Next Article in Journal
Potential Neuroprotective Role of Calretinin-N18 and Calbindin-D28k in the Retina of Adult Zebrafish Exposed to Different Wavelength Lights
Next Article in Special Issue
The Non-Invasive Assessment of Circulating D-Loop and mt-ccf Levels Opens an Intriguing Spyhole into Novel Approaches for the Tricky Diagnosis of NASH
Previous Article in Journal
α-Crystallin Domains of Five Human Small Heat Shock Proteins (sHsps) Differ in Dimer Stabilities and Ability to Incorporate Themselves into Oligomers of Full-Length sHsps
Previous Article in Special Issue
Iron- and Neuromelanin-Weighted Neuroimaging to Study Mitochondrial Dysfunction in Patients with Parkinson’s Disease
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Mitochondrial Dysfunction: The Hidden Player in the Pathogenesis of Atherosclerosis?

by
Giovanni Ciccarelli
1,2,*,
Stefano Conte
3,
Giovanni Cimmino
1,4,
Patrizia Maiorano
2,5,
Andrea Morrione
2 and
Antonio Giordano
2,5
1
Vanvitelli Cardiology Unit, Monaldi Hospital, 80131 Naples, Italy
2
Sbarro Institute for Cancer Research and Molecular Medicine, Center of Biotechnology, College of Science and Technology, Temple University, Philadelphia, PA 19122, USA
3
Department of Translational Medical Sciences, Section of Lung Diseases, University of Campania “Luigi Vanvitelli”, 80131 Naples, Italy
4
Department of Translational Medical Sciences, Section of Cardiology, University of Campania “Luigi Vanvitelli”, 80131 Naples, Italy
5
Department of Medical Biotechnologies, University of Siena, 53100 Siena, Italy
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(2), 1086; https://doi.org/10.3390/ijms24021086
Submission received: 15 November 2022 / Revised: 29 December 2022 / Accepted: 4 January 2023 / Published: 6 January 2023

Abstract

:
Atherosclerosis is a multifactorial inflammatory pathology that involves metabolic processes. Improvements in therapy have drastically reduced the prognosis of cardiovascular disease. Nevertheless, a significant residual risk is still relevant, and is related to unmet therapeutic targets. Endothelial dysfunction and lipid infiltration are the primary causes of atherosclerotic plaque progression. In this contest, mitochondrial dysfunction can affect arterial wall cells, in particular macrophages, smooth muscle cells, lymphocytes, and endothelial cells, causing an increase in reactive oxygen species (ROS), leading to oxidative stress, chronic inflammation, and intracellular lipid deposition. The detection and characterization of mitochondrial DNA (mtDNA) is crucial for assessing mitochondrial defects and should be considered the goal for new future therapeutic interventions. In this review, we will focus on a new idea, based on the analysis of data from many research groups, namely the link between mitochondrial impairment and endothelial dysfunction and, in particular, its effect on atherosclerosis and aging. Therefore, we discuss known and novel mitochondria-targeting therapies in the contest of atherosclerosis.

1. Introduction

Despite tremendous progresses in therapy in the last few years, cardiovascular diseases are still the leading cause of death worldwide [1,2]. To date, 18 million people die each year due to cardiovascular diseases [1,3,4,5,6,7]. These numbers are expected to increase, reaching 24 million yearly deaths worldwide from cardiovascular-related disease by 2030, with an average of over 66,000 per day, and a total global cost rising to over 1 trillion USD [1,3,4,5,6,7]. After a decline in mortality in recent decades, the numbers are rising again, reversing years of progress in both ischemic heart and cerebrovascular disease [1,4]. Thus, cardiovascular prevention is essential to make a 25% reduction in premature mortality from non-communicable diseases more realistic. There is an urgent need for plans and policies to reduce the burden of cardiovascular diseases which are a health challenge, starting from the territory and the role of cities in promoting health. There is also a need for improving our understanding of the basic mechanisms underlying cardiovascular diseases to design tailored therapy.
The atherosclerotic process is very often responsible for several cardiovascular and cerebrovascular diseases [8]. It is now well accepted and documented that atherosclerosis starts from endothelial dysfunction and lipid deposition, which progresses through macrophage infiltration, smooth muscle cell migration, and blood borne material deposition and becomes clinically relevant due to complications, eventually leading to local intravascular thrombus formation [8,9]. Modified lipoproteins, mainly oxidized low-density lipoproteins (oxLDL), are considered the major contributors to the genesis, progression, and immunological response occurring during the atherosclerotic process [10,11]. The first report linking mitochondria to atherosclerosis is from 1970 [12]. However, only in the last few years has increasing evidence really underlined the key role of mitochondrial dynamics in the pathogenesis of atherosclerosis [13]. Vascular cells, such as endothelial and smooth muscle cells, due to their metabolic functions and their barrier role are the main targets of mitochondrial dysfunction. In the atherosclerosis process, dysfunctional mitochondria might cause alterations in cellular metabolism and respiration resulting in the excessive production of reactive oxygen species (ROS), leading to oxidative stress [13]. While low levels of ROS exert important signaling functions [14,15], elevated ROS production induces the damage of cellular structures, alters DNA, proteins, and other molecules [16]. These conditions can become chronic, thereby favoring atherosclerosis progression and destabilization [17]. It is important to point out that mitochondrial dysfunction can be inherited and acquired [18]. Primary mitochondrial disease (PMD) is clinically diagnosed and confirmed by pathogenic mitochondrial DNA (mtDNA) or nuclear DNA (nDNA) mutations [18]. However, there are some disorders with a ‘mitochondrial’ phenotype without identifiable mtDNA or nDNA mutations, or with variants of unknown clinical significance [18]. Conversely, secondary mitochondrial dysfunction (SMD) can be caused by genes encoding neither function nor production of oxphos proteins, and associated with many hereditary non-mitochondrial diseases [18]. This SMD can also be related to nongenetic causes, such as environmental factors and ageing. The strong relationship between ageing, atherosclerosis, and cardiovascular diseases is well established [19], and a correlation between mitochondrial dysfunction–ageing, and vice-versa, cannot be ruled out [20]. In this review, we will particularly focus on the relationship between mitochondrial function and atherosclerosis, starting from its molecular basis to clinical perspectives and therapeutic options.

2. Mechanism of Mitochondrial Dysfunction and Molecular Implication for the Cardiovascular System

Mitochondria represent the engine of human cells generating ATP, and their function is regulated via mitophagy. However, mitochondria are also involved in the production of metabolites for protein assembling and signal transduction. Their metabolism is influenced by Ca2+ levels: low calcium concentrations can lead to mitochondrial dysfunction, while high levels of calcium can increase mitochondrial permeability [21].
As compared to other organs, the heart needs a high amount of energy; this is produced by oxidative metabolism in the mitochondria, which are one-third of the total volume of cardiomyocytes [22].
Mitochondrial dysfunctions are strongly related to cardiovascular diseases, in particular ischemic heart disease and atherosclerosis, cardiomyopathy, and hypertension [23]. This altered mitochondrial function results in reduced ATP, impairment of mitochondrial regulatory roles, reactive oxygen species (ROS) production and related signaling, cell growth and apoptosis, impaired mitochondrial electron transport chain activity, and an inflammatory response [24,25,26,27].
Mitochondrial fission and fusion impairments have been related to the development of cardiovascular diseases [28]. Novel mitochondrial biomarkers have been proposed to identify patients that can benefit from therapies specifically targeting mitochondria [29].
Mitochondrial homeostasis plays a crucial role in cells with high energy consumption, and in the pathological development of ischemic heart disease [30]. In particular, variation in the level of Ca2+ in mitochondria could lead to altered contraction coupling [31] and increased ROS levels during ischemic myocardial reperfusion, thereby inducing mitochondrial membrane damage, abnormal ATP synthesis, increased levels of Ca2+, and mPTP opening [32].
Thus, during ischemia, the mitochondrial succinate significantly increases in injured tissue [33,34], while during reperfusion, the accumulated metabolites are oxidized by the mitochondrial respiratory chain enzyme SDH (succinate dehydrogenase). This drives reactive oxygen species (ROS) production by reverse electron transport (RET) at mitochondrial complex I, stimulating mitochondrial permeability transition pore opening, and cell death associated with ischemia/reperfusion damage [35,36,37,38]. Sazanova et al. have demonstrated that specific mutations in the coding region of the mitochondrial genome, in particular m.14459G>A and m.5178C>A, are risk factors for occurrence and development of cardiac angina, while the mutation m.15059G>A instead had a protective effect [39].
Atherosclerosis is a chronic inflammatory condition characterized by impaired lipid metabolism, which stimulates innate and adaptive immune responses [9]. The presence of lipids in the intimal layer of arteries significantly affects mitochondrial activity/function. There are several mechanisms related to atherosclerosis progression and plaque instability, such as mutations in mitochondrial DNA (mtDNA), reactive oxidative species (ROS), and respiratory chain alterations, which induce hypertrophy of vascular smooth myocytes [40]. Mitochondrial ROS, associated with several risk factors, such as hyperglycemia and hypercholesterolemia [41], are involved in the expansion and potential damage of atherosclerotic plaques, particularly by inducing endothelial dysfunction, promoting monocyte infiltration, and increasing vascular smooth muscle cell and endothelial cell apoptosis [42].
Accumulation of ROS in the inner mitochondrial membrane can alter the mitochondrial cholesterol transporter, steroidogenic acute regulatory protein (StAR), which inhibits the storage of cholesterol through the mitochondrial membrane [43].
Under pathophysiological conditions, the dysfunctional mitochondria can stimulate the production of large amounts of ROS from adjacent mitochondria, modifying membrane potential, a process well known as “ROS-induced ROS”. Normally, this process of ROS is balanced by endogenous antioxidants. On the other hand, when there is an oxidative stress and hyper-production of ROS, the oxidative impairment in the arterial wall is increased [44], and a progression of the atherosclerotic plaque can be activated [45].
Mutations in mtDNA, related to ROS derived from mitochondrial dysfunction, are more frequent than in nuclear DNA and can cause impaired mitochondrial respiration in smooth muscle monocytes and macrophages, thereby contributing to atherosclerotic plaque progression [41]. In vitro studies have demonstrated that the damage induced byROS on mtDNA inhibits protein synthesis and alters gene expression, suggesting that ROS might mediate vascular cell dysfunction in the setting of atherogenesis [46] associated with changes in mitochondrial genome [47].
On the other hand, recent studies have put into question the correlation between ROS and mtDNA damage, highlighting that impaired mtDNA was observed in vascular and circulating cells before the development of atherosclerotic lesions [46]. Thus, these data suggested that the primary event is associated with mtDNA damage which increases ROS generation and variations in mitochondrial dysfunction and membrane potential, followed by activation of the apoptotic mechanism. Moreover, endogenous mitochondrial damage-associated molecular patterns (mtDAMPs) trigger sterile inflammation through various signaling pathways, including toll-like receptors (TLRs), nuclear factor kappa beta (NF-κβ), and the NOD-, LRR-, and pyrin domain-containing protein 3 (NLRP3) inflammasome [48]. In addition, damaged mtDNA can be seen as an endogenous damage-associated molecular pattern (DAMP), activating the inflammatory response [49].
Oxidized LDL (oxLDL) is involved in apoptosis through calcium-dependent mitochondrial pathways, mediated by cytochrome C and apoptosis-inducing factors [50]. In addition, oxLDL induces mitochondrial DNA (mtDNA) damage in atherosclerotic plaques, resulting in reduced aerobic respiration in the cap and core regions, thus decreasing the expression of complex I in vascular smooth muscle cells [45], which promotes the development of the necrotic core and the thickening of the fibrous cap in atherosclerotic plaques. Thus, oxLDL inhibits mitochondrial enzymes and modifies mitochondrial membrane potential causing production of MitoROS associated with endothelial apoptosis and atherosclerotic plaque progression [51,52]. Notably, damage to mitochondria of inflammatory cells, as macrophages or monocytes, induces plaque progression and complication, such as erosion or rupture [13].

3. Mitochondrial Dysfunction and Endothelial Dysfunction

Endothelial cells represent the barrier for the molecular transport between blood and tissue, regulating the exchange of nutrients [53] and the vascular tone by production of nitric oxide (NO) and contractile factors. Nevertheless, in addition to its function as a barrier, the endothelium modulates inflammation, lipid metabolism, vascular tone, and hemostasis, and prevents penetration of inflammatory cells into the bloodstream. It also secretes adhesion molecules and cytokines involved in the inflammatory cascade [54,55].
Endothelial cells’ dysfunction is characterized by ROS production, in particular from mitochondria, infiltrations of LDL, and subsequent oxidation to oxLDL [56], which might induce endothelial impairment. This enhances the expression of adhesion molecules ICAM-1, VCAM-1, and P-selectins, resulting in smooth muscle cell growth and activation of inflammatory cells, such as monocytes and macrophages, thereby causing a pro-inflammatory burden by releasing cytokines and increasing circulating immune cells’ adhesion to the endothelium [57].
Damaged endothelium produces growth factors, which activates smooth muscle cells (SMCs) in the vascular bed and the phagocytosis of lipids, with subsequent evolution into the fibrous cap [58].
Mitochondria are important regulators of apoptosis and NO production in vascular endothelial cells, stimulating the cellular response to stress [59]. At the same time, the amounts of mitochondria in the endothelium are not abundant, but despite that, mitochondria not only affect energy supply, but also regulate blood oxygen levels and mediate NO-mediated vasodilation [60,61]. The decreased NO production in atherosclerotic plaques is mainly due to endothelial NO synthase (eNOS) degradation, induced by ROS-mediated oxidative stress. In addition, eNOS decoupling dysfunction produces additional ROS, which interfere with endothelial function, inducing the progression of atherosclerosis [62].
Recent studies have highlighted that the transcription factor NF-E2 p45-related factor 2 (Nrf2; gene name NFE2L2) regulates some mitochondrial functions [63] via the mtROS pathway, such as antioxidant activity, autophagy, and metabolism [64]. Nrf2 has a crucial role in the maintenance of cellular redox homeostasis by regulating biosynthesis, utilization, and regeneration of glutathione, thioredoxin, and NADPH, and controlling the production of ROS by mitochondria and NADPH oxidase. Moreover, Nrf2 activation inhibits Drp1-mediated mitochondrial fission, improving endothelial dysfunction [65].
Incubation of endothelial cells with oxLDL leads to an increase in the activity of mitochondrial complex I and oxidative stress [66], thereby stimulating the transcription and expression of superoxide dismutase 2 (SOD2) in macrophages [67].
Activation of mitophagy, a lysosome-mediated selective mitochondria degradation [68], is one of the characteristic signs of mitochondrial dysfunction and damage and leads to membrane potential collapse, increased ROS production, and decreased ATP levels, oxidative stress, and apoptosis [69]. The balance between mitophagy and mitochondrial neogenesis is crucial for correct homeostasis in the human body [70]. Mitochondrial biogenesis is an important process for maintaining energy control and protecting endothelial cells’ survival in critical pathological scenarios [71,72]. Mitochondrial dysfunction might represent the first real step of atherosclerosis by determining the endothelial impairment, which is the starting point of the atherosclerotic process [21].
The endothelial dysfunction, in particular decreased vasodilatation, is also typical of MELAS (Mitochondrial myopathy, encephalopathy, lactic acidosis, and stroke-like episodes) patients [73].
These patients can have lower levels of l-arginine, which regulates the endothelial-dependent vascular relaxation [74], which was significantly lowered in both acute and interictal phases of MELAS as compared to control subjects [75].
A schematic view of the atherosclerotic mitochondria-related process is shown in Figure 1.

4. Clinical Perspectives and Therapeutic Implications: Something New?

In the last twenty years, several efforts have been made to evaluate pharmacological strategies to modulate mitochondrial function. Many preclinical and clinical studies have been designed based on the available literature, which support the notion that mitochondria might be crucial for the treatment of cardiovascular diseases. However, no specific studies are currently available on atherosclerosis. Starting from the concept that during cardiovascular stress mitochondrial oxidative function is enhanced, a primary therapeutic approach might be based on the reduction of this activation [23,76]. The first report targeting mitochondrial function with antioxidants (i.e., MitoQ) has reduced doxorubicin-induced cardiomyopathy in an animal model [77], and increased endothelial performance in the elderly [78]. Upcoming clinical trials (NCT03586414) will reveal the effect of MitoQ supplementation as a therapeutic modality for improving health outcomes in cardiac patients [79]. An alternative strategy to recover mitochondrial function is nicotinamide adenine dinucleotide (NAD+) supplementation [79]. It has been reported that both nicotinamide riboside and nicotinamide mononucleotide may improve cardiac function [80] and reduce stroke occurrence in several preclinical models [81]. Potential benefits of NAD+ supplementation have also been reported in heart failure patients [82]. Other general approaches might have the potential to preserve mitochondrial activity. First, it has been reported that resveratrol and curcumin may improve mitochondrial function in experimental models of heart diseases [83]. Moreover, some lifestyle interventions, such as caloric restriction and exercise, contribute to rescuing mitochondrial dysfunction [79,84]. A more specific modulation with elamipretide is under clinical investigation [85]. This is a cell-permeable small peptide SS-31 which improves homeostatic mitochondrial activities and cardiac function [85]. Despite the failure in limiting restenosis in the clinical contexts of coronary and renal angioplasty [86,87], promising results have been reported in an experimental model of heart failure [88], and it is actually under evaluation in clinical trials enrolling patients with heart failure [89].
Other specific agents targeting specific mitochondrial activities (such as mitophagy and/or autophagy, mitochondrial calcium uniporter protein, Ca2+-responsive enzymes, damage-associated molecular patterns, miRNA-targeting strategies) might have potential as direct or indirect therapeutical modulators of mitochondrial functions [85]. However, some of these new compounds have multiple targets, and thus it becomes difficult to establish the appropriate dose to specifically target mitochondrial dynamics [85]. Moreover, based on the available data, modulation of mitochondrial dynamics might be beneficial only in the short term, since detrimental cardiac effects may occur in the long term [85]. Thus, modulators of mitochondrial dynamics have not yet been tested on patients affected by cardiovascular diseases.
A new therapeutic perspective is characterized by neutralization of mitochondria by ROS, targeting ROS scavengers to the site of action, or by induction of mitochondrial biogenesis. Recently, new mitochondria targeted-antioxidants, preferentially delivered to the mitochondrial compartment to scavenge mitochondria-derived ROS, have been proposed to increase antioxidant levels specifically in mitochondria [90,91]. In particular, Karnewar et al. [92] have demonstrated that an an alkyl TPP + -tagged esculetin (mitochondria-targeted esculetin or Mito-Esc) reduces atherosclerosis by delaying vascular senescence and pro-inflammatory processes, and by increasing mitochondrial biogenesis. Furthermore, in humans, it has been demonstrated that pioglitazone, an anti-diabetic drug, stimulates mitochondrial biogenesis in the subcutaneous fat of patients with diabetes mellitus, a strong risk factor for atherosclerotic disease, contributing to the lipid-lowering effect of this drug [93].
Recently, Dorighello et al. [94] showed that reducing the endogenous production of mitochondria-related oxidants can counteract atherosclerosis progression, inducing a mild mitochondrial uncoupling and thus decreasing oxidant generation, resulting in a decrease in development within the hypercholesterolemic LDLr-/- model. Additionally, treatment with low doses of DNP (2,4-dinitrophenol) did not alter classical risk factors for atherosclerosis in the LDLr-/- mice, reinforcing that mitochondrial ROS in the arterial wall have a crucial pro-atherogenic role.

5. Current Gaps and Future Research

In view of the multiple mechanisms described above, the development of a drug that has the pharmacodynamic and pharmacokinetic characteristics that allow it to modulate mitochondrial function, with proven clinical efficacy, appears still far away.
The current knowledge of the mitochondrial mechanisms underlying cardiovascular diseases is still limited and fragmented; therefore, basic and clinical research have the primary goal to better define the pathways involved, in order to identify a more precise therapeutic target.
However, several issues should be taken into account. The first problem is mitochondrion specificity. Previously published research on antioxidants clearly indicates that in the absence of specificity, these agents do not act on mitochondrial activity only and inhibit reactive oxygen species (ROS) generated from multiple sources, with controversial clinical benefits. Improved pharmacological strategies now offer the possibility of targeting antioxidants to the mitochondria using cationic molecules which have the ability to accumulate spontaneously in the mitochondrial matrix [95,96]. However, the dysfunctional mitochondrion could present transmembrane potential abnormalities, thus altering the mitochondrial capacity to accumulate the drug delivered by cation-based strategies [97,98,99].
The second problem is the pharmacokinetic activity and biodistribution of the potential drug. To date, a tissue-specific strategy is missing. With systemic administration, any medication inevitably interacts with mitochondria, not only of the cardiovascular system, thus limiting the bioavailability, at the doses deemed safe, of some MTAs in the sites of interest [85].
It is known that several mitochondrial proteins are necessary for embryonic development and/or adult survival, especially for their bioenergetic functions. An example of this essential role is somatic cytochrome c (CYCS), which plays the role of electronic shuttle of the respiratory chain [100] and whose absence leads to in uterus death [101]. Unfortunately, many mitochondrial proteins exist in multiple isoforms and with multiple functions [102]. Genetic and functional redundancy complicates the generation of suitable experimental models [85].
Similarly, it is difficult to monitor mitochondrial function in vivo. Despite the possibility to measure the carbonylation of circulating proteins or lipoproteins to monitor oxidative stress in the context of CVD [103], this technique is not tissue specific and thus makes it impossible to clearly define the source of ROS. In the last few years, technological advancements have made it possible to measure, more specifically, the carbonylation of cardiac proteins, such as myosin binding protein C, cardiac type (MYBPC). However, it is not really specific for the identification of the ROS source, and it can be performed only post-mortem [104].
Recently, energy metabolites measured by mass spectrometry have been proposed as potential markers. However, the clinical applicability remains a matter of debate [105]. More promising results are instead linked to the MitoTimer mouse, which represents the most advanced experimental model of mitochondrial dysfunction in the preclinical cardiovascular field [106,107] providing the possibility of studying the mitochondrial structure, the redox state, and the mitophagy function [106,107].
Furthermore, a series of radioactive tracers seem to represent a good potential for monitoring mitochondrial functions in the CV field [108]. In this context, pre-specified studies could open a new frontier to better evaluate the link between mitochondrial dysfunction and multiple forms of CVD.
In the last decade, several experimental evidences on animal models have pointed out a significant gender related difference in CVD development [109]. These observations are also supported by real-life epidemiological data [110,111,112]. However, the evaluation of this aspect in correlation with mitochondrial dysfunction is completely missing.
Finally, the time factor should be considered. Atherosclerosis is a time-dependent chronic disease and at the moment of its clinical relevance, the etiopathogenetic factors have already reached a degree of expression which in some cases is irreversible [8]. Atherosclerosis-mediated CVDs mainly involve older individuals with other comorbidities beyond age, such as obesity, hypertension, diabetes, and immune frailty, thus defining progression and treatment sensitivity of CVD [113,114]. To date, few experimental models reproduce this clinical condition, further limiting our knowledge on the mixed contribution of these comorbidities on mitochondrial function.
Taking together all these limitations, future research should be directed to (a) improve the pharmacodynamic and pharmacokinetic properties of therapeutic strategies, (b) study in depth the mechanisms related to mitochondrial function, (c) reevaluate the pathogenesis of CVDs, or at least those such as atherosclerosis which strongly correlate with mitochondrial dysfunction.

6. Conclusions

Atherosclerosis is a multifactorial disease. Multiple clinical trials and basic studies have clearly demonstrated that the management of the known risk factors only is not enough to limit the burden of this condition which underlies most cardiovascular diseases. Increasing evidence suggests an important role for mitochondria in the initial steps of this process. They regulate the inflammatory response and oxidative stress, two key steps that, once dysfunctional, might modulate initiation and progression of the atherosclerotic lesion. Thus, the modulation of mitochondrial function could delay the development of endothelial dysfunction, which represents the primum movens of the atherosclerotic process. In this context, it will be important for research, at both preclinical and clinical levels, to define the precise therapeutic interventions focusing on mitochondrial functions, considering all the aforementioned factors such as age, sex, and relevant comorbidities.
In this review, we have summarized the latest data on the relationship between mitochondrial functions and atherosclerosis, with a look in depth at endothelial dysfunction. Although mitochondria have been recognized as a new therapeutic target in different pathological contests, the current gaps in therapeutic strategies to manage mitochondrial dysfunction do not allow for the provision of a clinical benefit in the development of atherosclerosis. Both antioxidants and gene therapy are attractive approaches for the treatment of atherosclerosis. However, further ongoing and new studies are needed to identify the correct strategy for reducing the impact of mitochondrial dysfunction on the progression of atherosclerosis.

Author Contributions

Conceptualization, G.C. (Giovanni Ciccarelli), A.G. and G.C. (Giovanni Cimmino); methodology, G.C. (Giovanni Ciccarelli) and G.C. (Giovanni Cimmino); resources, G.C. (Giovanni Ciccarelli), A.G. and G.C. (Giovanni Cimmino); data curation, G.C. (Giovanni Ciccarelli), A.G. and G.C. (Giovanni Cimmino); writing—original draft preparation, G.C. (Giovanni Ciccarelli) and S.C.; writing—review and editing, G.C. (Giovanni Ciccarelli), A.M., P.M. and G.C. (Giovanni Cimmino); visualization, G.C. (Giovanni Ciccarelli), G.C. (Giovanni Cimmino), S.C., P.M. and A.G.; supervision, A.G. and A.M.; project administration, A.G.; funding acquisition, A.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Written informed consent has been obtained from the patient(s) to publish this paper.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Virani, S.S.; Alonso, A.; Benjamin, E.J.; Bittencourt, M.S.; Callaway, C.W.; Carson, A.P.; Chamberlain, A.M.; Chang, A.R.; Cheng, S.; Delling, F.N.; et al. Heart Disease and Stroke Statistics-2020 Update: A Report From the American Heart Association. Circulation 2020, 141, e139–e596. [Google Scholar] [CrossRef]
  2. Roth, G.A.; Mensah, G.A.; Johnson, C.O.; Addolorato, G.; Ammirati, E.; Baddour, L.M.; Barengo, N.C.; Beaton, A.Z.; Benjamin, E.J.; Benziger, C.P.; et al. Global Burden of Cardiovascular Diseases and Risk Factors, 1990–2019: Update From the GBD 2019 Study. J. Am. Coll. Cardiol. 2020, 76, 2982–3021. [Google Scholar] [CrossRef] [PubMed]
  3. Sanchez-Vinas, A.; Corral-Partearroyo, C.; Gil-Girbau, M.; Penarrubia-Maria, M.T.; Gallardo-Gonzalez, C.; Olmos-Palenzuela, M.D.; Aznar-Lou, I.; Serrano-Blanco, A.; Rubio-Valera, M. Effectiveness and cost-effectiveness of an intervention to improve Initial Medication Adherence to treatments for cardiovascular diseases and diabetes in primary care: Study protocol for a pragmatic cluster randomised controlled trial and economic model (the IMA-cRCT study). BMC Prim. Care 2022, 23, 170. [Google Scholar] [CrossRef]
  4. Alizadeh, G.; Gholipour, K.; Azami-Aghdash, S.; Dehnavieh, R.; JafarAbadi, M.A.; Azmin, M.; Khodayari-Zarnaq, R. Social, Economic, Technological, and Environmental Factors Affecting Cardiovascular Diseases: A Systematic Review and Thematic Analysis. Int. J. Prev. Med. 2022, 13, 78. [Google Scholar] [CrossRef]
  5. Lee, C.B.; Liao, C.M.; Peng, L.H.; Lin, C.M. Economic fluctuations and cardiovascular diseases: A multiple-input time series analysis. PLoS ONE 2019, 14, e0219358. [Google Scholar] [CrossRef] [Green Version]
  6. Aminde, L.N.; Veerman, L. Interventions for the prevention of cardiovascular diseases: A protocol for a systematic review of economic evaluations in low-income and middle-income countries. BMJ Open 2016, 6, e013668. [Google Scholar] [CrossRef] [Green Version]
  7. Leal, J.; Luengo-Fernandez, R.; Gray, A.; Petersen, S.; Rayner, M. Economic burden of cardiovascular diseases in the enlarged European Union. Eur. Heart J. 2006, 27, 1610–1619. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Frak, W.; Wojtasinska, A.; Lisinska, W.; Mlynarska, E.; Franczyk, B.; Rysz, J. Pathophysiology of Cardiovascular Diseases: New Insights into Molecular Mechanisms of Atherosclerosis, Arterial Hypertension, and Coronary Artery Disease. Biomedicines 2022, 10, 1938. [Google Scholar] [CrossRef]
  9. Libby, P. The changing landscape of atherosclerosis. Nature 2021, 592, 524–533. [Google Scholar] [CrossRef]
  10. Boren, J.; Chapman, M.J.; Krauss, R.M.; Packard, C.J.; Bentzon, J.F.; Binder, C.J.; Daemen, M.J.; Demer, L.L.; Hegele, R.A.; Nicholls, S.J.; et al. Low-density lipoproteins cause atherosclerotic cardiovascular disease: Pathophysiological, genetic, and therapeutic insights: A consensus statement from the European Atherosclerosis Society Consensus Panel. Eur. Heart J. 2020, 41, 2313–2330. [Google Scholar] [CrossRef]
  11. Khatana, C.; Saini, N.K.; Chakrabarti, S.; Saini, V.; Sharma, A.; Saini, R.V.; Saini, A.K. Mechanistic Insights into the Oxidized Low-Density Lipoprotein-Induced Atherosclerosis. Oxidative Med. Cell. Longev. 2020, 2020, 5245308. [Google Scholar] [CrossRef] [PubMed]
  12. Greenberg, S.R. Atherosclerosis and mitochondria. Ann. Intern. Med. 1970, 73, 861–862. [Google Scholar] [CrossRef] [PubMed]
  13. Salnikova, D.; Orekhova, V.; Grechko, A.; Starodubova, A.; Bezsonov, E.; Popkova, T.; Orekhov, A. Mitochondrial Dysfunction in Vascular Wall Cells and Its Role in Atherosclerosis. Int. J. Mol. Sci. 2021, 22, 8990. [Google Scholar] [CrossRef] [PubMed]
  14. Sinenko, S.A.; Starkova, T.Y.; Kuzmin, A.A.; Tomilin, A.N. Physiological Signaling Functions of Reactive Oxygen Species in Stem Cells: From Flies to Man. Front. Cell Dev. Biol. 2021, 9, 714370. [Google Scholar] [CrossRef] [PubMed]
  15. Zhang, J.; Wang, X.; Vikash, V.; Ye, Q.; Wu, D.; Liu, Y.; Dong, W. ROS and ROS-Mediated Cellular Signaling. Oxidative Med. Cell. Longev. 2016, 2016, 4350965. [Google Scholar] [CrossRef] [Green Version]
  16. Yang, S.; Lian, G. ROS and diseases: Role in metabolism and energy supply. Mol. Cell. Biochem. 2020, 467, 1–12. [Google Scholar] [CrossRef] [PubMed]
  17. Suarez-Rivero, J.M.; Pastor-Maldonado, C.J.; Povea-Cabello, S.; Alvarez-Cordoba, M.; Villalon-Garcia, I.; Talaveron-Rey, M.; Suarez-Carrillo, A.; Munuera-Cabeza, M.; Sanchez-Alcazar, J.A. From Mitochondria to Atherosclerosis: The Inflammation Path. Biomedicines 2021, 9, 258. [Google Scholar] [CrossRef]
  18. Niyazov, D.M.; Kahler, S.G.; Frye, R.E. Primary Mitochondrial Disease and Secondary Mitochondrial Dysfunction: Importance of Distinction for Diagnosis and Treatment. Mol. Syndromol. 2016, 7, 122–137. [Google Scholar] [CrossRef] [Green Version]
  19. North, B.J.; Sinclair, D.A. The intersection between aging and cardiovascular disease. Circ. Res. 2012, 110, 1097–1108. [Google Scholar] [CrossRef]
  20. Haas, R.H. Mitochondrial Dysfunction in Aging and Diseases of Aging. Biology 2019, 8, 48. [Google Scholar] [CrossRef]
  21. Shemiakova, T.; Ivanova, E.; Wu, W.K.; Kirichenko, T.V.; Starodubova, A.V.; Orekhov, A.N. Atherosclerosis as Mitochondriopathy: Repositioning the Disease to Help Finding New Therapies. Front. Cardiovasc. Med. 2021, 8, 660473. [Google Scholar] [CrossRef] [PubMed]
  22. Jennings, R.B.; Ganote, C.E. Mitochondrial structure and function in acute myocardial ischemic injury. Circ. Res. 1976, 38, I80–I91. [Google Scholar] [PubMed]
  23. 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, 2060–2076. [Google Scholar] [CrossRef] [PubMed]
  24. Zhou, B.; Tian, R. Mitochondrial dysfunction in pathophysiology of heart failure. J. Clin. Investig. 2018, 128, 3716–3726. [Google Scholar] [CrossRef] [Green Version]
  25. Iglewski, M.; Hill, J.A.; Lavandero, S.; Rothermel, B.A. Mitochondrial fission and autophagy in the normal and diseased heart. Curr. Hypertens. Rep. 2010, 12, 418–425. [Google Scholar] [CrossRef] [Green Version]
  26. Lee, C.F.; Chavez, J.D.; Garcia-Menendez, L.; Choi, Y.; Roe, N.D.; Chiao, Y.A.; Edgar, J.S.; Goo, Y.A.; Goodlett, D.R.; Bruce, J.E.; et al. Normalization of NAD+ Redox Balance as a Therapy for Heart Failure. Circulation 2016, 134, 883–894. [Google Scholar] [CrossRef] [Green Version]
  27. Luongo, T.S.; Lambert, J.P.; Gross, P.; Nwokedi, M.; Lombardi, A.A.; Shanmughapriya, S.; Carpenter, A.C.; Kolmetzky, D.; Gao, E.; van Berlo, J.H.; et al. The mitochondrial Na+/Ca2+ exchanger is essential for Ca2+ homeostasis and viability. Nature 2017, 545, 93–97. [Google Scholar] [CrossRef] [Green Version]
  28. Chehaitly, A.; Guihot, A.L.; Proux, C.; Grimaud, L.; Aurriere, J.; Legouriellec, B.; Rivron, J.; Vessieres, E.; Tetaud, C.; Zorzano, A.; et al. Altered Mitochondrial Opa1-Related Fusion in Mouse Promotes Endothelial Cell Dysfunction and Atherosclerosis. Antioxidants 2022, 11, 1078. [Google Scholar] [CrossRef]
  29. Hubens, W.H.G.; Vallbona-Garcia, A.; de Coo, I.F.M.; van Tienen, F.H.J.; Webers, C.A.B.; Smeets, H.J.M.; Gorgels, T. Blood biomarkers for assessment of mitochondrial dysfunction: An expert review. Mitochondrion 2022, 62, 187–204. [Google Scholar] [CrossRef]
  30. Xin, Y.; Zhang, X.; Li, J.; Gao, H.; Li, J.; Li, J.; Hu, W.; Li, H. New Insights Into the Role of Mitochondria Quality Control in Ischemic Heart Disease. Front. Cardiovasc. Med. 2021, 8, 774619. [Google Scholar] [CrossRef]
  31. Garbincius, J.F.; Luongo, T.S.; Jadiya, P.; Hildebrand, A.N.; Kolmetzky, D.W.; Mangold, A.S.; Roy, R.; Ibetti, J.; Nwokedi, M.; Koch, W.J.; et al. Enhanced NCLX-dependent mitochondrial Ca2+ efflux attenuates pathological remodeling in heart failure. J. Mol. Cell. Cardiol. 2022, 167, 52–66. [Google Scholar] [CrossRef]
  32. Lesnefsky, E.J.; Chen, Q.; Tandler, B.; Hoppel, C.L. Mitochondrial Dysfunction and Myocardial Ischemia-Reperfusion: Implications for Novel Therapies. Annu. Rev. Pharmacol. Toxicol. 2017, 57, 535–565. [Google Scholar] [CrossRef]
  33. Martin, J.L.; Costa, A.S.H.; Gruszczyk, A.V.; Beach, T.E.; Allen, F.M.; Prag, H.A.; Hinchy, E.C.; Mahbubani, K.; Hamed, M.; Tronci, L.; et al. Succinate accumulation drives ischaemia-reperfusion injury during organ transplantation. Nat. Metab. 2019, 1, 966–974. [Google Scholar] [CrossRef] [Green Version]
  34. Zhang, J.; Wang, Y.T.; Miller, J.H.; Day, M.M.; Munger, J.C.; Brookes, P.S. Accumulation of Succinate in Cardiac Ischemia Primarily Occurs via Canonical Krebs Cycle Activity. Cell Rep. 2018, 23, 2617–2628. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Chouchani, E.T.; Pell, V.R.; James, A.M.; Work, L.M.; Saeb-Parsy, K.; Frezza, C.; Krieg, T.; Murphy, M.P. A Unifying Mechanism for Mitochondrial Superoxide Production during Ischemia-Reperfusion Injury. Cell Metab. 2016, 23, 254–263. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Milliken, A.S.; Kulkarni, C.A.; Brookes, P.S. Acid enhancement of ROS generation by complex-I reverse electron transport is balanced by acid inhibition of complex-II: Relevance for tissue reperfusion injury. Redox Biol. 2020, 37, 101733. [Google Scholar] [CrossRef]
  37. Yin, Z.; Burger, N.; Kula-Alwar, D.; Aksentijevic, D.; Bridges, H.R.; Prag, H.A.; Grba, D.N.; Viscomi, C.; James, A.M.; Mottahedin, A.; et al. Structural basis for a complex I mutation that blocks pathological ROS production. Nat. Commun. 2021, 12, 707. [Google Scholar] [CrossRef]
  38. Chouchani, E.T.; Pell, V.R.; Gaude, E.; Aksentijevic, D.; Sundier, S.Y.; Robb, E.L.; Logan, A.; Nadtochiy, S.M.; Ord, E.N.J.; Smith, A.C.; et al. Ischaemic accumulation of succinate controls reperfusion injury through mitochondrial ROS. Nature 2014, 515, 431–435. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Sazonova, M.A.; Ryzhkova, A.I.; Sinyov, V.V.; Sazonova, M.D.; Nikitina, N.N.; Shkurat, T.P.; Sobenin, I.A.; Orekhov, A.N. Mitochondrial mutations associated with cardiac angina. Vessel Plus 2019, 3, 8. [Google Scholar] [CrossRef]
  40. Shemiakova, T.; Ivanova, E.; Grechko, A.V.; Gerasimova, E.V.; Sobenin, I.A.; Orekhov, A.N. Mitochondrial Dysfunction and DNA Damage in the Context of Pathogenesis of Atherosclerosis. Biomedicines 2020, 8, 166. [Google Scholar] [CrossRef]
  41. Madamanchi, N.R.; Runge, M.S. Mitochondrial dysfunction in atherosclerosis. Circ. Res. 2007, 100, 460–473. [Google Scholar] [CrossRef] [Green Version]
  42. Hulsmans, M.; Van Dooren, E.; Holvoet, P. Mitochondrial reactive oxygen species and risk of atherosclerosis. Curr. Atheroscler. Rep. 2012, 14, 264–276. [Google Scholar] [CrossRef] [PubMed]
  43. Wen, R.; Banik, B.; Pathak, R.K.; Kumar, A.; Kolishetti, N.; Dhar, S. Nanotechnology inspired tools for mitochondrial dysfunction related diseases. Adv. Drug Deliv. Rev. 2016, 99, 52–69. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Naik, E.; Dixit, V.M. Mitochondrial reactive oxygen species drive proinflammatory cytokine production. J. Exp. Med. 2011, 208, 417–420. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Yu, E.P.K.; Reinhold, J.; Yu, H.; Starks, L.; Uryga, A.K.; Foote, K.; Finigan, A.; Figg, N.; Pung, Y.F.; Logan, A.; et al. Mitochondrial Respiration Is Reduced in Atherosclerosis, Promoting Necrotic Core Formation and Reducing Relative Fibrous Cap Thickness. Arterioscler. Thromb. Vasc. Biol. 2017, 37, 2322–2332. [Google Scholar] [CrossRef] [Green Version]
  46. Ballinger, S.W.; Patterson, C.; Yan, C.N.; Doan, R.; Burow, D.L.; Young, C.G.; Yakes, F.M.; Van Houten, B.; Ballinger, C.A.; Freeman, B.A.; et al. Hydrogen peroxide- and peroxynitrite-induced mitochondrial DNA damage and dysfunction in vascular endothelial and smooth muscle cells. Circ. Res. 2000, 86, 960–966. [Google Scholar] [CrossRef]
  47. Hu, H.; Lin, Y.; Xu, X.; Lin, S.; Chen, X.; Wang, S. The alterations of mitochondrial DNA in coronary heart disease. Exp. Mol. Pathol. 2020, 114, 104412. [Google Scholar] [CrossRef]
  48. Nakahira, K.; Hisata, S.; Choi, A.M. The Roles of Mitochondrial Damage-Associated Molecular Patterns in Diseases. Antioxid. Redox Signal. 2015, 23, 1329–1350. [Google Scholar] [CrossRef] [Green Version]
  49. Zakirov, F.H.; Zhang, D.; Grechko, A.V.; Wu, W.K.; Poznyak, A.V.; Orekhov, A.N. Lipid-based gene delivery to macrophage mitochondria for atherosclerosis therapy. Pharmacol. Res. Perspect. 2020, 8, e00584. [Google Scholar] [CrossRef]
  50. Vindis, C.; Elbaz, M.; Escargueil-Blanc, I.; Auge, N.; Heniquez, A.; Thiers, J.C.; Negre-Salvayre, A.; Salvayre, R. Two distinct calcium-dependent mitochondrial pathways are involved in oxidized LDL-induced apoptosis. Arterioscler. Thromb. Vasc. Biol. 2005, 25, 639–645. [Google Scholar] [CrossRef] [PubMed]
  51. Khwaja, B.; Thankam, F.G.; Agrawal, D.K. Mitochondrial DAMPs and altered mitochondrial dynamics in OxLDL burden in atherosclerosis. Mol. Cell. Biochem. 2021, 476, 1915–1928. [Google Scholar] [CrossRef]
  52. Li, D.; Yang, S.; Xing, Y.; Pan, L.; Zhao, R.; Zhao, Y.; Liu, L.; Wu, M. Novel Insights and Current Evidence for Mechanisms of Atherosclerosis: Mitochondrial Dynamics as a Potential Therapeutic Target. Front. Cell Dev. Biol. 2021, 9, 673839. [Google Scholar] [CrossRef]
  53. Liao, J.K. Linking endothelial dysfunction with endothelial cell activation. J. Clin. Investig. 2013, 123, 540–541. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Poredos, P.; Poredos, A.V.; Gregoric, I. Endothelial Dysfunction and Its Clinical Implications. Angiology 2021, 72, 604–615. [Google Scholar] [CrossRef] [PubMed]
  55. Sima, A.V.; Stancu, C.S.; Simionescu, M. Vascular endothelium in atherosclerosis. Cell Tissue Res. 2009, 335, 191–203. [Google Scholar] [CrossRef]
  56. Mundi, S.; Massaro, M.; Scoditti, E.; Carluccio, M.A.; van Hinsbergh, V.W.M.; Iruela-Arispe, M.L.; De Caterina, R. Endothelial permeability, LDL deposition, and cardiovascular risk factors—A review. Cardiovasc. Res. 2018, 114, 35–52. [Google Scholar] [CrossRef] [Green Version]
  57. Moriya, J. Critical roles of inflammation in atherosclerosis. J. Cardiol. 2019, 73, 22–27. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Bennett, M.R.; Sinha, S.; Owens, G.K. Vascular Smooth Muscle Cells in Atherosclerosis. Circ. Res. 2016, 118, 692–702. [Google Scholar] [CrossRef] [Green Version]
  59. Dominic, E.A.; Ramezani, A.; Anker, S.D.; Verma, M.; Mehta, N.; Rao, M. Mitochondrial cytopathies and cardiovascular disease. Heart 2014, 100, 611–618. [Google Scholar] [CrossRef]
  60. Tretter, L.; Ambrus, A. Measurement of ROS homeostasis in isolated mitochondria. Methods Enzymol. 2014, 547, 199–223. [Google Scholar] [CrossRef]
  61. Sun, Q.; Zhong, W.; Zhang, W.; Zhou, Z. Defect of mitochondrial respiratory chain is a mechanism of ROS overproduction in a rat model of alcoholic liver disease: Role of zinc deficiency. Am. J. Physiol. Gastrointest. Liver Physiol. 2016, 310, G205–G214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Frey, R.S.; Gao, X.; Javaid, K.; Siddiqui, S.S.; Rahman, A.; Malik, A.B. Phosphatidylinositol 3-kinase gamma signaling through protein kinase Czeta induces NADPH oxidase-mediated oxidant generation and NF-kappaB activation in endothelial cells. J. Biol. Chem. 2006, 281, 16128–16138. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Dinkova-Kostova, A.T.; Abramov, A.Y. The emerging role of Nrf2 in mitochondrial function. Free. Radic. Biol. Med. 2015, 88, 179–188. [Google Scholar] [CrossRef] [Green Version]
  64. Cheng, X.; Siow, R.C.; Mann, G.E. Impaired redox signaling and antioxidant gene expression in endothelial cells in diabetes: A role for mitochondria and the nuclear factor-E2-related factor 2-Kelch-like ECH-associated protein 1 defense pathway. Antioxid. Redox Signal. 2011, 14, 469–487. [Google Scholar] [CrossRef] [PubMed]
  65. Zhu, Y.; Li, M.; Lu, Y.; Li, J.; Ke, Y.; Yang, J. Ilexgenin A inhibits mitochondrial fission and promote Drp1 degradation by Nrf2-induced PSMB5 in endothelial cells. Drug Dev. Res. 2019, 80, 481–489. [Google Scholar] [CrossRef] [PubMed]
  66. Ceaser, E.K.; Ramachandran, A.; Levonen, A.L.; Darley-Usmar, V.M. Oxidized low-density lipoprotein and 15-deoxy-delta 12,14-PGJ2 increase mitochondrial complex I activity in endothelial cells. Am. J. Physiol. Heart Circ. Physiol. 2003, 285, H2298–H2308. [Google Scholar] [CrossRef]
  67. Kinscherf, R.; Deigner, H.P.; Usinger, C.; Pill, J.; Wagner, M.; Kamencic, H.; Hou, D.; Chen, M.; Schmiedt, W.; Schrader, M.; et al. Induction of mitochondrial manganese superoxide dismutase in macrophages by oxidized LDL: Its relevance in atherosclerosis of humans and heritable hyperlipidemic rabbits. FASEB J. Off. Publ. Fed. Am. Soc. Exp. Biol. 1997, 11, 1317–1328. [Google Scholar] [CrossRef]
  68. Yang, X.; Pan, W.; Xu, G.; Chen, L. Mitophagy: A crucial modulator in the pathogenesis of chronic diseases. Clin. Chim. Acta Int. J. Clin. Chem. 2020, 502, 245–254. [Google Scholar] [CrossRef]
  69. Garza-Lombo, C.; Pappa, A.; Panayiotidis, M.I.; Franco, R. Redox homeostasis, oxidative stress and mitophagy. Mitochondrion 2020, 51, 105–117. [Google Scholar] [CrossRef]
  70. Dymkowska, D. The involvement of autophagy in the maintenance of endothelial homeostasis: The role of mitochondria. Mitochondrion 2021, 57, 131–147. [Google Scholar] [CrossRef]
  71. Drabarek, B.; Dymkowska, D.; Szczepanowska, J.; Zablocki, K. TNFalpha affects energy metabolism and stimulates biogenesis of mitochondria in EA.hy926 endothelial cells. Int. J. Biochem. Cell Biol. 2012, 44, 1390–1397. [Google Scholar] [CrossRef] [PubMed]
  72. Dymkowska, D.; Drabarek, B.; Michalik, A.; Nowak, N.; Zablocki, K. TNFalpha stimulates NO release in EA.hy926 cells by activating the CaMKKbeta-AMPK-eNOS pathway. Int. J. Biochem. Cell Biol. 2019, 106, 57–67. [Google Scholar] [CrossRef]
  73. Koga, Y.; Akita, Y.; Junko, N.; Yatsuga, S.; Povalko, N.; Fukiyama, R.; Ishii, M.; Matsuishi, T. Endothelial dysfunction in MELAS improved by l-arginine supplementation. Neurology 2006, 66, 1766–1769. [Google Scholar] [CrossRef] [PubMed]
  74. Wang, X.L.; Sim, A.S.; Badenhop, R.F.; McCredie, R.M.; Wilcken, D.E. A smoking-dependent risk of coronary artery disease associated with a polymorphism of the endothelial nitric oxide synthase gene. Nat. Med. 1996, 2, 41–45. [Google Scholar] [CrossRef]
  75. Koga, Y.; Akita, Y.; Nishioka, J.; Yatsuga, S.; Povalko, N.; Tanabe, Y.; Fujimoto, S.; Matsuishi, T. L-arginine improves the symptoms of strokelike episodes in MELAS. Neurology 2005, 64, 710–712. [Google Scholar] [CrossRef]
  76. Koga, J.I.; Sun, X.; Ushio-Fukai, M. Editorial: Mitochondria, metabolism and cardiovascular diseases. Front. Cardiovasc. Med. 2022, 9, 996739. [Google Scholar] [CrossRef]
  77. Chandran, K.; Aggarwal, D.; Migrino, R.Q.; Joseph, J.; McAllister, D.; Konorev, E.A.; Antholine, W.E.; Zielonka, J.; Srinivasan, S.; Avadhani, N.G.; et al. Doxorubicin inactivates myocardial cytochrome c oxidase in rats: Cardioprotection by Mito-Q. Biophys. J. 2009, 96, 1388–1398. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Rossman, M.J.; Santos-Parker, J.R.; Steward, C.A.C.; Bispham, N.Z.; Cuevas, L.M.; Rosenberg, H.L.; Woodward, K.A.; Chonchol, M.; Gioscia-Ryan, R.A.; Murphy, M.P.; et al. Chronic Supplementation with a Mitochondrial Antioxidant (MitoQ) Improves Vascular Function in Healthy Older Adults. Hypertension 2018, 71, 1056–1063. [Google Scholar] [CrossRef]
  79. Lazou, A.; Ramachandra, C.J. Protecting the Mitochondria in Cardiac Disease. Int. J. Mol. Sci. 2022, 23, 8115. [Google Scholar] [CrossRef]
  80. Diguet, N.; Trammell, S.A.J.; Tannous, C.; Deloux, R.; Piquereau, J.; Mougenot, N.; Gouge, A.; Gressette, M.; Manoury, B.; Blanc, J.; et al. Nicotinamide Riboside Preserves Cardiac Function in a Mouse Model of Dilated Cardiomyopathy. Circulation 2018, 137, 2256–2273. [Google Scholar] [CrossRef]
  81. Forte, M.; Bianchi, F.; Cotugno, M.; Marchitti, S.; De Falco, E.; Raffa, S.; Stanzione, R.; Di Nonno, F.; Chimenti, I.; Palmerio, S.; et al. Pharmacological restoration of autophagy reduces hypertension-related stroke occurrence. Autophagy 2020, 16, 1468–1481. [Google Scholar] [CrossRef]
  82. Tong, D.; Schiattarella, G.G.; Jiang, N.; Altamirano, F.; Szweda, P.A.; Elnwasany, A.; Lee, D.I.; Yoo, H.; Kass, D.A.; Szweda, L.I.; et al. NAD(+) Repletion Reverses Heart Failure with Preserved Ejection Fraction. Circ. Res. 2021, 128, 1629–1641. [Google Scholar] [CrossRef]
  83. Forte, M.; Palmerio, S.; Bianchi, F.; Volpe, M.; Rubattu, S. Mitochondrial complex I deficiency and cardiovascular diseases: Current evidence and future directions. J. Mol. Med. 2019, 97, 579–591. [Google Scholar] [CrossRef] [PubMed]
  84. Dai, D.F.; Rabinovitch, P.S.; Ungvari, Z. Mitochondria and cardiovascular aging. Circ. Res. 2012, 110, 1109–1124. [Google Scholar] [CrossRef] [PubMed]
  85. Bonora, M.; Wieckowski, M.R.; Sinclair, D.A.; Kroemer, G.; Pinton, P.; Galluzzi, L. Targeting mitochondria for cardiovascular disorders: Therapeutic potential and obstacles. Nat. Rev. Cardiol. 2019, 16, 33–55. [Google Scholar] [CrossRef] [PubMed]
  86. Saad, A.; Herrmann, S.M.S.; Eirin, A.; Ferguson, C.M.; Glockner, J.F.; Bjarnason, H.; McKusick, M.A.; Misra, S.; Lerman, L.O.; Textor, S.C. Phase 2a Clinical Trial of Mitochondrial Protection (Elamipretide) During Stent Revascularization in Patients with Atherosclerotic Renal Artery Stenosis. Circ. Cardiovasc. Interv. 2017, 10, e005487. [Google Scholar] [CrossRef]
  87. Gibson, C.M.; Giugliano, R.P.; Kloner, R.A.; Bode, C.; Tendera, M.; Janosi, A.; Merkely, B.; Godlewski, J.; Halaby, R.; Korjian, S.; et al. EMBRACE STEMI study: A Phase 2a trial to evaluate the safety, tolerability, and efficacy of intravenous MTP-131 on reperfusion injury in patients undergoing primary percutaneous coronary intervention. Eur. Heart J. 2016, 37, 1296–1303. [Google Scholar] [CrossRef] [Green Version]
  88. Sabbah, H.N.; Gupta, R.C.; Singh-Gupta, V.; Zhang, K.; Lanfear, D.E. Abnormalities of Mitochondrial Dynamics in the Failing Heart: Normalization Following Long-Term Therapy with Elamipretide. Cardiovasc. Drugs Ther. 2018, 32, 319–328. [Google Scholar] [CrossRef] [Green Version]
  89. Butler, J.; Khan, M.S.; Anker, S.D.; Fonarow, G.C.; Kim, R.J.; Nodari, S.; O’Connor, C.M.; Pieske, B.; Pieske-Kraigher, E.; Sabbah, H.N.; et al. Effects of Elamipretide on Left Ventricular Function in Patients with Heart Failure with Reduced Ejection Fraction: The PROGRESS-HF Phase 2 Trial. J. Card. Fail. 2020, 26, 429–437. [Google Scholar] [CrossRef]
  90. Murphy, M.P.; Smith, R.A. Drug delivery to mitochondria: The key to mitochondrial medicine. Adv. Drug Deliv. Rev. 2000, 41, 235–250. [Google Scholar] [CrossRef]
  91. Mercer, J.R.; Yu, E.; Figg, N.; Cheng, K.K.; Prime, T.A.; Griffin, J.L.; Masoodi, M.; Vidal-Puig, A.; Murphy, M.P.; Bennett, M.R. The mitochondria-targeted antioxidant MitoQ decreases features of the metabolic syndrome in ATM+/-/ApoE-/- mice. Free. Radic. Biol. Med. 2012, 52, 841–849. [Google Scholar] [CrossRef] [PubMed]
  92. Karnewar, S.; Pulipaka, S.; Katta, S.; Panuganti, D.; Neeli, P.K.; Thennati, R.; Jerald, M.K.; Kotamraju, S. Mitochondria-targeted esculetin mitigates atherosclerosis in the setting of aging via the modulation of SIRT1-mediated vascular cell senescence and mitochondrial function in Apoe(-/-) mice. Atherosclerosis 2022, 356, 28–40. [Google Scholar] [CrossRef] [PubMed]
  93. Bogacka, I.; Xie, H.; Bray, G.A.; Smith, S.R. Pioglitazone induces mitochondrial biogenesis in human subcutaneous adipose tissue in vivo. Diabetes 2005, 54, 1392–1399. [Google Scholar] [CrossRef] [Green Version]
  94. Dorighello, G.G.; Rovani, J.C.; Paim, B.A.; Rentz, T.; Assis, L.H.P.; Vercesi, A.E.; Oliveira, H.C.F. Mild Mitochondrial Uncoupling Decreases Experimental Atherosclerosis, A Proof of Concept. J. Atheroscler. Thromb. 2022, 29, 825–838. [Google Scholar] [CrossRef] [PubMed]
  95. Kelso, G.F.; Porteous, C.M.; Coulter, C.V.; Hughes, G.; Porteous, W.K.; Ledgerwood, E.C.; Smith, R.A.; Murphy, M.P. Selective targeting of a redox-active ubiquinone to mitochondria within cells: Antioxidant and antiapoptotic properties. J. Biol. Chem. 2001, 276, 4588–4596. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Jiang, J.; Stoyanovsky, D.A.; Belikova, N.A.; Tyurina, Y.Y.; Zhao, Q.; Tungekar, M.A.; Kapralova, V.; Huang, Z.; Mintz, A.H.; Greenberger, J.S.; et al. A mitochondria-targeted triphenylphosphonium-conjugated nitroxide functions as a radioprotector/mitigator. Radiat. Res. 2009, 172, 706–717. [Google Scholar] [CrossRef] [Green Version]
  97. Galluzzi, L.; Vitale, I.; Kepp, O.; Seror, C.; Hangen, E.; Perfettini, J.L.; Modjtahedi, N.; Kroemer, G. Methods to dissect mitochondrial membrane permeabilization in the course of apoptosis. Methods Enzymol. 2008, 442, 355–374. [Google Scholar] [CrossRef]
  98. Horton, K.L.; Stewart, K.M.; Fonseca, S.B.; Guo, Q.; Kelley, S.O. Mitochondria-penetrating peptides. Chem. Biol. 2008, 15, 375–382. [Google Scholar] [CrossRef] [Green Version]
  99. Schmidt, O.; Pfanner, N.; Meisinger, C. Mitochondrial protein import: From proteomics to functional mechanisms. Nat. Rev. Mol. Cell Biol. 2010, 11, 655–667. [Google Scholar] [CrossRef]
  100. Galluzzi, L.; Joza, N.; Tasdemir, E.; Maiuri, M.C.; Hengartner, M.; Abrams, J.M.; Tavernarakis, N.; Penninger, J.; Madeo, F.; Kroemer, G. No death without life: Vital functions of apoptotic effectors. Cell Death Differ. 2008, 15, 1113–1123. [Google Scholar] [CrossRef]
  101. Hao, Z.; Duncan, G.S.; Chang, C.C.; Elia, A.; Fang, M.; Wakeham, A.; Okada, H.; Calzascia, T.; Jang, Y.; You-Ten, A.; et al. Specific ablation of the apoptotic functions of cytochrome C reveals a differential requirement for cytochrome C and Apaf-1 in apoptosis. Cell 2005, 121, 579–591. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Izzo, V.; Bravo-San Pedro, J.M.; Sica, V.; Kroemer, G.; Galluzzi, L. Mitochondrial Permeability Transition: New Findings and Persisting Uncertainties. Trends Cell Biol. 2016, 26, 655–667. [Google Scholar] [CrossRef]
  103. Mutlu-Turkoglu, U.; Akalin, Z.; Ilhan, E.; Yilmaz, E.; Bilge, A.; Nisanci, Y.; Uysal, M. Increased plasma malondialdehyde and protein carbonyl levels and lymphocyte DNA damage in patients with angiographically defined coronary artery disease. Clin. Biochem. 2005, 38, 1059–1065. [Google Scholar] [CrossRef] [PubMed]
  104. Aryal, B.; Jeong, J.; Rao, V.A. Doxorubicin-induced carbonylation and degradation of cardiac myosin binding protein C promote cardiotoxicity. Proc. Natl. Acad. Sci. USA 2014, 111, 2011–2016. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Ruiz, M.; Labarthe, F.; Fortier, A.; Bouchard, B.; Thompson Legault, J.; Bolduc, V.; Rigal, O.; Chen, J.; Ducharme, A.; Crawford, P.A.; et al. Circulating acylcarnitine profile in human heart failure: A surrogate of fatty acid metabolic dysregulation in mitochondria and beyond. Am. J. Physiol. Heart Circ. Physiol. 2017, 313, H768–H781. [Google Scholar] [CrossRef] [Green Version]
  106. Wilson, R.J.; Drake, J.C.; Cui, D.; Zhang, M.; Perry, H.M.; Kashatus, J.A.; Kusminski, C.M.; Scherer, P.E.; Kashatus, D.F.; Okusa, M.D.; et al. Conditional MitoTimer reporter mice for assessment of mitochondrial structure, oxidative stress, and mitophagy. Mitochondrion 2019, 44, 20–26. [Google Scholar] [CrossRef]
  107. Stotland, A.; Gottlieb, R.A. alpha-MHC MitoTimer mouse: In vivo mitochondrial turnover model reveals remarkable mitochondrial heterogeneity in the heart. J. Mol. Cell. Cardiol. 2016, 90, 53–58. [Google Scholar] [CrossRef] [Green Version]
  108. Li, J.; Lu, J.; Zhou, Y. Mitochondrial-Targeted Molecular Imaging in Cardiac Disease. BioMed Res. Int. 2017, 2017, 5246853. [Google Scholar] [CrossRef] [Green Version]
  109. Hu, X.X.; Fu, L.; Li, Y.; Lin, Z.B.; Liu, X.; Wang, J.F.; Chen, Y.X.; Wang, Z.P.; Zhang, X.; Ou, Z.J.; et al. The Cardioprotective Effect of Vitamin E (Alpha-Tocopherol) Is Strongly Related to Age and Gender in Mice. PLoS ONE 2015, 10, e0137405. [Google Scholar] [CrossRef]
  110. Crea, F.; Libby, P. Acute Coronary Syndromes: The Way Forward From Mechanisms to Precision Treatment. Circulation 2017, 136, 1155–1166. [Google Scholar] [CrossRef]
  111. Pagidipati, N.J.; Peterson, E.D. Acute coronary syndromes in women and men. Nat. Rev. Cardiol. 2016, 13, 471–480. [Google Scholar] [CrossRef] [PubMed]
  112. Dean, J.; Cruz, S.D.; Mehta, P.K.; Merz, C.N. Coronary microvascular dysfunction: Sex-specific risk, diagnosis, and therapy. Nat. Rev. Cardiol. 2015, 12, 406–414. [Google Scholar] [CrossRef] [PubMed]
  113. Andreadou, I.; Daiber, A.; Baxter, G.F.; Brizzi, M.F.; Di Lisa, F.; Kaludercic, N.; Lazou, A.; Varga, Z.V.; Zuurbier, C.J.; Schulz, R.; et al. Influence of cardiometabolic comorbidities on myocardial function, infarction, and cardioprotection: Role of cardiac redox signaling. Free. Radic. Biol. Med. 2021, 166, 33–52. [Google Scholar] [CrossRef] [PubMed]
  114. Sack, M.N.; Murphy, E. The role of comorbidities in cardioprotection. J. Cardiovasc. Pharmacol. Ther. 2011, 16, 267–272. [Google Scholar] [CrossRef]
Figure 1. A schematic view of the interaction between mitochondria dysfunction, atherosclerosis and oxidative stress.
Figure 1. A schematic view of the interaction between mitochondria dysfunction, atherosclerosis and oxidative stress.
Ijms 24 01086 g001
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ciccarelli, G.; Conte, S.; Cimmino, G.; Maiorano, P.; Morrione, A.; Giordano, A. Mitochondrial Dysfunction: The Hidden Player in the Pathogenesis of Atherosclerosis? Int. J. Mol. Sci. 2023, 24, 1086. https://doi.org/10.3390/ijms24021086

AMA Style

Ciccarelli G, Conte S, Cimmino G, Maiorano P, Morrione A, Giordano A. Mitochondrial Dysfunction: The Hidden Player in the Pathogenesis of Atherosclerosis? International Journal of Molecular Sciences. 2023; 24(2):1086. https://doi.org/10.3390/ijms24021086

Chicago/Turabian Style

Ciccarelli, Giovanni, Stefano Conte, Giovanni Cimmino, Patrizia Maiorano, Andrea Morrione, and Antonio Giordano. 2023. "Mitochondrial Dysfunction: The Hidden Player in the Pathogenesis of Atherosclerosis?" International Journal of Molecular Sciences 24, no. 2: 1086. https://doi.org/10.3390/ijms24021086

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

Article Metrics

Back to TopTop