagosomes during the IRI phase of the cell cycle in the heart and mediates autophagosome fusion with lysosomes [144]. The conditional combined ablation of Mfn1/Mfn2 in adult hearts induced mitochondrial fragmentation, cardiomyocyte and mitochondrial respiratory dysfunctions, and rapidly progressive and lethal dilated cardiomyopathy (DCM) [145]. In vitro findings obtained in HL-1 cardiac cells and neonatal rat ventricular myocytes (NRVMs) also demonstrated that the Mitofusins may have a cardioprotective role against cardiac dysfunction and IRI, since the overexpression of Mfn1 or Mfn2 can mitigate cell death, while knockdown of Mfn2 can worsen cell survival following acute IRI [119, 146]. Other evidence indicates that Mfn2 ablation reduces mitochondrial Ca2+ overload, interrupts fatal Drp1-induced mitochondrial fission and PINK1/Parkin-induced mitophagy, and reduces some of the damage caused by low oxygen levels in cardiac microvascular endothelial cells (CMEC) [147]. These effects, however, are highly dependent on the amount of Mfn2 present. In this regard, in the case of Mfn2 overexpression, Mfn2 facilitates the process by which mitochondria adhere to the sarcoplasmic reticulum (SR). As a result, there is an excess of free Ca2+ in the mitochondria, which ultimately leads to mitochondrial damage during IRI [148]. In addition, the hearts of adult mice with combined acute ablation of both Mfn1 and Mfn2, despite apparent mitochondrial dysfunction, were protected from in vivo acute MI and IRI; in particular, the cardioprotective phenotype of Mfn1 and Mfn2 depletion was due to beneficial effects on mitochondrial Ca2+ levels, oxidative stress, and mPTP opening [146, 149]. These contrasting results may depend on the pleiotropic actions of the Mitofusin proteins [150], as well as on the experimental/clinical context [151]; indeed, given the crucial role of Mfn2 in tethering SR and mitochondria to ensure the physiological inter-organellar Ca2+ signaling in the myocardium [152], the reduced interaction between mitochondria and SR during IRI reduces mitochondrial Ca2+ overload, resulting in cardioprotection [149]. These findings suggest that the role of Mitofusins in tethering the SR to mitochondria is more relevant for affording cardioprotection against acute IRI than changes in mitochondrial morphology per se. The conclusions of this study were further confirmed by Seidlmayer et al. [153], who generated cardio-specific tamoxifen-inducible Mfn2 KO mice to show that, when the physical linkage between SR and mitochondria by Mfn2 is disrupted, inositol 1,4,5-trisphosphate (IP3)-mediated SR Ca2+ release fails to induce ATP production. These findings may explain the decreased mitochondrial network excitability in HF cardiomyocytes, caused by disrupted mitochondrial network ultrastructure and impaired mitochondrial functional tethering, accompanied by decreased Mfn1 and Mfn2 levels [154].

On the other hand, several studies reported decreased levels of fusion proteins, including Mfn2 and OPA1, in failing hearts of small and large pre-clinical models of CVDs and patients with HF. In particular, depending on the aetiology and time course of hypertrophy, Mfn2 was downregulated in phenylephrine treated NRVMs, as well as and in hypertrophied failing hearts of spontaneously hypertensive rats and mice with pressure-overload hypertrophy. These findings establish a causal relationship between Mfn2 and cardiac hypertrophy [155]. Consistently, Sabbah et al. observed decreased levels of fusion-associated proteins (Mfn2 and OPA-1) concomitant to increased levels of fission-associated proteins (Fis-1 and Drp1) in left ventricular tissue from dogs and humans with HF (idiopathic dilated cardiomyopathy or ischemic cardiomyopathy), that were reverted by a long-term therapy with elamipretide (ELAM), a water-soluble tetrapeptide targeting mitochondrial dynamics [156]. Decreased levels of Mfn1 were also reported in cardiac tissues of HF patients presenting with idiopathic DCM who do not respond to established multidisciplinary treatment and associate with poor outcomes, that correlated with increased mitochondrial fragmentation [157]. As elegantly reviewed in [143], although additional studies are required for improving the knowledge regarding the involvement of Mitofusin proteins in cardiac hypertrophy and HF, enhancing Mfn2 function may be useful to obtain a therapeutic effect.

Similarly, the protein levels of OPA1 were decreased in rat and human ischaemic HF, and in H9c2 cells exposed to ischemia; on the other hand, reducing OPA1 through shRNA increased mitochondrial fragmentation and decreased tubularity of the mitochondria [158]. The crucial implication of OPA1 was confirmed by the lethal embryonic phenotype following its complete ablation resulting in small and fragmented mitochondria and altered fusion [158, 159], while OPA1+/– mice exhibited late-onset (12 months) left ventricular dysfunction associated with impaired mitochondrial function [50]. Reductions in OPA1 expression and mitochondrial respiratory complex downregulation have been found in hearts damaged by reperfusion. Reducing myocardial injury through drug pretreatment or OPA1 overexpression improves mitochondrial function and decreases mitochondrial fission and expression of the mitochondrial respiratory complex [160]. In reperfused heart tissue, OPA1 deletion contributes to increase the severity of mitochondrial fragmentation and cell damage [161]. Also in this case, additional studies in humans are required to extend the knowledge on OPA1 regulation to move towards the development of novel therapies for OPA1-specific cardiac disease progression [162].

Overall, these findings demonstrate that maintaining an adequate balance between mitochondrial fusion and fission, as well as an optimal mitophagy, is determinant for mitochondrial morphology and cardiac physiology and that, following mitochondrial stress, fusion and fission represent a key response to preserve a healthy mitochondrial network. The effects of fusion and fission are dependent by the magnitude of activation of each process and their balance, as well as by the molecular actors participating in their modulation. This may likely explain their adaptive or maladaptive response following cardiovascular stress [163]. However, the mechanisms by which mitochondrial fusion and fission machineries are altered during CVDs and how this disruption exactly promotes pathological states need to be further explored. The current scenario suggests that a decrease in fusion process, together with increased Drp1 activation, can shift the balance towards fission process, leading to excessive mitochondrial fission and favouring the onset and development of CVDs [164].

Another aspect that should not be underestimated, and that further proves the key role of balanced fusion and fission of mitochondria for an optimal cardiac function, regards the importance of a balanced OPA1 processing, regulated by the two mitochondrial proteases, OMA1 and the AAA protease YME1L. As demonstrated by Wai et al. [165], L-OPA1 is required for mitochondrial fusion and preserves cardiac function. The authors found that the cardiac-specific ablation of Yme1l can activate OMA1 inducing OPA1 proteolysis, resulting in mitochondrial fragmentation, altered cardiac metabolism and HF. On the other hand, cardiac function and mitochondrial morphology were restored by Oma1 deletion, which prevents OPA1 cleavage and mitochondrial fragmentation. Interestingly, this study also showed that mitochondrial fragmentation was accompanied by a metabolic switch from fatty acid to glucose utilization in the heart, a condition that was reverted by metabolic interventions able to preserve cardiac function despite the presence of mitochondrial fragmentation [165]. These findings not only demonstrate the crucial role of OPA1 processing for mitochondrial and heart function, but also suggest that the metabolic switch in the substrate utilization occurring in the failing adult heart may represent a maladaptive process, as previously reported [94].

Nan et al. demonstrated that epigallocatechin gallate, acting as an OMA1 inhibitor, protects cardiomyocytes from IRI damage by reducing OPA1 cleavage; OMA1 activity and S-OPA1 levels are important targets to attenuate myocardial IRI [166]. However, the mechanism that triggers their activation in response to myocardial damage is still unclear. Kent and colleagues found that neither ROS scavengers nor mPTP inhibitors play a causal role in OMA1 activation, even though both increase cardiac and mitochondrial function [167]. Jang et al. demonstrated that high Ca2+-mediated mitochondrial swelling facilitates L-OPA1 cleavage in cardiac mitochondria [168]. YME1L deletion in cardiomyocytes can also be triggered by activating OMA1 and triggering L-OPA1 proteolysis [165].

UPRmt in heart diseases

The involvement of UPRmt in cardiac pathophysiology has been widely investigated and, although there is an important consensus about its adaptive cardioprotective role in diverse CVDs (including IRI, chronic hemodynamic overload and HF), some studies suggested that UPRmt may promote the progression of heart diseases, since blocking several UPRmt elements can reduce the signs of HF of a different aetiology.

Smyrnias et al. found an activated UPRmt in isolated cardiac myocytes exposed to different stressful stimuli and in mice subjected to chronic pressure overload, while pharmacologically boosting UPRmt using nicotinamide riboside significantly reduced cardiac dysfunction [169]. In the hearts of patients with aortic stenosis, who often show left ventricular pressure overload, they also observed an increased expression of UPRmt-associated genes, that correlated with a reduction of cardiac dysfunction-related biomarkers (i.e., high-sensitivity troponin T and N-terminal pro–B-type natriuretic peptide) [169]. These results indicate the importance of UPRmt stimulation to promote cardioprotection, providing important evidence that UPRmt may serve as an adaptive process. Xu and collaborators demonstrated that choline can attenuate the mito-nuclear protein imbalance and that the activation of UPRmt is fundamental for choline-mediated cardioprotection, thereby preserving the ultrastructure and function of mitochondria during cardiac hypertrophy [170]. According to these findings, other recent evidence confirmed the cytoprotective significance of UPRmt signalling following mitochondrial dysfunction in heart diseases. In particular, the UPRmt inducers oligomycin or doxycycline protected the heart against IRI in ex vivo and in vivo settings in a mechanism strictly requiring ATF5, since UPRmt induced cardioprotection in wild-type but not ATF5-deficient mice, providing first evidence about the fundamental action of ATF5 for the effectiveness of mammalian UPRmt in vivo [171]. In addition, Zhang et al. have demonstrated that ATF5 is a downstream effector of PGC-1. UPRmt activation via the PGC-1/ATF5 axis mediates a cardioprotective effect in pathological cardiac hypertrophy [172]. On the other hand, by using both gain- and loss-of-function mouse models, together with isolated NRVMs, Venkatesh et al. demonstrated that LonP1 (i.e., an essential mitochondrial protease and component of UPRmt with key roles in maintaining mitochondrial proteostasis and mitigating cell stress) preserves mitochondrial redox status and reduces both oxidative protein damage and cardiomyocyte apoptosis during IRI by reducing Complex I activity [173]. Another investigation indicated that, in a rat model of neurogenic hypertension induced by intracerebroventricular infusion of angiotensin II, enhancing sympatho-excitation can block miR-18a-5p/HIF-1α signaling and increase mitochondrial stress proteotoxicity, reflected by decreased UPRmt and mitochondrial dynamics/OXPHOS/ΔΨm, and oxidative stress. These findings indicate that the consequent mitochondrial abnormalities drive pathological cardiac remodeling culminating in cardiomyopathy [174].

There is also evidence that UPRmt might be associated with harmful events in the heart. For instance, the localization of the mitochondrial chaperone molecule heat shock protein 60 (HSP60) to the cell surface triggered the innate immune system, inducing the release of tumor necrosis factor-α (TNF-α), which correlated with increased myocyte apoptosis in a rat coronary ligation model of HF. The excessive mitochondrial activation of Omi/HtrA2, a mitochondrial serine protease involved in mitochondrial homeostasis, whose levels are increased during UPRmt, decreased Δψm, induced cytochrome c release into cytosol through its serine protease activity, and promoted apoptosis in the hearts of aged rats [175]. Additionally, transgenic mice overexpressing cardiac-specific mitochondrial Omi/HtrA2 showed increased myocardial apoptosis and decreased systolic and diastolic function; these effects were prevented by Ucf-101, a specific Omi/HtrA2 inhibitor [125]. Other studies reported increased levels of numerous factors participating to the UPRmt and exerting maladaptive roles, in pre-clinical and clinical models of HF [176,177,178].

Overall, these studies suggest that the degree of UPRmt activation represents an important element in determining the beneficial or the detrimental cardiac effect of UPRmt. In particular, as an adaptive response, UPRmt induces a fundamental cardioprotective response following different cardiac stressful conditions when is moderately activated; under these conditions, UPRmt ensures the optimal control of damaged mitochondrial proteins and maintains the mitochondrial and cardiac functions. Conversely, when excessively activated, UPRmt may be responsible for an exaggerated cleavage of mitochondrial proteins, resulting in a cardiotoxic response and exacerbating cardiac dysfunction. Moreover, in examining the adaptive or maladaptive role of UPRmt during heart diseases, the different experimental settings, the type and duration of stressful conditions, as well as the complex aetiology and the multifactorial aspects of these pathologies, should be considered. This may provide other possible explanations behind the conflicting results on the dual (cardioprotective/cardiotoxic) role of UPRmt in the heart.

Noteworthy, these studies also indicate that endogenous activation of UPRmt appeared to be not sufficient to completely prevent myocardial injury, since the administration of UPRmt activators seems necessary to further improve UPRmt activity and afford additional cardioprotective action. Although other studies are required for explaining this mechanism, it should be noted that UPRmt is also involved in the removal of dysfunctional mitochondria by mitophagy. Indeed, the misfolding of mitochondrial proteins, that are critical for mitochondrial function, induces mitophagy and UPRmt co-activation, highlighting the importance of UPRmt as a regulator of mitochondrial quality control in the heart and indicating that myocardial stress activates endogenous UPRmt and mitophagy that work together to sustain mitochondrial performance and cardiac function [157]. The relationship between mitophagy and UPRmt was also recently demonstrated in a pre-clinical model of septic cardiomyopathy; here, the authors observed that mitophagy activation represents a direct consequence to the fact that endogenous UPRmt cannot completely repair mitochondrial damage under stress conditions; on the other hand, they also suggest that UPRmt may act as a compensatory mechanism in response to mitophagy repression [179].

Mitochondrial DNA mutations and CVDs

Mitochondrial dysfunctions leading to CVDs may also be due to mutations in either mitochondrial genome (mtDNA) or nuclear genome (nDNA). A primary risk factor for CVDs is indeed represented by mtDNA mutations, which can lead to disruption of mitochondrial homeostasis, oxidative stress, OXPHOS impairment, and energy metabolism damage. Therefore, genetic investigation of mtDNA mutation may be a useful therapeutic approach to detect and forecast CVDs [180]. mtDNA is a circular, double-stranded genome that of 16,569 base pairs in length and contains 37 genes necessary for aerobic respiration and the synthesis of cellular energy via the OXPHOS pathway. Since mtDNA is not recombined and is not protected by histones, its mutation rate is roughly 10–100-fold higher than nuclear DNA [181]. Mutations of mtDNA, which can occur as point mutations, deletions, fragment deletions, or large-scale mtDNA rearrangements, may directly impair OXPHOS [182]. Many cardiac diseases have been linked to mutations in mtDNA [183]. For instance, cardiomyopathy has been linked to mtDNA mutations in mitochondrial DNA, which supports the idea that proteins encoded by mtDNA are crucial to mitochondrial function in the heart [184]. HF may also be regarded as a bioenergetic disorder characterized by extensive mutations in mtDNA as well as by mitochondrial dysfunction, as outlined above [185]. The clinical manifestation of infantile cardiomyopathy is linked to the MRPL44-disorder, which disrupts the translation of a partial protein involved in OXPHOS [186], 187. mtDNA and functional integrity cannot be preserved in cardiomyocytes with myocardial hypertrophy [188] due to abnormal mitochondrial structure and dysfunction of mitophagy clearance. In nearly 40% of patients [189], hypertrophic cardiomyopathy is the most common form of cardiomyopathy associated with mtDNA diseases. Maternally inherited essential hypertension (MIEH) has been linked to mtDNA mutations in a number of studies [182, 190] suggesting that these changes may be one of the pathological mechanisms underlying MIEH. Furthermore, atherosclerosis is linked to mtDNA mutations [191]. The MT-RNR1 gene mutation m.A1555 G, the MT-TL1 gene mutation m.C3256 T, the MT-TL2 gene mutation m.G12315A, and the MT-CYB gene mutation m.G15059A have all been linked to atherosclerosis [192]. A crucial determinant of CVDs may also be represented by mtDNA mutations leading to OXPHOS impairment [96]. In individuals with CHF, it has been observed that a decrease in succinyl-CoA levels within myocardial mitochondrial leads to a reduction in OXPHOS [193]. In a recent analysis of samples harvested from human thoracic aortic aneurysm tissues, the suppression of mitochondrial OXPHOS-related gene expression resulted in an increase in chromatin OXPHOS-related genes. However, despite this increase, the production of ATP remained inadequate to sustain the contractile activity in human aortic smooth muscle cells (HAoSMCs) [194]. In a separate study examining the impact of NOTCH1 deletion on the contractile phenotype and mitochondrial dynamics of human HAoSMCs, it was observed that NOTCH1 deficiency can lead to mitochondrial dysfunction in HAoSMCs, reflected by a decrease in mitochondrial fusion, resulting in the loss of ΔΨm, an increase in ROS generation, inadequate ATP production, and an associated impairment in the contractile phenotype [195]. The deficiency of PGC-1β in the heart leads to the suppression of gene expression related to OXPHOS. This deficiency has the potential to hinder the progression from pressure overload-induced myocardial hypertrophy to HF by regulating the activity of PGC-1β [196].

Mitochondrial dysfunction in cancer: mechanisms and therapeutic targeting

Mitochondrial dynamics in cancer

Malignant transformation associated with cancer development has been associated with abnormal mitochondrial dynamics [197,198,199]. Mitochondria exist along fragmented and fused states regulating mitochondrial metabolism and cell death, and loss of proper mitochondrial dynamics impinges on cell growth and survival [200, 201]. Dysregulated expression ratio of Drp1/MFN1-2 proteins, triggered by multiple genetic or epigenetic events [202, 203] has been found to associate with enhanced cell proliferation and metastatic behaviour [204], as well as to poor prognosis in many cancer types [205]. Mitochondrial fission mediated by ROS imbalanced production represents a primary reason of hepatocellular carcinoma cell survival [73]. Mitochondrial fission also promotes chemotherapy resistance in numerous cancer types [206]. Damaged mitochondria are removed through mitophagy via an increase in mitochondrial fission in healthy cells, while enhanced mitochondrial fission results in OXPHOS defects, cellular and mitochondrial dysfunctions, thereby supporting high proliferation and invasiveness [207]. In several preclinical studies, fission-regulating proteins have become a functional target to restore the mitochondrial network observed in healthy cells, emerging as biomarker of therapeutic efficacy [208]. Indeed, DRP1 inhibitors appear useful anti-cancer therapeutics [208] targeting the survival, apoptosis, and drug resistance of breast cancer cells [204]. For instance, the DRP1 inhibitor Mdivi-1, as well as the novel ellipticine derivative Dripitor1a, decrease oxidative metabolism in cancer and promote cell growth inhibition in vitro and in vivo [209]. In prostate cancer cells, the Drp1-partner MFF was found to interact with the voltage-dependent anion channel-1 (VDAC1) at the OMM, and blocking such interaction led to a collapse of mitochondrial functions with increased MOMP, loss of inner membrane potential, Ca2+ unbalance, bioenergetics defects, and activation of cell death [210]. Worthy of note, targeting MFF–VDAC1 complex via an MFF peptidomimetic was safe towards health cells while triggering anti-cancer activity in several preclinical models [211, 212].

Mitofusin modulation could also represent a promising anticancer strategy. Recently, it has been shown that the mitofusin activator small molecules 7 (MASM7) induces mitochondrial fusion increasing ΔΨm, mitochondrial respiration, and ATP production. Conversely, mitochondrial fusion inhibitor 8 (MFI8) induces mitochondria fragmentation decreasing ΔΨm, mitochondrial respiration, and ATP production [213] (Fig. 3).

Fig. 3
figure 3

Schematic representation of dysregulated mitochondrial dynamics in cancer suggesting that mitochondrial-dependent mechanisms may serve as potential pharmacological/therapeutic target aimed at restoring mitochondrial network and function. A brief representation is provided for the central role of mitochondria in the adverse cardiovascular events following anticancer therapies. See text for details. ATP5F1C ATP synthase F1 subunit gamma, Drp1 dynamin-related protein 1, IDH1 isocitrate dehydrogenase [NADP(+] 1, IDH2 isocitrate dehydrogenase [NADP(+)] 2, MASM7 mitofusin activator small molecules 7, Mdivi-1 mitochondrial division inhibitor-1, MFI8 mitochondrial fusion inhibitor 8, SLC25A8 UCP2 (uncoupling protein 2)

Numerous drugs such as 4-bromo-resveratrol, nicotinamide, sirtinol, cambinol, salermide induce mitochondrial impairment through the inhibition of sirtuins which are directly involved in ATP production [214, 215]. These deacetylases play a key role not only as regulators in many metabolic processes such as glycolysis, OXPHOS, oxidative stress, but they are also implicated in many diseases, including cancer [216]. Indeed, SIRT3 activates the complex I of the electron transport chain, as well as directly regulates the peroxisome proliferator-activated receptor-gamma coactivator PGC1α alongside ERRα under oxidative stress conditions [43, 217]. PGC-1/ERRα axis controls mitochondrial OXPHOS activity and it has been reported as a new pathway driving tumor progression and resistance [218]. In accordance, Peroxisome proliferator-activated receptor gamma coactivator 1-alpha (PGC-1α), the master regulator of mitogenesis, finely regulates mitochondrial dynamics through autophagy and mitophagy [219, 220]; SIRT3 and PGC1α are also correlated by the activation of the AMPK signalling pathway, which increases PGC1α gene-expression [221]. PGC1/NRF1 gene signature have been shown to predict tumor recurrence, metastasis and poor overall survival in both estrogen receptor (ER +)/Luminal-A breast cancer patients [222].

Mitochondrial dynamics has been reported to influence survival and stemness features of cancer stem cells (CSCs) [222, 223], a small sub-population of cancer cells resistant to most anti-cancer therapies, and responsible for tumor recurrence, metastatic dissemination, treatment failure and poor clinical outcomes in nearly all cancer patients [224, 225]. In this context, PGC1α is necessary for the transcription of nuclear-encoded mitochondrial genes promoting mitochondrial biogenesis that is also required for the anchorage-independent survival and propagation of CSCs [226, 227].

UPRmt in cancer

The UPRmt plays a crucial role in cancer cell survival, anticancer drugs resistance and is emerging as a new therapeutic target [228, 229]. UPR activation has been reported in many human cancers, although it remains partially characterized [230, 231]. Several studies showed a dual role of the UPR in cancer, without a clear definition of the pro-survival and the pro-death effects [229]. Activation of the cytoprotective UPR signalling can promote tumorigenesis and resistance to oxidative and therapeutic stress [232]. Moreover, the UPRmt cellular response is generated by an unbalance between nuclear DNA-encoded OXPHOS proteins and mtDNA-encoded OXPHOS proteins [233]. The UPRmt stimulates mitochondrial protective genes and ROS defence antioxidants making a homeostatic activation within the mitochondrial protein-folding environment [234]. Indeed, UPRmt intricately interfaces with stress response pathways which drive autophagy and mitophagy acting as co-factors also in Sirtuin signalling [235, 236]. Moreover, UPRmt is activated during the disruption of electron transport balance of OXPHOS complexes that are encoded by both mitochondrial and nuclear genomes [237]. Downregulation of complex IV, complex III and ubiquinone synthesis also induces UPRmt [237]. Recent studies reported that the overexpression of misfolded proteins in the mitochondrial intermembrane space, promotes an additional UPRmt axis regulated by the estrogen receptor alpha (ERα) in cancer. The ERα- UPRmt axis is initiated by ROS, which in turn activate Akt, thereby leading to the phosphorylation of both ERα and the mitochondria biogenesis transcription factor NRF1 [67, 238]. The metabolic plasticity of tumors can be also ascribed to the different dependence on UPRmt proteins. Tumor types that are more incline to electrons transport unbalance generate high levels of ROS which determine the activation of UPRmt and the chaperone-mediated recruitment of the cytosolic misfolded proteins by the mitochondria. Targeting such protein recruitment with a new combined therapy could reduce UPRmt. In this regard, many clinical studies are investigating new potential inhibitors of UPRmt even in combination with chaperone-targeting compounds, to reduce tumor progression and drug resistance phenomena [239, 240].

Mitophagy in cancer

Mitophagy plays multiple roles in cancer development and progression [241, 242]. It can serve as a tumor suppressor, maintaining the balance between mitochondria amount/activity by removing damaged or dysfunctional mitochondria in certain cancer subtypes [243, 244]. Mitophagy also reduces ROS formation and limits tumour-initiating step mediated by ROS. However, mitophagy can also drive drug resistance, preventing chemotherapy-induced apoptosis during tumour progression [241, 245]. Functional mitophagy inhibits the accumulation of damaged mitochondria and selectively removes them through autophagosomes preventing carcinogenesis. PINK1 is accumulated in dysfunctional mitochondria activating the recruitment of the E3 ubiquitin protein ligase, Parkin [246,247,248]. Many studies showed that the activation of PINK1/Parkin pathway can induce mitophagy in cancer cells, removing dysfunctional mitochondria and reducing ROS formation [249]. On the other hand, the deletion or mutation of regulators genes, such as the Parkin family genes, can inhibit mitophagy, promoting carcinogenesis [243]. CSCs are regulated by the mitophagy process as well [225, 250, 251]. Indeed, mitophagy represents a pro-survival pathway for CSCs formation [252, 253], maintaining the stemness and self-renewal ability of CSCs. The inhibition of mitophagy downregulates the expression of essential nuclear genes, such as NANOG, involved in CSCs survival [254, 255].

The inhibition of specific mitogen-activated protein (MAP) kinase signaling pathways can reduce stemness through BNIP3 and BNIP3L activation. The MAPK7 (called also ERK5) inhibitor XMD8-92, suppressed CSCs formation in lung adenocarcinoma and glioblastoma [256]. Recently, a subpopulation of ATP-high cancer cells was identified in breast cancer, which are phenotypically aggressive showing an increased proliferation, stemness, cell migration, invasion and drug resistance, as well as high antioxidant capacity [257,258,259]. These results demonstrate a difference in OXPHOS regulation ad ATP production between normal and cancer cells. Indeed, the γ-subunit of the mitochondrial ATP-synthase (ATP5F1C) expression is increased in lymph-node metastases compared to primary tumors and is highly over-expressed in distant metastatic lesions [260], and can be targeted by the FDA-approved Bedaquiline in triple negative breast cancer cells [260, 261].

Mitochondria as pharmacological/therapeutic target in cancer

In the last few decades, the so called “Warburg effect” [262,263,264], by which cells predominantly use glucose that can be metabolized through glycolysis with generation of lactate in the cytoplasm, or by further metabolism of pyruvate via TCA cycle and OXPHOS [263, 265], has changed based on new evidence showing that mitochondria are not completely dysfunctional in cancer, but can produce similar or increased levels of TCA and fatty acid oxidation intermediates compared to non-tumorigenic cells [266, 267]. OXPHOS might continue its function even while aerobic glycolysis is increased, becoming a protagonist pathway in cancer development. Indeed, in response to oncogenic signals or even to drug therapy, mitochondria biogenesis and activity are upregulated, promoting increased TCA cycle intermediates that sustain metabolic demands of cancer cells [208, 268]. Both glycolysis and TCA are used by cells, although typically one pathway dominates. Cells within a given tumour are predominantly glycolytic, whereas others prefer an OXPHOS metabolic phenotype depending on nutrient availability [269, 270]. Many studies have demonstrated that tumor cells are vulnerable to the inhibition of OXPHOS suggesting that mitochondria float between differing metabolic status [271, 272]. It is now well documented that “metabolic rewiring” is necessary to meet the challenges of high energy demand requested by aggressive tumors for rapid cell division, migration and metastasis formation indicating that mitochondrial energy reprogramming also regulates oncogenic pathways and post-translational modifications of oncoproteins [273]. In the “reverse Warburg effect”, cancer cells induce oxidative stress by inducing aerobic glycolysis neighbour fibroblasts, which in turn secrete pyruvate and lactate taken up by cancer cells to augment mitochondrial OXPHOS [274,275,276]. This mechanism sheds light on previously unappreciated metabolic heterogeneity and plasticity of cancer cells [277, 278]. Experimental evidence demonstrates that PGC1, a key transcription factor that is regulated by MYC which drives mitochondrial biogenesis and fission, is highly expressed and is poor prognostic marker in numerous cancer types; increasing OXPHOS and establishing drug resistance [206, 279]. PGC1 inhibitors reduce the invasive migratory phenotype of cancer cells by altering the mitochondrial metabolism and the ATP content of the cancer cells [280, 281]. The upregulation of PGC1α induce overexpression of OXPHOS gene set, by providing a structural basis for enhanced OXPHOS in multiple myeloma cells; SR18292, a new PGC1α inhibitor, exerts potent antimyeloma effects becoming a new therapeutic strategy [281]. SR18292 showed efficacy also in pediatric acute myeloid leukemia cells, where PGC1α inhibition decreased the mitochondrial DNA (mtDNA) copy number. PGC1α contributes to enhanced mtDNA copy number predicting disease aggressiveness and poor survival outcome [282]. The mtDNA transferred by non-cancer cells to cancer cells lack functional mitochondria and increases tumour efficacy in animal models [283,284,285]. Moreover, breast-cancer cells with high telomerase transcriptional activity and stemcell-like phenotypes showed increased mitochondrial mass and membrane potentials [286, 287]. Indeed, the mitochondrial encoded proteins are critically important for cancer aggressiveness; interfering with them determine anticancer effects [288, 289]. CSCs have been shown to rely on OXPHOS instead on glycolysis [251, 290], depending by mitochondrial metabolism for CSCs features maintenance. New or repurposing drugs that target mitochondria and inhibit OXPHOS have been studied as potential anticancer agents [250, 291, 292]. Atovaquone and Metformin, which interfere with OXPHOS by inhibiting Q-cytochrome c oxidoreductase (complex III) and NADH-coenzyme Q oxidoreductase (complex I), respectively, showed a clear efficacy against breast cancer progression and CSCs formation in vitro and in vivo [293, 294].

Preclinical studies in pancreatic and ovarian cancer showed that Atovaquone reduced cancer formation and dissemination in vivo [295, 296] acting as platinum sensitizing agent [297]. Metformin reduces gluconeogenesis in the liver, and is an approved drug for the treatment of type 2 diabetes mellitus [298], demonstrating anti-proliferative activity in vitro and in vivo, and showing anti-cancer effects in the context of clinical trials [299, 300]. Metformin has been shown to affect CSCs that are more dependent by OXPHOS than non-staminal tumor cells [301, 302]. Metformin as complex I inhibitor, is being assessed in over 100 ongoing clinical trials in nondiabetic patients with cancer [303, 304].

Interfering with ATP synthase (complex V), or its ecto-α-subunit [305]), interrupts the electron chain transport and ATP formation. A previously mentioned drug Bedaquiline, targets the mitochondrial ATP-synthase, leading to mitochondrial dysfunction and ATP depletion. Indeed, by affecting the F0-F1 ATP-synthase and in particular the ATP5F1C protein, Bedaquiline showed anticancer activity, reducing CSCs formation, inhibiting migration, invasion and metastasis formation in vitro and in vivo [260, 261]. Mitochondrial metabolic enzymes offer a variety of attractive targets for anticancer treatment [306, 307]. Isocitrate dehydrogenase (IDH) enzymes catalyse the conversion of isocitrate into α-ketoglutarate, which also has roles in cellular processes such as the hypoxic response and gene expression modulation. IDH1 and IDH2 mutants promote oncometabolite formation by disruption of canonical α-ketoglutarate production [308, 309]. Targeting the IDH mutants is a new promising target therapy to prevent glioma or acute myeloid leukaemia (AML) generated by IDH-mutated clones [310,311,312]. The mitochondria inner membrane transporters, also known as solute carrier family SLC25, are a class of proteins that regulate the traffic between mitochondria and cytosol [313, 314]. Recent studies have shown the involvement of mitochondria membrane transporter in the regulation of cancer plasticity and adaptability. Indeed, many SLC25 transporters are involved in phenomena of drug resistance and cancer failure therapy [315, 316]. Recently, the SLC25A8 protein called mitochondrial uncoupling protein 2 (UCP2), inducing proton leak and involved in the transport of C4 metabolites out of mitochondria [317, 318], has been demonstrated to support growth of KRAS mutated pancreatic cancer [319].

Anticancer therapies and cardiotoxicity: a role for mitochondria

Adverse cardiovascular events are frequently experienced in cancer patients as a direct consequence of the neoplasm or as a side effect induced by treatments. With the formulation of more advanced anti-cancer regimens, the global improvement of prognostic parameters, and the development of better tools for early diagnosis, an overall prolongment of survival has been observed; on the other hand, cardiovascular side effects may manifest latently, even decades after the completion of the anti-cancer therapeutic scheme. The urge to address this relevant clinical issue has led to deepen the knowledge on the molecular, biochemical, and biological mechanisms of cardiac injury by anti-cancer drugs. Among the multiple obstacles encountered, the wide portfolio of available pharmacological strategies in cancer therapies implies that different mechanisms could converge on the establishment of cardiac toxicity. While this general assumption suggests categorizing cardiac side effects on the basis of the drug class, it should be noted that a unifying perspective is retrieved when looking at mitochondrial aberrations as the main triggers for cardiac damage. Therefore, mitochondria could represent suitable targets of pharmacological interventions aimed at preserving cardiac function in cancer patients (Fig. 3).

A high degree of mitochondrial impairment triggers irreversible cardiac injury, which is usually caused by traditional chemotherapies; these agents dramatically affect the ultrastructural organization of cardiomyocytes, with large vacuolized areas, cytoskeleton disarrays and mitochondrial morphological aberrations; in contrast, newly developed and tailored-cut pharmacological agents induce lower cardiomyocytes damage consistent with benign ultrastructural changes, and overall fewer cardiac side effects.

Anticancer therapy-induced cardiotoxicity

Chemotherapy: the paradigmatic example of doxorubicin

For certain tumors, which are unresponsive to more advanced therapies, chemotherapy still represents the mainstay of treatment. Although it represents an irreplaceable option in some cases, the use of certain chemotherapy agents may determine cardiac damage, particularly for anthracyclines [320, 321]. This class of drugs, which elicit potent antibiotic activity, also exhibits well-acknowledged anti-cancer properties in breast, urogenital, brain and stomach cancers as well as in leukemia, lymphomas, and sarcomas [322,323,324,325]. Despite their effectiveness in halting cancer progression, left ventricular alterations, cardiomyopathy, and heart failure are included within the spectrum of possible cardiac manifestations in response to anthracyclines treatment [326,327,328,329,330].

Doxorubicin represents the prototypical and most frequently prescribed member of anthracyclines. Its anti-cancer mechanism relies on the ability to interfere with DNA replication and transcription at multiple levels [331,332,333]. Cardiovascular and dose-dependent side effects of doxorubicin include hypotension, tachycardia, and arrythmias [334], whereas the most serious cardiac events are represented by cardiomyopathy and congestive heart failure [12, 327, 335,336,337].

At the molecular level, it has been demonstrated that Doxorubicin enters mitochondria due to its cationic nature, and accumulates within the inner membrane, where it binds to the phospholipid cardiolipin [338]. Not only cardiolipin plays a role in mitochondrial membrane structure, but it regulates the activity and function of diverse mitochondrial proteins, including the enzyme complexes of the electron transport chain [339], determining an increased generation of oxidative stress and a diminished production of ATP [340, 341]. In turn, the impairment of mitochondrial respiration and fatty acid oxidation paves the way for a metabolic reprograming toward the glycolytic pathway [342], possibly anticipating heart failure. The transcriptional profiling of mitochondria-related genes revealed a substantial decline in the expression of genes controlling oxidative phosphorylation and fatty acid metabolism upon Doxorubicin exposure [343]. Additionally, it was shown that the perturbations of energy metabolism reflect a shift toward protein catabolism to obtain fuels [344]. Likewise, Doxorubicin suppresses AMPK activity [345], simultaneously lowering ATP consumption [346].

Together with alterations in functionality, a dramatic reduction in mitochondrial mass is observed upon treatment with Doxorubicin, which is known to activate mitophagy [347] and inhibit mitochondrial biogenesis [348]. Furthermore, ROS accumulation within mitochondria paves the way to mPTP opening and apoptosis [340, 349], further contributing to cardiomyocytes loss. Additional studies have demonstrated that administration of Doxorubicin interferes with the natural antioxidant defense response [350]. In line with this evidence, Zeng and colleagues (2020) show that, in mouse and human cardiac tissues, Doxorubicin increased the expression of NOX1 and NOX4 and induced mitochondrial fission through DRP1 activation, DRP1 translocation into mitochondria with an increase in DRP1 phosphorylation at Ser616 (p-Ser616) and a decrease in phosphorylation at Ser637 (p-Ser637). Along with the ability of Doxorubicin to impair cardiac mitochondrial function decreasing the expression of MFN1, MFN2, and OPA1, and increasing Drp1, activating autophagy and mitophagy in vivo, additional evidence indicates that this anticancer drug can also sensitize the propensity for mPTP to open, inducing apoptosis and necroptosis [351]. Doxorubicin can dose-dependently stimulate mPTP, as a consequence of mitochondrial oxidative damage and Ca2+ overload (i.e., both potent inducers of mPTP) typically involved in its acute, sub-chronic, and chronic cardiotoxic actions. Indeed, a recent study reported, in NRVMs, adult mouse ventricular myocytes (AMVMs), and in mice, that inhibiting monoamine oxidase (MAO)-dependent ROS generation abolishes Doxorubicin-induced oxidative stress that generates lipid peroxidation, alteration in mitochondrial membrane potential, mPTP, and redox potential, abnormalities of mitochondrial distribution and perimeter, sarcomere organization, and intracellular Ca2+ homeostasis, indicating that mitochondrial ROS formation represents an upstream event of Doxorubicin-induced cardiotoxicity [352]. Moreover, mitochondria isolated from Doxorubicin-treated human atrial trabeculae displayed reduced transmembrane potential and Ca2+ retention capacity, while the immune-suppressant cyclosporin A, a blocker of mPTP opening, prevented mitochondrial dysfunction and impaired contractile performance induced by Doxorubicin [349]. This finding was confirmed in vivo, where cyclosporine A mitigates the alteration of mitochondrial bioenergetics and mitochondrial fragmentation secondary to Doxorubicin exposure [353]. On the other hand, Doxorubicin has been shown to stimulate the receptor-interacting protein 3 (RIPK3)-induced activation of Ca2+-calmodulin-dependent protein kinase (CaMKII), triggering mPTP opening, thus inducing myocardial necroptosis [354], clearly supporting the role of mPTP opening in Doxorubicin cardiotoxicity in both experimental models and human hearts.

Non-anthracyclines drugs have been shown to trigger cardiovascular side effects [355] also through the alteration of mitochondrial activity and function. This is the case for cisplatin, an alkylating agent used alone or combination with other antineoplastic agents to treat diverse types of tumors [356]. By prompting organelles’ structural and ultrastructural modifications and alterations of ΔΨm, together enhanced ROS generation, cisplatin elicits cumulative cardiotoxicity, which is frequently associated with remarkable nephrological side effects [357]). Among the alkylating agents, high doses of cyclophosphamide and ifosfamide are known to induce acute myopericarditis and severe arrhythmias, which appear to be associated also with the increase of oxidative stress, alterations of IMM, impairment of mitochondrial transport and oxidation of long chain fatty acids, as well as aberrant mitochondrial respiration [358,359,360].

5-fluorouracile belongs to the class of antimetabolites. Through the inhibition of thymidylate synthase and the impairment of DNA replication and RNA synthesis, 5-fluorouracile elicits antineoplastic action in a broad spectrum of cancers [361]. However, cardiac side effects have been reported [362]. Despite the pathogenetic mechanisms have not been fully clarified, several studies have demonstrated that the drug-induced cardiotoxicity may be associated at least in part to the increased formation of ROS resulting from lipid peroxidation and depletion of glutathione pool, ultimately leading to apoptosis [363]. Furthermore, the accumulation of the highly toxic intermediate fluoroacetate from the biotransformation of 5-fluorouracil may impact on mitochondrial function by inhibiting the Krebs cycle [364].

Other antineoplastic agents with reported cardiovascular toxicity include the antibiotics mitoxantrone (inhibitor of type II topoisomerase), mitomycin C (inhibitor of DNA synthesis and cross-linker) and bleomycin (inhibitor of DNA biosynthesis through the intercalation into DNA helix). One of the most common and shared pathogenetic mechanism of these chemotherapeutics is the induction of oxidative stress, associated with mitochondrial dysfunctions [355, 365, 366]. For instance, mitoxantrone is active in certain solid tumors (metastatic breast cancer and metastatic hormone-refractory prostate cancer), as well as in myeloid leukemia, non-Hodgkin lymphoma and acute lymphoblastic leukemia [367]. The irreversible cardiomyopathy observed after long-term treatment with mitoxantrone is associated with remarkable metabolic impairments that are consistent with aberrant mitochondrial activity, such as decreased ATP levels, alterations of mitochondrial membrane potential, inhibition of ATP synthase expression, together with dysfunction of complexes IV and V [365, 368].

Arrhythmias and conduction disorders have been observed in a small percentage of patients undergoing systemic treatment with the antimicrotubule agent paclitaxel [369]; while the involvement of mitochondrial damage in these side effects has been postulated, drug-induced alterations within the Purkinje fibers and histamine release are more likely to be responsible for the chronotropic actions of paclitaxel [370, 371]

In addition, the vinca alkaloids vincristine and vinblastine, that block microtubule formation in mitotic spindle leading to the arrest of cell division of metaphase, may induce hypertension, angina, and ischemic heart disease [372]. Among the mechanisms proposed, Herradón and co-workers found that alterations of TNFα, eNOS and iNOS in cardiac tissue after chronic treatment with vincristine may be responsible for alterations of heart function; moreover, the increase in mitochondrial connexin 43 expression induced by vincristine suggests that this important cardioprotective regulator counteracts the detrimental increase in free radical to protect the heart tissue from apoptotic damage [372, 373].

Radiotherapy

Despite the enormous advances in the clinical management of cancer patients, radiation therapy is still currently indicated in more than 50% of cases, both in the neoadjuvant and adjuvant setting [374, 375]. However, radiation-induced heart disease (RIDH) includes arrhythmias, valvular dysfunction, ischemic heart disease, non-ischemic cardiomyopathy, and pericardial disease [376].

The first evidence that radiation induces relevant subcellular morphological changes in mitochondria was provided more than 50 years ago by means of electron microscopy, which showed a high degree of swelling, disorganization and clumping of the cristae, together with fusion of the inner and outer membranes [377]. mtDNA is a highly susceptible target of ionizing radiation, as the ability to promote single and double strand breaks, nucleotide damage and cross-link, and gene deletion, is not efficiently counteracted by an efficient DNA repair system [378, 379]; in addition, the lack of histones further facilitate the damaging action of radiation on mtDNA [380].

As a consequence of the alterations detected in mitochondrial structure and in mtDNA, a reduced functionality of these organelles is expected, together with an increase in ROS generation. Likewise, irradiated cardiac mitochondria exhibit lower respiratory capacity, consistent with alterations in respiratory chain proteins [381]. A comprehensive proteomic analysis of C57BL/6 mice performed after total body exposure to radiation revealed molecular responses of inflammation, antioxidant defense and cytoskeleton perturbations [382], with mitochondrial proteins being the most strikingly affected by ionizing radiation [382].

Targeted therapies

Opposite to chemo/radiotherapy, targeted therapies recognize specific molecular targets implicated in tumor progression, thus warranting higher specificity and fewer overall side effects. However, certain agents belonging to this class exhibit cardiovascular toxicity which may be occasionally critic, particularly when drug combinations are administered, rather than monotherapy regimens [383, 384].

The most frequently evidenced cardiac side effects of targeted therapies include hypertension, left ventricular dysfunction, thromboembolism, and heart failure [383]. Trastuzumab, an antibody targeting the Human epidermal growth factor receptor 2 (HER2), Trastuzumab may trigger an asymptomatic decreased in LV ejection fraction, or in some cases, congestive heart failure [385]. It should be mentioned that Trastuzumab is mostly reversible with the termination of the therapy [386].

Considering the importance of receptors tyrosine kinase (RTKs) in cancer, small-molecule drugs have been designed with the aim to target the ATP binding pockets of TKs, thus hindering the access to the kinase site within the receptor molecule and the subsequent activation of the stimulatory transduction cascade. Despite RTK inhibitors are quite well tolerated at the cardiovascular levels, imatinib and sunitinib are two representative examples of on-target [387] and off-target [388] cardiotoxicity, respectively. Imatinib allows to control chronic myeloid leukemia (CML) and certain gastrointestinal tumors [389, 390]. The drug targets specifically BCR-ABL, a fusion product which activates stimulatory cascades involved in cell proliferation and survival like RAS, PI3K/AKT and STAT5 [390]. As BCR-ABL aberrations occurs in more than 90% of chronic myeloid leukemic cells, the drug is highly specific and effective. Cardiac side events are rarely detected after use of imatinib, totalling on average 1% of the patients receiving the treatment [391]; however, a strict follow up is required for patients with a prior history of cardiovascular disease.

Immunotherapy

The wise manipulation of the immune system toward the instigation of an efficient anti-cancer response is the main objective of immunotherapies. In cancer patients the immune system is highly dysregulated, as tumor cells are enabled with peculiar mechanisms to evade immunosurveillance [392]. The impaired recognition of tumor cells sets up the stage for a tolerogenic environment that facilitates disease progression. In recent years, striking breakthrough in the field of clinical immunotherapy have revolutionized anti-cancer approaches [393], as targeting the immune system rather than the tumor cells only, now represents a feasible and successful therapeutic strategy. Immunotherapeutic agents are highly heterogeneous: the modification of patients’ own dendritic-cells (DCs) and T-cells and their subsequent reinfusion has led to the development of DC-based and T-cells based anti-cancer vaccines [394, 395]; in addition, monoclonal antibodies, immune checkpoint inhibitors (ICIs), cytokines and interleukins, together with bi-specific T-cell engagers are included among the portfolio of available anti-cancer immunotherapies. Despite the enormous clinical success obtained using these approaches in certain types of tumors, a long-term assessment of the potential side effects in response to these strategies is still missing, due to their relatively recent clinical implementation. On the other hand, cardiovascular side effects have been reported after administration of ICIs. This class of agents includes monoclonal antibodies directed at the checkpoint inhibitors CLTA4, PD1 and PDL1, a system of receptors and ligands that suppresses T cells activation and inhibits T cell function, thus protecting from an aberrant immune activation [396, 397]. ICIs are able to block the immune checkpoints, thus relieving their inhibitory effect on the immune system, ultimately leading to an efficient anti-cancer immune response. Few cases of myocarditis, HF and impairment of normal hemodynamics have been observed upon administration of ICIs [398,399,400]. Despite the incidence of cardiovascular toxicity is not currently high, the percentage of patients affected by cardiac disorders is expected to increase in the upcoming years, when more individuals will likely be treated with ICIs.

The mechanisms implicated in the cardiotoxicity of ICIs are far from being fully elucidated, however, it is clear that the PD1/PDL1 signalling is essential in cardiomyocytes immune homeostasis, as it prevents aberrant T-cells activation [401], hence blocking this pathway may elicit detrimental effects like the formation of autoantibodies. In addition, dysregulated immune cells may target normal cardiomyocytes components like cardiolipin, thus compromising cell activity and function [398]. Moreover, several changes in the immune microenvironment may be detected upon administration of ICIs, including imbalanced Treg and Teff cells, abnormal release of cytokines, induction of macrophages polarization and T-cell infiltration [399]. Not surprisingly, immune inflammation, ROS accumulation and mitochondrial impairment may represent key effectors of ICI-induced cardiotoxicity [402, 403].

Mitochondria-based approaches to relieve the cardiotoxicity of anti-cancer therapies

Considering the crucial role played by mitochondria in the cardiotoxicity of onco-therapies, the pharmacological preservation of mitochondrial function represents a feasible and suitable option for cardioprotective purposes. Therefore, it is not surprising that several anti-oxidant agents may elicit beneficial effects by relieving ROS accumulation and preventing the subsequent mitochondrial damage. In this regard, dexrazoxane is among the most investigated drug for preventing Doxorubicin-induced cardiotoxicity, approved by both FDA and European Medicines Agency (EMA) [404]. After being hydrolyzed into its active compound, dexrazoxane acts as an iron chelator, thereby preventing the formation of peroxide radicals and alleviating oxidative stress. In fact, mitochondrial structure and function appeared preserved in cardiomyocytes stimulated with Doxorubicin in the presence of dexrazoxane, despite certain alterations in mtDNA content were still evidenced [404]. Similarly, other FDA-approved drugs have been shown to decrease the oxidant load and thereby obliterate Doxorubicin-induced cardiotoxicity. This is the case for carvedilol, a β-blocker and vasodilating agent used in the treatment of systemic hypertension, left ventricular dysfunctions and congestive heart failure [405]. By reducing oxidative stress and lipid oxidation, carvedilol has been shown to contrast Doxorubicin-induced cardiac events [405, 406], whereas the drug was ineffective in hampering the cardiotoxicity of other anti-cancer molecules like trastuzumab [407].

Additional agents, of both natural and synthetic origin, were found to reduce ROS production and mitochondrial impairment by doxorubicin, thus promoting cardioprotection; among these, the plant-derived compound Ginsenoside Rh2 [408], the flavonoid Luteolin-7-o-β-d-glucopyranoside [409], the garlic derivative allicin [410], the antihyperlipidemic agent probucol [411], the adiponectin agonist ADP355 [412], the Phosphodiesterase-5 inhibitor sildenafil [413], the alternate G-protein estrogen receptor agonist G-1 [414] and the lipid-lowering statin Fluvastatin [415]. Due to its anti-oxidant properties, vitamin E was also suggested to prevent cardiac damage, while increasing doxorubicin anti-cancer efficacy [416, 417]; however, a recent thorough analysis of trials and systemic reviews highlights certain controversies on the benefit on vitamin E against Dox-induced cardiotoxicity; therefore, further investigations are warranted to turn the inconclusiveness of these findings [418].

Interestingly, Beak and collaborators found that the α1 adrenergic agonist Dabuzalgron, an orally bioavailable and well-tolerated drug previously tested for the treatment of urinary incontinence, efficiently enhances contractile function and restores energy production in Doxorubicin-treated mice [419]; these effects rely on the restoration of mitochondrial function, as evidenced by the increased expression of PGC1α implicated in mitochondrial biogenesis, and other mitochondrial genes regulators of energetic metabolism [419].

Mitochondrial transfer may be used as an efficient tool to hamper doxorubicin-dependent negative effects in the heart. More specifically, the Authors employed patient-specific induced pluripotent stem cell-derived cardiomyocytes (iCMs), which were subjected to doxorubicin injury and then co-cultured with extracellular vesicles (EVs) from mesenchymal stem cells (MSCs). Interestingly, large EVs enriched in mitochondria protected iCMs from injury, whereas small EVs, poorly enriched in mitochondria, failed in their protective effect [420]. iCMs stimulated with mitochondria-enriched EVs exhibited improved contractility, ATP production and mitochondrial biogenesis, whereas a mitochondrial inhibitor (1-methyl-4-phenylpyridinium) hampered this effect [420]. This interesting study sheds new light into the potential of stem cells and mitochondrial transfer as a useful novel approach to preserve cardiac function in doxorubicin-treated cancer patients.

Undoubtedly, mitochondria-preserving strategies could thwart also the toxicity of targeted therapies. In this regard, data from the MANTICORE-101 study (Multidisciplinary Approach to Novel Therapies in Cardiology Oncology Research) have clarified the effects of the angiotensin-converting enzyme (ACE) inhibitor perindopril versus the β-blocker bisprolol in patients with early diagnosis of HER2 positive breast cancer undergoing treatment with trastuzumab [421]. While both drugs were well-tolerated and provided protection against therapy-induced decline in LVEF, an inhibition of left ventricular remodeling was not achieved [421]. However, Shirmard et al. demonstrated that the impairment of mitochondrial aberration is sufficient to protect cardiac function in male adult Wistar rats treated with trastuzumab in combination with curcumin, chrysin and thymoquinone [422].

Likewise, manipulating mitochondrial dynamics could represent a feasible strategy to counteract trastuzumab cardiotoxicity. For instance, the transmembrane protein Connexin-43 (Cx43), if localized at the mitochondrial membranes, regulates these organelles’ structure and function in terms of volume and respiratory capability [423]. By controlling mitochondrial calcium homeostasis and ROS generation, Cx43 is engaged in cardiomyocytes mitochondria to prompt apoptotic effects upon trastuzumab administration [424]; hence, controlling Cx43 recruitment at mitochondria could circumvent trastuzumab-dependent cardiomyocytes death.

Furthermore, the antidiabetic drug metformin, which regulates energy metabolism mainly through the activation of AMPK, ameliorates trastuzumab dysfunctions possibly restoring the energetic capacity of cardiomyocytes [425]; this action could involve both the increased uptake of glucose mediated by insulin, together with the correction of energetic imbalances [425].

Extending these findings, a retrospective analysis performed in a large cohort of almost 7000 women affected by early breast cancer revealed that administration of metformin, but not thiazolidinediones, reduces the risk of cardiac events in response to adjuvant radiotherapy, thus suggesting that administration of AMPK regulators might be a useful post-surgery strategy when radiations are required [426].

It should be mentioned that the sensitivity or resistance to cardiac damage induced by radiations may depend on the myocardial expression of certain inherited variants of mitochondrial genes, thus suggesting that complex genetic modifiers may impact on patients’ cardiac tolerance to radiotherapy [427].

Furthermore, it has been shown that overexpression of certain protective factors like G protein-coupled receptor kinase 2 (GRK2) protects mitochondria from radiation-induced cardiomyocytes injury, whereas silencing GRK2 triggers mitochondrial aberrations suggestive of damage, such as reduced membrane potential together with structural and functional impairments [428]. Considering the evidence that GRK2 interacts with MFN1 and MFN2 to regulate mitochondrial dynamics of fission and fusion [428, 429], the role of GRK2 in regulating mitochondrial function in stressful conditions and the potential of this protein to serve as pharmacological target of intervention in cardioprotective strategies should be explored more in depth. Table 1 summarizes the cardiovascular or anticancer effect of drugs targeting mitochondrial functions.

Table 1 Cardiovascular or anticancer effect of drugs targeting mitochondrial functions

Conclusions and future perspectives

A large body of evidence indicates that cancer and CVD share numerous pathophysiological mechanisms, such as neuro‐hormonal activation, inflammation, oxidative stress, and immune system alteration, as well as diverse risk factors, including hypertension, obesity, smoking and air pollution, and cardiometabolic disease. The enhanced survival of cancer patients achieved thanks to the progress of screening programmes, cancer therapies, and diagnostic approaches has determined an increase of CVD, including, but not limited to, atherosclerosis and atrial fibrillation. Similarly, an augmented prevention, specific dietary regimens, together with combined pharmacological regimens, improved ventricular assist devices and early percutaneous cardiac intervention, have led to improved prognosis and decreased mortality for HF and other CVD, while determining a simultaneous increase of cancer‐related mortality. On the other hand, different approaches for treating cancer have been associated with cardiovascular complications, including coronary artery disease and HF, conduction defects, and valvular dysfunction [3, 430]. This is the case of radiation exposure (especially to the thoracic region), used in the treatment of various types of cancers, several classes of cytotoxic chemotherapy, displaying potent anticancer activity while causing cardiotoxicity, cardiomyopathy, and vascular toxicity. Also monoclonal antibodies, such as bevacizumab, or tyrosine kinase inhibition (sunitinib, pazopanib, and sorafenib) can be responsible for increased arterial thrombotic events, including myocardial infarction [12, 14, 431]

The complexity of different interacting pathways, that drive both cancer and CVDs and determine the risk of CVDs in cancer patients, as well as the risk of cancer in patients presenting with CVDs, makes it difficult the identification of specific drugs that, through counteracting cardiovascular risk factors promoting cancer development, should reduce both cardiovascular‐ and cancer-related mortality. Therefore, comprehensive and detailed mechanistic studies are needed to further elucidate the molecular determinants driving cancer and CVDs and their common pathophysiological features.

Over the last years there has been increasing focus on the role of mitochondrial dysfunction, in terms of alteration in mitochondrial dynamics and quality control (i.e., key mechanisms to maintain the optimal performance of mitochondria) and on how these processes can participate to cancer development and CVDs progression. Even if the pathophysiological complexity underlying cancer and CVDs cannot be addressed by a single pharmacological approach designed for preventing or resolving mitochondrial damage and dysfunction, improving our understanding of mitochondria-dependent mechanisms in cancer and CVDs may become a useful option to develop additional potential therapeutic targets for the treatment of CVDs and/or cancer. However, if targeting dysfunctional mitochondria for therapeutic purposes proved to be an effective strategy for treating CVDs in pre-clinical studies, the clinical translation of mitochondrial targeting agents (MTAs) has hitherto failed due to several pitfalls that mainly concern pharmacodynamic and pharmacokinetic issues [432]. For instance, blocking mPTP with cyclosporine A, which successfully interfered with rodent models of CVDs [433], is therapeutically unfeasible in cardiovascular patients due to the potent immunosuppressant effect of this drug [432]. Novel cyclosporine A analogues that lack immunosuppressant activity, such as Debio025 and NIM811, are currently under investigation. The efficacy of untargeted antioxidant strategies to attenuate ROS overproduction in mitochondria is also limited by their off-targeted effects on cytosolic ROS, which stimulate cellular survival and other primary functions at moderate levels [432, 434]. It has been suggested to exploit the highly hyperpolarized ΔΨm to specifically deliver cationic antioxidant molecules into the mitochondrial matrix [435]; unfortunately, the ΔΨm is rather depolarized in many CVDs [432]. Moreover, systemically administered MTAs could target the mitochondria that are abundantly located in other cellular compartments, such as brain, skeletal muscle, and vascular endothelium [432]. This limitation could, however, be circumvented by incorporating the MTAs within poly(lactic coglycolic acid) (PLGA)-decorated nanoparticles to specifically target the myocardium [432].

Therefore, the quest for more specific and less toxic therapies, aimed at improving therapies efficacy and reducing adverse drug events, still remains an important issue and an open challenge in the field of cardio-oncology.

Availability of data and materials

Not applicable.

Abbreviations

ATF:

Activating transcription factor

BNIP3:

Bcl-2 interacting protein 3

CHF:

Chronic heart failure

CMEC:

Cardiac microvascular endothelial cells

CML:

Chronic myeloid leukemia

CSCs:

Cancer stem cells

CVDs:

Cardiovascular diseases;

Cx43:

Connexin-43

DCM:

Dilated cardiomyopathy

DRP1:

Dynamin-related protein 1

ER:

Estrogen receptor

EVs:

Extracellular vesicles

FIS1:

Mitochondrial fission protein 1

GRK2 G:

Protein-coupled receptor kinase 2

HAoSMCs:

Aortic smooth muscle cells

HER2:

Human epidermal growth factor receptor 2

HF:

Heart failure

HIF-1α:

Hypoxia-inducible factor-1α

I/R:

Ischemia/reperfusion

ICMs:

Induced pluripotent stem cell-derived cardiomyocytes

IMM:

Inner mitochondrial membrane

IRI:

I/R injury

ISR:

Integrated stress response

MASM7:

Mitofusin activator small molecules 7

MTAs:

Mitochondrial targeting agents

Mdivi-1:

Mitochondrial division inhibitor-1

MFF:

Mitochondrial fission factor

MFI8:

Mitochondrial fusion inhibitor 8

Mfn1:

Mitofusin 1

Mfn2:

Mitofusin 2

mPTP:

Mitochondrial permeability transition pore

mtDNA:

Mitochondrial DNA

NRVMs:

Neonatal rat ventricular myocytes

OMM:

Outer mitochondrial membrane

OPA1:

Optic atrophy 1

OXPHOS:

Oxidative phosphorylation

PGC-1α:

Peroxisome proliferator-activated receptor gamma coactivator 1-alpha

PINK1:

PTEN-induced putative kinase 1

RIPK3:

Receptor-interacting serine/threonine kinase 3

ROS:

Reactive oxygen species

SR:

Sarcoplasmic reticulum

TNF-α:

Tumor necrosis factor-α

TRAF2:

Tumor necrosis factor receptor-associated factor-2

UCP2:

Uncoupling protein 2

UPRmt :

Mitochondrial unfolded protein response

UPS:

Ubiquitin proteasome system

VDAC:

Voltage-dependent anion channel

ΔΨm:

Mitochondrial membrane potential

References

  1. Masoudkabir F, Sarrafzadegan N, Gotay C, Ignaszewski A, Krahn AD, Davis MK, Franco C, Mani A. Cardiovascular disease and cancer: evidence for shared disease pathways and pharmacologic prevention. Atherosclerosis. 2017;263:343–51.

    CAS  PubMed  PubMed Central  Google Scholar 

  2. Coviello JS. Cardiovascular and cancer risk: the role of cardio-oncology. J Adv Pract Oncol. 2018;9:160–76.

    PubMed  PubMed Central  Google Scholar 

  3. de Boer RA, Meijers WC, van der Meer P, van Veldhuisen DJ. Cancer and heart disease: associations and relations. Eur J Heart Fail. 2019;21:1515–25.

    PubMed  Google Scholar 

  4. Bouhamida E, Morciano G, Perrone M, Kahsay AE, Della Sala M, Wieckowski MR, Fiorica F, Pinton P, Giorgi C, Patergnani S. The interplay of hypoxia signaling on mitochondrial dysfunction and inflammation in cardiovascular diseases and cancer: from molecular mechanisms to therapeutic approaches. Biology. 2022;11:300.

    CAS  PubMed  PubMed Central  Google Scholar 

  5. Rocca C, Pasqua T, Boukhzar L, Anouar Y, Angelone T. Progress in the emerging role of selenoproteins in cardiovascular disease: focus on endoplasmic reticulum-resident selenoproteins. Cell Mol Life Sci. 2019;76:3969–85.

    CAS  PubMed  Google Scholar 

  6. Gernaat SAM, Boer JMA, van den Bongard DHJ, Maas A, van der Pol CC, Bijlsma RM, Grobbee DE, Verkooijen HM, Peeters PH. The risk of cardiovascular disease following breast cancer by Framingham risk score. Breast Cancer Res Treat. 2018;170:119–27.

    PubMed  PubMed Central  Google Scholar 

  7. Takada S, Kinugawa S, Handa H, Yokota T, Sabe H. Cross-disease communication between cancer and heart failure provides a rational approach to prevention and treatment of both diseases. Front Oncol. 2022;12:1006322.

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Strongman H, Gadd S, Matthews A, Mansfield KE, Stanway S, Lyon AR, Dos-Santos-Silva I, Smeeth L, Bhaskaran K. Medium and long-term risks of specific cardiovascular diseases in survivors of 20 adult cancers: a population-based cohort study using multiple linked UK electronic health records databases. Lancet. 2019;394:1041–54.

    PubMed  PubMed Central  Google Scholar 

  9. Sturgeon KM, Deng L, Bluethmann SM, Zhou S, Trifiletti DM, Jiang C, Kelly SP, Zaorsky NG. A population-based study of cardiovascular disease mortality risk in US cancer patients. Eur Heart J. 2019;40:3889–97.

    PubMed  PubMed Central  Google Scholar 

  10. Chen M, Xue J, Wang M, Yang J, Chen T. Cardiovascular complications of pan-cancer therapies: the need for cardio-oncology. Cancers. 2023;15:3055.

    CAS  PubMed  PubMed Central  Google Scholar 

  11. Bikomeye JC, Terwoord JD, Santos JH, Beyer AM. Emerging mitochondrial signaling mechanisms in cardio-oncology: beyond oxidative stress. Am J Physiol Heart Circ Physiol. 2022;323:H702–20.

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Rocca C, Pasqua T, Cerra MC, Angelone T. Cardiac damage in anthracyclines therapy: focus on oxidative stress and inflammation. Antioxid Redox Signal. 2020;32:1081–97.

    CAS  PubMed  Google Scholar 

  13. Herrmann J. Adverse cardiac effects of cancer therapies: cardiotoxicity and arrhythmia. Nat Rev Cardiol. 2020;17:474–502.

    CAS  PubMed  PubMed Central  Google Scholar 

  14. Rocca C, De Francesco EM, Pasqua T, Granieri MC, De Bartolo A, Gallo Cantafio ME, Muoio MG, Gentile M, Neri A, Angelone T, et al. Mitochondrial determinants of anti-cancer drug-induced cardiotoxicity. Biomedicines. 2022;10:520.

    CAS  PubMed  PubMed Central  Google Scholar 

  15. Murphy E, Ardehali H, Balaban RS, DiLisa F, Dorn GW 2nd, Kitsis RN, Otsu K, Ping P, Rizzuto R, Sack MN, et al. Mitochondrial function, biology, and role in disease: a scientific statement from the american heart association. Circ Res. 2016;118:1960–91.

    CAS  PubMed  PubMed Central  Google Scholar 

  16. Modesti L, Danese A, Angela Maria Vitto V, Ramaccini D, Aguiari G, Gafa R, Lanza G, Giorgi C, Pinton P. Mitochondrial Ca(2+) signaling in health, disease and therapy. Cells. 2021;10:1317.

    CAS  PubMed  PubMed Central  Google Scholar 

  17. Giorgi C, Danese A, Missiroli S, Patergnani S, Pinton P. Calcium dynamics as a machine for decoding signals. Trends Cell Biol. 2018;28:258–73.

    CAS  PubMed  Google Scholar 

  18. Ma K, Chen G, Li W, Kepp O, Zhu Y, Chen Q. Mitophagy, mitochondrial homeostasis, and cell fate. Front Cell Dev Biol. 2020;8:467.

    PubMed  PubMed Central  Google Scholar 

  19. Spinelli JB, Haigis MC. The multifaceted contributions of mitochondria to cellular metabolism. Nat Cell Biol. 2018;20:745–54.

    CAS  PubMed  PubMed Central  Google Scholar 

  20. Luo Z, Yao J, Wang Z, Xu J. Mitochondria in endothelial cells angiogenesis and function: current understanding and future perspectives. J Transl Med. 2023;21:441.

    CAS  PubMed  PubMed Central  Google Scholar 

  21. Szabadkai G, Rizzuto R. Kalphalambdaovarsigma kappaalphaiota Agammaalphathetaovarsigma: how mitochondrial beauty translates into biological virtue. Curr Opin Cell Biol. 2013;25:477–82.

    CAS  PubMed  Google Scholar 

  22. Dorn GW 2nd, Vega RB, Kelly DP. Mitochondrial biogenesis and dynamics in the developing and diseased heart. Genes Dev. 2015;29:1981–91.

    CAS  PubMed  PubMed Central  Google Scholar 

  23. Morciano G, Boncompagni C, Ramaccini D, Pedriali G, Bouhamida E, Tremoli E, Giorgi C, Pinton P. Comprehensive analysis of mitochondrial dynamics alterations in heart diseases. Int J Mol Sci. 2023;24:3414.

    CAS  PubMed  PubMed Central  Google Scholar 

  24. Soda T, Brunetti V, Berra-Romani R, Moccia F. The emerging role of n-methyl-d-aspartate (NMDA) receptors in the cardiovascular system: physiological implications, pathological consequences, and therapeutic perspectives. Int J Mol Sci. 2023;24:3914.

    CAS  PubMed  PubMed Central  Google Scholar 

  25. Lionetti V, Bollini S, Coppini R, Gerbino A, Ghigo A, Iaccarino G, Madonna R, Mangiacapra F, Miragoli M, Moccia F, et al. Understanding the heart-brain axis response in COVID-19 patients: a suggestive perspective for therapeutic development. Pharmacol Res. 2021;168:105581.

    CAS  PubMed  PubMed Central  Google Scholar 

  26. Ul Fatima N, Ananthanarayanan V. Mitochondrial movers and shapers: recent insights into regulators of fission, fusion and transport. Curr Opin Cell Biol. 2023;80:102150.

    CAS  PubMed  Google Scholar 

  27. San-Millan I. The key role of mitochondrial function in health and disease. Antioxidants. 2023;12:782.

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Tong S, Xia M, Xu Y, Sun Q, Ye L, Yuan F, Wang Y, Cai J, Ye Z, Tian D. Identification and validation of a novel prognostic signature based on mitochondria and oxidative stress related genes for glioblastoma. J Transl Med. 2023;21:136.

    CAS  PubMed  PubMed Central  Google Scholar 

  29. Negri S, Faris P, Moccia F. Reactive oxygen species and endothelial Ca(2+) signaling: brothers in arms or partners in crime? Int J Mol Sci. 2021;22:9821.

    CAS  PubMed  PubMed Central  Google Scholar 

  30. Du F, Yang LH, Liu J, Wang J, Fan L, Duangmano S, Liu H, Liu M, Wang J, Zhong X, et al. The role of mitochondria in the resistance of melanoma to PD-1 inhibitors. J Transl Med. 2023;21:345.

    PubMed  PubMed Central  Google Scholar 

  31. Wang S, Tan J, Miao Y, Zhang Q. Mitochondrial dynamics, mitophagy, and mitochondria-endoplasmic reticulum contact sites crosstalk under hypoxia. Front Cell Dev Biol. 2022;10:848214.

    PubMed  PubMed Central  Google Scholar 

  32. Reddy PH, Reddy TP, Manczak M, Calkins MJ, Shirendeb U, Mao P. Dynamin-related protein 1 and mitochondrial fragmentation in neurodegenerative diseases. Brain Res Rev. 2011;67:103–18.

    CAS  PubMed  Google Scholar 

  33. Liu YJ, McIntyre RL, Janssens GE, Houtkooper RH. Mitochondrial fission and fusion: a dynamic role in aging and potential target for age-related disease. Mech Ageing Dev. 2020;186:111212.

    CAS  PubMed  Google Scholar 

  34. Adebayo M, Singh S, Singh AP, Dasgupta S. Mitochondrial fusion and fission: the fine-tune balance for cellular homeostasis. FASEB J. 2021;35:e21620.

    CAS  PubMed  Google Scholar 

  35. Arnoult D, Grodet A, Lee YJ, Estaquier J, Blackstone C. Release of OPA1 during apoptosis participates in the rapid and complete release of cytochrome c and subsequent mitochondrial fragmentation. J Biol Chem. 2005;280:35742–50.

    CAS  PubMed  Google Scholar 

  36. Casellas-Diaz S, Larramona-Arcas R, Rique-Pujol G, Tena-Morraja P, Muller-Sanchez C, Segarra-Mondejar M, Gavalda-Navarro A, Villarroya F, Reina M, Martinez-Estrada OM, Soriano FX. Mfn2 localization in the ER is necessary for its bioenergetic function and neuritic development. EMBO Rep. 2021;22:e51954.

    CAS  PubMed  PubMed Central  Google Scholar 

  37. Chen Y, Dorn GW 2nd. PINK1-phosphorylated mitofusin 2 is a Parkin receptor for culling damaged mitochondria. Science. 2013;340:471–5.

    CAS  PubMed  PubMed Central  Google Scholar 

  38. Lee H, Yoon Y. Mitochondrial membrane dynamics-functional positioning of OPA1. Antioxidants. 2018;7:186.

    PubMed  PubMed Central  Google Scholar 

  39. Wang R, Mishra P, Garbis SD, Moradian A, Sweredoski MJ, Chan DC. Identification of new OPA1 cleavage site reveals that short isoforms regulate mitochondrial fusion. Mol Biol Cell. 2021;32:157–68.

    CAS  PubMed  PubMed Central  Google Scholar 

  40. Song Z, Chen H, Fiket M, Alexander C, Chan DC. OPA1 processing controls mitochondrial fusion and is regulated by mRNA splicing, membrane potential, and Yme1L. J Cell Biol. 2007;178:749–55.

    CAS  PubMed  PubMed Central  Google Scholar 

  41. Jiang X, Jiang H, Shen Z, Wang X. Activation of mitochondrial protease OMA1 by Bax and Bak promotes cytochrome c release during apoptosis. Proc Natl Acad Sci U S A. 2014;111:14782–7.

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Lee H, Smith SB, Yoon Y. The short variant of the mitochondrial dynamin OPA1 maintains mitochondrial energetics and cristae structure. J Biol Chem. 2017;292:7115–30.

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Samant SA, Zhang HJ, Hong Z, Pillai VB, Sundaresan NR, Wolfgeher D, Archer SL, Chan DC, Gupta MP. SIRT3 deacetylates and activates OPA1 to regulate mitochondrial dynamics during stress. Mol Cell Biol. 2014;34:807–19.

    PubMed  PubMed Central  Google Scholar 

  44. Gollmer J, Zirlik A, Bugger H. Mitochondrial mechanisms in diabetic cardiomyopathy. Diabetes Metab J. 2020;44:33–53.

    PubMed  PubMed Central  Google Scholar 

  45. Pyakurel A, Savoia C, Hess D, Scorrano L. Extracellular regulated kinase phosphorylates mitofusin 1 to control mitochondrial morphology and apoptosis. Mol Cell. 2015;58:244–54.

    CAS  PubMed  PubMed Central  Google Scholar 

  46. Joo HK, Lee YR, Choi S, Park MS, Kang G, Kim CS, Jeon BH. Protein kinase C beta II upregulates intercellular adhesion molecule-1 via mitochondrial activation in cultured endothelial cells. Korean J Physiol Pharmacol. 2017;21:377–84.

    CAS  PubMed  PubMed Central  Google Scholar 

  47. Yoo SM, Jung YK. A molecular approach to mitophagy and mitochondrial dynamics. Mol Cells. 2018;41:18–26.

    CAS  PubMed  PubMed Central  Google Scholar 

  48. Doblado L, Lueck C, Rey C, Samhan-Arias AK, Prieto I, Stacchiotti A, Monsalve M. Mitophagy in human diseases. Int J Mol Sci. 2021;22:3903.

    CAS  PubMed Central  Google Scholar 

  49. Dai SH, Wu QC, Zhu RR, Wan XM, Zhou XL. Notch1 protects against myocardial ischaemia-reperfusion injury via regulating mitochondrial fusion and function. J Cell Mol Med. 2020;24:3183–91.

    CAS  PubMed  PubMed Central  Google Scholar 

  50. Chen L, Liu T, Tran A, Lu X, Tomilov AA, Davies V, Cortopassi G, Chiamvimonvat N, Bers DM, Votruba M, Knowlton AA. OPA1 mutation and late-onset cardiomyopathy: mitochondrial dysfunction and mtDNA instability. J Am Heart Assoc. 2012;1:e003012.

    PubMed  PubMed Central  Google Scholar 

  51. Quintana-Cabrera R, Manjarres-Raza I, Vicente-Gutierrez C, Corrado M, Bolanos JP, Scorrano L. Opa1 relies on cristae preservation and ATP synthase to curtail reactive oxygen species accumulation in mitochondria. Redox Biol. 2021;41:101944.

    CAS  PubMed  PubMed Central  Google Scholar 

  52. Serasinghe MN, Chipuk JE. Mitochondrial fission in human diseases. Handb Exp Pharmacol. 2017;240:159–88.

    CAS  PubMed  PubMed Central  Google Scholar 

  53. Zerihun M, Sukumaran S, Qvit N. The Drp1-mediated mitochondrial fission protein interactome as an emerging core player in mitochondrial dynamics and cardiovascular disease therapy. Int J Mol Sci. 2023;24:5785.

    CAS  PubMed  PubMed Central  Google Scholar 

  54. Loson OC, Song Z, Chen H, Chan DC. Fis1, Mff, MiD49, and MiD51 mediate Drp1 recruitment in mitochondrial fission. Mol Biol Cell. 2013;24:659–67.

    CAS  PubMed  PubMed Central  Google Scholar 

  55. Zhu PP, Patterson A, Stadler J, Seeburg DP, Sheng M, Blackstone C. Intra- and intermolecular domain interactions of the C-terminal GTPase effector domain of the multimeric dynamin-like GTPase Drp1. J Biol Chem. 2004;279:35967–74.

    CAS  PubMed  Google Scholar 

  56. Jin JY, Wei XX, Zhi XL, Wang XH, Meng D. Drp1-dependent mitochondrial fission in cardiovascular disease. Acta Pharmacol Sin. 2021;42:655–64.

    CAS  PubMed  Google Scholar 

  57. Bragoszewski P, Turek M, Chacinska A. Control of mitochondrial biogenesis and function by the ubiquitin-proteasome system. Open Biol. 2017;7:170007.

    PubMed  PubMed Central  Google Scholar 

  58. Sulkshane P, Ram J, Glickman MH. Ubiquitination of intramitochondrial proteins: implications for metabolic adaptability. Biomolecules. 2020;10:1559.

    CAS  PubMed  PubMed Central  Google Scholar 

  59. Penna C, Sorge M, Femmino S, Pagliaro P, Brancaccio M. Redox aspects of chaperones in cardiac function. Front Physiol. 2018;9:216.

    PubMed  PubMed Central  Google Scholar 

  60. Voos W, Pollecker K. The Mitochondrial Lon protease: novel functions off the beaten track? Biomolecules. 2020;10:253.

    CAS  PubMed  PubMed Central  Google Scholar 

  61. Wachoski-Dark E, Zhao T, Khan A, Shutt TE, Greenway SC. Mitochondrial protein homeostasis and cardiomyopathy. Int J Mol Sci. 2022;23:3353.

    CAS  PubMed  PubMed Central  Google Scholar 

  62. Arnould T, Michel S, Renard P. Mitochondria retrograde signaling and the UPR mt: where are we in mammals? Int J Mol Sci. 2015;16:18224–51.

    CAS  PubMed  PubMed Central  Google Scholar 

  63. Melber A, Haynes CM. UPR(mt) regulation and output: a stress response mediated by mitochondrial-nuclear communication. Cell Res. 2018;28:281–95.

    CAS  PubMed  PubMed Central  Google Scholar 

  64. Munoz-Carvajal F, Sanhueza M. The mitochondrial unfolded protein response: a hinge between healthy and pathological aging. Front Aging Neurosci. 2020;12:581849.

    CAS  PubMed  PubMed Central  Google Scholar 

  65. Wang Y, Li J, Zhang Z, Wang R, Bo H, Zhang Y. Exercise improves the coordination of the mitochondrial unfolded protein response and mitophagy in aging skeletal muscle. Life. 2023;13:1006.

    CAS  PubMed  PubMed Central  Google Scholar 

  66. Ng MYW, Wai T, Simonsen A. Quality control of the mitochondrion. Dev Cell. 2021;56:881–905.

    CAS  PubMed  Google Scholar 

  67. Keerthiga R, Pei DS, Fu A. Mitochondrial dysfunction, UPR(mt) signaling, and targeted therapy in metastasis tumor. Cell Biosci. 2021;11:186.

    CAS  PubMed  PubMed Central  Google Scholar 

  68. Suarez-Rivero JM, Pastor-Maldonado CJ, Povea-Cabello S, Alvarez-Cordoba M, Villalon-Garcia I, Talaveron-Rey M, Suarez-Carrillo A, Munuera-Cabeza M, Reche-Lopez D, Cilleros-Holgado P, et al. Activation of the mitochondrial unfolded protein response: a new therapeutic target? Biomedicines. 2022;10:1611.

    CAS  PubMed  PubMed Central  Google Scholar 

  69. Naresh NU, Haynes CM. Signaling and regulation of the mitochondrial unfolded protein response. Cold Spring Harb Perspect Biol. 2019;11:a033944.

    CAS  PubMed  PubMed Central  Google Scholar 

  70. Suarez-Rivero JM, Pastor-Maldonado CJ, Povea-Cabello S, Alvarez-Cordoba M, Villalon-Garcia I, Talaveron-Rey M, Suarez-Carrillo A, Munuera-Cabeza M, Reche-Lopez D, Cilleros-Holgado P, et al. UPR(mt) activation improves pathological alterations in cellular models of mitochondrial diseases. Orphanet J Rare Dis. 2022;17:204.

    PubMed  PubMed Central  Google Scholar 

  71. Duncker DJ, Koller A, Merkus D, Canty JM Jr. Regulation of coronary blood flow in health and ischemic heart disease. Prog Cardiovasc Dis. 2015;57:409–22.

    PubMed  Google Scholar 

  72. Guo X, Aviles G, Liu Y, Tian R, Unger BA, Lin YT, Wiita AP, Xu K, Correia MA, Kampmann M. Mitochondrial stress is relayed to the cytosol by an OMA1-DELE1-HRI pathway. Nature. 2020;579:427–32.

    CAS  PubMed  PubMed Central  Google Scholar 

  73. Zorov DB, Juhaszova M, Sollott SJ. Mitochondrial reactive oxygen species (ROS) and ROS-induced ROS release. Physiol Rev. 2014;94:909–50.

    CAS  PubMed  PubMed Central  Google Scholar 

  74. Peoples JN, Saraf A, Ghazal N, Pham TT, Kwong JQ. Mitochondrial dysfunction and oxidative stress in heart disease. Exp Mol Med. 2019;51:1–13.

    CAS  PubMed  Google Scholar 

  75. Pakos-Zebrucka K, Koryga I, Mnich K, Ljujic M, Samali A, Gorman AM. The integrated stress response. EMBO Rep. 2016;17:1374–95.

    CAS  PubMed  PubMed Central  Google Scholar 

  76. Teske BF, Fusakio ME, Zhou D, Shan J, McClintick JN, Kilberg MS, Wek RC. CHOP induces activating transcription factor 5 (ATF5) to trigger apoptosis in response to perturbations in protein homeostasis. Mol Biol Cell. 2013;24:2477–90.

    CAS  PubMed  PubMed Central  Google Scholar 

  77. Cuervo AM, Bergamini E, Brunk UT, Droge W, Ffrench M, Terman A. Autophagy and aging: the importance of maintaining “clean” cells. Autophagy. 2005;1:131–40.

    PubMed  Google Scholar 

  78. Parzych KR, Klionsky DJ. An overview of autophagy: morphology, mechanism, and regulation. Antioxid Redox Signal. 2014;20:460–73.

    CAS  PubMed  PubMed Central  Google Scholar 

  79. Tekirdag K, Cuervo AM. Chaperone-mediated autophagy and endosomal microautophagy: joint by a chaperone. J Biol Chem. 2018;293:5414–24.

    CAS  PubMed  Google Scholar 

  80. Pickles S, Vigie P, Youle RJ. Mitophagy and quality control mechanisms in mitochondrial maintenance. Curr Biol. 2018;28:R170–85.

    CAS  PubMed  PubMed Central  Google Scholar 

  81. Fu T, Ma Y, Li Y, Wang Y, Wang Q, Tong Y. Mitophagy as a mitochondrial quality control mechanism in myocardial ischemic stress: from bench to bedside. Cell Stress Chaperones. 2023;28:239–51.

    PubMed  PubMed Central  Google Scholar 

  82. Wang Y, Liu HH, Cao YT, Zhang LL, Huang F, Yi C. The role of mitochondrial dynamics and mitophagy in carcinogenesis, metastasis and therapy. Front Cell Dev Biol. 2020;8:413.

    PubMed  PubMed Central  Google Scholar 

  83. Sekine S, Youle RJ. PINK1 import regulation; a fine system to convey mitochondrial stress to the cytosol. BMC Biol. 2018;16:2.

    PubMed  PubMed Central  Google Scholar 

  84. Durcan TM, Fon EA. The three ’P’s of mitophagy: PARKIN, PINK1, and post-translational modifications. Genes Dev. 2015;29:989–99.

    CAS  PubMed  PubMed Central  Google Scholar 

  85. Kazlauskaite A, Kondapalli C, Gourlay R, Campbell DG, Ritorto MS, Hofmann K, Alessi DR, Knebel A, Trost M, Muqit MM. Parkin is activated by PINK1-dependent phosphorylation of ubiquitin at Ser65. Biochem J. 2014;460:127–39.

    CAS  PubMed  Google Scholar 

  86. Geisler S, Holmstrom KM, Skujat D, Fiesel FC, Rothfuss OC, Kahle PJ, Springer W. PINK1/Parkin-mediated mitophagy is dependent on VDAC1 and p62/SQSTM1. Nat Cell Biol. 2010;12:119–31.

    CAS  PubMed  Google Scholar 

  87. Pankiv S, Clausen TH, Lamark T, Brech A, Bruun JA, Outzen H, Overvatn A, Bjorkoy G, Johansen T. p62/SQSTM1 binds directly to Atg8/LC3 to facilitate degradation of ubiquitinated protein aggregates by autophagy. J Biol Chem. 2007;282:24131–45.

    CAS  PubMed  Google Scholar 

  88. Hale AN, Ledbetter DJ, Gawriluk TR, Rucker EB 3rd. Autophagy: regulation and role in development. Autophagy. 2013;9:951–72.

    CAS  PubMed  PubMed Central  Google Scholar 

  89. Fritsch LE, Moore ME, Sarraf SA, Pickrell AM. Ubiquitin and receptor-dependent mitophagy pathways and their implication in neurodegeneration. J Mol Biol. 2020;432:2510–24.

    CAS  PubMed  Google Scholar 

  90. Teresak P, Lapao A, Subic N, Boya P, Elazar Z, Simonsen A. Regulation of PRKN-independent mitophagy. Autophagy. 2022;18:24–39.

    CAS  PubMed  Google Scholar 

  91. Liu L, Sakakibara K, Chen Q, Okamoto K. Receptor-mediated mitophagy in yeast and mammalian systems. Cell Res. 2014;24:787–95.

    CAS  PubMed  PubMed Central  Google Scholar 

  92. Zhang J, Ney PA. Role of BNIP3 and NIX in cell death, autophagy, and mitophagy. Cell Death Differ. 2009;16:939–46.

    CAS  PubMed  Google Scholar 

  93. Chistiakov DA, Shkurat TP, Melnichenko AA, Grechko AV, Orekhov AN. The role of mitochondrial dysfunction in cardiovascular disease: a brief review. Ann Med. 2018;50:121–7.

    CAS  PubMed  Google Scholar 

  94. Pasqua T, Rocca C, Giglio A, Angelone T. Cardiometabolism as an interlocking puzzle between the healthy and diseased heart: new frontiers in therapeutic applications. J Clin Med. 2021;10:721.

    CAS  PubMed  PubMed Central  Google Scholar 

  95. Baines CP. The molecular composition of the mitochondrial permeability transition pore. J Mol Cell Cardiol. 2009;46:850–7.

    CAS  PubMed  PubMed Central  Google Scholar 

  96. Stamerra CA, Di Giosia P, Giorgini P, Ferri C, Sukhorukov VN, Sahebkar A. Mitochondrial dysfunction and cardiovascular disease: pathophysiology and emerging therapies. Oxid Med Cell Longev. 2022;2022:9530007.

    PubMed  PubMed Central  Google Scholar 

  97. Virani SS, Alonso A, Aparicio HJ, Benjamin EJ, Bittencourt MS, Callaway CW, Carson AP, Chamberlain AM, Cheng S, Delling FN, et al. Heart disease and stroke statistics-2021 update: a report from the American Heart Association. Circulation. 2021;143:e254–743.

    PubMed  Google Scholar 

  98. Lin J, Duan J, Wang Q, Xu S, Zhou S, Yao K. Mitochondrial dynamics and mitophagy in cardiometabolic disease. Front Cardiovasc Med. 2022;9:917135.

    CAS  PubMed  PubMed Central  Google Scholar 

  99. Vasquez-Trincado C, Garcia-Carvajal I, Pennanen C, Parra V, Hill JA, Rothermel BA, Lavandero S. Mitochondrial dynamics, mitophagy and cardiovascular disease. J Physiol. 2016;594:509–25.

    CAS  PubMed  PubMed Central  Google Scholar 

  100. Forte M, Schirone L, Ameri P, Basso C, Catalucci D, Modica J, Chimenti C, Crotti L, Frati G, Rubattu S, et al. The role of mitochondrial dynamics in cardiovascular diseases. Br J Pharmacol. 2021;178:2060–76.

    CAS  PubMed  Google Scholar 

  101. Bravo-San Pedro JM, Kroemer G, Galluzzi L. Autophagy and mitophagy in cardiovascular disease. Circ Res. 2017;120:1812–24.

    CAS  PubMed  Google Scholar 

  102. Liu M, Wu Y. Role of mitophagy in coronary heart disease: targeting the mitochondrial dysfunction and inflammatory regulation. Front Cardiovasc Med. 2022;9:819454.

    CAS  PubMed  PubMed Central  Google Scholar 

  103. Yang Y, Li T, Li Z, Liu N, Yan Y, Liu B. Role of mitophagy in cardiovascular disease. Aging Dis. 2020;11:419–37.

    PubMed  PubMed Central  Google Scholar 

  104. Yang M, Linn BS, Zhang Y, Ren J. Mitophagy and mitochondrial integrity in cardiac ischemia-reperfusion injury. Biochim Biophys Acta Mol Basis Dis. 2019;1865:2293–302.

    CAS  PubMed  Google Scholar 

  105. Tedgui A, Mallat Z. Cytokines in atherosclerosis: pathogenic and regulatory pathways. Physiol Rev. 2006;86:515–81.

    CAS  PubMed  Google Scholar 

  106. Pasqua T, Pagliaro P, Rocca C, Angelone T, Penna C. Role of NLRP-3 inflammasome in hypertension: a potential therapeutic target. Curr Pharm Biotechnol. 2018;19:708–14.

    CAS  PubMed  Google Scholar 

  107. Chen M, Chen Z, Wang Y, Tan Z, Zhu C, Li Y, Han Z, Chen L, Gao R, Liu L, Chen Q. Mitophagy receptor FUNDC1 regulates mitochondrial dynamics and mitophagy. Autophagy. 2016;12:689–702.

    CAS  PubMed  PubMed Central  Google Scholar 

  108. Ma X, Rawnsley DR, Kovacs A, Islam M, Murphy JT, Zhao C, Kumari M, Foroughi L, Liu H, Qi K, et al. TRAF2, an innate immune sensor, reciprocally regulates mitophagy and inflammation to maintain cardiac myocyte homeostasis. JACC Basic Transl Sci. 2022;7:223–43.

    PubMed  Google Scholar 

  109. Luan Y, Luan Y, Feng Q, Chen X, Ren KD, Yang Y. Emerging role of mitophagy in the heart: therapeutic potentials to modulate mitophagy in cardiac diseases. Oxid Med Cell Longev. 2021;2021:3259963.

    PubMed  PubMed Central  Google Scholar 

  110. Liu Y, Wang Y, Bi Y, Zhao Z, Wang S, Lin S, Yang Z, Wang X, Mao J. Emerging role of mitophagy in heart failure: from molecular mechanism to targeted therapy. Cell Cycle. 2023;22:906–18.

    CAS  PubMed  Google Scholar 

  111. Zhu H, Tannous P, Johnstone JL, Kong Y, Shelton JM, Richardson JA, Le V, Levine B, Rothermel BA, Hill JA. Cardiac autophagy is a maladaptive response to hemodynamic stress. J Clin Invest. 2007;117:1782–93.

    CAS  PubMed  PubMed Central  Google Scholar 

  112. Morciano G, Patergnani S, Bonora M, Pedriali G, Tarocco A, Bouhamida E, Marchi S, Ancora G, Anania G, Wieckowski MR, et al. Mitophagy in cardiovascular diseases. J Clin Med. 2020;9:892.

    CAS  PubMed  PubMed Central  Google Scholar 

  113. Ikeda Y, Shirakabe A, Maejima Y, Zhai P, Sciarretta S, Toli J, Nomura M, Mihara K, Egashira K, Ohishi M, et al. Endogenous Drp1 mediates mitochondrial autophagy and protects the heart against energy stress. Circ Res. 2015;116:264–78.

    CAS  PubMed  Google Scholar 

  114. Shirakabe A, Zhai P, Ikeda Y, Saito T, Maejima Y, Hsu CP, Nomura M, Egashira K, Levine B, Sadoshima J. Drp1-dependent mitochondrial autophagy plays a protective role against pressure overload-induced mitochondrial dysfunction and heart failure. Circulation. 2016;133:1249–63.

    CAS  PubMed  PubMed Central  Google Scholar 

  115. Lee Y, Lee HY, Hanna RA, Gustafsson AB. Mitochondrial autophagy by Bnip3 involves Drp1-mediated mitochondrial fission and recruitment of Parkin in cardiac myocytes. Am J Physiol Heart Circ Physiol. 2011;301:H1924-1931.

    CAS  PubMed  PubMed Central  Google Scholar 

  116. Song M, Gong G, Burelle Y, Gustafsson AB, Kitsis RN, Matkovich SJ, Dorn GW 2nd. Interdependence of Parkin-mediated mitophagy and mitochondrial fission in adult mouse hearts. Circ Res. 2015;117:346–51.

    CAS  PubMed  PubMed Central  Google Scholar 

  117. Wakabayashi J, Zhang Z, Wakabayashi N, Tamura Y, Fukaya M, Kensler TW, Iijima M, Sesaki H. The dynamin-related GTPase Drp1 is required for embryonic and brain development in mice. J Cell Biol. 2009;186:805–16.

    CAS  PubMed  PubMed Central  Google Scholar 

  118. Kageyama Y, Hoshijima M, Seo K, Bedja D, Sysa-Shah P, Andrabi SA, Chen W, Hoke A, Dawson VL, Dawson TM, et al. Parkin-independent mitophagy requires Drp1 and maintains the integrity of mammalian heart and brain. EMBO J. 2014;33:2798–813.

    CAS  PubMed  PubMed Central  Google Scholar 

  119. Ong SB, Subrayan S, Lim SY, Yellon DM, Davidson SM, Hausenloy DJ. Inhibiting mitochondrial fission protects the heart against ischemia/reperfusion injury. Circulation. 2010;121:2012–22.

    CAS  PubMed  Google Scholar 

  120. Sharp WW, Fang YH, Han M, Zhang HJ, Hong Z, Banathy A, Morrow E, Ryan JJ, Archer SL. Dynamin-related protein 1 (Drp1)-mediated diastolic dysfunction in myocardial ischemia-reperfusion injury: therapeutic benefits of Drp1 inhibition to reduce mitochondrial fission. FASEB J. 2014;28:316–26.

    CAS  PubMed  PubMed Central  Google Scholar 

  121. Din S, Mason M, Volkers M, Johnson B, Cottage CT, Wang Z, Joyo AY, Quijada P, Erhardt P, Magnuson NS, et al. Pim-1 preserves mitochondrial morphology by inhibiting dynamin-related protein 1 translocation. Proc Natl Acad Sci USA. 2013;110:5969–74.

    CAS  PubMed  PubMed Central  Google Scholar 

  122. Maneechote C, Palee S, Kerdphoo S, Jaiwongkam T, Chattipakorn SC, Chattipakorn N. Balancing mitochondrial dynamics via increasing mitochondrial fusion attenuates infarct size and left ventricular dysfunction in rats with cardiac ischemia/reperfusion injury. Clin Sci. 2019;133:497–513.

    CAS  Google Scholar 

  123. Su ZD, Li CQ, Wang HW, Zheng MM, Chen QW. Inhibition of DRP1-dependent mitochondrial fission by Mdivi-1 alleviates atherosclerosis through the modulation of M1 polarization. J Transl Med. 2023;21:427.

    CAS  PubMed  PubMed Central  Google Scholar 

  124. Hu Q, Zhang H, Gutierrez Cortes N, Wu D, Wang P, Zhang J, Mattison JA, Smith E, Bettcher LF, Wang M, et al. Increased Drp1 acetylation by lipid overload induces cardiomyocyte death and heart dysfunction. Circ Res. 2020;126:456–70.

    CAS  PubMed  PubMed Central  Google Scholar 

  125. Wang K, Yuan Y, Liu X, Lau WB, Zuo L, Wang X, Ma L, Jiao K, Shang J, Wang W, et al. Cardiac specific overexpression of mitochondrial Omi/HtrA2 induces myocardial apoptosis and cardiac dysfunction. Sci Rep. 2016;6:37927.

    CAS  PubMed  PubMed Central  Google Scholar 

  126. Tao A, Xu X, Kvietys P, Kao R, Martin C, Rui T. Experimental diabetes mellitus exacerbates ischemia/reperfusion-induced myocardial injury by promoting mitochondrial fission: role of down-regulation of myocardial Sirt1 and subsequent Akt/Drp1 interaction. Int J Biochem Cell Biol. 2018;105:94–103.

    CAS  PubMed  Google Scholar 

  127. Wang X, Winter D, Ashrafi G, Schlehe J, Wong YL, Selkoe D, Rice S, Steen J, LaVoie MJ, Schwarz TL. PINK1 and Parkin target Miro for phosphorylation and degradation to arrest mitochondrial motility. Cell. 2011;147:893–906.

    CAS  PubMed  PubMed Central  Google Scholar 

  128. Zaja I, Bai X, Liu Y, Kikuchi C, Dosenovic S, Yan Y, Canfield SG, Bosnjak ZJ. Cdk1, PKCdelta and calcineurin-mediated Drp1 pathway contributes to mitochondrial fission-induced cardiomyocyte death. Biochem Biophys Res Commun. 2014;453:710–21.

    CAS  PubMed  PubMed Central  Google Scholar 

  129. Li A, Gao M, Liu B, Qin Y, Chen L, Liu H, Wu H, Gong G. Mitochondrial autophagy: molecular mechanisms and implications for cardiovascular disease. Cell Death Dis. 2022;13:444.

    PubMed  PubMed Central  Google Scholar 

  130. Kamga Pride C, Mo L, Quesnelle K, Dagda RK, Murillo D, Geary L, Corey C, Portella R, Zharikov S, St Croix C, et al. Nitrite activates protein kinase A in normoxia to mediate mitochondrial fusion and tolerance to ischaemia/reperfusion. Cardiovasc Res. 2014;101:57–68.

    PubMed  Google Scholar 

  131. He J, Cheng J, Wang T. SUMOylation-mediated response to mitochondrial stress. Int J Mol Sci. 2020;21:5657.

    CAS  PubMed  PubMed Central  Google Scholar 

  132. Sheng Z, Zhu J, Deng YN, Gao S, Liang S. SUMOylation modification-mediated cell death. Open Biol. 2021;11:210050.

    CAS  PubMed  PubMed Central  Google Scholar 

  133. Kim EY, Zhang Y, Beketaev I, Segura AM, Yu W, Xi Y, Chang J, Wang J. SENP5, a SUMO isopeptidase, induces apoptosis and cardiomyopathy. J Mol Cell Cardiol. 2015;78:154–64.

    CAS  PubMed  Google Scholar 

  134. Guo C, Wilkinson KA, Evans AJ, Rubin PP, Henley JM. SENP3-mediated deSUMOylation of Drp1 facilitates interaction with Mff to promote cell death. Sci Rep. 2017;7:43811.

    PubMed  PubMed Central  Google Scholar 

  135. Lu YT, Li LZ, Yang YL, Yin X, Liu Q, Zhang L, Liu K, Liu B, Li J, Qi LW. Succinate induces aberrant mitochondrial fission in cardiomyocytes through GPR91 signaling. Cell Death Dis. 2018;9:672.

    PubMed  PubMed Central  Google Scholar 

  136. Zhou H, Wang J, Zhu P, Zhu H, Toan S, Hu S, Ren J, Chen Y. Retraction Note to: NR4A1 aggravates the cardiac microvascular ischemia reperfusion injury through suppressing FUNDC1-mediated mitophagy and promoting Mff-required mitochondrial fission by CK2alpha. Basic Res Cardiol. 2023;118:24.

    PubMed  Google Scholar 

  137. Jin Q, Li R, Hu N, Xin T, Zhu P, Hu S, Ma S, Zhu H, Ren J, Zhou H. DUSP1 alleviates cardiac ischemia/reperfusion injury by suppressing the Mff-required mitochondrial fission and Bnip3-related mitophagy via the JNK pathways. Redox Biol. 2018;14:576–87.

    CAS  PubMed  Google Scholar 

  138. Markaki M, Tsagkari D, Tavernarakis N. Mitophagy mechanisms in neuronal physiology and pathology during ageing. Biophys Rev. 2021;13:955–65.

    CAS  PubMed  PubMed Central  Google Scholar 

  139. Zhou H, Hu S, Jin Q, Shi C, Zhang Y, Zhu P, Ma Q, Tian F, Chen Y. Mff-dependent mitochondrial fission contributes to the pathogenesis of cardiac microvasculature ischemia/reperfusion injury via induction of mROS-mediated cardiolipin oxidation and HK2/VDAC1 disassociation-involved mPTP opening. J Am Heart Assoc. 2017;6:e005328.

    PubMed  PubMed Central  Google Scholar 

  140. Chen H, Ren S, Clish C, Jain M, Mootha V, McCaffery JM, Chan DC. Titration of mitochondrial fusion rescues Mff-deficient cardiomyopathy. J Cell Biol. 2015;211:795–805.

    CAS  PubMed  PubMed Central  Google Scholar 

  141. Haileselassie B, Mukherjee R, Joshi AU, Napier BA, Massis LM, Ostberg NP, Queliconi BB, Monack D, Bernstein D, Mochly-Rosen D. Drp1/Fis1 interaction mediates mitochondrial dysfunction in septic cardiomyopathy. J Mol Cell Cardiol. 2019;130:160–9.

    CAS  PubMed  PubMed Central  Google Scholar 

  142. Sun S, Yu W, Xu H, Li C, Zou R, Wu NN, Wang L, Ge J, Ren J, Zhang Y. TBC1D15-Drp1 interaction-mediated mitochondrial homeostasis confers cardioprotection against myocardial ischemia/reperfusion injury. Metabolism. 2022;134:155239.

    CAS  PubMed  Google Scholar 

  143. Chen L, Liu B, Qin Y, Li A, Gao M, Liu H, Gong G. Mitochondrial fusion protein Mfn2 and its role in heart failure. Front Mol Biosci. 2021;8:681237.

    CAS  PubMed  PubMed Central  Google Scholar 

  144. Zhao T, Huang X, Han L, Wang X, Cheng H, Zhao Y, Chen Q, Chen J, Cheng H, Xiao R, Zheng M. Central role of mitofusin 2 in autophagosome-lysosome fusion in cardiomyocytes. J Biol Chem. 2012;287:23615–25.

    CAS  PubMed  PubMed Central  Google Scholar 

  145. Chen Y, Liu Y, Dorn GW 2nd. Mitochondrial fusion is essential for organelle function and cardiac homeostasis. Circ Res. 2011;109:1327–31.

    CAS  PubMed  PubMed Central  Google Scholar 

  146. Papanicolaou KN, Khairallah RJ, Ngoh GA, Chikando A, Luptak I, O’Shea KM, Riley DD, Lugus JJ, Colucci WS, Lederer WJ, et al. Mitofusin-2 maintains mitochondrial structure and contributes to stress-induced permeability transition in cardiac myocytes. Mol Cell Biol. 2011;31:1309–28.

    CAS  PubMed  PubMed Central  Google Scholar 

  147. Filadi R, Greotti E, Turacchio G, Luini A, Pozzan T, Pizzo P. Mitofusin 2 ablation increases endoplasmic reticulum-mitochondria coupling. Proc Natl Acad Sci U S A. 2015;112:E2174-2181.

    CAS  PubMed  PubMed Central  Google Scholar 

  148. Negri S, Faris P, Moccia F. Endolysosomal Ca(2+) signaling in cardiovascular health and disease. Int Rev Cell Mol Biol. 2021;363:203–69.

    CAS  PubMed  Google Scholar 

  149. Hall AR, Burke N, Dongworth RK, Kalkhoran SB, Dyson A, Vicencio JM, Dorn GW II, Yellon DM, Hausenloy DJ. Hearts deficient in both Mfn1 and Mfn2 are protected against acute myocardial infarction. Cell Death Dis. 2016;7:e2238.

    CAS  PubMed  PubMed Central  Google Scholar 

  150. de Brito OM, Scorrano L. Mitofusin 2: a mitochondria-shaping protein with signaling roles beyond fusion. Antioxid Redox Signal. 2008;10:621–33.

    PubMed  Google Scholar 

  151. Zhang X, Xu X, Lu L, Wan X, Qin Y, Ruan W, Lv C, He L, Guo X. A new Mfn-2 related synthetic peptide promotes vascular smooth muscle cell apoptosis via regulating the mitochondrial apoptotic pathway by inhibiting Akt signaling. J Transl Med. 2021;19:395.

    CAS  PubMed  PubMed Central  Google Scholar 

  152. Chen Y, Csordas G, Jowdy C, Schneider TG, Csordas N, Wang W, Liu Y, Kohlhaas M, Meiser M, Bergem S, et al. Mitofusin 2-containing mitochondrial-reticular microdomains direct rapid cardiomyocyte bioenergetic responses via interorganelle Ca(2+) crosstalk. Circ Res. 2012;111:863–75.

    CAS  PubMed  PubMed Central  Google Scholar 

  153. Seidlmayer LK, Mages C, Berbner A, Eder-Negrin P, Arias-Loza PA, Kaspar M, Song M, Dorn GW II, Kohlhaas M, Frantz S, et al. Mitofusin 2 is essential for IP(3)-mediated SR/mitochondria metabolic feedback in ventricular myocytes. Front Physiol. 2019;10:733.

    PubMed  PubMed Central  Google Scholar 

  154. Goh KY, Qu J, Hong H, Liu T, Dell’Italia LJ, Wu Y, O’Rourke B, Zhou L. Impaired mitochondrial network excitability in failing guinea-pig cardiomyocytes. Cardiovasc Res. 2016;109:79–89.

    CAS  PubMed  Google Scholar 

  155. Fang L, Moore XL, Gao XM, Dart AM, Lim YL, Du XJ. Down-regulation of mitofusin-2 expression in cardiac hypertrophy in vitro and in vivo. Life Sci. 2007;80:2154–60.

    CAS  PubMed  Google Scholar 

  156. Sabbah HN, Gupta RC, Singh-Gupta V, Zhang K, Lanfear DE. Abnormalities of mitochondrial dynamics in the failing heart: normalization following long-term therapy with elamipretide. Cardiovasc Drugs Ther. 2018;32:319–28.

    CAS  PubMed  PubMed Central  Google Scholar 

  157. Hsiao YT, Shimizu I, Wakasugi T, Yoshida Y, Ikegami R, Hayashi Y, Suda M, Katsuumi G, Nakao M, Ozawa T, et al. Cardiac mitofusin-1 is reduced in non-responding patients with idiopathic dilated cardiomyopathy. Sci Rep. 2021;11:6722.

    CAS  PubMed  PubMed Central  Google Scholar 

  158. Chen L, Gong Q, Stice JP, Knowlton AA. Mitochondrial OPA1, apoptosis, and heart failure. Cardiovasc Res. 2009;84:91–9.

    CAS  PubMed  PubMed Central  Google Scholar 

  159. Davies VJ, Hollins AJ, Piechota MJ, Yip W, Davies JR, White KE, Nicols PP, Boulton ME, Votruba M. Opa1 deficiency in a mouse model of autosomal dominant optic atrophy impairs mitochondrial morphology, optic nerve structure and visual function. Hum Mol Genet. 2007;16:1307–18.

    CAS  PubMed  Google Scholar 

  160. Wang J, Zhou H. Mitochondrial quality control mechanisms as molecular targets in cardiac ischemia-reperfusion injury. Acta Pharm Sin B. 2020;10:1866–79.

    PubMed  PubMed Central  Google Scholar 

  161. Le Page S, Niro M, Fauconnier J, Cellier L, Tamareille S, Gharib A, Chevrollier A, Loufrani L, Grenier C, Kamel R, et al. Increase in cardiac ischemia-reperfusion injuries in Opa1+/- mouse model. PLoS ONE. 2016;11:e0164066.

    PubMed  PubMed Central  Google Scholar 

  162. Noone J, O’Gorman DJ, Kenny HC. OPA1 regulation of mitochondrial dynamics in skeletal and cardiac muscle. Trends Endocrinol Metab. 2022;33:710–21.

    CAS  PubMed  Google Scholar 

  163. Ajoolabady A, Chiong M, Lavandero S, Klionsky DJ, Ren J. Mitophagy in cardiovascular diseases: molecular mechanisms, pathogenesis, and treatment. Trends Mol Med. 2022;28:836–49.

    CAS  PubMed  PubMed Central  Google Scholar 

  164. Quiles JM, Gustafsson AB. The role of mitochondrial fission in cardiovascular health and disease. Nat Rev Cardiol. 2022;19:723–36.

    CAS  PubMed  PubMed Central  Google Scholar 

  165. Wai T, Garcia-Prieto J, Baker MJ, Merkwirth C, Benit P, Rustin P, Ruperez FJ, Barbas C, Ibanez B, Langer T. Imbalanced OPA1 processing and mitochondrial fragmentation cause heart failure in mice. Science. 2015;350:aad0116.

    PubMed  Google Scholar 

  166. Nan J, Nan C, Ye J, Qian L, Geng Y, Xing D, Rahman MSU, Huang M. EGCG protects cardiomyocytes against hypoxia-reperfusion injury through inhibition of OMA1 activation. J Cell Sci. 2019;132:jcs220871.

    CAS  PubMed  Google Scholar 

  167. Kent AC, El Baradie KBY, Hamrick MW. Targeting the mitochondrial permeability transition pore to prevent age-associated cell damage and neurodegeneration. Oxid Med Cell Longev. 2021;2021:6626484.

    PubMed  PubMed Central  Google Scholar 

  168. Jang S, Javadov S. OPA1 regulates respiratory supercomplexes assembly: the role of mitochondrial swelling. Mitochondrion. 2020;51:30–9.

    CAS  PubMed  Google Scholar 

  169. Smyrnias I, Gray SP, Okonko DO, Sawyer G, Zoccarato A, Catibog N, Lopez B, Gonzalez A, Ravassa S, Diez J, Shah AM. Cardioprotective effect of the mitochondrial unfolded protein response during chronic pressure overload. J Am Coll Cardiol. 2019;73:1795–806.

    CAS  PubMed  PubMed Central  Google Scholar 

  170. Xu M, Xue RQ, Lu Y, Yong SY, Wu Q, Cui YL, Zuo XT, Yu XJ, Zhao M, Zang WJ. Choline ameliorates cardiac hypertrophy by regulating metabolic remodelling and UPRmt through SIRT3-AMPK pathway. Cardiovasc Res. 2019;115:530–45.

    CAS  PubMed  Google Scholar 

  171. Wang YT, Lim Y, McCall MN, Huang KT, Haynes CM, Nehrke K, Brookes PS. Cardioprotection by the mitochondrial unfolded protein response requires ATF5. Am J Physiol Heart Circ Physiol. 2019;317:H472–8.

    CAS  PubMed  PubMed Central  Google Scholar 

  172. Zhang B, Tan Y, Zhang Z, Feng P, Ding W, Wang Q, Liang H, Duan W, Wang X, Yu S, et al. Novel PGC-1alpha/ATF5 axis partly activates UPR(mt) and mediates cardioprotective role of tetrahydrocurcumin in pathological cardiac hypertrophy. Oxid Med Cell Longev. 2020;2020:9187065.

    PubMed  PubMed Central  Google Scholar 

  173. Venkatesh S, Li M, Saito T, Tong M, Rashed E, Mareedu S, Zhai P, Barcena C, Lopez-Otin C, Yehia G, et al. Mitochondrial LonP1 protects cardiomyocytes from ischemia/reperfusion injury in vivo. J Mol Cell Cardiol. 2019;128:38–50.

    CAS  PubMed  Google Scholar 

  174. Nandi SS, Katsurada K, Mahata SK, Patel KP. Neurogenic hypertension mediated mitochondrial abnormality leads to cardiomyopathy: contribution of UPR(mt) and norepinephrine-miR- 18a–5p-HIF-1alpha Axis. Front Physiol. 2021;12:718982.

    PubMed  PubMed Central  Google Scholar 

  175. Liu X, Lei J, Wang K, Ma L, Liu D, Du Y, Wu Y, Zhang S, Wang W, Ma X, Liu H. Mitochondrial Omi/HtrA2 promotes caspase activation through cleavage of HAX-1 in aging heart. Rejuvenation Res. 2017;20:183–92.

    CAS  PubMed  Google Scholar 

  176. Svagusa T, Martinic M, Martinic M, Kovacevic L, Sepac A, Milicic D, Bulum J, Starcevic B, Sirotkovic-Skerlev M, Seiwerth F, et al. Mitochondrial unfolded protein response, mitophagy and other mitochondrial quality control mechanisms in heart disease and aged heart. Croat Med J. 2020;61:126–38.

    CAS  PubMed  PubMed Central  Google Scholar 

  177. Liu M, Lv J, Pan Z, Wang D, Zhao L, Guo X. Mitochondrial dysfunction in heart failure and its therapeutic implications. Front Cardiovasc Med. 2022;9:945142.

    CAS  PubMed  PubMed Central  Google Scholar 

  178. Duan Q, Chen C, Yang L, Li N, Gong W, Li S, Wang DW. MicroRNA regulation of unfolded protein response transcription factor XBP1 in the progression of cardiac hypertrophy and heart failure in vivo. J Transl Med. 2015;13:363.

    PubMed  PubMed Central  Google Scholar 

  179. Wang Y, Jasper H, Toan S, Muid D, Chang X, Zhou H. Mitophagy coordinates the mitochondrial unfolded protein response to attenuate inflammation-mediated myocardial injury. Redox Biol. 2021;45:102049.

    CAS  PubMed  PubMed Central  Google Scholar 

  180. Dabravolski SA, Khotina VA, Sukhorukov VN, Kalmykov VA, Mikhaleva LM, Orekhov AN. The role of mitochondrial DNA mutations in cardiovascular diseases. Int J Mol Sci. 2022;23:952.

    CAS  PubMed  PubMed Central  Google Scholar 

  181. Luo J, Shen S, Xia J, Wang J, Gu Z. Mitochondria as the essence of Yang Qi in the human body. Phenomics. 2022;2:336–48.

    PubMed  PubMed Central  Google Scholar 

  182. Zhu Y, You J, Xu C, Gu X. Associations of mitochondrial DNA 3777–4679 region mutations with maternally inherited essential hypertensive subjects in China. BMC Med Genet. 2020;21:105.

    CAS  PubMed  PubMed Central  Google Scholar 

  183. Aravintha Siva M, Mahalakshmi R, Bhakta-Guha D, Guha G. Gene therapy for the mitochondrial genome: purging mutations, pacifying ailments. Mitochondrion. 2019;46:195–208.

    CAS  PubMed  Google Scholar 

  184. DiMauro S, Schon EA. Mitochondrial respiratory-chain diseases. N Engl J Med. 2003;348:2656–68.

    CAS  PubMed  Google Scholar 

  185. Bray AW, Ballinger SW. Mitochondrial DNA mutations and cardiovascular disease. Curr Opin Cardiol. 2017;32:267–74.

    PubMed  PubMed Central  Google Scholar 

  186. Friederich MW, Geddes GC, Wortmann SB, Punnoose A, Wartchow E, Knight KM, Prokisch H, Creadon-Swindell G, Mayr JA, Van Hove JLK. Pathogenic variants in MRPL44 cause infantile cardiomyopathy due to a mitochondrial translation defect. Mol Genet Metab. 2021;133:362–71.

    CAS  PubMed  PubMed Central  Google Scholar 

  187. Tang J, Tang QX, Liu S. METTL3-modified lncRNA-SNHG8 binds to PTBP1 to regulate ALAS2 expression to increase oxidative stress and promote myocardial infarction. Mol Cell Biochem. 2023;478:1217–29.

    CAS  PubMed  Google Scholar 

  188. Kumar AA, Kelly DP, Chirinos JA. Mitochondrial dysfunction in heart failure with preserved ejection fraction. Circulation. 2019;139:1435–50.

    PubMed  PubMed Central  Google Scholar 

  189. Lioncino M, Monda E, Caiazza M, Fusco A, Cirillo A, Dongiglio F, Simonelli V, Sampaolo S, Ruggiero L, Scarano G, et al. Cardiovascular involvement in mtDNA disease: diagnosis, management, and therapeutic options. Heart Fail Clin. 2022;18:51–60.

    PubMed  Google Scholar 

  190. Zhu Y, Gu X, Xu C. Mitochondrial DNA 7908–8816 region mutations in maternally inherited essential hypertensive subjects in China. BMC Med Genomics. 2018;11:89.

    CAS  PubMed  PubMed Central  Google Scholar 

  191. Markin AM, Khotina VA, Zabudskaya XG, Bogatyreva AI, Starodubova AV, Ivanova E, Nikiforov NG, Orekhov AN. Disturbance of mitochondrial dynamics and mitochondrial therapies in atherosclerosis. Life. 2021;11:165.

    CAS  PubMed  PubMed Central  Google Scholar 

  192. Sobenin IA, Sazonova MA, Postnov AY, Bobryshev YV, Orekhov AN. Changes of mitochondria in atherosclerosis: possible determinant in the pathogenesis of the disease. Atherosclerosis. 2013;227:283–8.

    CAS  PubMed  Google Scholar 

  193. Takada S, Maekawa S, Furihata T, Kakutani N, Setoyama D, Ueda K, Nambu H, Hagiwara H, Handa H, Fumoto Y, et al. Succinyl-CoA-based energy metabolism dysfunction in chronic heart failure. Proc Natl Acad Sci USA. 2022;119:e2203628119.

    CAS  PubMed  PubMed Central  Google Scholar 

  194. Li Y, Ren P, Dawson A, Vasquez HG, Ageedi W, Zhang C, Luo W, Chen R, Li Y, Kim S, et al. Single-cell transcriptome analysis reveals dynamic cell populations and differential gene expression patterns in control and aneurysmal human aortic tissue. Circulation. 2020;142:1374–88.

    CAS  PubMed  PubMed Central  Google Scholar 

  195. Abudupataer M, Zhu S, Yan S, Xu K, Zhang J, Luo S, Ma W, Alam MF, Tang Y, Huang H, et al. Aorta smooth muscle-on-a-chip reveals impaired mitochondrial dynamics as a therapeutic target for aortic aneurysm in bicuspid aortic valve disease. Elife. 2021;10:e69310.

    CAS  PubMed  PubMed Central  Google Scholar 

  196. Riehle C, Wende AR, Zaha VG, Pires KM, Wayment B, Olsen C, Bugger H, Buchanan J, Wang X, Moreira AB, et al. PGC-1beta deficiency accelerates the transition to heart failure in pressure overload hypertrophy. Circ Res. 2011;109:783–93.

    CAS  PubMed  PubMed Central  Google Scholar 

  197. Porporato PE, Filigheddu N, Pedro JMB, Kroemer G, Galluzzi L. Mitochondrial metabolism and cancer. Cell Res. 2018;28:265–80.

    CAS  PubMed  Google Scholar 

  198. Boland ML, Chourasia AH, Macleod KF. Mitochondrial dysfunction in cancer. Front Oncol. 2013;3:292.

    PubMed  PubMed Central  Google Scholar 

  199. Liu AR, Lv Z, Yan ZW, Wu XY, Yan LR, Sun LP, Yuan Y, Xu Q. Association of mitochondrial homeostasis and dynamic balance with malignant biological behaviors of gastrointestinal cancer. J Transl Med. 2023;21:27.

    PubMed  PubMed Central  Google Scholar 

  200. Anderson GR, Wardell SE, Cakir M, Yip C, Ahn YR, Ali M, Yllanes AP, Chao CA, McDonnell DP, Wood KC. Dysregulation of mitochondrial dynamics proteins are a targetable feature of human tumors. Nat Commun. 2018;9:1677.

    PubMed  PubMed Central  Google Scholar 

  201. Sanchez-Alvarez R, De Francesco EM, Fiorillo M, Sotgia F, Lisanti MP. Mitochondrial fission factor (MFF) inhibits mitochondrial metabolism and reduces breast cancer stem cell (CSC) activity. Front Oncol. 2020;10:1776.

    PubMed  PubMed Central  Google Scholar 

  202. D’Aquila P, Ronchetti D, Gallo Cantafio ME, Todoerti K, Taiana E, Fabiani F, Montesanto A, Neri A, Passarino G, Viglietto G, et al. Epigenetic regulation of mitochondrial quality control genes in multiple myeloma: a sequenom massARRAY pilot investigation on HMCLs. J Clin Med. 2021;10:1295.

    CAS  PubMed  PubMed Central  Google Scholar 

  203. Xu X, Wang X, Chen Q, Zheng A, Li D, Meng Z, Li X, Cai H, Li W, Huang S, Wang F. Sp1 promotes tumour progression by remodelling the mitochondrial network in cervical cancer. J Transl Med. 2023;21:307.

    CAS  PubMed  PubMed Central  Google Scholar 

  204. Zhao J, Zhang J, Yu M, Xie Y, Huang Y, Wolff DW, Abel PW, Tu Y. Mitochondrial dynamics regulates migration and invasion of breast cancer cells. Oncogene. 2013;32:4814–24.

    CAS  PubMed  Google Scholar 

  205. Boulton DP, Caino MC. Mitochondrial fission and fusion in tumor progression to metastasis. Front Cell Dev Biol. 2022;10:849962.

    PubMed  PubMed Central  Google Scholar 

  206. Jin P, Jiang J, Zhou L, Huang Z, Nice EC, Huang C, Fu L. Mitochondrial adaptation in cancer drug resistance: prevalence, mechanisms, and management. J Hematol Oncol. 2022;15:97.

    PubMed  PubMed Central  Google Scholar 

  207. Chen H, Chan DC. Mitochondrial dynamics in regulating the unique phenotypes of cancer and stem cells. Cell Metab. 2017;26:39–48.

    CAS  PubMed  PubMed Central  Google Scholar 

  208. Roth KG, Mambetsariev I, Kulkarni P, Salgia R. The mitochondrion as an emerging therapeutic target in cancer. Trends Mol Med. 2020;26:119–34.

    CAS  PubMed  Google Scholar 

  209. Bordt EA, Clerc P, Roelofs BA, Saladino AJ, Tretter L, Adam-Vizi V, Cherok E, Khalil A, Yadava N, Ge SX, et al. The putative Drp1 inhibitor mdivi-1 is a reversible mitochondrial complex I inhibitor that modulates reactive oxygen species. Dev Cell. 2017;40:583-594.e586.

    CAS  PubMed  PubMed Central  Google Scholar 

  210. Seo JH, Agarwal E, Chae YC, Lee YG, Garlick DS, Storaci AM, Ferrero S, Gaudioso G, Gianelli U, Vaira V, Altieri DC. Mitochondrial fission factor is a novel Myc-dependent regulator of mitochondrial permeability in cancer. EBioMedicine. 2019;48:353–63.

    CAS  PubMed  PubMed Central  Google Scholar 

  211. Seo JH, Chae YC, Kossenkov AV, Lee YG, Tang HY, Agarwal E, Gabrilovich DI, Languino LR, Speicher DW, Shastrula PK, et al. MFF regulation of mitochondrial cell death is a therapeutic target in cancer. Cancer Res. 2019;79:6215–26.

    CAS  PubMed  PubMed Central  Google Scholar 

  212. Gallo Cantafio ME, Torcasio R, Viglietto G, Amodio N. Non-coding RNA-dependent regulation of mitochondrial dynamics in cancer pathophysiology. Noncoding RNA. 2023;9:16.

    CAS  PubMed  PubMed Central  Google Scholar 

  213. Zacharioudakis E, Agianian B, Kumar Mv V, Biris N, Garner TP, Rabinovich-Nikitin I, Ouchida AT, Margulets V, Nordstrom LU, Riley JS, et al. Modulating mitofusins to control mitochondrial function and signaling. Nat Commun. 2022;13:3775.

    CAS  PubMed  PubMed Central  Google Scholar 

  214. Verdin E, Hirschey MD, Finley LW, Haigis MC. Sirtuin regulation of mitochondria: energy production, apoptosis, and signaling. Trends Biochem Sci. 2010;35:669–75.

    CAS  PubMed  PubMed Central  Google Scholar 

  215. Villalba JM, Alcain FJ. Sirtuin activators and inhibitors. BioFactors. 2012;38:349–59.

    CAS  PubMed  PubMed Central  Google Scholar 

  216. Zhu S, Dong Z, Ke X, Hou J, Zhao E, Zhang K, Wang F, Yang L, Xiang Z, Cui H. The roles of sirtuins family in cell metabolism during tumor development. Semin Cancer Biol. 2019;57:59–71.

    CAS  PubMed  Google Scholar 

  217. Zhang X, Ren X, Zhang Q, Li Z, Ma S, Bao J, Li Z, Bai X, Zheng L, Zhang Z, et al. PGC-1alpha/ERRalpha-Sirt3 pathway regulates DAergic neuronal death by directly deacetylating sod2 and ATP synthase beta. Antioxid Redox Signal. 2016;24:312–28.

    CAS  PubMed  PubMed Central  Google Scholar 

  218. Brindisi M, Fiorillo M, Frattaruolo L, Sotgia F, Lisanti MP, Cappello AR. Cholesterol and mevalonate: two metabolites involved in breast cancer progression and drug resistance through the ERRalpha pathway. Cells. 1819;2020:9.

    Google Scholar 

  219. Puigserver P, Spiegelman BM. Peroxisome proliferator-activated receptor-gamma coactivator 1 alpha (PGC-1 alpha): transcriptional coactivator and metabolic regulator. Endocr Rev. 2003;24:78–90.

    CAS  PubMed  Google Scholar 

  220. Liang H, Ward WF. PGC-1alpha: a key regulator of energy metabolism. Adv Physiol Educ. 2006;30:145–51.

    PubMed  Google Scholar 

  221. Canto C, Auwerx J. PGC-1alpha, SIRT1 and AMPK, an energy sensing network that controls energy expenditure. Curr Opin Lipidol. 2009;20:98–105.

    CAS  PubMed  PubMed Central  Google Scholar 

  222. Ertel A, Tsirigos A, Whitaker-Menezes D, Birbe RC, Pavlides S, Martinez-Outschoorn UE, Pestell RG, Howell A, Sotgia F, Lisanti MP. Is cancer a metabolic rebellion against host aging? In the quest for immortality, tumor cells try to save themselves by boosting mitochondrial metabolism. Cell Cycle. 2012;11:253–63.

    CAS  PubMed  PubMed Central  Google Scholar 

  223. Praharaj PP, Panigrahi DP, Bhol CS, Patra S, Mishra SR, Mahapatra KK, Behera BP, Singh A, Patil S, Bhutia SK. Mitochondrial rewiring through mitophagy and mitochondrial biogenesis in cancer stem cells: a potential target for anti-CSC cancer therapy. Cancer Lett. 2021;498:217–28.

    CAS  PubMed  Google Scholar 

  224. Ayob AZ, Ramasamy TS. Cancer stem cells as key drivers of tumour progression. J Biomed Sci. 2018;25:20.

    PubMed  PubMed Central  Google Scholar 

  225. Yu Z, Pestell TG, Lisanti MP, Pestell RG. Cancer stem cells. Int J Biochem Cell Biol. 2012;44:2144–51.

    CAS  PubMed  PubMed Central  Google Scholar 

  226. Tan Z, Luo X, Xiao L, Tang M, Bode AM, Dong Z, Cao Y. The role of PGC1alpha in cancer metabolism and its therapeutic implications. Mol Cancer Ther. 2016;15:774–82.

    CAS  PubMed  Google Scholar 

  227. De Luca A, Fiorillo M, Peiris-Pages M, Ozsvari B, Smith DL, Sanchez-Alvarez R, Martinez-Outschoorn UE, Cappello AR, Pezzi V, Lisanti MP, Sotgia F. Mitochondrial biogenesis is required for the anchorage-independent survival and propagation of stem-like cancer cells. Oncotarget. 2015;6:14777–95.

    PubMed  PubMed Central  Google Scholar 

  228. Wang M, Law ME, Castellano RK, Law BK. The unfolded protein response as a target for anticancer therapeutics. Crit Rev Oncol Hematol. 2018;127:66–79.

    PubMed  Google Scholar 

  229. Kim I, Xu W, Reed JC. Cell death and endoplasmic reticulum stress: disease relevance and therapeutic opportunities. Nat Rev Drug Discov. 2008;7:1013–30.

    CAS  PubMed  Google Scholar 

  230. Corazzari M, Gagliardi M, Fimia GM, Piacentini M. Endoplasmic reticulum stress, unfolded protein response, and cancer cell fate. Front Oncol. 2017;7:78.

    PubMed  PubMed Central  Google Scholar 

  231. Yadav RK, Chae SW, Kim HR, Chae HJ. Endoplasmic reticulum stress and cancer. J Cancer Prev. 2014;19:75–88.

    PubMed  PubMed Central  Google Scholar 

  232. Ojha R, Amaravadi RK. Targeting the unfolded protein response in cancer. Pharmacol Res. 2017;120:258–66.

    CAS  PubMed  PubMed Central  Google Scholar 

  233. Koopman WJ, Distelmaier F, Smeitink JA, Willems PH. OXPHOS mutations and neurodegeneration. EMBO J. 2013;32:9–29.

    CAS  PubMed  Google Scholar 

  234. Wang Z, Bo H, Song Y, Li C, Zhang Y. Mitochondrial ROS produced by skeletal muscle mitochondria promote the decisive signal for UPRmt activation. Biomed Res Int. 2022;2022:7436577.

    PubMed  PubMed Central  Google Scholar 

  235. Munch C. The different axes of the mammalian mitochondrial unfolded protein response. BMC Biol. 2018;16:81.

    PubMed  PubMed Central  Google Scholar 

  236. Papa L, Germain D. SirT3 regulates the mitochondrial unfolded protein response. Mol Cell Biol. 2014;34:699–710.

    PubMed  PubMed Central  Google Scholar 

  237. Jovaisaite V, Mouchiroud L, Auwerx J. The mitochondrial unfolded protein response, a conserved stress response pathway with implications in health and disease. J Exp Biol. 2014;217:137–43.

    CAS  PubMed  PubMed Central  Google Scholar 

  238. Papa L, Germain D. Estrogen receptor mediates a distinct mitochondrial unfolded protein response. J Cell Sci. 2011;124:1396–402.

    CAS  PubMed  PubMed Central  Google Scholar 

  239. Shevtsov M, Multhoff G, Mikhaylova E, Shibata A, Guzhova I, Margulis B. Combination of anti-cancer drugs with molecular chaperone inhibitors. Int J Mol Sci. 2019;20:5284.

    CAS  PubMed  PubMed Central  Google Scholar 

  240. Inigo JR, Kumar R, Chandra D. Targeting the mitochondrial unfolded protein response in cancer: opportunities and challenges. Trends Cancer. 2021;7:1050–3.

    CAS  PubMed  PubMed Central  Google Scholar 

  241. Chourasia AH, Boland ML, Macleod KF. Mitophagy and cancer. Cancer Metab. 2015;3:4.

    PubMed  PubMed Central  Google Scholar 

  242. Vara-Perez M, Felipe-Abrio B, Agostinis P. Mitophagy in cancer: a tale of adaptation. Cells. 2019;8:493.

    CAS  PubMed  PubMed Central  Google Scholar 

  243. Guan Y, Wang Y, Li B, Shen K, Li Q, Ni Y, Huang L. Mitophagy in carcinogenesis, drug resistance and anticancer therapeutics. Cancer Cell Int. 2021;21:350.

    PubMed  PubMed Central  Google Scholar 

  244. Chang JY, Yi HS, Kim HW, Shong M. Dysregulation of mitophagy in carcinogenesis and tumor progression. Biochim Biophys Acta Bioenerg. 2017;1858:633–40.

    CAS  PubMed  Google Scholar 

  245. Iommarini L, Ghelli A, Gasparre G, Porcelli AM. Mitochondrial metabolism and energy sensing in tumor progression. Biochim Biophys Acta Bioenerg. 2017;1858:582–90.

    CAS  PubMed  Google Scholar 

  246. Ding WX, Yin XM. Mitophagy: mechanisms, pathophysiological roles, and analysis. Biol Chem. 2012;393:547–64.

    CAS  PubMed  PubMed Central  Google Scholar 

  247. Vives-Bauza C, Zhou C, Huang Y, Cui M, de Vries RL, Kim J, May J, Tocilescu MA, Liu W, Ko HS, et al. PINK1-dependent recruitment of Parkin to mitochondria in mitophagy. Proc Natl Acad Sci U S A. 2010;107:378–83.

    CAS  PubMed  Google Scholar 

  248. Matsuda N, Sato S, Shiba K, Okatsu K, Saisho K, Gautier CA, Sou YS, Saiki S, Kawajiri S, Sato F, et al. PINK1 stabilized by mitochondrial depolarization recruits Parkin to damaged mitochondria and activates latent Parkin for mitophagy. J Cell Biol. 2010;189:211–21.

    CAS  PubMed  PubMed Central  Google Scholar 

  249. Ashrafi G, Schwarz TL. The pathways of mitophagy for quality control and clearance of mitochondria. Cell Death Differ. 2013;20:31–42.

    CAS  PubMed  Google Scholar 

  250. De Francesco EM, Sotgia F, Lisanti MP. Cancer stem cells (CSCs): metabolic strategies for their identification and eradication. Biochem J. 2018;475:1611–34.

    PubMed  Google Scholar 

  251. Peiris-Pages M, Martinez-Outschoorn UE, Pestell RG, Sotgia F, Lisanti MP. Cancer stem cell metabolism. Breast Cancer Res. 2016;18:55.

    PubMed  PubMed Central  Google Scholar 

  252. Nazio F, Bordi M, Cianfanelli V, Locatelli F, Cecconi F. Autophagy and cancer stem cells: molecular mechanisms and therapeutic applications. Cell Death Differ. 2019;26:690–702.

    CAS  PubMed  PubMed Central  Google Scholar 

  253. Batlle E, Clevers H. Cancer stem cells revisited. Nat Med. 2017;23:1124–34.

    CAS  PubMed  Google Scholar 

  254. Vlashi E, Pajonk F. The metabolic state of cancer stem cells-a valid target for cancer therapy? Free Radic Biol Med. 2015;79:264–8.

    CAS  PubMed  Google Scholar 

  255. Najafzadeh B, Asadzadeh Z, Motafakker Azad R, Mokhtarzadeh A, Baghbanzadeh A, Alemohammad H, Abdoli Shadbad M, Vasefifar P, Najafi S, Baradaran B. The oncogenic potential of NANOG: an important cancer induction mediator. J Cell Physiol. 2021;236:2443–58.

    CAS  PubMed  Google Scholar 

  256. Song C, Xu Q, Jiang K, Zhou G, Yu X, Wang L, Zhu Y, Fang L, Yu Z, Lee JD, et al. Inhibition of BMK1 pathway suppresses cancer stem cells through BNIP3 and BNIP3L. Oncotarget. 2015;6:33279–89.

    PubMed  PubMed Central  Google Scholar 

  257. Fiorillo M, Sotgia F, Lisanti MP. “Energetic” cancer stem cells (e-CSCs): a new hyper-metabolic and proliferative tumor cell phenotype, driven by mitochondrial energy. Front Oncol. 2018;8:677.

    PubMed  Google Scholar 

  258. Ozsvari B, Bonuccelli G, Sanchez-Alvarez R, Foster R, Sotgia F, Lisanti MP. Targeting flavin-containing enzymes eliminates cancer stem cells (CSCs), by inhibiting mitochondrial respiration: vitamin B2 (Riboflavin) in cancer therapy. Aging. 2017;9:2610–28.

    CAS  PubMed  PubMed Central  Google Scholar 

  259. Miranda-Lorenzo I, Dorado J, Lonardo E, Alcala S, Serrano AG, Clausell-Tormos J, Cioffi M, Megias D, Zagorac S, Balic A, et al. Intracellular autofluorescence: a biomarker for epithelial cancer stem cells. Nat Methods. 2014;11:1161–9.

    CAS  PubMed  Google Scholar 

  260. Fiorillo M, Scatena C, Naccarato AG, Sotgia F, Lisanti MP. Bedaquiline, an FDA-approved drug, inhibits mitochondrial ATP production and metastasis in vivo, by targeting the gamma subunit (ATP5F1C) of the ATP synthase. Cell Death Differ. 2021;28:2797–817.

    CAS  PubMed  PubMed Central  Google Scholar 

  261. Fiorillo M, Lamb R, Tanowitz HB, Cappello AR, Martinez-Outschoorn UE, Sotgia F, Lisanti MP. Bedaquiline, an FDA-approved antibiotic, inhibits mitochondrial function and potently blocks the proliferative expansion of stem-like cancer cells (CSCs). Aging. 2016;8:1593–607.

    CAS  PubMed  PubMed Central  Google Scholar 

  262. Liberti MV, Locasale JW. The Warburg effect: how does it benefit cancer cells? Trends Biochem Sci. 2016;41:211–8.

    CAS  PubMed  PubMed Central  Google Scholar 

  263. Pavlides S, Vera I, Gandara R, Sneddon S, Pestell RG, Mercier I, Martinez-Outschoorn UE, Whitaker-Menezes D, Howell A, Sotgia F, Lisanti MP. Warburg meets autophagy: cancer-associated fibroblasts accelerate tumor growth and metastasis via oxidative stress, mitophagy, and aerobic glycolysis. Antioxid Redox Signal. 2012;16:1264–84.

    CAS  PubMed  PubMed Central  Google Scholar 

  264. DeBerardinis RJ, Chandel NS. We need to talk about the Warburg effect. Nat Metab. 2020;2:127–9.

    PubMed  Google Scholar 

  265. Vander Heiden MG, Cantley LC, Thompson CB. Understanding the Warburg effect: the metabolic requirements of cell proliferation. Science. 2009;324:1029–33.

    CAS  PubMed  PubMed Central  Google Scholar 

  266. Hayes JD, Dinkova-Kostova AT, Tew KD. Oxidative stress in cancer. Cancer Cell. 2020;38:167–97.

    CAS  PubMed  PubMed Central  Google Scholar 

  267. Gaude E, Frezza C. Defects in mitochondrial metabolism and cancer. Cancer Metab. 2014;2:10.

    PubMed  PubMed Central  Google Scholar 

  268. Arnandis T, Monteiro P, Adams SD, Bridgeman VL, Rajeeve V, Gadaleta E, Marzec J, Chelala C, Malanchi I, Cutillas PR, Godinho SA. Oxidative stress in cells with extra centrosomes drives non-cell-autonomous invasion. Dev Cell. 2018;47:409-424.e409.

    CAS  PubMed  PubMed Central  Google Scholar 

  269. Zheng J. Energy metabolism of cancer: glycolysis versus oxidative phosphorylation (Review). Oncol Lett. 2012;4:1151–7.

    CAS  PubMed  PubMed Central  Google Scholar 

  270. Martinez-Outschoorn UE, Peiris-Pages M, Pestell RG, Sotgia F, Lisanti MP. Cancer metabolism: a therapeutic perspective. Nat Rev Clin Oncol. 2017;14:11–31.

    CAS  PubMed  Google Scholar 

  271. Palorini R, Simonetto T, Cirulli C, Chiaradonna F. Mitochondrial complex I inhibitors and forced oxidative phosphorylation synergize in inducing cancer cell death. Int J Cell Biol. 2013;2013:243876.

    PubMed  PubMed Central  Google Scholar 

  272. Fiorillo M, Ozsvari B, Sotgia F, Lisanti MP. High ATP production fuels cancer drug resistance and metastasis: implications for mitochondrial ATP depletion therapy. Front Oncol. 2021;11:740720.

    CAS  PubMed  PubMed Central  Google Scholar 

  273. Wei Q, Qian Y, Yu J, Wong CC. Metabolic rewiring in the promotion of cancer metastasis: mechanisms and therapeutic implications. Oncogene. 2020;39:6139–56.

    CAS  PubMed  PubMed Central  Google Scholar 

  274. Bonuccelli G, Whitaker-Menezes D, Castello-Cros R, Pavlides S, Pestell RG, Fatatis A, Witkiewicz AK, Vander Heiden MG, Migneco G, Chiavarina B, et al. The reverse Warburg effect: glycolysis inhibitors prevent the tumor promoting effects of caveolin-1 deficient cancer associated fibroblasts. Cell Cycle. 2010;9:1960–71.

    CAS  PubMed  Google Scholar 

  275. Goodwin ML, Gladden LB, Nijsten MW, Jones KB. Lactate and cancer: revisiting the Warburg effect in an era of lactate shuttling. Front Nutr. 2014;1:27.

    PubMed  Google Scholar 

  276. Putignani L, Raffa S, Pescosolido R, Aimati L, Signore F, Torrisi MR, Grammatico P. Alteration of expression levels of the oxidative phosphorylation system (OXPHOS) in breast cancer cell mitochondria. Breast Cancer Res Treat. 2008;110:439–52.

    CAS  PubMed  Google Scholar 

  277. Wilde L, Roche M, Domingo-Vidal M, Tanson K, Philp N, Curry J, Martinez-Outschoorn U. Metabolic coupling and the Reverse Warburg Effect in cancer: implications for novel biomarker and anticancer agent development. Semin Oncol. 2017;44:198–203.

    CAS  PubMed  PubMed Central  Google Scholar 

  278. Fu Y, Liu S, Yin S, Niu W, Xiong W, Tan M, Li G, Zhou M. The reverse Warburg effect is likely to be an Achilles’ heel of cancer that can be exploited for cancer therapy. Oncotarget. 2017;8:57813–25.

    PubMed  PubMed Central  Google Scholar 

  279. Sancho P, Burgos-Ramos E, Tavera A, Bou Kheir T, Jagust P, Schoenhals M, Barneda D, Sellers K, Campos-Olivas R, Grana O, et al. MYC/PGC-1alpha balance determines the metabolic phenotype and plasticity of pancreatic cancer stem cells. Cell Metab. 2015;22:590–605.

    CAS  PubMed  Google Scholar 

  280. Gentric G, Kieffer Y, Mieulet V, Goundiam O, Bonneau C, Nemati F, Hurbain I, Raposo G, Popova T, Stern MH, et al. PML-regulated mitochondrial metabolism enhances chemosensitivity in human ovarian cancers. Cell Metab. 2019;29(156–173):e110.

    Google Scholar 

  281. Xiang Y, Fang B, Liu Y, Yan S, Cao D, Mei H, Wang Q, Hu Y, Guo T. SR18292 exerts potent antitumor effects in multiple myeloma via inhibition of oxidative phosphorylation. Life Sci. 2020;256:117971.

    CAS  PubMed  Google Scholar 

  282. Chaudhary S, Ganguly S, Palanichamy JK, Singh A, Bakhshi R, Jain A, Chopra A, Bakhshi S. PGC1A driven enhanced mitochondrial DNA copy number predicts outcome in pediatric acute myeloid leukemia. Mitochondrion. 2021;58:246–54.

    CAS  PubMed  Google Scholar 

  283. Wallace DC. Mitochondria and cancer. Nat Rev Cancer. 2012;12:685–98.

    CAS  PubMed  PubMed Central  Google Scholar 

  284. Tan AS, Baty JW, Dong LF, Bezawork-Geleta A, Endaya B, Goodwin J, Bajzikova M, Kovarova J, Peterka M, Yan B, et al. Mitochondrial genome acquisition restores respiratory function and tumorigenic potential of cancer cells without mitochondrial DNA. Cell Metab. 2015;21:81–94.

    CAS  PubMed  Google Scholar 

  285. Moschoi R, Imbert V, Nebout M, Chiche J, Mary D, Prebet T, Saland E, Castellano R, Pouyet L, Collette Y, et al. Protective mitochondrial transfer from bone marrow stromal cells to acute myeloid leukemic cells during chemotherapy. Blood. 2016;128:253–64.

    CAS  PubMed  Google Scholar 

  286. Lamb R, Bonuccelli G, Ozsvari B, Peiris-Pages M, Fiorillo M, Smith DL, Bevilacqua G, Mazzanti CM, McDonnell LA, Naccarato AG, et al. Mitochondrial mass, a new metabolic biomarker for stem-like cancer cells: understanding WNT/FGF-driven anabolic signaling. Oncotarget. 2015;6:30453–71.

    PubMed  PubMed Central  Google Scholar 

  287. Lamb R, Ozsvari B, Bonuccelli G, Smith DL, Pestell RG, Martinez-Outschoorn UE, Clarke RB, Sotgia F, Lisanti MP. Dissecting tumor metabolic heterogeneity: telomerase and large cell size metabolically define a sub-population of stem-like, mitochondrial-rich, cancer cells. Oncotarget. 2015;6:21892–905.

    PubMed  PubMed Central  Google Scholar 

  288. Criscuolo D, Avolio R, Matassa DS, Esposito F. Targeting mitochondrial protein expression as a future approach for cancer therapy. Front Oncol. 2021;11:797265.

    CAS  PubMed  PubMed Central  Google Scholar 

  289. Scheid AD, Beadnell TC, Welch DR. Roles of mitochondria in the hallmarks of metastasis. Br J Cancer. 2021;124:124–35.

    CAS  PubMed  Google Scholar 

  290. Sancho P, Barneda D, Heeschen C. Hallmarks of cancer stem cell metabolism. Br J Cancer. 2016;114:1305–12.

    CAS  PubMed  PubMed Central  Google Scholar 

  291. Bonuccelli G, Sotgia F, Lisanti MP. Identification of natural products and FDA-approved drugs for targeting cancer stem cell (CSC) propagation. Aging. 2022;14:9466–83.

    CAS  PubMed  PubMed Central  Google Scholar 

  292. Sotgia F, Ozsvari B, Fiorillo M, De Francesco EM, Bonuccelli G, Lisanti MP. A mitochondrial based oncology platform for targeting cancer stem cells (CSCs): MITO-ONC-RX. Cell Cycle. 2018;17:2091–100.

    CAS  PubMed  PubMed Central  Google Scholar 

  293. Fiorillo M, Lamb R, Tanowitz HB, Mutti L, Krstic-Demonacos M, Cappello AR, Martinez-Outschoorn UE, Sotgia F, Lisanti MP. Repurposing atovaquone: targeting mitochondrial complex III and OXPHOS to eradicate cancer stem cells. Oncotarget. 2016;7:34084–99.

    PubMed  PubMed Central  Google Scholar 

  294. Gupta N, Srivastava SK. Atovaquone suppresses the growth of metastatic triple-negative breast tumors in lungs and brain by inhibiting integrin/FAK signaling axis. Pharmaceuticals. 2021;14:521.

    CAS  PubMed  PubMed Central  Google Scholar 

  295. Guo Y, Hu B, Fu B, Zhu H. Atovaquone at clinically relevant concentration overcomes chemoresistance in ovarian cancer via inhibiting mitochondrial respiration. Pathol Res Pract. 2021;224:153529.

    CAS  PubMed  Google Scholar 

  296. Kapur A, Mehta P, Simmons AD, Ericksen SS, Mehta G, Palecek SP, Felder M, Stenerson Z, Nayak A, Dominguez JMA, et al. Atovaquone: an inhibitor of oxidative phosphorylation as studied in gynecologic cancers. Cancers. 2022;14:2297.

    CAS  PubMed  PubMed Central  Google Scholar 

  297. Coates JTT, Rodriguez-Berriguete G, Puliyadi R, Ashton T, Prevo R, Wing A, Granata G, Pirovano G, McKenna GW, Higgins GS. The anti-malarial drug atovaquone potentiates platinum-mediated cancer cell death by increasing oxidative stress. Cell Death Discov. 2020;6:110.

    CAS  PubMed  PubMed Central  Google Scholar 

  298. Madiraju AK, Erion DM, Rahimi Y, Zhang XM, Braddock DT, Albright RA, Prigaro BJ, Wood JL, Bhanot S, MacDonald MJ, et al. Metformin suppresses gluconeogenesis by inhibiting mitochondrial glycerophosphate dehydrogenase. Nature. 2014;510:542–6.

    CAS  PubMed  PubMed Central  Google Scholar 

  299. Vancura A, Bu P, Bhagwat M, Zeng J, Vancurova I. Metformin as an anticancer agent. Trends Pharmacol Sci. 2018;39:867–78.

    CAS  PubMed  PubMed Central  Google Scholar 

  300. Dowling RJ, Goodwin PJ, Stambolic V. Understanding the benefit of metformin use in cancer treatment. BMC Med. 2011;9:33.

    CAS  PubMed  PubMed Central  Google Scholar 

  301. Hirsch HA, Iliopoulos D, Tsichlis PN, Struhl K. Metformin selectively targets cancer stem cells, and acts together with chemotherapy to block tumor growth and prolong remission. Cancer Res. 2009;69:7507–11.

    CAS  PubMed  PubMed Central  Google Scholar 

  302. Sanchez-Alvarez R, Martinez-Outschoorn UE, Lamb R, Hulit J, Howell A, Gandara R, Sartini M, Rubin E, Lisanti MP, Sotgia F. Mitochondrial dysfunction in breast cancer cells prevents tumor growth: understanding chemoprevention with metformin. Cell Cycle. 2013;12:172–82.

    CAS  PubMed  PubMed Central  Google Scholar 

  303. Chen K, Li Y, Guo Z, Zeng Y, Zhang W, Wang H. Metformin: current clinical applications in nondiabetic patients with cancer. Aging. 2020;12:3993–4009.

    CAS  PubMed  PubMed Central  Google Scholar 

  304. Chae YK, Arya A, Malecek MK, Shin DS, Carneiro B, Chandra S, Kaplan J, Kalyan A, Altman JK, Platanias L, Giles F. Repurposing metformin for cancer treatment: current clinical studies. Oncotarget. 2016;7:40767–80.

    PubMed  PubMed Central  Google Scholar 

  305. Pan J, Sun LC, Tao YF, Zhou Z, Du XL, Peng L, Feng X, Wang J, Li YP, Liu L, et al. ATP synthase ecto-alpha-subunit: a novel therapeutic target for breast cancer. J Transl Med. 2011;9:211.

    CAS  PubMed  PubMed Central  Google Scholar 

  306. Sainero-Alcolado L, Liano-Pons J, Ruiz-Perez MV, Arsenian-Henriksson M. Targeting mitochondrial metabolism for precision medicine in cancer. Cell Death Differ. 2022;29:1304–17.

    CAS  PubMed  PubMed Central  Google Scholar 

  307. Weinberg SE, Chandel NS. Targeting mitochondria metabolism for cancer therapy. Nat Chem Biol. 2015;11:9–15.

    CAS  PubMed  PubMed Central  Google Scholar 

  308. Dang L, White DW, Gross S, Bennett BD, Bittinger MA, Driggers EM, Fantin VR, Jang HG, Jin S, Keenan MC, et al. Cancer-associated IDH1 mutations produce 2-hydroxyglutarate. Nature. 2009;462:739–44.

    CAS  PubMed  PubMed Central  Google Scholar 

  309. Yang H, Ye D, Guan KL, Xiong Y. IDH1 and IDH2 mutations in tumorigenesis: mechanistic insights and clinical perspectives. Clin Cancer Res. 2012;18:5562–71.

    CAS  PubMed  PubMed Central  Google Scholar 

  310. Dang L, Yen K, Attar EC. IDH mutations in cancer and progress toward development of targeted therapeutics. Ann Oncol. 2016;27:599–608.

    CAS  PubMed  Google Scholar 

  311. Turkalp Z, Karamchandani J, Das S. IDH mutation in glioma: new insights and promises for the future. JAMA Neurol. 2014;71:1319–25.

    PubMed  Google Scholar 

  312. Han S, Liu Y, Cai SJ, Qian M, Ding J, Larion M, Gilbert MR, Yang C. IDH mutation in glioma: molecular mechanisms and potential therapeutic targets. Br J Cancer. 2020;122:1580–9.

    PubMed  PubMed Central  Google Scholar 

  313. Palmieri F. The mitochondrial transporter family SLC25: identification, properties and physiopathology. Mol Aspects Med. 2013;34:465–84.

    CAS  PubMed  Google Scholar 

  314. Palmieri F. The mitochondrial transporter family (SLC25): physiological and pathological implications. Pflugers Arch. 2004;447:689–709.

    CAS  PubMed  Google Scholar 

  315. Lauria G, Curcio R, Lunetti P, Tiziani S, Coppola V, Dolce V, Fiermonte G, Ahmed A. Role of mitochondrial transporters on metabolic rewiring of pancreatic adenocarcinoma: a comprehensive review. Cancers. 2023;15:411.

    CAS  PubMed  PubMed Central  Google Scholar 

  316. Fernandez HR, Gadre SM, Tan M, Graham GT, Mosaoa R, Ongkeko MS, Kim KA, Riggins RB, Parasido E, Petrini I, et al. The mitochondrial citrate carrier, SLC25A1, drives stemness and therapy resistance in non-small cell lung cancer. Cell Death Differ. 2018;25:1239–58.

    CAS  PubMed  PubMed Central  Google Scholar 

  317. Vozza A, Parisi G, De Leonardis F, Lasorsa FM, Castegna A, Amorese D, Marmo R, Calcagnile VM, Palmieri L, Ricquier D, et al. UCP2 transports C4 metabolites out of mitochondria, regulating glucose and glutamine oxidation. Proc Natl Acad Sci U S A. 2014;111:960–5.

    CAS  PubMed  PubMed Central  Google Scholar 

  318. Fleury C, Neverova M, Collins S, Raimbault S, Champigny O, Levi-Meyrueis C, Bouillaud F, Seldin MF, Surwit RS, Ricquier D, Warden CH. Uncoupling protein-2: a novel gene linked to obesity and hyperinsulinemia. Nat Genet. 1997;15:269–72.

    CAS  PubMed  Google Scholar 

  319. Raho S, Capobianco L, Malivindi R, Vozza A, Piazzolla C, De Leonardis F, Gorgoglione R, Scarcia P, Pezzuto F, Agrimi G, et al. KRAS-regulated glutamine metabolism requires UCP2-mediated aspartate transport to support pancreatic cancer growth. Nat Metab. 2020;2:1373–81.

    CAS  PubMed  Google Scholar 

  320. Dong J, Chen H. Cardiotoxicity of anticancer therapeutics. Front Cardiovasc Med. 2018;5:9.

    PubMed  PubMed Central  Google Scholar 

  321. McGowan JV, Chung R, Maulik A, Piotrowska I, Walker JM, Yellon DM. Anthracycline chemotherapy and cardiotoxicity. Cardiovasc Drugs Ther. 2017;31:63–75.

    CAS  PubMed  PubMed Central  Google Scholar 

  322. Bonadonna G, Monfardini S, De Lena M, Fossati-Bellani F. Clinical evaluation of adriamycin, a new antitumour antibiotic. Br Med J. 1969;3:503–6.

    CAS  PubMed  PubMed Central  Google Scholar 

  323. van Dyk JJ, van der Merwe AM, Falkson HC, Falkson G. Adriamycin in the treatment of cancer. S Afr Med J. 1976;50:61–6.

    PubMed  Google Scholar 

  324. Bonadonna G, Monfardini S, De Lena M, Fossati-Bellani F, Beretta G. Phase I and preliminary phase II evaluation of adriamycin (NSC 123127). Cancer Res. 1970;30:2572–82.

    CAS  PubMed  Google Scholar 

  325. Hortobagyi GN. Anthracyclines in the treatment of cancer: an overview. Drugs. 1997;54(Suppl 4):1–7.

    CAS  PubMed  Google Scholar 

  326. Rawat PS, Jaiswal A, Khurana A, Bhatti JS, Navik U. Doxorubicin-induced cardiotoxicity: an update on the molecular mechanism and novel therapeutic strategies for effective management. Biomed Pharmacother. 2021;139:111708.

    CAS  PubMed  Google Scholar 

  327. Swain SM, Whaley FS, Ewer MS. Congestive heart failure in patients treated with doxorubicin: a retrospective analysis of three trials. Cancer. 2003;97:2869–79.

    CAS  PubMed  Google Scholar 

  328. Gaudin PB, Hruban RH, Beschorner WE, Kasper EK, Olson JL, Baughman KL, Hutchins GM. Myocarditis associated with doxorubicin cardiotoxicity. Am J Clin Pathol. 1993;100:158–63.

    CAS  PubMed  Google Scholar 

  329. Kilickap S, Barista I, Akgul E, Aytemir K, Aksoy S, Tekuzman G. Early and late arrhythmogenic effects of doxorubicin. South Med J. 2007;100:262–5.

    PubMed  Google Scholar 

  330. Jordan JH, Castellino SM, Melendez GC, Klepin HD, Ellis LR, Lamar Z, Vasu S, Kitzman DW, Ntim WO, Brubaker PH, et al. Left ventricular mass change after anthracycline chemotherapy. Circ Heart Fail. 2018;11:e004560.

    CAS  PubMed  PubMed Central  Google Scholar 

  331. Pang B, Qiao X, Janssen L, Velds A, Groothuis T, Kerkhoven R, Nieuwland M, Ovaa H, Rottenberg S, van Tellingen O, et al. Drug-induced histone eviction from open chromatin contributes to the chemotherapeutic effects of doxorubicin. Nat Commun. 1908;2013:4.

    Google Scholar 

  332. Tarr M, van Helden PD. Inhibition of transcription by adriamycin is a consequence of the loss of negative superhelicity in DNA mediated by topoisomerase II. Mol Cell Biochem. 1990;93:141–6.

    CAS  PubMed  Google Scholar 

  333. de Jong S, Zijlstra JG, de Vries EG, Mulder NH. Reduced DNA topoisomerase II activity and drug-induced DNA cleavage activity in an adriamycin-resistant human small cell lung carcinoma cell line. Cancer Res. 1990;50:304–9.

    PubMed  Google Scholar 

  334. Buzdar AU, Marcus C, Smith TL, Blumenschein GR. Early and delayed clinical cardiotoxicity of doxorubicin. Cancer. 1985;55:2761–5.

    CAS  PubMed  Google Scholar 

  335. Villani F, Meazza R, Materazzo C. Non-invasive monitoring of cardiac hemodynamic parameters in doxorubicin-treated patients: comparison with echocardiography. Anticancer Res. 2006;26:797–801.

    CAS  PubMed  Google Scholar 

  336. Steinberg JS, Cohen AJ, Wasserman AG, Cohen P, Ross AM. Acute arrhythmogenicity of doxorubicin administration. Cancer. 1987;60:1213–8.

    CAS  PubMed  Google Scholar 

  337. Lefrak EA, Pitha J, Rosenheim S, Gottlieb JA. A clinicopathologic analysis of adriamycin cardiotoxicity. Cancer. 1973;32:302–14.

    CAS  PubMed  Google Scholar 

  338. Goormaghtigh E, Chatelain P, Caspers J, Ruysschaert JM. Evidence of a specific complex between adriamycin and negatively-charged phospholipids. Biochim Biophys Acta. 1980;597:1–14.

    CAS  PubMed  Google Scholar 

  339. Schlame M, Rua D, Greenberg ML. The biosynthesis and functional role of cardiolipin. Prog Lipid Res. 2000;39:257–88.

    CAS  PubMed  Google Scholar 

  340. Xu MF, Tang PL, Qian ZM, Ashraf M. Effects by doxorubicin on the myocardium are mediated by oxygen free radicals. Life Sci. 2001;68:889–901.

    CAS  PubMed  Google Scholar 

  341. Cunha-Oliveira T, Ferreira LL, Coelho AR, Deus CM, Oliveira PJ. Doxorubicin triggers bioenergetic failure and p53 activation in mouse stem cell-derived cardiomyocytes. Toxicol Appl Pharmacol. 2018;348:1–13.

    CAS  PubMed  Google Scholar 

  342. Carvalho RA, Sousa RP, Cadete VJ, Lopaschuk GD, Palmeira CM, Bjork JA, Wallace KB. Metabolic remodeling associated with subchronic doxorubicin cardiomyopathy. Toxicology. 2010;270:92–8.

    CAS  PubMed  Google Scholar 

  343. Vijay V, Moland CL, Han T, Fuscoe JC, Lee T, Herman EH, Jenkins GR, Lewis SM, Cummings CA, Gao Y, et al. Early transcriptional changes in cardiac mitochondria during chronic doxorubicin exposure and mitigation by dexrazoxane in mice. Toxicol Appl Pharmacol. 2016;295:68–84.

    CAS  PubMed  Google Scholar 

  344. Chen Y, Tang Y, Zhang YC, Huang XH, Xie YQ, Xiang Y. A metabolomic study of rats with doxorubicin-induced cardiomyopathy and Shengmai injection treatment. PLoS ONE. 2015;10:e0125209.

    PubMed  PubMed Central  Google Scholar 

  345. Ni C, Ma P, Wang R, Lou X, Liu X, Qin Y, Xue R, Blasig I, Erben U, Qin Z. Doxorubicin-induced cardiotoxicity involves IFNgamma-mediated metabolic reprogramming in cardiomyocytes. J Pathol. 2019;247:320–32.

    CAS  PubMed  Google Scholar 

  346. Herzig S, Shaw RJ. AMPK: guardian of metabolism and mitochondrial homeostasis. Nat Rev Mol Cell Biol. 2018;19:121–35.

    CAS  PubMed  Google Scholar 

  347. Yin J, Guo J, Zhang Q, Cui L, Zhang L, Zhang T, Zhao J, Li J, Middleton A, Carmichael PL, Peng S. Doxorubicin-induced mitophagy and mitochondrial damage is associated with dysregulation of the PINK1/parkin pathway. Toxicol In Vitro. 2018;51:1–10.

    CAS  PubMed  Google Scholar 

  348. Jirkovsky E, Popelova O, Krivakova-Stankova P, Vavrova A, Hroch M, Haskova P, Brcakova-Dolezelova E, Micuda S, Adamcova M, Simunek T, et al. Chronic anthracycline cardiotoxicity: molecular and functional analysis with focus on nuclear factor erythroid 2-related factor 2 and mitochondrial biogenesis pathways. J Pharmacol Exp Ther. 2012;343:468–78.

    CAS  PubMed  Google Scholar 

  349. Montaigne D, Marechal X, Preau S, Baccouch R, Modine T, Fayad G, Lancel S, Neviere R. Doxorubicin induces mitochondrial permeability transition and contractile dysfunction in the human myocardium. Mitochondrion. 2011;11:22–6.

    CAS  PubMed  Google Scholar 

  350. Nordgren KK, Wallace KB. Keap1 redox-dependent regulation of doxorubicin-induced oxidative stress response in cardiac myoblasts. Toxicol Appl Pharmacol. 2014;274:107–16.

    CAS  PubMed  Google Scholar 

  351. Morelli MB, Bongiovanni C, Da Pra S, Miano C, Sacchi F, Lauriola M, D’Uva G. Cardiotoxicity of anticancer drugs: molecular mechanisms and strategies for cardioprotection. Front Cardiovasc Med. 2022;9:847012.

    CAS  PubMed  PubMed Central  Google Scholar 

  352. Antonucci S, Di Sante M, Tonolo F, Pontarollo L, Scalcon V, Alanova P, Menabo R, Carpi A, Bindoli A, Rigobello MP, et al. The determining role of mitochondrial reactive oxygen species generation and monoamine oxidase activity in doxorubicin-induced cardiotoxicity. Antioxid Redox Signal. 2021;34:531–50.

    CAS  PubMed  PubMed Central  Google Scholar 

  353. Wallace KB, Sardao VA, Oliveira PJ. Mitochondrial determinants of doxorubicin-induced cardiomyopathy. Circ Res. 2020;126:926–41.

    CAS  PubMed  PubMed Central  Google Scholar 

  354. Zhang T, Zhang Y, Cui M, Jin L, Wang Y, Lv F, Liu Y, Zheng W, Shang H, Zhang J, et al. CaMKII is a RIP3 substrate mediating ischemia- and oxidative stress-induced myocardial necroptosis. Nat Med. 2016;22:175–82.

    PubMed  Google Scholar 

  355. Madeddu C, Deidda M, Piras A, Cadeddu C, Demurtas L, Puzzoni M, Piscopo G, Scartozzi M, Mercuro G. Pathophysiology of cardiotoxicity induced by nonanthracycline chemotherapy. J Cardiovasc Med. 2016;17(Suppl 1):e12–8.

    Google Scholar 

  356. Minerva, Bhat A, Verma S, Chander G, Jamwal RS, Sharma B, Bhat A, Katyal T, Kumar R, Shah R. Cisplatin-based combination therapy for cancer. J Cancer Res Ther. 2023;19:530–6.

    CAS  PubMed  Google Scholar 

  357. Ma H, Jones KR, Guo R, Xu P, Shen Y, Ren J. Cisplatin compromises myocardial contractile function and mitochondrial ultrastructure: role of endoplasmic reticulum stress. Clin Exp Pharmacol Physiol. 2010;37:460–5.

    CAS  PubMed  Google Scholar 

  358. Kamphuis JAM, Linschoten M, Cramer MJ, Gort EH, van Rhenen A, Asselbergs FW, Doevendans PA, Teske AJ. Cancer therapy-related cardiac dysfunction of nonanthracycline chemotherapeutics: what is the evidence? JACC CardioOncol. 2019;1:280–90.

    PubMed  PubMed Central  Google Scholar 

  359. Sayed-Ahmed MM, Aldelemy ML, Al-Shabanah OA, Hafez MM, Al-Hosaini KA, Al-Harbi NO, Al-Sharary SD, Al-Harbi MM. Inhibition of gene expression of carnitine palmitoyltransferase I and heart fatty acid binding protein in cyclophosphamide and ifosfamide-induced acute cardiotoxic rat models. Cardiovasc Toxicol. 2014;14:232–42.

    CAS  PubMed  Google Scholar 

  360. Iqubal A, Iqubal MK, Sharma S, Ansari MA, Najmi AK, Ali SM, Ali J, Haque SE. Molecular mechanism involved in cyclophosphamide-induced cardiotoxicity: old drug with a new vision. Life Sci. 2019;218:112–31.

    CAS  PubMed  Google Scholar 

  361. Ghafouri-Fard S, Abak A, Tondro Anamag F, Shoorei H, Fattahi F, Javadinia SA, Basiri A, Taheri M. 5-Fluorouracil: a narrative review on the role of regulatory mechanisms in driving resistance to this chemotherapeutic agent. Front Oncol. 2021;11:658636.

    CAS  PubMed  PubMed Central  Google Scholar 

  362. Sorrentino MF, Kim J, Foderaro AE, Truesdell AG. 5-fluorouracil induced cardiotoxicity: review of the literature. Cardiol J. 2012;19:453–8.

    PubMed  Google Scholar 

  363. Eskandari MR, Moghaddam F, Shahraki J, Pourahmad J. A comparison of cardiomyocyte cytotoxic mechanisms for 5-fluorouracil and its pro-drug capecitabine. Xenobiotica. 2015;45:79–87.

    CAS  PubMed  Google Scholar 

  364. Arellano M, Malet-Martino M, Martino R, Gires P. The anti-cancer drug 5-fluorouracil is metabolized by the isolated perfused rat liver and in rats into highly toxic fluoroacetate. Br J Cancer. 1998;77:79–86.

    CAS  PubMed  PubMed Central  Google Scholar 

  365. Rossato LG, Costa VM, Dallegrave E, Arbo M, Silva R, Ferreira R, Amado F, Dinis-Oliveira RJ, Duarte JA, de Lourdes BM, et al. Mitochondrial cumulative damage induced by mitoxantrone: late onset cardiac energetic impairment. Cardiovasc Toxicol. 2014;14:30–40.

    CAS  PubMed  Google Scholar 

  366. Doroshow JH. Mitomycin C-enhanced superoxide and hydrogen peroxide formation in rat heart. J Pharmacol Exp Ther. 1981;218:206–11.

    CAS  PubMed  Google Scholar 

  367. Shenkenberg TD, Von Hoff DD. Mitoxantrone: a new anticancer drug with significant clinical activity. Ann Intern Med. 1986;105:67–81.

    CAS  PubMed  Google Scholar 

  368. Douarre C, Sourbier C, Dalla Rosa I, Brata Das B, Redon CE, Zhang H, Neckers L, Pommier Y. Mitochondrial topoisomerase I is critical for mitochondrial integrity and cellular energy metabolism. PLoS ONE. 2012;7:e41094.

    CAS  PubMed  PubMed Central  Google Scholar 

  369. Senkus E, Jassem J. Cardiovascular effects of systemic cancer treatment. Cancer Treat Rev. 2011;37:300–11.

    CAS  PubMed  Google Scholar 

  370. Kitamura N, Sakamoto K, Ono T, Kimura J. The inhibitory effect of paclitaxel on (Kv2.1) K+ current in H9c2 cells. Fukushima J Med Sci. 2015;61:47–53.

    CAS  PubMed  PubMed Central  Google Scholar 

  371. Decorti G, Bartoli Klugmann F, Candussio L, Baldini L. Effect of paclitaxel and Cremophor EL on mast cell histamine secretion and their interaction with adriamycin. Anticancer Res. 1996;16:317–20.

    CAS  PubMed  Google Scholar 

  372. Herradon E, Gonzalez C, Gonzalez A, Uranga JA, Lopez-Miranda V. Cardiovascular toxicity induced by chronic vincristine treatment. Front Pharmacol. 2021;12:692970.

    CAS  PubMed  PubMed Central  Google Scholar 

  373. Rodriguez-Sinovas A, Ruiz-Meana M, Denuc A, Garcia-Dorado D. Mitochondrial Cx43, an important component of cardiac preconditioning. Biochim Biophys Acta Biomembr. 2018;1860:174–81.

    CAS  PubMed  Google Scholar 

  374. Royce TJ, Qureshi MM, Truong MT. Radiotherapy utilization and fractionation patterns during the first course of cancer treatment in the united states from 2004 to 2014. J Am Coll Radiol. 2018;15:1558–64.

    PubMed  Google Scholar 

  375. Overgaard J. Radiotherapy: gazing at the crystal ball of European radiotherapy. Nat Rev Clin Oncol. 2015;12:5–6.

    PubMed  Google Scholar 

  376. Koutroumpakis E, Palaskas NL, Lin SH, Abe JI, Liao Z, Banchs J, Deswal A, Yusuf SW. Modern radiotherapy and risk of cardiotoxicity. Chemotherapy. 2020;65:65–76.

    CAS  PubMed  Google Scholar 

  377. Burch GE, Sohal RS, Sun SC, Miller GC, Colcolough HL. Effects of radiation on the human heart: an electron microscopic study. Arch Intern Med. 1968;121:230–4.

    CAS  PubMed  Google Scholar 

  378. Wang L, Kuwahara Y, Li L, Baba T, Shin RW, Ohkubo Y, Ono K, Fukumoto M. Analysis of common deletion (CD) and a novel deletion of mitochondrial DNA induced by ionizing radiation. Int J Radiat Biol. 2007;83:433–42.

    CAS  PubMed  Google Scholar 

  379. Schilling-Toth B, Sandor N, Kis E, Kadhim M, Safrany G, Hegyesi H. Analysis of the common deletions in the mitochondrial DNA is a sensitive biomarker detecting direct and non-targeted cellular effects of low dose ionizing radiation. Mutat Res. 2011;716:33–9.

    CAS  PubMed  Google Scholar 

  380. Ljungman M, Hanawalt PC. Efficient protection against oxidative DNA damage in chromatin. Mol Carcinog. 1992;5:264–9.

    CAS  PubMed  Google Scholar 

  381. Barjaktarovic Z, Shyla A, Azimzadeh O, Schulz S, Haagen J, Dorr W, Sarioglu H, Atkinson MJ, Zischka H, Tapio S. Ionising radiation induces persistent alterations in the cardiac mitochondrial function of C57BL/6 mice 40 weeks after local heart exposure. Radiother Oncol. 2013;106:404–10.

    CAS  PubMed  Google Scholar 

  382. Azimzadeh O, Scherthan H, Sarioglu H, Barjaktarovic Z, Conrad M, Vogt A, Calzada-Wack J, Neff F, Aubele M, Buske C, et al. Rapid proteomic remodeling of cardiac tissue caused by total body ionizing radiation. Proteomics. 2011;11:3299–311.

    CAS  PubMed  Google Scholar 

  383. Chen ZI, Ai DI. Cardiotoxicity associated with targeted cancer therapies. Mol Clin Oncol. 2016;4:675–81.

    CAS  PubMed  PubMed Central  Google Scholar 

  384. McLean BA, Hansen R, Paterson DI, White JA, Oudit GY. Breast cancer patients receiving anthracycline chemotherapy and trastuzumab have biventricular dysfunction and reduced heart mass. J Am Coll Cardiol. 2018;72:1872–3.

    PubMed  Google Scholar 

  385. Onitilo AA, Engel JM, Stankowski RV. Cardiovascular toxicity associated with adjuvant trastuzumab therapy: prevalence, patient characteristics, and risk factors. Ther Adv Drug Saf. 2014;5:154–66.

    PubMed  PubMed Central  Google Scholar 

  386. Nemeth BT, Varga ZV, Wu WJ, Pacher P. Trastuzumab cardiotoxicity: from clinical trials to experimental studies. Br J Pharmacol. 2017;174:3727–48.

    CAS  PubMed  Google Scholar 

  387. Orphanos GS, Ioannidis GN, Ardavanis AG. Cardiotoxicity induced by tyrosine kinase inhibitors. Acta Oncol. 2009;48:964–70.

    CAS  PubMed  Google Scholar 

  388. Kerkela R, Woulfe KC, Durand JB, Vagnozzi R, Kramer D, Chu TF, Beahm C, Chen MH, Force T. Sunitinib-induced cardiotoxicity is mediated by off-target inhibition of AMP-activated protein kinase. Clin Transl Sci. 2009;2:15–25.

    CAS  PubMed  PubMed Central  Google Scholar 

  389. Kantarjian H, Sawyers C, Hochhaus A, Guilhot F, Schiffer C, Gambacorti-Passerini C, Niederwieser D, Resta D, Capdeville R, Zoellner U, et al. Hematologic and cytogenetic responses to imatinib mesylate in chronic myelogenous leukemia. N Engl J Med. 2002;346:645–52.

    CAS  PubMed  Google Scholar 

  390. Goldman JM, Melo JV. Targeting the BCR-ABL tyrosine kinase in chronic myeloid leukemia. N Engl J Med. 2001;344:1084–6.

    CAS  PubMed  Google Scholar 

  391. Trent JC, Patel SS, Zhang J, Araujo DM, Plana JC, Lenihan DJ, Fan D, Patel SR, Benjamin RS, Khakoo AY. Rare incidence of congestive heart failure in gastrointestinal stromal tumor and other sarcoma patients receiving imatinib mesylate. Cancer. 2010;116:184–92.

    CAS  PubMed  Google Scholar 

  392. Hu X, Chen Z, Wang Z, Xiao Q. Cancer evolution: special focus on the immune aspect of cancer. Semin Cancer Biol. 2022;86:420–35.

    PubMed  Google Scholar 

  393. Couzin-Frankel J. Breakthrough of the year 2013. Cancer Immunother Sci. 2013;342:1432–3.

    CAS  Google Scholar 

  394. Gilboa E. DC-based cancer vaccines. J Clin Invest. 2007;117:1195–203.

    CAS  PubMed  PubMed Central  Google Scholar 

  395. Grupp SA, Kalos M, Barrett D, Aplenc R, Porter DL, Rheingold SR, Teachey DT, Chew A, Hauck B, Wright JF, et al. Chimeric antigen receptor-modified T cells for acute lymphoid leukemia. N Engl J Med. 2013;368:1509–18.

    CAS  PubMed  PubMed Central  Google Scholar 

  396. Karwacz K, Bricogne C, MacDonald D, Arce F, Bennett CL, Collins M, Escors D. PD-L1 co-stimulation contributes to ligand-induced T cell receptor down-modulation on CD8+ T cells. EMBO Mol Med. 2011;3:581–92.

    CAS  PubMed  PubMed Central  Google Scholar 

  397. Shrikant P, Khoruts A, Mescher MF. CTLA-4 blockade reverses CD8+ T cell tolerance to tumor by a CD4+ T cell- and IL-2-dependent mechanism. Immunity. 1999;11:483–93.

    CAS  PubMed  Google Scholar 

  398. Heinzerling L, Ott PA, Hodi FS, Husain AN, Tajmir-Riahi A, Tawbi H, Pauschinger M, Gajewski TF, Lipson EJ, Luke JJ. Cardiotoxicity associated with CTLA4 and PD1 blocking immunotherapy. J Immunother Cancer. 2016;4:50.

    PubMed  PubMed Central  Google Scholar 

  399. Dong H, Qi Y, Kong X, Wang Z, Fang Y, Wang J. PD-1/PD-L1 inhibitor-associated myocarditis: epidemiology, characteristics, diagnosis, treatment, and potential mechanism. Front Pharmacol. 2022;13:835510.

    CAS  PubMed  PubMed Central  Google Scholar 

  400. Voskens CJ, Goldinger SM, Loquai C, Robert C, Kaehler KC, Berking C, Bergmann T, Bockmeyer CL, Eigentler T, Fluck M, et al. The price of tumor control: an analysis of rare side effects of anti-CTLA-4 therapy in metastatic melanoma from the ipilimumab network. PLoS ONE. 2013;8:e53745.

    CAS  PubMed  PubMed Central  Google Scholar 

  401. Grabie N, Lichtman AH, Padera R. T cell checkpoint regulators in the heart. Cardiovasc Res. 2019;115:869–77.

    CAS  PubMed  PubMed Central  Google Scholar 

  402. Suzuki S, Ishikawa N, Konoeda F, Seki N, Fukushima S, Takahashi K, Uhara H, Hasegawa Y, Inomata S, Otani Y, et al. Nivolumab-related myasthenia gravis with myositis and myocarditis in Japan. Neurology. 2017;89:1127–34.

    CAS  PubMed  Google Scholar 

  403. Tajiri K, Aonuma K, Sekine I. Immune checkpoint inhibitor-related myocarditis. Jpn J Clin Oncol. 2018;48:7–12.

    PubMed  Google Scholar 

  404. Seifert CF, Nesser ME, Thompson DF. Dexrazoxane in the prevention of doxorubicin-induced cardiotoxicity. Ann Pharmacother. 1994;28:1063–72.

    CAS  PubMed  Google Scholar 

  405. Nabati M, Janbabai G, Baghyari S, Esmaili K, Yazdani J. Cardioprotective effects of carvedilol in inhibiting doxorubicin-induced cardiotoxicity. J Cardiovasc Pharmacol. 2017;69:279–85.

    CAS  PubMed  Google Scholar 

  406. Matsui H, Morishima I, Numaguchi Y, Toki Y, Okumura K, Hayakawa T. Protective effects of carvedilol against doxorubicin-induced cardiomyopathy in rats. Life Sci. 1999;65:1265–74.

    CAS  PubMed  Google Scholar 

  407. Esfandbod M, Naderi M, Sadatnaseri A, Ahmadi A, Noroozi M, Sadeghi Joni S. Evaluation of the preventive effects of carvedilol on trastuzumab-induced cardiotoxicity in early-stage and locally advanced HER2-positive breast cancer patients. Int J Hematol Oncol Stem Cell Res. 2021;15:206–12.

    PubMed  PubMed Central  Google Scholar 

  408. Wang H, Yu P, Gou H, Zhang J, Zhu M, Wang ZH, Tian JW, Jiang YT, Fu FH. Cardioprotective effects of 20(S)-ginsenoside Rh2 against doxorubicin-induced cardiotoxicity in vitro and in vivo. Evid Based Complement Alternat Med. 2012;2012:506214.

    PubMed  PubMed Central  Google Scholar 

  409. Wang SQ, Han XZ, Li X, Ren DM, Wang XN, Lou HX. Flavonoids from Dracocephalum tanguticum and their cardioprotective effects against doxorubicin-induced toxicity in H9c2 cells. Bioorg Med Chem Lett. 2010;20:6411–5.

    CAS  PubMed  Google Scholar 

  410. Laird-Fick HS, Tokala H, Kandola S, Kehdi M, Pelosi A, Wang L, Grondahl B. Early morphological changes in cardiac mitochondria after subcutaneous administration of trastuzumab in rabbits: possible prevention with oral selenium supplementation. Cardiovasc Pathol. 2020;44:107159.

    PubMed  Google Scholar 

  411. Siveski-Iliskovic N, Hill M, Chow DA, Singal PK. Probucol protects against adriamycin cardiomyopathy without interfering with its antitumor effect. Circulation. 1995;91:10–5.

    CAS  PubMed  Google Scholar 

  412. Zhao D, Xue C, Li J, Feng K, Zeng P, Chen Y, Duan Y, Zhang S, Li X, Han J, Yang X. Adiponectin agonist ADP355 ameliorates doxorubicin-induced cardiotoxicity by decreasing cardiomyocyte apoptosis and oxidative stress. Biochem Biophys Res Commun. 2020;533:304–12.

    CAS  PubMed  Google Scholar 

  413. Das A, Durrant D, Mitchell C, Mayton E, Hoke NN, Salloum FN, Park MA, Qureshi I, Lee R, Dent P, Kukreja RC. Sildenafil increases chemotherapeutic efficacy of doxorubicin in prostate cancer and ameliorates cardiac dysfunction. Proc Natl Acad Sci USA. 2010;107:18202–7.

    CAS  PubMed  PubMed Central  Google Scholar 

  414. De Francesco EM, Rocca C, Scavello F, Amelio D, Pasqua T, Rigiracciolo DC, Scarpelli A, Avino S, Cirillo F, Amodio N, et al. Protective role of GPER agonist G-1 on cardiotoxicity induced by doxorubicin. J Cell Physiol. 2017;232:1640–9.

    PubMed  Google Scholar 

  415. Riad A, Bien S, Westermann D, Becher PM, Loya K, Landmesser U, Kroemer HK, Schultheiss HP, Tschope C. Pretreatment with statin attenuates the cardiotoxicity of Doxorubicin in mice. Cancer Res. 2009;69:695–9.

    CAS  PubMed  Google Scholar 

  416. Bjelogrlic SK, Radic J, Jovic V, Radulovic S. Activity of d, l-alpha-tocopherol (vitamin E) against cardiotoxicity induced by doxorubicin and doxorubicin with cyclophosphamide in mice. Basic Clin Pharmacol Toxicol. 2005;97:311–9.

    CAS  PubMed  Google Scholar 

  417. Danhier F, Kouhe TT, Duhem N, Ucakar B, Staub A, Draoui N, Feron O, Preat V. Vitamin E-based micelles enhance the anticancer activity of doxorubicin. Int J Pharm. 2014;476:9–15.

    CAS  PubMed  Google Scholar 

  418. Moustafa I, Saka S, Viljoen M, Oosthuizen F. Vitamin E and levocarnitine as prophylaxis against doxorubicin-induced cardio toxicity in the adult cancer patient: a review. J Oncol Pharm Pract. 2022;28:1388–99.

    CAS  PubMed  Google Scholar 

  419. Beak J, Huang W, Parker JS, Hicks ST, Patterson C, Simpson PC, Ma A, Jin J, Jensen BC. An oral selective alpha-1A adrenergic receptor agonist prevents doxorubicin cardiotoxicity. JACC Basic Transl Sci. 2017;2:39–53.

    PubMed  PubMed Central  Google Scholar 

  420. O’Brien CG, Ozen MO, Ikeda G, Vaskova E, Jung JH, Bayardo N, Santoso MR, Shi L, Wahlquist C, Jiang Z, et al. Mitochondria-rich extracellular vesicles rescue patient-specific cardiomyocytes from doxorubicin injury: insights into the SENECA trial. JACC CardioOncol. 2021;3:428–40.

    PubMed  PubMed Central  Google Scholar 

  421. Pituskin E, Mackey JR, Koshman S, Jassal D, Pitz M, Haykowsky MJ, Pagano JJ, Chow K, Thompson RB, Vos LJ, et al. Multidisciplinary approach to novel therapies in cardio-oncology research (MANTICORE 101-Breast): a randomized trial for the prevention of trastuzumab-associated cardiotoxicity. J Clin Oncol. 2017;35:870–7.

    CAS  PubMed  Google Scholar 

  422. Shirmard LR, Shabani M, Moghadam AA, Zamani N, Ghanbari H, Salimi A. Protective effect of curcumin, chrysin and thymoquinone injection on trastuzumab-induced cardiotoxicity via mitochondrial protection. Cardiovasc Toxicol. 2022;22:663–75.

    CAS  PubMed  Google Scholar 

  423. Denuc A, Nunez E, Calvo E, Loureiro M, Miro-Casas E, Guaras A, Vazquez J, Garcia-Dorado D. New protein-protein interactions of mitochondrial connexin 43 in mouse heart. J Cell Mol Med. 2016;20:794–803.

    CAS  PubMed  PubMed Central  Google Scholar 

  424. Pecoraro M, Pinto A, Popolo A. Trastuzumab-induced cardiotoxicity and role of mitochondrial connexin43 in the adaptive response. Toxicol In Vitro. 2020;67:104926.

    CAS  PubMed  Google Scholar 

  425. Ajzashokouhi AH, Bostan HB, Jomezadeh V, Hayes AW, Karimi G. A review on the cardioprotective mechanisms of metformin against doxorubicin. Hum Exp Toxicol. 2020;39:237–48.

    CAS  PubMed  Google Scholar 

  426. Yu JM, Hsieh MC, Qin L, Zhang J, Wu SY. Metformin reduces radiation-induced cardiac toxicity risk in patients having breast cancer. Am J Cancer Res. 2019;9:1017–26.

    PubMed  PubMed Central  Google Scholar 

  427. Schlaak RA, Frei A, SenthilKumar G, Tsaih SW, Wells C, Mishra J, Flister MJ, Camara AKS, Bergom C. Differences in expression of mitochondrial complexes due to genetic variants may alter sensitivity to radiation-induced cardiac dysfunction. Front Cardiovasc Med. 2020;7:23.

    CAS  PubMed  PubMed Central  Google Scholar 

  428. Franco A, Sorriento D, Gambardella J, Pacelli R, Prevete N, Procaccini C, Matarese G, Trimarco B, Iaccarino G, Ciccarelli M. GRK2 moderates the acute mitochondrial damage to ionizing radiation exposure by promoting mitochondrial fission/fusion. Cell Death Discov. 2018;4:25.

    PubMed  PubMed Central  Google Scholar 

  429. Sorriento D, Gambardella J, Fiordelisi A, Iaccarino G, Illario M. GRKs and beta-Arrestins: “Gatekeepers” of mitochondrial function in the failing heart. Front Pharmacol. 2019;10:64.

    CAS  PubMed  PubMed Central  Google Scholar 

  430. Leiva O, AbdelHameid D, Connors JM, Cannon CP, Bhatt DL. Common pathophysiology in cancer, atrial fibrillation, atherosclerosis, and thrombosis: JACC: cardiooncology state-of-the-art review. JACC CardioOncol. 2021;3:619–34.

    PubMed  PubMed Central  Google Scholar 

  431. Banfill K, Giuliani M, Aznar M, Franks K, McWilliam A, Schmitt M, Sun F, Vozenin MC, Faivre Finn C, Committee IART. Cardiac toxicity of thoracic radiotherapy: existing evidence and future directions. J Thorac Oncol. 2021;16:216–27.

    CAS  PubMed  PubMed Central  Google Scholar 

  432. Bonora M, Wieckowski MR, Sinclair DA, Kroemer G, Pinton P, Galluzzi L. Targeting mitochondria for cardiovascular disorders: therapeutic potential and obstacles. Nat Rev Cardiol. 2019;16:33–55.

    CAS  PubMed  PubMed Central  Google Scholar 

  433. Conrad M, Angeli JP, Vandenabeele P, Stockwell BR. Regulated necrosis: disease relevance and therapeutic opportunities. Nat Rev Drug Discov. 2016;15:348–66.

    CAS  PubMed  PubMed Central  Google Scholar 

  434. Berra-Romani R, Brunetti V, Pellavio G, Soda T, Laforenza U, Scarpellino G, Moccia F. Allyl isothiocianate induces Ca(2+) signals and nitric oxide release by inducing reactive oxygen species production in the human cerebrovascular endothelial cell line hCMEC/D3. Cells. 2023;12:1732.

    CAS  PubMed  PubMed Central  Google Scholar 

  435. Kelso GF, Porteous CM, Coulter CV, Hughes G, Porteous WK, Ledgerwood EC, Smith RA, Murphy MP. Selective targeting of a redox-active ubiquinone to mitochondria within cells: antioxidant and antiapoptotic properties. J Biol Chem. 2001;276:4588–96.

    CAS  PubMed  Google Scholar 

  436. Wang T, Ma F, Qian HL. Defueling the cancer: ATP synthase as an emerging target in cancer therapy. Mol Ther Oncolytics. 2021;23:82–95.

    CAS  PubMed  PubMed Central  Google Scholar 

  437. Lamb R, Fiorillo M, Chadwick A, Ozsvari B, Reeves KJ, Smith DL, Clarke RB, Howell SJ, Cappello AR, Martinez-Outschoorn UE, et al. Doxycycline down-regulates DNA-PK and radiosensitizes tumor initiating cells: Implications for more effective radiation therapy. Oncotarget. 2015;6:14005–25.

    PubMed  PubMed Central  Google Scholar 

  438. Scatena C, Roncella M, Di Paolo A, Aretini P, Menicagli M, Fanelli G, Marini C, Mazzanti CM, Ghilli M, Sotgia F, et al. Doxycycline, an inhibitor of mitochondrial biogenesis, effectively reduces cancer stem cells (CSCs) in early breast cancer patients: a clinical pilot study. Front Oncol. 2018;8:452.

    PubMed  PubMed Central  Google Scholar 

  439. Luo M, Zhou W, Patel H, Srivastava AP, Symersky J, Bonar MM, Faraldo-Gomez JD, Liao M, Mueller DM. Bedaquiline inhibits the yeast and human mitochondrial ATP synthases. Commun Biol. 2020;3:452.

    CAS  PubMed  PubMed Central  Google Scholar 

  440. Hu J, Jing H, Lin H. Sirtuin inhibitors as anticancer agents. Future Med Chem. 2014;6:945–66.

    CAS  PubMed  Google Scholar 

  441. George J, Nihal M, Singh CK, Ahmad N. 4ʹ-Bromo-resveratrol, a dual Sirtuin-1 and Sirtuin-3 inhibitor, inhibits melanoma cell growth through mitochondrial metabolic reprogramming. Mol Carcinog. 2019;58:1876–85.

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We acknowledge co-funding from Next Generation EU, in the context of the National Recovery and Resilience Plan, Investment PE8 – Project Age-It: “Ageing Well in an Ageing Society”. This resource was co-financed by the Next Generation EU [DM 1557 11.10.2022]. The views and opinions expressed are only those of the authors and do not necessarily reflect those of the European Union or the European Commission. Neither the European Union nor the European Commission can be held responsible for them.

Funding

This research was supported by grants from (i) the Italian Association for Cancer Research (IG24449) and the Italian Ministry of Health (GR-2016-02361523) to N.A.; (ii) “SI.F.I.PA.CRO.DE.—Sviluppo e industrializzazione farmaci innovativi per terapia molecolare personalizzata PA.CRO.DE, (PON ARS01_00568) to T.A.; (iii) CALabria HUB per Ricerca Innovativa ed Avanzata (CAL.HUB.RIA), (T4-AN-09), Piano Operativo Salute—Traiettoria 4 “Biotecnologie, Bioinformatica e Sviluppo Farmaceutico” to C.R. and T.A.; (iv) the PNRR M4C2-Investimento 1.4-CN00000041 finanziato dall’Unione europea—Next Generation EU, to G.V. and N.A ; (v) the Italian Association for Cancer Research under Start-Up 2018-ID. 21651- P.I. E.M.D.F.

Author information

Authors and Affiliations

Authors

Contributions

Design, organization and conceptualization: CR, TS, TA and NA; writing—original draft preparation: CR, TS, EMDF, MF, FM, GV, TA and NA; writing—review and editing: CR, TS, FM, GV, TA, and NA. All authors discussed and commented on the manuscript, provided critical feedback, helped shape the final version of the manuscript, and approved the submitted version.

Corresponding authors

Correspondence to Tommaso Angelone or Nicola Amodio.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare no competing or financial interests.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Rocca, C., Soda, T., De Francesco, E.M. et al. Mitochondrial dysfunction at the crossroad of cardiovascular diseases and cancer. J Transl Med 21, 635 (2023). https://doi.org/10.1186/s12967-023-04498-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12967-023-04498-5

Keywords