Next Article in Journal
Treatment Response Predictors of Neoadjuvant Therapy for Locally Advanced Gastric Cancer: Current Status and Future Perspectives
Next Article in Special Issue
Mitochondrial Ca2+ Signaling and Bioenergetics in Alzheimer’s Disease
Previous Article in Journal
The Impact of Neutrophil-to-Lymphocyte Ratio after Two Courses of Pembrolizumab for Oncological Outcomes in Patients with Metastatic Urothelial Carcinoma
Previous Article in Special Issue
HMTM-Mediated Enhancement of Brain Bioenergetics in a Mouse Tauopathy Model Is Blocked by Chronic Administration of Rivastigmine
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Activation of the Mitochondrial Unfolded Protein Response: A New Therapeutic Target?

by
Juan M. Suárez-Rivero
1,
Carmen J. Pastor-Maldonado
1,
Suleva Povea-Cabello
1,
Mónica Álvarez-Córdoba
1,
Irene Villalón-García
1,
Marta Talaverón-Rey
1,
Alejandra Suárez-Carrillo
1,
Manuel Munuera-Cabeza
1,
Diana Reche-López
1,
Paula Cilleros-Holgado
1,
Rocío Piñero-Pérez
1 and
José A. Sánchez-Alcázar
1,2,*
1
Centro Andaluz de Biología del Desarrollo (CABD-CSIC-Universidad Pablo de Olavide), Centro de Investigación Biomédica en Red: Enfermedades Raras, Instituto de Salud Carlos III, 41013 Sevilla, Spain
2
Centro Andaluz de Biología del Desarrollo (CABD), Consejo Superior de Investigaciones Científicas, Universidad Pablo de Olavide, Carretera de Utrera Km 1, 41013 Sevilla, Spain
*
Author to whom correspondence should be addressed.
Biomedicines 2022, 10(7), 1611; https://doi.org/10.3390/biomedicines10071611
Submission received: 20 May 2022 / Revised: 1 July 2022 / Accepted: 4 July 2022 / Published: 6 July 2022
(This article belongs to the Special Issue Mitochondria and Brain Disease 2.0)

Abstract

:
Mitochondrial dysfunction is a key hub that is common to many diseases. Mitochondria’s role in energy production, calcium homeostasis, and ROS balance makes them essential for cell survival and fitness. However, there are no effective treatments for most mitochondrial and related diseases to this day. Therefore, new therapeutic approaches, such as activation of the mitochondrial unfolded protein response (UPRmt), are being examined. UPRmt englobes several compensation processes related to proteostasis and antioxidant mechanisms. UPRmt activation, through an hormetic response, promotes cell homeostasis and improves lifespan and disease conditions in biological models of neurodegenerative diseases, cardiopathies, and mitochondrial diseases. Although UPRmt activation is a promising therapeutic option for many conditions, its overactivation could lead to non-desired side effects, such as increased heteroplasmy of mitochondrial DNA mutations or cancer progression in oncologic patients. In this review, we present the most recent UPRmt activation therapeutic strategies, UPRmt’s role in diseases, and its possible negative consequences in particular pathological conditions.

1. Mitochondria and Homeostasis

Homeostasis is defined as any self-regulating process by which biological systems tend to maintain stability while adjusting to conditions that are optimal for survival. In short, homeostasis is the capacity of organisms to adapt to external or internal changes. The cellular responses must be tightly regulated and fast in order to allow for adaptation to environmental or internal changes and the maintenance of cellular survival. These mechanisms are the basis of cellular homeostasis, organized at several molecular levels, and require a perfect coordination: epigenetic remodeling [1], transcriptional responses [2], post transcriptional modifications [3], organelle reorganization [4,5], osmotic control [6], and protein regulation [7], among many other things, determine if the cells will survive or die depending on whether their responses are flexible enough to adapt.
Mitochondria are a key organelle for cell homeostasis [8]. They are vital compartments for every nucleated cell, given that energy production through oxidative phosphorylation (OXPHOS) is their main function. Additionally, mitochondria are necessary for steroid and heme biosynthesis [9], numerous metabolites’ production [10], ion regulation [11,12], redox signaling [13], programmed cell death [14], and innate immunity [15]. Because of mitochondria’s importance for cell homeostasis, mitochondrial dysfunctions cause a great variety of diseases which can affect almost all the tissues in the body. Mitochondrial dysfunction is generally associated with reactive oxygen species (ROS) overproduction, being closely correlated and present in all mitochondrial diseases and aging [16]. Due to the high ROS production of these organelles, their components tend to degenerate, resulting in underperforming mitochondrial function [17]. On the other hand, mutations in the mitochondrial or nuclear genome produce aberrant proteins that contribute to ROS overproduction and functional loss [18]. These alterations might not only lead to mitochondrial diseases but also to the development of neurodegenerative diseases [19], aging [20], fertility loss [21], cancer [22], diabetes [23], and a vast spectrum of related pathologies.
In order to maintain a healthy mitochondrial network, cells must possess mechanisms to recycle and repair defective mitochondria. Mitochondrial quality control is an important aspect of cellular homeostasis, especially in postmitotic tissues, since it protects against the release of pro-apoptotic proteins, ROS production, and ineffective generation of ATP by damaged or aged mitochondria. The regulation of the mitochondrial life cycle and the maintenance of a robust functional network within cells are achieved through the maintenance of a precise balance between mitochondrial turnover and biogenesis [24]. Mitophagy, the selective degradation of damaged mitochondria by autophagy [25], is necessary for mitochondrial quality control and is the only known pathway to enable the turnover of whole mitochondrial genomes [26]. Mitochondria could be targeted for autophagic degradation for a variety of reasons, including basal turnover for recycling, starvation induced degradation, and degradation due to damage [25]. The core autophagic machinery is evolutionarily highly conserved, and the signaling cascades regulating autophagy are common to both bulk and selective autophagic processes [27]. Damaged or depolarized mitochondrial fragments are engulfed by autophagosomes, double-membraned organelles that are central to macroautophagy. Autophagy initiation depends on the recruitment and maturation of microtubule-associated protein-1 light chain-3 (LC3) to the autophagosomal membrane. In selective autophagy processes, such as mitophagy, the recognition of the cargo is mediated by cargo receptors that interact with matured LC3 (LC3-ll) through the LC3-interacting region (LIR) domain. Acidification and degradation of the cargo follow the formation of autolysosomes by fusion of the autophagosomes with lysosomes. Under physiological conditions, mitochondrial removal and replenishment reaches an equilibrium, allowing the maintenance of a constant mitochondrial volume. This balance is established according to tissue type, energy demand, activity level, stress, mutations, and abundance of nutrients [28].
Contrary to mitophagy, the mitochondrial unfolded protein response (UPRmt) is a mechanism aimed to preserve or repair damaged mitochondria. UPRmt is responsible for maintaining mitochondrial proteostasis via mitochondrial activation of a transcriptional program in the nuclear DNA [29]. This compensation system can be divided into three main pathways: activation of (1) chaperones, which boost refolding of misfolded proteins to restore them to their functional conformation and assist the folding of newly synthesized proteins; (2) proteases that are able to degrade aberrant proteins or aggregates; and (3) an antioxidant system that palliates ROS overproduction [30]. Although human UPRmt is not fully understood, it is gaining relevance in a variety of physiological processes on top of its canonical function, such as ageing, oxidative stress resistance, hematopoietic stem cell maintenance, glycolysis, antibacterial immunity, coenzyme Q biosynthesis, and mitochondrial fission [31,32]. Loss of mitochondrial proteostasis is the main UPRmt inducer. Accumulation of damaged proteins exceeding the protein-processing capacity of the chaperones and proteases in mitochondria would activate UPRmt, for instance [33]. Additionally, factors interfering with mitochondrial function promote UPRmt induction. Examples of these are the inhibition of complex I by rotenone [34], bacterial toxins [35], knockdown of quality control proteins [36], or generation of excess ROS by paraquat [37].
Nonetheless, if the UPRmt is unable to repair the mitochondrial damage, elimination of the entire mitochondria is promoted. Mitophagy is likely the last-resort response because it requires the cell to replace the organelles. The challenge remains to delineate the exact signaling components and the temporal characteristics of these two processes in the recovery of the mitochondrial network [38]. Finally, if mitophagy is impaired or the damage persists, cells undergo senescence and/or apoptosis [39] (Figure 1).
In this review, we focus on the growing interest of UPRmt induction as potential therapy for several mitochondria-related diseases and its possible consequences.

2. What Is the UPRmt?

Although UPRmt was discovered in mammalian cells [40,41], it has been thoroughly studied in C. elegans [42]. The stress-activated transcription factor 1 (ATFS-1) was identified in C. elegans as a key regulator of the UPRmt. ATFS-1 is normally imported into mitochondria, but during mitochondrial dysfunction, a percentage of ATFS-1 accumulates in the cytosol and then traffics to the nucleus, where it induces transcription of mitochondrial chaperones, proteases, and antioxidants [42,43]. Consistent with mediating a protective transcriptional response, worms lacking ATFS-1 incur respiratory defects during mitochondrial stress and higher susceptibility to mitochondrial perturbations [35,43]. This process is evolutionarily conserved [44], and there are three key protein regulators of human UPRmt: Activation Transcription Factor 5 (ATF5), Activation Transcription Factor 4 (ATF4), and C/EBP homologous protein (CHOP) [45,46]. There is still no consensus about how these proteins interact with each other or of their exact role in UPRmt regulation. What is indeed known is that all three are upregulated after mitochondrial damage, either by exposure to rotenone, mtDNA depletion, or mitochondrial proteins over-aggregation [45,47,48]. All three transcription factors are also involved in the integrated stress response (ISR), which is a conserved adaptive response that is activated by a wide variety of stressors [33]. In fact, in mammals, the ISR acts as an essential precursor of UPRmt activation. Sensing, surviving, and adapting to mitochondrial dysfunction depends on attenuated protein synthesis, privileged translation of transcription factors, and the activity of the ISR eucaryotic initiation factor 2 alpha (eIF2α) kinases [47].
Even though each of these transcription factors is implicated in promoting transcription of genes required to overcome mitochondrial dysfunction, it is unclear whether they function individually or in concert. Studies show that they regulate the expression of one another [48]. There are many examples of their role in restoring mitochondrial function: HeLa cells lacking a functional copy of ATF4 failed to upregulate several mitochondrial enzymes and exhibited a reduction in ATP-dependent respiration [49]; gene expression analysis of human and mouse tissues revealed a tight correlation between ATF4 induction and UPRmt responsive genes [45]; worms lacking ATFS-1 failed to induce chaperones’ expression in response to mitochondrial dysfunction were rescued by ATF5 expression but not ATF4 [44]; knockdown of ATF5 in HEK293 cells dampened induction of UPRmt responsive genes [50]; and elevated transcript levels of ATF4 and ATF5 were detected in mice harboring a mutation in the mitochondrial DNA helicase [51]. In addition, global transcriptomic analyses have validated the presence of CHOP-binding elements in many UPRmt gene promoters, while also revealing abundant ATF4-binding motifs [49].
In general, the most accepted pathway describes that ATF5 shuttles between the mitochondria and nuclei and induces the UPRmt to promote cell proliferation and mitochondrial functional recovery under mitochondrial stress. In addition, mitochondrial dysfunction upregulates CHOP, which can act as transcription factor that binds to the promoters of UPRmt-related genes, thereby inducing the expression of mitochondrial chaperones, proteases, and antioxidants [52]. Furthermore, ISR induction leads to the increased translation of ATF4, which activates many genes, including CHOP [53]. Nonetheless, recent studies expose that there are many other proteins implicated in the UPRmt, such as the whole sirtuin family [54]. For an in-depth description of the UPRmt molecular pathways, we recommend the recent reviews from Yun et al. [55], Naresh et al. [47], and Anderson et al. [33].
From a clinical point of view, UPRmt induction could be linked to a novel concept derived from an old statement: mitohormesis. Hormesis is defined as any adaptive mechanism exhibiting a biphasic dose response. Mitohormesis is a broad and diverse cytosolic and nuclear response that can be triggered by any of a variety of insults leading to mitochondrial stress. Interestingly, this response seems to induce a wide-ranging cytoprotective state, resulting in long-lasting metabolic and biochemical changes. Remarkably, rather than being harmful, these changes may reduce susceptibility to disease, as well as potentially determine lifespan [55]. In the context of mitohormesis, the activation of the UPRmt might provide biologically important short- and long-term adaptation against several pathologies. In fact, there is a growing body of evidence suggesting that activation of the UPRmt might be an important determinant of lifespan [56]. The clearest example of hormesis is physical exercise. From an objective point of view, physical exercise is considered a form of stress (ROS production, lactic fermentation, and muscle fibers’ damage), yet physical exercise by itself increases the production of antioxidant proteins. In fact, this response was not observed in individuals taking a combination of vitamin C and vitamin E. The salutary benefits of exercise appeared to be inhibited in subjects given antioxidant supplements [57]. Therefore, the mild oxidative stress caused by exercise promotes cells to produce antioxidants which will protect the organism; however, antioxidants will block the oxidative signal making the organism prone to oxidative damage in the long term. Ranging down to mitochondria, mild and isolated stress could lead to the activation of several compensation mechanisms, optimize mitochondrial function, and therefore improve cellular fitness. A schematical representation is shown in Figure 2.
However, the contribution of particular genes in the UPRmt response pathway should be carefully evaluated, taking into account that many of them are constitutive genes that are required for the maintenance of basic cellular function, and their up- or downregulation may produce numerous changes in the cells, including the mentioned UPRmt.

3. Mitochondrial Diseases

Mitochondrial diseases encompass a broad spectrum of muscular and neurodegenerative disorders, both chronic and progressive, caused by mutations in nuclear (nDNA) or mitochondrial (mtDNA) DNA [58]. The prevalence of these diseases has been established at 1:5000 [59]; however, this number is increasing due to the standardization of whole-exome sequencing [18]. Most OXPHOS disorders in children are a consequence of the mutation of nuclear DNA, and they are transmitted as autosomal recessive traits, usually with severe phenotypes and a fatal outcome. Among the maternally inherited pathogenic mtDNA mutations, more than 50% have been identified in genes encoded by mitochondrial transfer RNAs (mt-tRNA) (MTT genes). At first sight, it might seem paradoxical to induce a mild mitochondria stress in a general mitochondrial dysfunction context; however, recent studies present this strategy as a new alternative treatment for mitochondrial diseases.
Perry et al. showed that the tetracycline antibiotics family increased cell survival and fitness in MELAS cybrids and Rieske cells (Knockout Complex III mouse fibroblasts) under glucose restriction [60]. Specifically, doxycycline improved survival in wild-type cells treated with piericidin (complex I inhibitor) or antimycin (complex III inhibitor) during glucose deprivation. In addition to tetracyclines, the anti-parasitic agent pentamidine and the antibiotic retapamulin also scored positive on the screening in MELAS and mutant ND1 cybrid cells. They propose a “mitohormetic effect” by the induction of ATF4 (UPRmt) and p-eIF2α (ISR) as mechanism of action; however, in some cases, cell survival was independent of ATF4, suggesting the high complexity of this pathway. Suarez-Rivero et al. also demonstrated that tetracycline treatment boosts the production of UPRmt-related proteins and promotes the activation of pathways involving cAMP and cGMP, which might be implicated in mitochondrial compensatory mechanisms comprising sirtuins and chaperones’ activity [61,62,63]. Tetracycline treatment and the subsequent activation of UPRmt, the increased number of chaperones, and mitochondrial auxiliary proteins may promote the stability of a fraction of mutated mitochondrial proteins which would carry out their function to some extent [64]. All of these antibiotics have one factor in common: they are inhibitors of the mitochondrial translation at different degrees [65,66,67].
Given the bacterial origin of mammalian cells’ mitochondria and the fact that they conserve prokaryotic features such as the 55S or 60S ribosomes, it is understandable that mitochondria are exceptionally sensitive to antibiotics [68]. In this way, mitohormesis can be triggered by a partial inhibition of mitochondrial translation promoting a beneficial retrograde signaling response including the modulation of mitochondrial dynamics, the expression of nuclear and mitochondrial-encoded genes, the antioxidant response, stimulating mitochondrial function, and boosting cellular defense mechanisms that increase stress resistance [69].
Although the research of Perry et al. and Suarez-Rivero et al. is promising, antibiotic use in mitochondrial diseases is still in debate [70]. Several antibiotics can worsen the conditions of individuals bearing mtDNA mutations. Aminoglycosides, for instance, induce ototoxic hearing loss in subjects with mutations in the 12S rRNA gene [71]. Although aminoglycosides typically display specificity toward prokaryotes over eukaryotes, human mitochondrial ribosomes that have A1408 and G1491 at analogous positions exhibit higher resemblance to their bacterial counterparts. This similarity is likely responsible for some of the adverse effects shown by aminoglycosides [72].
It has recently been demonstrated that pterostilbene in combination with mitochondrial cofactors treatment activates SIRT3 and UPRmt as compensatory mechanisms, as well as enhances sirtuins’ levels and mitochondrial activity in several cell models of mitochondrial diseases [73]. In contrast to long-term antibiotic treatment, pterostilbene is considered safe for human consumption and presents numerous well-known features, such as a prominent antioxidant activity and a high anti-inflammatory potential [74]. Furthermore, it has been reported to prolong lifespan in several animal models [75] due to its neuroprotective [76] and cardioprotective [77] properties. At the molecular level, pterostilbene activates sirtuins [75] and has been linked to AMP-activated protein kinase (AMPK) [78] and Nuclear factor erythroid 2-related factor 2 (Nrf2) by recent studies [79], suggesting that it could be a promising compound to maintain mitochondrial homeostasis.

4. Neurodegeneration

Neurodegenerative diseases are a widely heterogeneous group of disorders characterized by the progressive degeneration of the structure and function of the central and peripheral nervous system. Although they affect millions of patients worldwide and have a profound impact on families and their community, the pathogenic mechanisms underlying these conditions remain unclear. The therapeutic options for patients are scarce and merely palliative. Nevertheless, there is currently great interest among the scientific community in regard to the identification of biomarkers or concrete genetic mutations that might help clinicians anticipate the onset of these diseases in a way to either treat them when still reversible or slow down their development [80]. This is especially relevant given the fact that the protein abnormalities that are characteristic of these diseases are present in patients long before the clinical symptoms become noticeable [81,82].
Most neurodegenerative diseases share in common the accumulation of misfolded proteins; however, they also possess traits associated with progressive neuronal dysfunction and death, such as proteotoxic stress, abnormalities in autophagy, neuroinflammation, apoptosis, ROS production, and mitochondrial dysfunction [83]. Mitochondrial function is of utmost importance for neurons because limited glycolysis causes them to rely exclusively on OXPHOS for energy production. Given that the long neuronal axons require energy transport over long distances, and because synaptic transmission is dependent on calcium signaling, mitochondrial performance needs to be tightly regulated [84]. Taken together, there is compelling evidence to suggest a relevant mitochondrial involvement in the pathogenesis of several neurodegenerative diseases, with their role being particularly significant in Parkinson’s disease (PD). According to several studies, patients of this condition present a selective deficiency of respiratory chain complex I, which is most remarkable in the substantia nigra [85,86]. In Alzheimer’s diseases (AD), it has been demonstrated that patients present impaired oxygen consumption in the brain, further adding to the notion that bioenergetic dysfunction and mitochondrial impairment are common features [87]. In fact, impairment of several mitochondrial enzymes has been detected in AD patients, where respiratory chain complex IV is most affected [88]. Complex IV activity was found to be significantly reduced in the brain tissue of AD patients, which presented a general decrease of electron transport chain function [89]. Apart from AD and PD, mitochondrial perturbations have been reported in many other neurodegenerative diseases, among which amyotrophic lateral sclerosis (ALS) and Huntington’s disease (HD) deserve special mention [90]. Nevertheless, a critical question that remains unanswered is whether mitochondrial dysfunction contributes to the initial stages of neurodegeneration, being primarily responsible for the onset of pathogenesis or just a secondary feature arising from alternative phenomena such as the accumulation of misfolded proteins or cellular stress. Either way, it has been demonstrated that targeting mitochondrial dysfunction is a promising therapeutic strategy for neurodegenerative conditions [91].

4.1. Parkinson’s Diseases

PD is an age-associated neurodegenerative movement disorder that is mainly caused by the death of dopaminergic neurons in the brain substantia nigra [92]. UPRmt may regulate the occurrence and development of PD. It has been shown that a PINK1 mutation can activate ATFS-1-dependent UPRmt and promote dopaminergic neuron survival in a PD worm model [93]. Ginseng protein protected against neurodegeneration by inducing UPRmt in a PINK1 fly model of PD [94]. Dastidar SG et al. demonstrated that activation of 4E-BP1 correlates with UPRmt induction, which reduces PD associated neurotoxicity in mouse neurons [95]. In fact, mutant LRRK2G2019S, the most common PD-causing allele in humans, results in reduced 4E-BP1 function and may contribute to PD pathogenesis by UPRmt underactivation [96]. In addition, altered omi and HtrA2 protein, UPRmt-related proteases, causes neuropathy with the same characteristics as PD [97].

4.2. Alzheimer’s Disease

Alzheimer’s disease (AD) is the most common form of dementia [98]. Clinically, AD is defined by cognitive impairment that is pervasive enough to interfere with a person’s ability to work or complete daily activities. The main histologic characteristics include amyloid plaques that largely contain amyloid beta (Aβ) protein and neurofibrillary tangles (NFTs) that consist of hyperphosphorylated tau protein, both of which cause premature neuronal death. Current AD treatments confer a slight benefit but ultimately do not prevent progression [99].
Mitochondria in AD show numerous changes [100]. First, complex-IV-reduced activity has consistently been observed across numerous tissues. In brain tissue specifically, the decrease in complex IV subunits has been linked with disease progression [101]. Mitochondrial surface area also decreases, cristae structures are altered, and an increased variability in mitochondrial shape has been observed [102]. There are changes in fusion and fission mitochondrial proteins that appear to affect mitochondrial localization in AD neurons [103]. Indeed, neuronal cultures with AD-relevant fission/fusion protein alterations recapitulate AD-like changes in mitochondrial distribution [104].
Compared to cognitively intact controls, the expression of UPRmt-related genes is increased by 40–60% in sporadic AD subjects and 70–90% in familial AD, respectively [105]. Sorrentino et al. showed that increasing mitochondrial proteostasis by targeting mitochondrial translation and mitophagy both pharmacologically and genetically increases the fitness and lifespan of worms, cells, and mouse models of AD and reduces amyloid aggregation [106]. In addition, UPRmt is strongly activated and exerts a protective role against Aβ protein toxicity in PITRM1-knockout human cells. In line with this, pharmacological inhibition of UPRmt exacerbates the Aβ proteotoxicity in PITRM1-knockout human cells [107]. Consistent with this study, UPRmt is activated in the mouse brains and human SHSY5Y cells after Aβ treatment, while the inhibition of UPRmt aggravates cytotoxic effects of Aβ [108]. Treatments which increase the NAD+ pool, a sirtuin cofactor, alleviate protein toxicity and improve the memory of AD mouse by activating the UPRmt. Taken together, these studies provide the evidence of the protective role of UPRmt on AD. Furthermore, the activation of UPRmt induced by Aβ may depend on the regulation of two signaling pathways: the mevalonic acid pathway and ceramide pathway [109]. Therefore, the activation of UPRmt can relieve the symptoms of AD.

4.3. Huntington’s Disease

Huntington’s disease (HD) is a fatal and inherited neurodegenerative disorder that progresses for 15–20 years after the initial onset [110]. The main cause of HD is the expanded CAG repeats encoding polyglutamine (polyQ) in the N-terminus of the huntingtin (Htt) protein [111]. The genetic cause of HD was identified more than 20 years ago. However, the underlying mechanisms leading to the pathogenesis of HD remain elusive.
Accumulating evidence suggests that mitochondrial dysfunction plays an important role in the pathogenesis of HD [112]. For instance, mutant Htt associates with the outer mitochondrial membrane in different HD models, resulting in mitochondrial permeability transition pore opening, calcium disturbance, reduced ATP production, mitochondrial membrane potential loss, increased ROS production, and premature release of cytochrome c [113]. How mutant Htt can affect selectively medium spiny striatal neurons despite its ubiquitous expression is a key question that still remains unanswered. One of the several theories that have been proposed points again to mitochondria, suggesting that medium spiny neurons, characterized by particularly high-energy demands, are especially susceptible to mutant Huntingtin-induced mitochondrial dysfunction and inhibition of respiration [114].
Fu et al. [115] found that mutant Htt suppressed the expression of ABCB10, a component of the UPRmt [116], in various HD models by impairing its mRNA stability. Deletion of ABCB10 induced ROS production and cell death in HD mouse striatal cells. Moreover, ABCB10 was required for CHOP activation under mitochondrial stress. They also showed that chaperone HSP60 and protease Clpp, two downstream genes of CHOP [41], were decreased in HD cells.

4.4. Amyotrophic Lateral Sclerosis

Amyotrophic lateral sclerosis (ALS) is caused by the progressive degeneration of motor neurons in the spinal cord, the brain stem, and the motor cortex. Motor neuron death prompts muscle weakness and paralysis, causing death in 1–5 years from the time of symptoms’ onset. The primary identified ALS-linked gene was superoxide dismutase 1 (SOD1), an antioxidant protein [117]. However, mutations in several other genes have been reported, and various molecular pathways have been associated to neuronal death in ALS [118]. The only two disease-modifying potential therapies currently approved for ALS are riluzole and edaravone, but the exact neuroprotective action of these drugs is unknown, and they present no obvious improvement in ALS patients’ health [119]. The absence of alternative drugs for the treatment of ALS indicates the need for the implementation of novel therapeutic strategies
ALS-associated mitochondrial dysfunction comes in many shapes, including defective OXPHOS, ROS production, impaired calcium buffering capacity, and defective mitochondrial dynamics. In addition, mitochondrial dysfunction appears to be directly or indirectly linked to all mechanisms of toxicity associated with ALS, including excitotoxicity, loss of protein homeostasis, and defective axonal transport [120].
TDP-43 proteinopathy is characterized by the presence of TDP-43 immunoreactive inclusion bodies in the affected tissues. This alteration is present in several neurodegenerative diseases with high predominance in ALS [121,122]. Peng et al. showed that LonP1, one of the key mitochondrial proteases constituting the UPRmt response, protects against TDP-43-induced cytotoxicity and neurodegeneration. They suggest the activation of LonP1 via UPRmt as a treatment for TDP-43 proteinopathy, protecting against or reversing mitochondrial damage as a potential therapeutic approach to these neurodegenerative disorders [123]. Qi et al. demonstrated that the mitochondrial function pathway was disrupted in the brain of SOD1G93A mice, and the replenishment of intracellular NAD+, by providing nicotinamide, could reduce neurotoxic protein aggregates of mitochondria in the brain of SOD1G93A mice [124]. Moreover, they concluded that nicotinamide might modulate mitochondrial proteostasis and improve adult neurogenesis through activation of the UPRmt signaling in the brain of SOD1G93A mice. Curiously, it has been reported that UPRmt is transiently activated in the spinal cord of ALS mice models in the late symptomatic phase. However, there is a significant sex difference in the activation of the UPRmt: it was significantly activated in female SOD1G93A mice but not in males [125]. Finally, Gomez et al. hypothesized that SOD1 may have multiple functions: antioxidant, UPRmt regulator, and gene transcription enhancer [126]. Taken together, these functions place SOD1 as a key regulator of the communication between the nucleus and mitochondria.

5. Heart Diseases

Maintaining mitochondrial quality in cardiomyocytes is essential, given that they produce 8% of the total ATP consumed by the organism and power the constant contraction and relaxation of the myocardium [127]. In fact, mitochondrial abnormalities are a common feature to all types of cardiomyopathies [128]. Mitochondria with aberrant structures are commonly observed in cardiac cells of all forms of heart disease [129]. Furthermore, cardiomyopathy has been reported due to mitochondrial disease [130], and mitochondrial dysfunction has been linked with hypertension. It is nowadays clear that the importance of functional mitochondria for cardiac health is undeniable. One of the main mechanisms ensuring mitochondrial fitness is proteostasis. It finetunes biogenesis, folding, and degradation of mitochondrial proteins, processes which are commonly altered during cardiac stress. It has been demonstrated that stimulation of UPRmt improves mitochondrial function and reduces cardiac damage in response to stress. In this regard, a recent study proved that UPRmt enhancement with small-molecule agents ameliorates mitochondrial and contractile dysfunction in the murine heart. This is the reason why its authors proposed UPRmt as a potential therapeutic target in heart failure [131].
Wang et al. demonstrated the relevance of mitophagy and UPRmt in myocardial injuries and stress [132]. They proposed UPRmt activation with oligomycin, a complex V inhibitor, as a mechanism to reduce sepsis-mediated mitochondrial injury and myocardial dysfunction; however, this cardioprotective effect was imperceptible in mitophagy-inhibited mice. On the other hand, when UPRmt was inhibited, mitophagy-mediated protection of mitochondria and cardiomyocytes was partly blunted. Taken together, their observations suggest that both UPRmt and mitophagy are slightly activated by myocardial stress and that they work together to sustain mitochondrial performance and cardiac function.

6. Lifespan

Upregulation of UPRmt has been proposed as the common pathway in lifespan extension induced by mitochondrial defects [133]. Studies in C. elegans revealed that the induction of mitochondrial stress at an early age resulted in lifespan extension and UPRmt upregulation which persisted even after the stress inductor was withdrawn [134]. This observation suggests that UPRmt upregulation is directly responsible for lifespan extension in worms. Supporting this idea, Yokoyama et al. demonstrated that transgenic worms expressing the nematode UPRmt protein Hsp70F live over 40% more than wild-type animals [135]. To prove whether UPRmt upregulation also influenced lifespan extension in mammals, a recent study assessed UPRmt activation in a long-lived mouse model, Snell dwarf mice. These mice have a single point mutation in Pit1 and show more than 40% increase in mean and maximal lifespan [136]. The results of the study show that UPRmt was upregulated in both living mice and cultured cells. Moreover, the authors reported elevated basal and post-stress UPRmt levels in primary fibroblasts derived from Snell mice. However, at the organismal level, UPRmt was upregulated only after doxycycline-induced mitochondrial stress exposure [137].
The ability of UPRmt activation to prolong lifespan can be explained by several reasons: its positive impact on mitochondrial function, its ability to reduce mitochondrial metabolic by-products, its modulation of insulin/IGF and mTOR signaling, and its involvement in hormetic cellular maintenance [138]. Following UPRmt activation, mitochondria undergo structural and functional changes. The mitochondrial network becomes fragmented [133], and cellular oxygen consumption is significantly reduced [139,140]. This particular phenotype may be a coordinated response to stress that favors mitophagy and lowers respiration to enable mitochondrial repair. Taken together, these measures would ultimately lead to enhanced homeostasis and cell survival, consequently increasing lifespan.
On top of this, UPRmt is responsible for changes in the production of a series of metabolic by-products with relevant physiological functions. Amidst these, ROS is the most widely studied. It elicits varied cellular stress responses, and by doing so, it influences homeostasis and longevity [141]. Additionally, the UPRmt has an impact on the mevalonate pathway, an important target of statins, and on NAD+, which is the substrate for poly (ADP-ribose) polymerases (PARPs) and sirtuin deacetylases and acts as major longevity regulator [56,142]. The link of UPRmt to longevity might also be explained by its effect on insulin signaling. Reduced insulin/IGF-1 signaling has been directly linked to a series of protective mechanisms such as regulation of endoplasmic reticulum unfolded protein response (UPRER) or activation of heat shock factor 1 (HIF-1), which boosts the expression of small heat shock proteins, leading to lifespan extension in worms [143]. In this line, it has been proven that impairment of the insulin/IGF-1 signaling induces a transient ROS stress signal that potentially activates UPRmt and results in prolonged lifespan in C. elegans [144]. Improving the level of cellular NAD+ in mice not only increases mitochondrial function but also induces the expression of UPRmt-associated genes, preventing skeletal muscle stem cells from aging and enhancing life span [145].
Although several different routes have been proposed to explain UPRmt implication in longevity, there is the possibility that its lifespan-prolonging ability is the result of improved cellular housekeeping. It is extensively known that loss of proteostasis and the consequent accumulation of misfolded proteins are a fundamental cause of ageing. UPRmt activation would palliate cellular damage in this context by restoring protein quality control. Not only does it assist correct protein folding via enhancement of chaperones’ expression, but it also promotes refolding of aberrant products and sequestration of protein aggregates in less cytotoxic states [146].

7. Therapeutic Concerns

Although UPRmt activation seems to be a promising therapeutic approach for mitochondria- related disorders, it does not come without a risk. Prolonged overactivation of UPRmt can also have severe side effects, as proven by recent studies. Lin et al. demonstrated that, in the context of mtDNA heteroplasmy, activation of the UPRmt and constant ATFS-1 signaling maintain and propagate deleterious mitochondrial genomes in C. elegans. This observation led the authors to dispute the role of UPRmt activation in longevity, especially in individuals with a heteroplasmic background, such as inherited mitochondrial mutations or cancer cells. Nevertheless, they point out that it is still to be determined whether UPRmt in mammals acts in a similar way to in C. elegans [147].
In this line, Martinez et al. reported a detrimental effect of constitutive UPRmt signaling in C. elegans dopaminergic neurons after overexpression or overactivation of ATFS-1. Interestingly, they could not find a clear link between the cell death phenotype observed in dopaminergic neurons expressing ATFS-1 and mitophagy, which in C. elegans is primarily regulated by PINK1. Knockdown PINK1 worms did not present signs of neuronal cell death. This led the authors to point out that mitophagy impairment does not seem to be the mechanism responsible for neuronal death elicited by ATFS-1 overactivation. Rotenone treatment, nevertheless, had a neuroprotective attenuation effect dependent on PINK1. Overall, these results indicate that ATFS-1 overexpression might promote the accumulation of defective mitochondria in dopaminergic neurons, leading to abnormal cellular homeostasis and cell death. The phenotype can be reversed by increasing aberrant mitochondria clearance via mitophagy [148].
Regarding aging, most studies agree that UPRmt contributes to health and longevity [133], including neurons where UPRmt can activate protective cell non-autonomous signals and epigenetically rewire animal models to live longer [149]. However, in some circumstances, UPRmt seems to promote aging, as pointed out by Angeli S et al. [150]. They discovered in C. elegans that certain mitochondrial insults during development lead to the lasting activation of the UPRmt and are associated with longevity; however, the same stimuli during adulthood induce premature aging. In this case, suppression of the UPRmt via genetic or pharmacological interventions is protective. These results, again, evidence the complexity of UPRmt pathways.
Constitutive UPRmt activation can also be extremely detrimental for cancer patients. It is widely known that cancer cells make use of numerous stress response pathways to counteract and survive endogenous, exogenous, and environmental stresses. Among them, UPRmt may help cancer cells clear excessive cellular damage that could eventually lead to apoptosis. Similarly to cardiomyocytes, cancer cells use this stress response to ensure their survival and proliferation [151]. In fact, HSP60, a key chaperone of UPRmt activation process across different species, is overexpressed in several cancer types such as acute myeloid leukemia, pancreatic ductal adenocarcinoma, ovarian carcinoma, breast ductal invasive carcinoma, prostate adenocarcinoma, and others [48,152]. By folding and refolding oncoproteins and denatured/misfolded proteins within mitochondria, UPRmt facilitates cancer growth and increases the apoptotic threshold of cancer cells [153]. To better understand the implication of UPRmt and HSP60 in cancer, Tsai et al. proved that HSP60 overexpression significantly increases cellular migration and invasion in vitro, as well as increased tumor volume and metastasis in vivo [154]. Moreover, it has been reported that HSP60 plays a crucial role in the metastasis of pancreatic cancer cells [155] and increases motility in breast cancer cells [156]. Thus, in cancer cells, the UPRmt is exploited for mitochondrial repair and tumor growth, invasion, and metastasis promotion. In fact, disrupting proteostasis in cancer cells by targeting UPRmt constitutes a novel anticancer therapeutic strategy [157]. In addition, cancer stem cells (CSCs) represent a highly tumorigenic group of cells in primary tumors [158]. Mitochondrial changes in CSCs, including morphological changes, abnormal activation of signaling pathways, mitochondrial dysfunctions, production of ROS, enhanced mitophagy, and UPRmt modulation, are key regulators of CSC proliferation and apoptosis and are also among the reasons for the failure of antitumor treatments [159]. Therefore, targeting UPRmt in CSCs can be essential for the effective treatment of cancer [160].

8. Conclusions

The potential of UPRmt activation therapies is undeniable. The UPRmt is a fundamental stress response that improves cellular fitness and proteostasis, both of which are leading causes of very diverse conditions when impaired. From mitochondrial diseases to neurodegeneration or cardiovascular disorders, mitochondrial dysfunction has been identified as a pivotal factor for disease development (Table 1). Studies have demonstrated the therapeutic efficacy of UPRmt, yet no specifical activators of UPRmt have been applied for patient treatment so far. The side effects and deleterious repercussions of the sustained activation of UPRmt in patients are important concerns for clinicians. Therefore, efforts have been made to identify safe, well-studied drugs that succeed in activating this mitochondrial response. Nevertheless, further studies tackling the safety profile of UPRmt activators and their possible long-term side effects need to be conducted to ascertain that patients will not undergo risks when adopting these novel therapeutic approaches.

Author Contributions

Conceptualization and writing, J.A.S.-A., J.M.S.-R., C.J.P.-M. and M.Á.-C.; review and editing, S.P.-C., I.V.-G., M.T.-R., A.S.-C., M.M.-C., D.R.-L., P.C.-H. and R.P.-P. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by FIS PI16/00786 (2016) and FIS PI19/00377 (2019) grants; the Ministerio de Sanidad, Spain; and the Fondo Europeo de Desarrollo Regional (FEDER Unión Europea), Spanish Ministry of Education, Culture, and Sport. This activity was co-financed by the European Regional Development Fund (ERDF) and by the Regional Ministry of Economic Transformation, Industry, Knowledge, and Universities of the Junta de Andalucía, within the framework of the ERDF Andalusia operational program 2014–2020 Thematic objective “01—Reinforcement of research, technological development and innovation” through the reference research project CTS-5725 and PY18-850.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We acknowledge the support of “Ayudas para la Formación de Profesorado Universitario” (FPU/MINECO), AEPMI (Asociación de Enfermos de Patología Mitocondrial), Fúndación Mencia, Fundación Antonio Guerrero, FEDER (Federación Española de Enfermedades Raras), ASANOL association, Yo Nemálínica Association, KAT6A Association, Superauténticos Association, AMPELA Association, Fundación MERCK Salud, and Fundación MEHUER/Colegio Oficial de Farmacéuticos.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

4E-BP1Eukaryotic translation initiation factor 4E (eIF4E)-binding protein 1
ABCB10ATP-binding cassette sub-family B member 10
ADAlzheimer’s disease
ALSamyotrophic lateral sclerosis
AMPKAMP-activated protein kinase
ATF4Activating Transcription Factor 4
ATF5Activating Transcription Factor 5
ATFS-1Activated transcription factor 1
ATPadenosine triphosphate
amyloid beta protein
cAMPcyclic adenosine monophosphate
cGMPcyclic guanosine monophosphate
CHOPC/EBP homologous protein
CSCscancer stem cells
Eif2αeucaryotic initiation factor 2 alpha
HDHuntington’s disease
HSPheat shock protein
HtrA2HtrA Serine Peptidase 2
Htthuntingtin
IGF-1insulin growth factor 1
ISRintegrated stress response
LC3microtubule-associated protein-1 light chain-3
LIRLC3-interacting region
MELASmitochondrial encephalomyopathy, lactic acidosis, and stroke-like episodes
mtDNAmitochondrial DNA
ND1NADH–ubiquinone oxidoreductase chain 1
NFTneurofibrillary tangles
Nrf2Nuclear factor erythroid 2–related factor 2
OXPHOSoxidation phosphorylation
P-Eif2αPhosphorylated Eukaryotic Initiation Factor 2 alpha
PDParkinson’s disease
PINK1PTEN-induced kinase 1
PITRM1pitrilysin metallopeptidase 1
ROSreactive oxygen species
SIRTsirtuin
TDP-43TAR DNA-binding protein 43
UPRmtmitochondrial unfolded protein response

References

  1. Gut, P.; Verdin, E. The nexus of chromatin regulation and intermediary metabolism. Nature 2013, 502, 489–498. [Google Scholar] [CrossRef] [PubMed]
  2. Hatting, M.; Tavares, C.D.J.; Sharabi, K.; Rines, A.K.; Puigserver, P. Insulin regulation of gluconeogenesis. Ann. N. Y. Acad. Sci. 2018, 1411, 21–35. [Google Scholar] [CrossRef] [PubMed]
  3. Park, J.-M.; Jo, S.-H.; Kim, M.-Y.; Kim, T.-H.; Ahn, Y.-H. Role of transcription factor acetylation in the regulation of metabolic homeostasis. Protein Cell 2015, 6, 804–813. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Ballabio, A.; Bonifacino, J.S. Lysosomes as dynamic regulators of cell and organismal homeostasis. Nat. Rev. Mol. Cell Biol. 2020, 21, 101–118. [Google Scholar] [CrossRef] [PubMed]
  5. Marchi, S.; Patergnani, S.; Missiroli, S.; Morciano, G.; Rimessi, A.; Wieckowski, M.R.; Giorgi, C.; Pinton, P. Mitochondrial and endoplasmic reticulum calcium homeostasis and cell death. Cell Calcium 2018, 69, 62–72. [Google Scholar] [CrossRef]
  6. Danziger, J.; Zeidel, M.L. Osmotic homeostasis. Clin. J. Am. Soc. Nephrol. 2015, 10, 852–862. [Google Scholar] [CrossRef] [Green Version]
  7. Fernández-Fernández, M.R.; Valpuesta, J.M. Hsp70 chaperone: A master player in protein homeostasis. F1000Research 2018, 7, 1497. [Google Scholar] [CrossRef] [Green Version]
  8. Dan Dunn, J.; Alvarez, L.A.; Zhang, X.; Soldati, T. Reactive oxygen species and mitochondria: A nexus of cellular homeostasis. Redox Biol. 2015, 6, 472–485. [Google Scholar] [CrossRef]
  9. Hederstedt, L. Heme A biosynthesis. Biochim. Biophys. Acta 2012, 1817, 920–927. [Google Scholar] [CrossRef] [Green Version]
  10. Nakhle, J.; Rodriguez, A.-M.; Vignais, M.-L. Multifaceted roles of mitochondrial components and metabolites in metabolic diseases and cancer. Int. J. Mol. Sci. 2020, 21, 4405. [Google Scholar] [CrossRef]
  11. Wang, J.; Pantopoulos, K. Regulation of cellular iron metabolism. Biochem. J. 2011, 434, 365–381. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Delierneux, C.; Kouba, S.; Shanmughapriya, S.; Potier-Cartereau, M.; Trebak, M.; Hempel, N. Mitochondrial calcium regulation of redox signaling in cancer. Cells 2020, 9, 432. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Yin, F.; Boveris, A.; Cadenas, E. Mitochondrial energy metabolism and redox signaling in brain aging and neurodegeneration. Antioxid. Redox Signal. 2014, 20, 353–371. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Estaquier, J.; Vallette, F.; Vayssiere, J.L.; Mignotte, B. The mitochondrial pathways of apoptosis. Adv. Exp. Med. Biol. 2012, 942, 157–183. [Google Scholar]
  15. Riley, J.S.; Tait, S.W. Mitochondrial DNA in inflammation and immunity. EMBO Rep. 2020, 21, e49799. [Google Scholar] [CrossRef]
  16. Bratic, A.; Larsson, N.-G. The role of mitochondria in aging. J. Clin. Investig. 2013, 123, 951–957. [Google Scholar] [CrossRef] [Green Version]
  17. Stefanatos, R.; Sanz, A. The role of mitochondrial ROS in the aging brain. FEBS Lett. 2018, 592, 743–758. [Google Scholar] [CrossRef] [Green Version]
  18. Molnar, M.J.; Kovacs, G.G. Mitochondrial diseases. Handb. Clin. Neurol. 2017, 145, 147–155. [Google Scholar]
  19. Bose, A.; Beal, M.F. Mitochondrial dysfunction in Parkinson’s disease. J. Neurochem. 2016, 139 (Suppl. 1), 216–231. [Google Scholar] [CrossRef]
  20. Balaban, R.S.; Nemoto, S.; Finkel, T. Mitochondria, oxidants, and aging. Cell 2005, 120, 483–495. [Google Scholar] [CrossRef] [Green Version]
  21. Vertika, S.; Singh, K.K.; Rajender, S. Mitochondria, spermatogenesis, and male infertility—An update. Mitochondrion 2020, 54, 26–40. [Google Scholar] [CrossRef] [PubMed]
  22. Wallace, D.C. Mitochondria and cancer. Nat. Rev. Cancer 2012, 12, 685–698. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Rovira-Llopis, S.; Bañuls, C.; Diaz-Morales, N.; Hernandez-Mijares, A.; Rocha, M.; Victor, V.M. Mitochondrial dynamics in type 2 diabetes: Pathophysiological implications. Redox Biol. 2017, 11, 637–645. [Google Scholar] [CrossRef]
  24. Scheibye-Knudsen, M.; Fang, E.F.; Croteau, D.L.; Wilson, D.M.; Bohr, V.A. Protecting the mitochondrial powerhouse. Trends Cell Biol. 2015, 25, 158–170. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Dombi, E.; Mortiboys, H.; Poulton, J. Modulating mitophagy in mitochondrial disease. Curr. Med. Chem. 2018, 25, 5597–5612. [Google Scholar] [CrossRef] [PubMed]
  26. Valenci, I.; Yonai, L.; Bar-Yaacov, D.; Mishmar, D.; Ben-Zvi, A. Parkin modulates heteroplasmy of truncated mtDNA in Caenorhabditis elegans. Mitochondrion 2015, 20, 64–70. [Google Scholar] [CrossRef]
  27. Youle, R.J.; Narendra, D.P. Mechanisms of mitophagy. Nat. Rev. Mol. Cell Biol. 2011, 12, 9–14. [Google Scholar] [CrossRef]
  28. Pickles, S.; Vigié, P.; Youle, R.J. Mitophagy and quality control mechanisms in mitochondrial maintenance. Curr. Biol. 2018, 28, R170–R185. [Google Scholar] [CrossRef] [Green Version]
  29. Zhu, L.; Zhou, Q.; He, L.; Chen, L. Mitochondrial unfolded protein response: An emerging pathway in human diseases. Free Radic. Biol. Med. 2021, 163, 125–134. [Google Scholar] [CrossRef]
  30. Muñoz-Carvajal, F.; Sanhueza, M. The mitochondrial unfolded protein response: A hinge between healthy and pathological. Aging Front. Aging Neurosci. 2020, 12, 581849. [Google Scholar] [CrossRef]
  31. Yi, H.-S.; Chang, J.Y.; Shong, M. The mitochondrial unfolded protein response and mitohormesis: A perspective on metabolic diseases. J. Mol. Endocrinol. 2018, 61, R91–R105. [Google Scholar] [CrossRef] [PubMed]
  32. Wang, S.; Gao, K.; Liu, Y. UPR mt coordinates immunity to maintain mitochondrial homeostasis and animal fitness. Mitochondrion 2018, 41, 9–13. [Google Scholar] [CrossRef] [PubMed]
  33. Anderson, N.S.; Haynes, C.M. Folding the mitochondrial UPR into the integrated stress response. Trends Cell Biol. 2020, 30, 428–439. [Google Scholar] [CrossRef]
  34. Cabezas, R.; El-Bachá, R.S.; González, J.; Barreto, G.E. Mitochondrial functions in astrocytes: Neuroprotective implications from oxidative damage by rotenone. Neurosci. Res. 2012, 74, 80–90. [Google Scholar] [CrossRef] [PubMed]
  35. Pellegrino, M.W.; Nargund, A.M.; Kirienko, N.; Gillis, R.; Fiorese, C.; Haynes, C.M. Mitochondrial UPR-regulated innate immunity provides resistance to pathogen infection. Nature 2014, 516, 414–417. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Yoneda, T.; Benedetti, C.; Urano, F.; Clark, S.G.; Harding, H.; Ron, D. Compartment-specific perturbation of protein handling activates genes encoding mitochondrial chaperones. J. Cell Sci. 2004, 117, 4055–4066. [Google Scholar] [CrossRef] [Green Version]
  37. Runkel, E.D.; Liu, S.; Baumeister, R.; Schulze, E. Surveillance-activated defenses block the ros-induced mitochondrial unfolded protein response. PLoS Genet. 2013, 9, e1003346. [Google Scholar] [CrossRef] [Green Version]
  38. Smyrnias, I. The mitochondrial unfolded protein response and its diverse roles in cellular stress. Int. J. Biochem. Cell Biol. 2021, 133, 105934. [Google Scholar] [CrossRef]
  39. Onishi, M.; Yamano, K.; Sato, M.; Matsuda, N.; Okamoto, K. Molecular mechanisms and physiological functions of mitophagy. EMBO J. 2021, 40, e104705. [Google Scholar] [CrossRef]
  40. Martinus, R.D.; Garth, G.P.; Webster, T.L.; Cartwright, P.; Naylor, D.J.; Høj, P.B.; Hoogenraad, N.J. Selective Induction of Mitochondrial Chaperones in Response to Loss of the Mitochondrial Genome. JBIC J. Biol. Inorg. Chem. 1996, 240, 98–103. [Google Scholar] [CrossRef]
  41. Zhao, Q.; Wang, J.; Levichkin, I.V.; Stasinopoulos, S.; Ryan, M.; Hoogenraad, N.J. A mitochondrial specific stress response in mammalian cells. EMBO J. 2002, 21, 4411–4419. [Google Scholar] [CrossRef] [PubMed]
  42. Nargund, A.M.; Pellegrino, M.W.; Fiorese, C.J.; Baker, B.M.; Haynes, C.M. Mitochondrial import efficiency of ATFS-1 regulates mitochondrial UPR activation. Science 2012, 337, 587–590. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Nargund, A.M.; Fiorese, C.J.; Pellegrino, M.W.; Deng, P.; Haynes, C.M. Mitochondrial and nuclear accumulation of the transcription factor ATFS-1 promotes OXPHOS recovery during the UPR(mt). Mol. Cell. 2015, 58, 123–133. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Fiorese, C.; Schulz, A.M.; Lin, Y.-F.; Rosin, N.; Pellegrino, M.W.; Haynes, C.M. The transcription factor ATF5 mediates a mammalian mitochondrial UPR. Curr. Biol. 2016, 26, 2037–2043. [Google Scholar] [CrossRef] [Green Version]
  45. Quiros, P.M.; Prado, M.A.; Zamboni, N.; D’Amico, D.; Williams, R.W.; Finley, D.; Gygi, S.P.; Auwerx, J. Multi-omics analysis identifies ATF4 as a key regulator of the mitochondrial stress response in mammals. J. Cell Biol. 2017, 216, 2027–2045. [Google Scholar] [CrossRef]
  46. Horibe, T.; Hoogenraad, N.J. The chop gene contains an element for the positive regulation of the mitochondrial unfolded protein response. PLoS ONE 2007, 2, e835. [Google Scholar] [CrossRef] [Green Version]
  47. Naresh, N.U.; Haynes, C.M. Signaling and regulation of the mitochondrial unfolded protein response. Cold Spring Harb. Perspect. Biol. 2019, 11, a033944. [Google Scholar] [CrossRef] [Green Version]
  48. Zhou, D.; Palam, L.R.; Jiang, L.; Narasimhan, J.; Staschke, K.A.; Wek, R.C. Phosphorylation of eIF2 Directs ATF5 translational control in response to diverse stress conditions. J. Biol. Chem. 2008, 283, 7064–7073. [Google Scholar] [CrossRef] [Green Version]
  49. Münch, C.; Harper, J.W. Mitochondrial unfolded protein response controls matrix pre-RNA processing and translation. Nature 2016, 534, 710–713. [Google Scholar] [CrossRef] [Green Version]
  50. Sun, X.; Angelastro, J.M.; Merino, D.; Zhou, Q.; Siegelin, M.D.; Greene, L.A. Dominant-negative ATF5 rapidly depletes survivin in tumor cells. Cell Death Dis. 2019, 10, 709. [Google Scholar] [CrossRef]
  51. Khan, N.A.; Nikkanen, J.; Yatsuga, S.; Jackson, C.; Wang, L.; Pradhan, S.; Kivelä, R.; Pessia, A.; Velagapudi, V.; Suomalainen, A. mTORC1 regulates mitochondrial integrated stress response and mitochondrial myopathy progression. Cell Metab. 2017, 26, 419–428. [Google Scholar] [CrossRef] [PubMed]
  52. Aldridge, J.E.; Horibe, T.; Hoogenraad, N.J. Discovery of genes activated by the mitochondrial unfolded protein response (mtUPR) and cognate promoter elements. PLoS ONE 2007, 2, e874. [Google Scholar] [CrossRef] [PubMed]
  53. Pakos-Zebrucka, K.; Koryga, I.; Mnich, K.; Ljujic, M.; Samali, A.; Gorman, A.M. The integrated stress response. EMBO Rep. 2016, 17, 1374–1395. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Weng, H.; Ma, Y.; Chen, L.; Cai, G.; Chen, Z.; Zhang, S.; Ye, Q. A New Vision of Mitochondrial Unfolded Protein Response to the Sirtuin Family. Curr. Neuropharmacol. 2020, 18, 613–623. [Google Scholar] [CrossRef]
  55. Yun, J.; Finkel, T. Mitohormesis. Cell Metab. 2014, 19, 757–766. [Google Scholar] [CrossRef] [Green Version]
  56. Mouchiroud, L.; Houtkooper, R.H.; Moullan, N.; Katsyuba, E.; Ryu, D.; Cantó, C.; Mottis, A.; Jo, Y.-S.; Viswantahan, M.; Schoonjans, K.; et al. The NAD(+)/sirtuin pathway modulates longevity through activation of mitochondrial UPR and FOXO signaling. Cell 2013, 154, 430–441. [Google Scholar] [CrossRef] [Green Version]
  57. Ristow, M.; Zarse, K.; Oberbach, A.; Klöting, N.; Birringer, M.; Kiehntopf, M.; Stumvoll, M.; Kahn, C.R.; Blüher, M. Antioxidants prevent health-promoting effects of physical exercise in humans. Proc. Natl. Acad. Sci. USA 2009, 106, 8665–8670. [Google Scholar] [CrossRef] [Green Version]
  58. Zeviani, M.; Carelli, V. Mitochondrial disorders. Curr. Opin. Neurol. 2007, 20, 564–571. [Google Scholar] [CrossRef]
  59. Gorman, G.S.; Schaefer, A.M.; Ng, Y.; Gomez, N.; Blakely, E.L.; Alston, C.L.; Feeney, C.; Horvath, R.; Yu-Wai-Man, P.; Chinnery, P.F.; et al. Prevalence of nuclear and mitochondrial DNA mutations related to adult mitochondrial disease. Ann. Neurol. 2015, 77, 753–759. [Google Scholar] [CrossRef] [Green Version]
  60. Perry, E.A.; Bennett, C.F.; Luo, C.; Balsa, E.; Jedrychowski, M.; O’Malley, K.E.; Latorre-Muro, P.; Ladley, R.P.; Reda, K.; Wright, P.M.; et al. Tetracyclines promote survival and fitness in mitochondrial disease models. Nat. Metab. 2021, 3, 33–42. [Google Scholar] [CrossRef]
  61. Zhang, F.; Zhang, L.; Qi, Y.; Xu, H. Mitochondrial cAMP signaling. Cell. Mol. Life Sci. 2016, 73, 4577–4590. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Scarpulla, R.C.; Vega, R.B.; Kelly, D.P. Transcriptional integration of mitochondrial biogenesis. Trends Endocrinol. Metab. 2012, 23, 459–466. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Ventura-Clapier, R.; Garnier, A.; Veksler, V. Transcriptional control of mitochondrial biogenesis: The central role of PGC-1 alpha. Cardiovasc. Res. 2008, 79, 208–217. [Google Scholar] [CrossRef] [Green Version]
  64. Suárez-Rivero, J.M.; Pastor-Maldonado, C.J.; Povea-Cabello, S.; Álvarez-Córdoba, M.; Villalón-García, I.; Talaverón-Rey, M.; Suárez-Carrillo, A.; Munuera-Cabeza, M.; Reche-López, 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. [Google Scholar] [CrossRef] [PubMed]
  65. Chukwudi, C.U. rRNA binding sites and the molecular mechanism of action of the tetracyclines. Antimicrob. Agents Chemother. 2016, 60, 4433–4441. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Meydan, S.; Marks, J.; Klepacki, D.; Sharma, V.; Baranov, P.V.; Firth, A.E.; Margus, T.; Kefi, A.; Vázquez-Laslop, N.; Mankin, A.S. Retapamulin-assisted ribosome profiling reveals the alternative bacterial proteome. Mol. Cell 2019, 74, 481–493. [Google Scholar] [CrossRef] [PubMed]
  67. Sun, T.; Zhang, Y. Pentamidine binds to tRNA through non-specific hydrophobic interactions and inhibits aminoacylation and translation. Nucleic Acids Res. 2008, 36, 1654–1664. [Google Scholar] [CrossRef] [Green Version]
  68. Wang, X.; Ryu, D.; Houtkooper, R.H.; Auwerx, J. Antibiotic use and abuse: A threat to mitochondria and chloroplasts with impact on research, health, and environment. BioEssays 2015, 37, 1045–1053. [Google Scholar] [CrossRef]
  69. Palmeira, C.; Teodoro, J.; Amorim, J.A.; Steegborn, C.; Sinclair, D.; Rolo, A.P. Mitohormesis and metabolic health: The interplay between ROS, cAMP and sirtuins. Free Radic. Biol. Med. 2019, 141, 483–491. [Google Scholar] [CrossRef]
  70. Suárez-Rivero, J.M.; Pastor-Maldonado, C.J.; Povea-Cabello, S.; Álvarez-Córdoba, M.; Villalón-García, I.; Talaverón-Rey, M.; Suárez-Carrillo, A.; Munuera-Cabeza, M.; Sánchez-Alcázar, J.A. Mitochondria and antibiotics: For good or for evil? Biomolecules 2021, 11, 1050. [Google Scholar] [CrossRef]
  71. Gao, Z.; Chen, Y.; Guan, M.-X. Mitochondrial DNA mutations associated with aminoglycoside induced ototoxicity. J. Otol. 2017, 12, 1–8. [Google Scholar] [CrossRef] [PubMed]
  72. McCoy, L.S.; Xie, Y.; Tor, Y. Antibiotics that target protein synthesis. Wiley Interdiscip. Rev. RNA 2011, 2, 209–232. [Google Scholar] [CrossRef] [PubMed]
  73. Suárez-Rivero, J.M.; Pastor-Maldonado, C.J.; Romero-González, A.; Gómez-Fernandez, D.; Povea-Cabello, S.; Álvarez-Córdoba, M.; Villalón-García, I.; Talaverón-Rey, M.; Suárez-Carrillo, A.; Munuera-Cabeza, M.; et al. Pterostilbene in combination with mitochondrial cofactors improve mitochondrial function in cellular models of mitochondrial diseases. Front. Pharmacol. 2022, 13, 862085. [Google Scholar] [CrossRef]
  74. Lange, K.W.; Li, S. Resveratrol, pterostilbene, and dementia. Biofactors 2018, 44, 83–90. [Google Scholar] [CrossRef] [PubMed]
  75. Li, Y.-R.; Li, S.; Lin, C.-C. Effect of resveratrol and pterostilbene on aging and longevity. BioFactors 2018, 44, 69–82. [Google Scholar] [CrossRef]
  76. Li, Q.; Chen, L.; Liu, X.; Li, X.; Cao, Y.; Bai, Y.; Qi, F. Pterostilbene inhibits amyloid-beta-induced neuroinflammation in a microglia cell line by inactivating the NLRP3/caspase-1 inflammasome pathway. J. Cell Biochem. 2018, 119, 7053–7062. [Google Scholar] [CrossRef] [PubMed]
  77. Natalin, H.M.; Garcia, A.F.E.; Ramalho, L.N.Z.; Restini, C.B.A. Resveratrol improves vasoprotective effects of captopril on aortic remodeling and fibrosis triggered by renovascular hypertension. Cardiovasc. Pathol. 2016, 25, 116–119. [Google Scholar] [CrossRef]
  78. Malik, S.A.; Acharya, J.D.; Mehendale, N.K.; Kamat, S.S.; Ghaskadbi, S.S. Pterostilbene reverses palmitic acid mediated insulin resistance in HepG2 cells by reducing oxidative stress and triglyceride accumulation. Free Radic. Res. 2019, 53, 815–827. [Google Scholar] [CrossRef]
  79. Zhou, J.; Ci, X.; Ma, X.; Yu, Q.; Cui, Y.; Zhen, Y.; Li, S. Pterostilbene Activates the Nrf2-Dependent Antioxidant Response to Ameliorate Arsenic-Induced Intracellular Damage and Apoptosis in Human Keratinocytes. Front. Pharmacol. 2019, 10, 497. [Google Scholar] [CrossRef]
  80. Beach, T.G. A review of biomarkers for neurodegenerative disease: Will they swing us across the valley? Neurol. Ther. 2017, 6 (Suppl. 1), 5–13. [Google Scholar] [CrossRef] [Green Version]
  81. Evidente, V.G.H.; Adler, C.H.; Sabbagh, M.N.; Connor, D.J.; Hentz, J.G.; Caviness, J.N.; Sue, L.I.; Beach, T.G. Neuropathological findings of PSP in the elderly without clinical PSP: Possible incidental PSP? Parkinsonism Relat. Disord. 2011, 17, 365–371. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Frigerio, R.; Fujishiro, H.; Ahn, T.B.; Josephs, K.A.; Maraganore, D.M.; DelleDonne, A.; Parisi, J.E.; Klos, K.J.; Boeve, B.F.; Disckson, D.W.; et al. Incidental Lewy body disease: Do some cases represent a preclinical stage of dementia with Lewy bodies? Neurobiol. Aging 2011, 32, 857–863. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Katsnelson, A.; De Strooper, B.; Zoghbi, H.Y. Neurodegeneration: From cellular concepts to clinical applications. Sci. Transl. Med. 2016, 8, 364. [Google Scholar] [CrossRef]
  84. Burchell, V.S.; Gandhi, S.; Deas, E.; Wood, N.; Abramov, A.; Plun-Favreau, H. Targeting mitochondrial dysfunction in neurodegenerative disease: Part I. Expert Opin. Ther. Targets 2010, 14, 369–385. [Google Scholar] [CrossRef]
  85. Schapira, A.H.V.; Cooper, J.M.; Dexter, D.; Clark, J.B.; Jenner, P.; Marsden, C.D. Mitochondrial complex I deficiency in Parkinson’s disease. Lancet 1989, 333, 1269. [Google Scholar] [CrossRef]
  86. Schapira, A.H.V.; Mann, V.M.; Cooper, J.M.; Dexter, D.; Daniel, S.E.; Jenner, P.; Clark, J.B.; Marsden, C.D. Anatomic and Disease Specificity of NADH CoQ1 Reductase (Complex I) Deficiency in Parkinson’s Disease. J. Neurochem. 1990, 55, 2142–2145. [Google Scholar] [CrossRef]
  87. Tohgi, H.; Yonezawa, H.; Takahashi, S.; Sato, N.; Kato, E.; Kudo, M.; Hatano, K.; Sasaki, T. Cerebral blood flow and oxygen metabolism in senile dementia of Alzheimer’s type and vascular dementia with deep white matter changes. Neuroradiology 1998, 40, 131. [Google Scholar] [CrossRef]
  88. Musatov, A.; Robinson, N.C. Susceptibility of mitochondrial electron-transport complexes to oxidative damage. Focus on cytochrome c oxidase. Free Radic. Res. 2012, 46, 1313–1326. [Google Scholar] [CrossRef]
  89. Parker, W.D.; Parks, J.; Filley, C.M.; Kleinschmidt-DeMasters, B.K. Electron transport chain defects in Alzheimer’s disease brain. Neurology 1994, 44, 1090. [Google Scholar] [CrossRef]
  90. Lezi, E.; Swerdlow, R.H. Mitochondria in neurodegeneration. Adv. Exp. Med. Biol. 2012, 942, 269–286. [Google Scholar]
  91. Bader, V.; Winklhofer, K.F. Mitochondria at the interface between neurodegeneration and neuroinflammation. Semin. Cell Dev. Biol. 2020, 99, 163–171. [Google Scholar] [CrossRef] [PubMed]
  92. Puspita, L.; Chung, S.Y.; Shim, J.-W. Oxidative stress and cellular pathologies in Parkinson’s disease. Mol. Brain 2017, 10, 53. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Cooper, J.F.; Machiela, E.; Dues, D.J.; Spielbauer, K.K.; Senchuk, M.M.; Van Raamsdonk, J.M. Activation of the mitochondrial unfolded protein response promotes longevity and dopamine neuron survival in Parkinson’s disease models. Sci. Rep. 2017, 7, 16441. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Liu, M.; Yu, S.; Wang, J.; Qiao, J.; Liu, Y.; Wang, S.; Zhao, Y. Ginseng protein protects against mitochondrial dysfunction and neurodegeneration by inducing mitochondrial unfolded protein response in Drosophila melanogaster PINK1 model of Parkinson’s disease. J. Ethnopharmacol. 2020, 247, 112213. [Google Scholar] [CrossRef]
  95. Dastidar, S.G.; Pham, M.T.; Mitchell, M.B.; Yeom, S.G.; Jordan, S.; Chang, A.; Sopher, B.L.; La Spada, A.R. 4E-BP1 Protects neurons from misfolded protein stress and Parkinson’s disease toxicity by inducing the mitochondrial unfolded protein response. J. Neurosci. 2020, 40, 8734–8745. [Google Scholar] [CrossRef]
  96. Imai, Y.; Gehrke, S.; Wang, H.-Q.; Takahashi, R.; Hasegawa, K.; Oota, E.; Lu, B. Phosphorylation of 4E-BP by LRRK2 affects the maintenance of dopaminergic neurons in Drosophila. EMBO J. 2008, 27, 2432–2443. [Google Scholar] [CrossRef] [Green Version]
  97. Strauss, K.M.; Martins, L.M.; Plun-Favreau, H.; Marx, F.P.; Kautzmann, S.; Berg, D.; Gasser, T.; Wszolek, Z.; Müller, T.; Bornemann, A.; et al. Loss of function mutations in the gene encoding Omi/HtrA2 in Parkinson’s disease. Hum. Mol. Genet. 2005, 14, 2099–3111. [Google Scholar] [CrossRef] [Green Version]
  98. 2020 Alzheimer’s disease facts and figures. Alzheimers Dement. 2020, 16, 391–460. [CrossRef]
  99. Grossberg, G.T.; Tong, G.; Burke, A.D.; Tariot, P.N. Present algorithms and future treatments for Alzheimer’s disease. J. Alzheimer’s Dis. 2019, 67, 1157–1171. [Google Scholar] [CrossRef]
  100. Swerdlow, R.H. Mitochondria and cell bioenergetics: Increasingly recognized components and a possible etiologic cause of Alzheimer’s disease. Antioxid. Redox Signal 2012, 16, 1434–1455. [Google Scholar] [CrossRef]
  101. Kish, S.J.; Mastrogiacomo, F.; Guttman, M.; Furukawa, Y.; Taanman, J.W.; Dozic, S.; Pandolfo, M.; Lamarche, J.; DiStefano, L.; Chang, L.-J. Decreased brain protein levels of cytochrome oxidase subunits in Alzheimer’s disease and in hereditary spinocerebellar ataxia disorders: A nonspecific change? J. Neurochem. 1999, 72, 700–707. [Google Scholar] [CrossRef] [PubMed]
  102. Baloyannis, S.J. Mitochondrial alterations in Alzheimer’s disease. J. Alzheimers Dis. 2006, 9, 119–126. [Google Scholar] [CrossRef] [PubMed]
  103. Manczak, M.; Calkins, M.J.; Reddy, P.H. Impaired mitochondrial dynamics and abnormal interaction of amyloid beta with mitochondrial protein Drp1 in neurons from patients with Alzheimer’s disease: Implications for neuronal damage. Hum. Mol. Genet. 2011, 20, 2495–2509. [Google Scholar] [CrossRef]
  104. Wang, X.; Su, B.O.; Lee, H.G.; Li, X.; Perry, G.; Smith, M.A.; Zhu, X. Impaired balance of mitochondrial fission and fusion in Alzheimer’s disease. J. Neurosci. 2009, 29, 9090–9103. [Google Scholar] [CrossRef] [PubMed]
  105. Beck, J.S.; Mufson, E.J.; Counts, S.E. Evidence for mitochondrial UPR gene activation in familial and sporadic Alzheimer’s Disease. Curr. Alzheimer Res. 2016, 13, 610–614. [Google Scholar] [CrossRef] [PubMed]
  106. Sorrentino, V.; Romani, M.; Mouchiroud, L.; Beck, J.S.; Zhang, H.; D’Amico, D.; Moullan, N.; Potenza, F.; Schmid, W.A.; Rietsch, S.; et al. Enhancing mitochondrial proteostasis reduces amyloid-beta proteotoxicity. Nature 2017, 552, 187–193. [Google Scholar] [CrossRef] [PubMed]
  107. Pérez, M.J.; Ivanyuk, D.; Panagiotakopoulou, V.; Di Napoli, G.; Kalb, S.; Brunetti, D.; Al-Shaana, R.; Kaeser, S.A.; Fraschka, S.A.-K.; Jucker, M.; et al. Loss of function of the mitochondrial peptidase PITRM1 induces proteotoxic stress and Alzheimer’s disease-like pathology in human cerebral organoids. Mol. Psych. 2021, 26, 5733–5750. [Google Scholar] [CrossRef]
  108. Shen, Y.; Ding, M.; Xie, Z.; Liu, X.; Yang, H.; Jin, S.; Xu, S.; Zhu, Z.; Wang, Y.; Wang, D.; et al. Activation of Mitochondrial Unfolded Protein Response in SHSY5Y Expressing APP Cells and APP/PS1 Mice. Front. Cell. Neurosci. 2019, 13, 568. [Google Scholar] [CrossRef]
  109. Weidling, I.W.; Swerdlow, R.H. Mitochondria in Alzheimer’s disease and their potential role in Alzheimer’s proteostasis. Exp. Neurol. 2020, 330, 113321. [Google Scholar] [CrossRef]
  110. Landles, C.; Bates, G.P. Huntingtin and the molecular pathogenesis of Huntington’s disease: Fourth in molecular medicine review series. EMBO Rep. 2004, 5, 958–963. [Google Scholar] [CrossRef]
  111. MacDonald, M.E.; Ambrose, C.M.; Duyao, M.P.; Myers, R.H.; Lin, C.; Srinidhi, L.; Barnes, G.; Taylor, S.A.; James, M.; Groot, N.; et al. A novel gene containing a trinucleotide repeat that is expanded and unstable on Huntington’s disease chromosomes. The Huntington’s Disease Collaborative Research Group. Cell 1993, 72, 971–983. [Google Scholar] [CrossRef]
  112. Costa, V.; Scorrano, L. Shaping the role of mitochondria in the pathogenesis of Huntington’s disease. EMBO J. 2012, 31, 1853–1864. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Bossy-Wetzel, E.; Petrilli, A.; Knott, A.B. Mutant huntingtin and mitochondrial dysfunction. Trends Neurosci. 2008, 31, 609–616. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Pickrell, A.M.; Fukui, H.; Wang, X.; Pinto, M.; Moraes, C.T. The Striatum Is Highly Susceptible to Mitochondrial Oxidative Phosphorylation Dysfunctions. J. Neurosci. 2011, 31, 9895–9904. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Fu, Z.; Liu, F.; Liu, C.; Jin, B.; Jiang, Y.; Tang, M.; Qi, X.; Guo, X. Mutant huntingtin inhibits the mitochondrial unfolded protein response by impairing ABCB10 mRNA stability. Biochim. Biophys. Acta—Mol. Basis Dis. 2019, 1865, 1428–1435. [Google Scholar] [CrossRef] [PubMed]
  116. Yano, M. ABCB10 depletion reduces unfolded protein response in mitochondria. Biochem. Biophys. Res. Commun. 2017, 486, 465–469. [Google Scholar] [CrossRef]
  117. Rosen, D.R.; Siddique, T.; Patterson, D.; Figlewicz, D.A.; Sapp, P.; Hentati, A.; Donaldson, D.; Goto, J.; O’Regan, J.P.; Deng, H.-X.; et al. Mutations in Cu/Zn superoxide dismutase gene are associated with familial amyotrophic lateral sclerosis. Nature 1993, 362, 59–62. [Google Scholar] [CrossRef]
  118. Renton, A.E.; Chio, A.; Traynor, B.J. State of play in amyotrophic lateral sclerosis genetics. Nat. Neurosci. 2014, 17, 17–23. [Google Scholar] [CrossRef]
  119. Jawaid, A.; Khan, R.; Polymenidou, M.; Schulz, P.E. Disease-modifying effects of metabolic perturbations in ALS/FTLD. Mol. Neurodegener. 2018, 13, 63. [Google Scholar] [CrossRef]
  120. Smith, E.F.; Shaw, P.J.; De Vos, K.J. The role of mitochondria in amyotrophic lateral sclerosis. Neurosci. Lett. 2019, 710, 132933. [Google Scholar] [CrossRef]
  121. Prasad, A.; Bharathi, V.; Sivalingam, V.; Girdhar, A.; Patel, B.K. Molecular Mechanisms of TDP-43 Misfolding and Pathology in Amyotrophic Lateral Sclerosis. Front. Mol. Neurosci. 2019, 12, 25. [Google Scholar] [CrossRef] [PubMed]
  122. Scotter, E.L.; Chen, H.J.; Shaw, C.E. TDP-43 Proteinopathy and ALS: Insights into disease mechanisms and therapeutic targets. Neurotherapeutics 2015, 12, 352–363. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Wang, P.; Deng, J.; Dong, J.; Liu, J.; Bigio, E.H.; Mesulam, M.; Wang, T.; Sun, L.; Wang, L.; Lee, A.Y.-L.; et al. TDP-43 induces mitochondrial damage and activates the mitochondrial unfolded protein response. PLoS Genet. 2019, 15, e1007947. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Zhou, Q.; Zhu, L.; Qiu, W.; Liu, Y.; Yang, F.; Chen, W.; Xu, R. Nicotinamide riboside enhances mitochondrial proteostasis and adult neurogenesis through activation of mitochondrial unfolded protein response signaling in the brain of ALS SOD1(G93A) mice. Int. J. Biol. Sci. 2020, 16, 284–297. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Riar, A.K.; Burstein, S.R.; Palomo, G.M.; Arreguin, A.; Manfredi, G.; Germain, D. Sex specific activation of the ERalpha axis of the mitochondrial UPR (UPRmt) in the G93A-SOD1 mouse model of familial ALS. Hum. Mol. Genet. 2017, 26, 1318–1327. [Google Scholar] [CrossRef] [PubMed]
  126. Gomez, M.; Germain, D. Cross talk between SOD1 and the mitochondrial UPR in cancer and neurodegeneration. Mol. Cell. Neurosci. 2019, 98, 12–18. [Google Scholar] [CrossRef]
  127. Sciarretta, S.; Maejima, Y.; Zablocki, D.; Sadoshima, J. The Role of Autophagy in the Heart. Annu. Rev. Physiol. 2018, 80, 1–26. [Google Scholar] [CrossRef]
  128. Nandi, S.S.; Katsurada, K.; Sharma, N.M.; Mahata, S.K.; Patel, K.P. Abstract 15288: Mitochondrial Injury in Cardiomyopathy of Neurogenic Hypertension: Role of MiR-18a-5p/HIF-1a Axis. Circulation 2020, 142 (Suppl. 3), A15288. [Google Scholar] [CrossRef]
  129. Wrobel, L.; Topf, U.; Bragoszewski, P.; Wiese, S.; Sztolsztener, M.E.; Oeljeklaus, S.; Varabyova, A.; Lirski, M.; Chroscicki, P.; Mroczek, S.; et al. Mistargeted mitochondrial proteins activate a proteostatic response in the cytosol. Nature 2015, 524, 485–488. [Google Scholar] [CrossRef]
  130. García-Díaz, L.; Coserria, F.; Antiñolo, G. Hypertrophic Cardiomyopathy due to Mitochondrial Disease: Prenatal Diagnosis, Management, and Outcome. Case Rep. Obstet. Gynecol. 2013, 2013, 472356. [Google Scholar] [CrossRef]
  131. Smyrnias, I.; Gray, S.P.; Okonko, D.O.; Sawyer, G.; Zoccarato, A.; Catibog, N.; López, B.; Gonzalez, A.; Ravassa, S.; Díez, J.; et al. Cardioprotective Effect of the Mitochondrial Unfolded Protein Response During Chronic Pressure Overload. J. Am. Coll. Cardiol. 2019, 73, 1795–1806. [Google Scholar] [CrossRef] [PubMed]
  132. 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. [Google Scholar] [CrossRef] [PubMed]
  133. Houtkooper, R.; Mouchiroud, L.; Ryu, D.; Moullan, N.; Katsyuba, E.; Knott, G.W.; Williams, R.W.; Auwerx, J. Mitonuclear protein imbalance as a conserved longevity mechanism. Nature 2013, 497, 451–457. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Durieux, J.; Wolff, S.; Dillin, A. The Cell-Non-Autonomous Nature of Electron Transport Chain-Mediated Longevity. Cell 2011, 144, 79–91. [Google Scholar] [CrossRef] [Green Version]
  135. Yokoyama, K.; Fukumoto, K.; Murakami, T.; Harada, S.I.; Hosono, R.; Wadhwa, R.; Mitsui, Y.; Ohkuma, S. Extended longevity of Caenorhabditis elegans by knocking in extra copies of hsp70F, a homolog of mot-2 (mortalin)/mthsp70/Grp75. FEBS Lett. 2002, 516, 53–57. [Google Scholar] [CrossRef] [Green Version]
  136. Flurkey, K.; Papaconstantinou, J.; Miller, R.A.; Harrison, D.E. Lifespan extension and delayed immune and collagen aging in mutant mice with defects in growth hormone production. Proc. Natl. Acad. Sci. USA 2001, 98, 6736–6741. [Google Scholar] [CrossRef] [Green Version]
  137. Ozkurede, U.; Miller, R.A. Improved mitochondrial stress response in long-lived Snell dwarf mice. Aging Cell 2019, 18, e13030. [Google Scholar] [CrossRef] [Green Version]
  138. Wei, Y.; Zhang, Y.J.; Cai, Y.; Xu, M.H. The role of mitochondria in mTOR-regulated longevity. Biol. Rev. Camb. Philos. Soc. 2015, 90, 167–181. [Google Scholar] [CrossRef]
  139. Jensen, M.B.; Jasper, H. Mitochondrial Proteostasis in the Control of Aging and Longevity. Cell Metab. 2014, 20, 214–225. [Google Scholar] [CrossRef] [Green Version]
  140. Haynes, C.M.; Yang, Y.; Blais, S.P.; Neubert, T.A.; Ron, D. The Matrix Peptide Exporter HAF-1 Signals a Mitochondrial UPR by Activating the Transcription Factor ZC376.7 in C. elegans. Mol. Cell 2010, 37, 529–540. [Google Scholar] [CrossRef] [Green Version]
  141. Ristow, M.; Schmeisser, S. Extending life span by increasing oxidative stress. Free Radic. Biol. Med. 2011, 51, 327–336. [Google Scholar]
  142. Schmeisser, K.; Mansfeld, J.; Kuhlow, D.; Weimer, S.; Priebe, S.; Heiland, I.; Birringer, M.; Groth, M.; Segref, A.; Kanfi, Y.; et al. Role of sirtuins in lifespan regulation is linked to methylation of nicotinamide. Nat. Chem. Biol. 2013, 9, 693–700. [Google Scholar] [CrossRef] [Green Version]
  143. Hsu, A.-L.; Murphy, C.T.; Kenyon, C. Regulation of aging and age-related disease by DAF-16 and heat-shock factor. Science 2003, 300, 1142–1145. [Google Scholar] [CrossRef] [Green Version]
  144. Yang, W.; Hekimi, S. Two modes of mitochondrial dysfunction lead independently to lifespan extension in Caenorhabditis elegans. Aging Cell 2010, 9, 433–447. [Google Scholar] [CrossRef] [PubMed]
  145. Zhang, H.; Ryu, D.; Wu, Y.; Gariani, K.; Wang, X.; Luan, P.; D’Amico, D.; Ropelle, E.R.; Lutolf, M.P.; Aebersold, R.; et al. NAD+ repletion improves mitochondrial and stem cell function and enhances life span in mice. Science 2016, 352, 1436–1443. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Behrends, C.; Langer, C.A.; Boteva, R.; Böttcher, U.M.; Stemp, M.J.; Schaffar, G.; Rao, B.V.; Giese, A.; Kretzschmar, H.; Siegers, K.; et al. Chaperonin TRiC Promotes the Assembly of polyQ Expansion Proteins into Nontoxic Oligomers. Mol. Cell 2006, 23, 887–897. [Google Scholar] [CrossRef]
  147. Lin, Y.-F.; Schulz, A.M.; Pellegrino, M.W.; Lu, Y.; Shaham, S.; Haynes, C.M. Maintenance and propagation of a deleterious mitochondrial genome by the mitochondrial unfolded protein response. Nature 2016, 533, 416–419. [Google Scholar] [CrossRef]
  148. Martinez, B.A.; Petersen, D.A.; Gaeta, A.L.; Stanley, S.P.; Caldwell, G.A.; Caldwell, K.A. Dysregulation of the mitochondrial unfolded protein response induces non-apoptotic dopaminergic neuro-degeneration in C. elegans models of Parkinson’s disease. J. Neurosci. 2017, 37, 11085–11100. [Google Scholar] [CrossRef] [Green Version]
  149. Shao, L.-W.; Peng, Q.; Dong, M.; Gao, K.; Li, Y.; Li, Y.; Li, C.-Y.; Liu, Y. Histone deacetylase HDA-1 modulates mitochondrial stress response and longevity. Nat. Commun. 2020, 1, 4639. [Google Scholar] [CrossRef]
  150. Angeli, S.; Foulger, A.; Chamoli, M.; Peiris, T.H.; Gerencser, A.; Shahmirzadi, A.A.; Andersen, J.; Lithgow, G. The mitochondrial permeability transition pore activates the mitochondrial unfolded protein response and promotes aging. eLife 2021, 10, 63453. [Google Scholar] [CrossRef]
  151. O’Malley, J.; Kumar, R.; Inigo, J.; Yadava, N.; Chandra, D. Mitochondrial stress response and cancer. Trends Cancer 2020, 6, 688–701. [Google Scholar] [CrossRef] [PubMed]
  152. Cappello, F.; Conway de Macario, E.; Marasà, L.; Zummo, G.; Macario, A.J. Hsp60 expression, new locations, functions and perspectives for cancer diagnosis and therapy. Cancer Biol. Ther. 2008, 7, 801–809. [Google Scholar] [CrossRef] [PubMed]
  153. Macario, A.J.L.; De Macario, E.C. Chaperonopathies by defect, excess, or mistake. Ann. N. Y. Acad. Sci. 2007, 1113, 178–191. [Google Scholar] [CrossRef] [PubMed]
  154. Tsai, Y.P.; Yang, M.H.; Huang, C.H.; Chang, S.Y.; Chen, P.M.; Liu, C.J.; Teng, S.-C.; Wu, K.J. Interaction between HSP60 and beta-catenin promotes metastasis. Carcinogenesis 2009, 30, 1049–1057. [Google Scholar] [CrossRef]
  155. Piselli, P.; Vendetti, S.; Vismara, D.; Cicconi, R.; Poccia, F.; Colizzi, V.; Delpino, A. Different expression of CD44, ICAM-1, and HSP60 on primary tumor and metastases of a human pancreatic carcinoma growing in scid mice. Anticancer Res 2000, 20, 825–831. [Google Scholar]
  156. O’Barazi, H.; Zhou, L.; Templeton, N.S.; Krutzsch, H.C.; Roberts, D.D. Identification of heat shock protein 60 as a molecular mediator of alpha 3 beta 1 integrin activation. Cancer Res. 2002, 62, 1541–1548. [Google Scholar]
  157. Wang, G.; Fan, Y.; Cao, P.; Tan, K. Insight into the mitochondrial unfolded protein response and cancer: Opportunities and challenges. Cell Biosci. 2022, 12, 1–23. [Google Scholar] [CrossRef]
  158. Desgrosellier, J.S.; Cheresh, D.A. Integrins in cancer: Biological implications and therapeutic opportunities. Nat. Cancer 2010, 10, 9–22. [Google Scholar] [CrossRef] [Green Version]
  159. Du, F.-Y.; Zhou, Q.-F.; Sun, W.-J.; Chen, G.-L. Targeting cancer stem cells in drug discovery: Current state and future perspectives. World J. Stem Cells 2019, 11, 398–420. [Google Scholar] [CrossRef]
  160. Gu, L.-F.; Chen, J.-Q.; Lin, Q.-Y.; Yang, Y.-Z. Roles of mitochondrial unfolded protein response in mammalian stem cells. World J. Stem Cells 2021, 13, 737–752. [Google Scholar] [CrossRef]
Figure 1. Mitochondrial stress response. The most common mitochondrial stressors are ROS, unfolded proteins, and protein aggregates. To deal with such insults and their consequences, mitochondria possess several mechanisms: First, the UPRmt will promote the expression of antioxidant proteins, chaperones, or proteases. If damage persists and/or increases, the cell will try to remove highly damaged mitochondria with mitophagic processes. Mitophagy degrade dysfunctional mitochondria by a combination of mitochondrial fission and autophagy; then mitochondrial biogenesis and fusion are enhanced to restore energetic balance. All of these quality-control mechanisms are non-exclusives, and cells can regulate them according to the situation. Finally, if the cell is unable to overcome the energetic crisis, it will undergo apoptosis through mitochondrial release of pro-apoptotic factors. Figure was created with BioRender (BioRender.com).
Figure 1. Mitochondrial stress response. The most common mitochondrial stressors are ROS, unfolded proteins, and protein aggregates. To deal with such insults and their consequences, mitochondria possess several mechanisms: First, the UPRmt will promote the expression of antioxidant proteins, chaperones, or proteases. If damage persists and/or increases, the cell will try to remove highly damaged mitochondria with mitophagic processes. Mitophagy degrade dysfunctional mitochondria by a combination of mitochondrial fission and autophagy; then mitochondrial biogenesis and fusion are enhanced to restore energetic balance. All of these quality-control mechanisms are non-exclusives, and cells can regulate them according to the situation. Finally, if the cell is unable to overcome the energetic crisis, it will undergo apoptosis through mitochondrial release of pro-apoptotic factors. Figure was created with BioRender (BioRender.com).
Biomedicines 10 01611 g001
Figure 2. The UPRmt is a conserved adaptive mechanism for improving cell survival under mitochondrial stress. The UPRmt may act as a first line of defense against mitochondrial stress and involves communication between the stressed mitochondria and the nucleus. Activation of the UPRmt aims to restore protein homeostasis and function within the mitochondria, therefore preserving cellular functions. UPRmt is mainly regulated for a triad of proteins—ATF4, ATF5, and CHOP; however, although their interactions are not well stablished yet. Different reports have described the protective effects of UPRmt activation in various disease conditions, such as mitochondrial diseases, neurodegeneration, or heart diseases. However, after trespassing a certain stress threshold and to prevent the toxic effects of dysfunctional cell processes, affected mitochondria are eliminated via mitophagy, which ultimately will lead to apoptosis if the damage is irreversible. Figure was created with BioRender (BioRender.com).
Figure 2. The UPRmt is a conserved adaptive mechanism for improving cell survival under mitochondrial stress. The UPRmt may act as a first line of defense against mitochondrial stress and involves communication between the stressed mitochondria and the nucleus. Activation of the UPRmt aims to restore protein homeostasis and function within the mitochondria, therefore preserving cellular functions. UPRmt is mainly regulated for a triad of proteins—ATF4, ATF5, and CHOP; however, although their interactions are not well stablished yet. Different reports have described the protective effects of UPRmt activation in various disease conditions, such as mitochondrial diseases, neurodegeneration, or heart diseases. However, after trespassing a certain stress threshold and to prevent the toxic effects of dysfunctional cell processes, affected mitochondria are eliminated via mitophagy, which ultimately will lead to apoptosis if the damage is irreversible. Figure was created with BioRender (BioRender.com).
Biomedicines 10 01611 g002
Table 1. Summary of UPRmt-related conditions and studies.
Table 1. Summary of UPRmt-related conditions and studies.
ConditionRelated Studies
Mitochondrial diseases[60,64,73]
Parkinson’s disease[93,94,95,96,148]
Alzheimer’s disease[105,106,107,108,109]
Huntington’s disease[115]
Amyotrophic lateral sclerosis[121,122,123,124]
Heart diseases[132]
Aging[139,140,141,142]
Cancer[157,160]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Suárez-Rivero, J.M.; Pastor-Maldonado, C.J.; Povea-Cabello, S.; Álvarez-Córdoba, M.; Villalón-García, I.; Talaverón-Rey, M.; Suárez-Carrillo, A.; Munuera-Cabeza, M.; Reche-López, D.; Cilleros-Holgado, P.; et al. Activation of the Mitochondrial Unfolded Protein Response: A New Therapeutic Target? Biomedicines 2022, 10, 1611. https://doi.org/10.3390/biomedicines10071611

AMA Style

Suárez-Rivero JM, Pastor-Maldonado CJ, Povea-Cabello S, Álvarez-Córdoba M, Villalón-García I, Talaverón-Rey M, Suárez-Carrillo A, Munuera-Cabeza M, Reche-López D, Cilleros-Holgado P, et al. Activation of the Mitochondrial Unfolded Protein Response: A New Therapeutic Target? Biomedicines. 2022; 10(7):1611. https://doi.org/10.3390/biomedicines10071611

Chicago/Turabian Style

Suárez-Rivero, Juan M., Carmen J. Pastor-Maldonado, Suleva Povea-Cabello, Mónica Álvarez-Córdoba, Irene Villalón-García, Marta Talaverón-Rey, Alejandra Suárez-Carrillo, Manuel Munuera-Cabeza, Diana Reche-López, Paula Cilleros-Holgado, and et al. 2022. "Activation of the Mitochondrial Unfolded Protein Response: A New Therapeutic Target?" Biomedicines 10, no. 7: 1611. https://doi.org/10.3390/biomedicines10071611

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